GENE DUPLICATION AND FUSION: STRATEGY FOR ACTIVE SITE CONTROL AND STARTING POINT FOR NEW CATALYSTS by Yu Wang A dissertation submitted to the Graduate Faculty of Auburn University in partial fulfillment of the requirements for the Degree of Doctor of Philosophy Auburn, Alabama Dec 8, 2012 Keywords: catalase-peroxidase, KatG, folding platform, heme protein, C-terminal domain, gene duplication and fusion Copyright 2012 by Yu Wang Approved by Douglas Goodwin, Chair, Professor of Biochemistry Holly Ellis, Professor of Biochemistry Evert Duin, Professor of Biochemistry Christian Goldsmith, Professor of Chemistry iv Abstract Catalase?peroxidases (KatGs) have two peroxidase-like domains. The N-terminal domain contains the heme-dependent, bifunctional active site. Though the C-terminal domain lacks the ability to bind heme or directly catalyze any reaction, it has been proposed to serve as a platform to direct the folding of the N-terminal domain. Toward such a purpose, its I?-helix is highly conserved and appears at the interface between the two domains. Single and multiple substitution variants targeting highly conserved residues of the I?-helix were generated for intact KatG as well as the stand-alone C-terminal domain (KatG C ). Single variants of intact KatG produced only subtle variations in spectroscopic and catalytic properties of the enzyme. However, the double and quadruple variants showed spectroscopic and catalytic properties similar to that observed for the N-terminal domain on its own (KatG N ). The analogous variants of KatG C showed a much more profound loss of function as evaluated by their ability to return KatG N to its active conformation. These results suggest that the I?-helix is central to direct structural adjustments in the adjacent N-terminal domain. In particular, substitution of E695, a strictly conserved residue of the I?helix was especially destructive to KatG active site integrity. Available structures of KatG indicate that E695 is a central component of several hydrogen-bonded networks that also include strictly conserved R126 and W159 from the adjacent N-terminal domain. Both W159 (N- terminus of the D-helix) and R126 (BC helical connecting loop) are part of core structures of the peroxidase-like N-terminal domain with connections to the active site. This points to a potential connection between the I?-helix and its influence on active site conformation and function. To v evaluate this hypothesis, we replaced R126, W159 and E695 with alanine, singly and in combination, not only for intact KatG, but also for the stand-alone KatG N or KatG C , as appropriate. Single variants of intact KatG showed a substantial loss of stability, particularly at the active site. The analogous variants of KatG C and KatG N showed a profound loss of function as evaluated by the return KatG N (or its variants) to a functional conformation, suggesting C- terminal domain through these interactions directs active site structural adjustments that are essential for catalytic function in the active site 25 ? away. The C-terminal domain not only directs conformational adjustments in its N-terminal domain partner, but due to its origin from gene duplication, also still retains the helical architecture of a typical peroxidase active site. This site lacks the residues necessary for catalytic activity or even heme binding. As such, the C- terminal domain appears to provide an ideal ?blank slate? for engineering new heme-dependent catalysts. Spectroscopic measurements showed that a M616G/R617H variant of KatG C was sufficient to restore heme binding, producing a hexacoordinate low-spin ferric state. Additional modifications produced spectra very similar to those observed for KatG N . vi Acknowledgments The research conducted in this dissertation would not have been possible without the help and support of many people whom I wish to thank. Listed here are only a few of those. First, I would like to thank my advisor Dr. Douglas Goodwin. He guided me through my whole research, trained me to be an independent researcher. He had a significant influence on my future career and personal life. To me he is more than an advisor; without his encouragement, I could not have reached today?s achievement. I wish I will continue to make him proud. Second, I would like to thank Dr. Holly Ellis for taking the time to review every manuscript I have written. Dr. Evert Duin has been of great assistance with EPR. Third, I would like thank all my co- workers at Auburn University, especially Robert Moore and Carma Cook. They provided great help and assistance when I first started my research. I also feel lucky to have worked with Shalley Kudalkar, Ellizebeth Ngwane and Haijun Duan. Especially I thank my family for their continuous understanding, help, support and encouragement for these five years. Finally, I would like to thank NSF, the Department of Chemistry and Biochemistry, Auburn University?s graduate school, and COSAM for their funding support. vii Table of Contents Abstract ......................................................................................................................................... iv Acknowledgments ......................................................................................................................... vi CHAPTER ONE ............................................................................................................................ 1 1.1 Hydrogen peroxide and its scavengers in cels ...................................................................................... 2 1.2 Typical (i.e., monofunctional) catalases ................................................................................................ 3 1.3 Peroxidases ............................................................................................................................................. 6 1.4 Catalase-peroxidases (KatGs) .............................................................................................................. 40 CHAPTER TWO ......................................................................................................................... 54 2.1 Introduction .......................................................................................................................................... 54 2.2 Materials and methods ......................................................................................................................... 57 2.3 Results .................................................................................................................................................. 66 2.4 Discusions ........................................................................................................................................... 85 CHAPTER THREE ..................................................................................................................... 94 3.1 Introduction .......................................................................................................................................... 94 3.2 Materials and methods ......................................................................................................................... 96 3.3 Results ................................................................................................................................................ 102 3.5 Discusion .......................................................................................................................................... 121 CHAPTER FOUR ...................................................................................................................... 134 4.1 Introduction ........................................................................................................................................ 134 viii 4.2 Materials and methods ....................................................................................................................... 138 4.3 Results ................................................................................................................................................ 141 4.4 Discusion and conclusion ................................................................................................................. 152 CHAPTER FIVE ....................................................................................................................... 157 5.1 Introduction ........................................................................................................................................ 157 5.2 Methods .............................................................................................................................................. 159 5.3 Results ................................................................................................................................................ 160 5.4 Conclusion ......................................................................................................................................... 165 CHAPTER SIX .......................................................................................................................... 167 REFERENCES .......................................................................................................................... 174 ix List of Figures Figure 1.1: Overall structure (A), one subunit (B) and the active site (C) of the E.coli catalase (HP II) ............................................................................................................................5? Figure 1.2: Catalytic mechanism of typical catalases ......................................................................7? Figure 1.3 Overview of the classification of heme peroxidases (super) families ..........................10? Figure 1.4: Catalytic cycle of bacterial diheme cytochrome c peroxidase ....................................12? Figure 1.5: Structure of di-heme-cytochrome c peroxidase and conformational change during catalysis ........................................................................................................................14? Figure 1.6 Structures of DyP and DyP-type peroxidase ................................................................16? Figure 1.7: Schematic diagram of the proposed formation of compound I by DyP ......................18? Figure 1.8: A modified mechanism of compound I formation by DyP .........................................18? Figure 1.9: Structure of CPO .........................................................................................................20? Figure 1.10: Catalytic cycle of haloperoxidases ............................................................................22? Figure 1.11: Proposed mechanism of compound I formation by chloroperoxidase ......................23? Figure 1.12: Structure of human MPO ..........................................................................................26? Figure 1.13: Catalytic cycles of MPO ............................................................................................27? Figure 1.14: Structure of HRP .......................................................................................................31? Figure 1.15: Peroxidase cycle of HRP ...........................................................................................32? Figure 1.16: Structure of Class II peroxidases. ..............................................................................36? Figure 1.17: Catalytic cycle of manganese peroxidase (MnP) ......................................................37? x Figure 1.18: Structure of Class I peroxidase. .................................................................................41? Figure 1.19: Comparisons of the overall structures and the active sites of monofunctional peroxidases, KatG and monofunctional catalases. .......................................................44? Figure 1.20: Two novel disulfide bridges between the N-terminal domains of two subunits in KatG .............................................................................................................................46? Figure 1.21: Structure of catalase-peroxidase. ...............................................................................48? Figure 1.22: The covalent adduct in KatG. ....................................................................................49? Figure 1.23: Classical model of catalytic cycle of KatG (A), modified model of catalase mechanism of KatG (B). ..............................................................................................50? Figure 2.1: (A) Far-UV circular dichroism of wtKatG, and I?-helix variants. (B) Far-UV circular dichroism of C-terminal domain I?helix variants and KatG C ......................................67? Figure 2.2: Absorption spectra of ferric wtKatG, its I'-helix substitution variants, and KatG N ....68? Figure 2.3: MCD spectra of ferrous wtKatG, KatG N and I?-helix substitution variants. ...............70? Figure 2.4: EPR spectra recorded for wtKatG, its I'-helix substitution variants, and KatG N . .......71? Figure 2.5: GuHCl-mediated unfolding of wtKatG, its I'-helix variants, and KatG N as monitored by far-UV CD spectroscopy .........................................................................................78? Figure 2.6: Thermal stability of heme cavity of wtKatG and its I'-helix variants .........................79 Figure 2.7: UV?visible absorption spectra of KatG N following incubation with KatG C or its I'- helix substitution variants ............................................................................................81? Figure 2.8: EPR spectra of KatG N following incubation with KatG C or its I'-helix substitution variants .........................................................................................................................83? Figure 2.9: Time course for recovery of KatG N catalase activity upon addition of equimolar KatG C or one of its I'-helix substitution variants .........................................................86? xi Figure 2.10: Interfaces between the KatG N-terminal domain and I'-helix from the C-terminal domain ..........................................................................................................................92? Figure 3.1: Far-UV circular dichroism of wtKatG, and its E695, R126 and W159 substitution variants .......................................................................................................................103? Figure 3.2: UV-visible absorption spectra of ferric wtKatG and its E695, R126 and W159 substitution variants ...................................................................................................104? Figure 3.3: MCD spectra of ferrous wtKatG and its E695A, R126A and W159A .....................106 Figure 3.4: EPR spectra of wtKatG and its E695, R126 and W159 substitution variants ...........107? Figure 3.5: GuHCl-mediated unfolding of wtKatG and E695, R126 and W159 substitution variants as monitored by far-UV circular dichroism (CD) spectroscopy. .................113? Figure 3.6: Thermal stability of the heme cavity of wtKatG, E695, R126 and W159 substitution variants as monitored by UV-visible absorption due to the heme retention ..............115? Figure 3.7: UV-visible absorption spectra of KatG N (or its R126A, W159A substitution variants) following incubation with KatG C (or its E695A, E695Q substitution variants). .......118? Figure 3.8: EPR spectra of KatG N (or its R126A, W159A substitution variants) following incubation with KatG C (or its E695A, E695Q substitution variants) ........................119? Figure 3.9: Time course for recovery of KatG N (or its R126A, W159A) catalase activity upon addition of equimolar KatG C (or its E695A, E695Q) ................................................124 Figure 3.10: Interactions between the N-terminal domain BC loop and the B?C? loop from the intrasubunit C-terminal domain .................................................................................127? Figure 3.11: Interactions between R126, W159 and E695 and water-mediated hydrogen-bonded networks along BC loop; BC loop in manganese peroxidase and Ca 2+ ....................130 Figure 3.12: Overlaid structures of KatGs and the water molecules ...........................................131? xii Figure 4.1: Overlaid structures of KatG C (green) and CCP (cyan) with a bound heme (red). ....136? Figure 4.2: Comparison of the ?active site? environments of APX and KatG C ..........................137? Figure 4.3: Far-UV circular dichroism of KatG C and KatG C variants. ........................................144? Figure 4.4: UV-vis spectra of ferric KatG C , KatG N and KatG C variants (HB 1) .........................146? Figure 4.5: UV-vis spectra of ferric KatG C , KatG N and KatG C variants .....................................147? Figure 4.6: Magnetic circular dichroism (MCD) spectra for KatG C and KatG N variants ...........150? Figure 4.7: EPR spectra for ferric state of KatG C and KatG N variants. .......................................151? Figure 5.1: Distal active site M-Y-W covalent adduct in KatG from M. tuberculosis ................158? Figure 5.2: Evaluation of M-Y-W covalent adduct formation by SDS-PAGE ...........................161? Figure 5.3: The influence of peripheral structures on the formation of M-Y-W adduct. ............163? xiii List of Tables Table 2.1: Primer sets for site-directed mutagenesis of I?-helix. ...................................................59? Table 2.2: Ratios of EPR signals observed for wtKatG, its I'-helix variants, and KatG N . ............73? Table 2.3: Apparent catalase kinetic parameters for wtKatG and its I'-helix variants .................74? Table 2.4: Apparent peroxidase kinetic parameters for wtKatG and its I?-helix variants ..............75? Table 2.5: Parameters GuHCl-mediated unfolding of wtKatG and I?-helix variants ....................77? Table 2.6: Ratios of EPR signals observed following KatG N incubation with KatG C or its I'-helix substitution variants .......................................................................................................84? Table 2.7: KatG N reactivation k obs and apparent catalase parameters for reactivated enzyme. .....87? Table 2.8: Apparent Peroxidase parameters for reactivated KatG N by KatG C or its I?-helix variants ..........................................................................................................................88? Table 3.1: Primer sets for site-directed mutagenesis .....................................................................98? Table 3.2: Ratios of EPR signals observed in wtKatG and its E695, R126 and W159 substitution variants. .......................................................................................................................109? Table 3.3: Apparent catalase kinetic parameters for wtKatG and its E695, R126 and W159 substitution variants ....................................................................................................110? Table 3.4: Apparent peroxidase kinetic parameters for wtKatG and its E695, R126 and W159 substitution variants ....................................................................................................111? Table 3.5: Parameters of GuHCl-mediated unfolding wtKatG and E695, R126 and W159 substitution variants. ....................................................................................................114? xiv Table 3.6: Ratios of EPR signals observed following KatG N (or its R126A, W159A) following incubation with KatG C (or its E695A, E695Q) ...........................................................120? Table 3.7: Apparent catalase kinetic parameters for KatG N (or its R126A, W159A) following incubation with KatG C (or its E695A, E695Q) ..........................................................122? Table 3.8: Apparent peroxidase kinetic parameters for KatG N (or its R126A, W159A) following incubation with KatG C (or its E695A, E695Q substitution variants) .........................123? Table 4.1: Primer sets for site-directed mutagenesis of KatG C ...................................................140? Table 4.2: Variants name corresponding to the detailed mutagenesis strategy ...........................143? Table 4.3: Absorption characteristics of KatG N and KatG C variants ...........................................148? 1 CHAPTER ONE LITERATURE REVIEW The central proposes of the research described in this dissertation are to evaluate the role of the C-terminal domain in catalase-peroxidase (KatG) enzymes, and to borrow this structure as a template for engineering new heme-dependent catalysts. The two-domain structure of KatG is a novel feature in comparison to other members of its superfamily as is its bifunctional catalytic capability. The connection between KatG?s structure and its function remains poorly defined, but its contrast against other structurally similar but functionally distinct relatives, provides an excellent opportunity to evaluate the mechanism by which structures peripheral to and distant from an active site influence structure and function. More than this, it affords an ideal system to apply the information and strategy toward engineering new heme enzymes. This chapter will first establish the importance of catalases and peroxidases in nature, and then discuss the catalytic mechanisms employed by these H 2 O 2 -degrading enzymes by comparing their structures and functions in order to understand the importance of the two-domain structure, bifunctionality and molecular phylogeny of the catalase-peroxidases. Furthermore, discussion focused on the structure and function relationships of various heme-dependent peroxidases will shed light on the structural basis and strategy for engineering the KatG C-terminal domain to a new heme- dependent catalyst, building on its unique origin and features. 2 1.1 Hydrogen peroxide and its scavengers in cells: Hydrogen peroxide is one of the most common reactive oxygen species that aerobic organisms encounter. For example, in well-fed aerobic E.coli cells, H 2 O 2 is formed at a constant rate of 10-15 ?M/s [1]. This endogenous H 2 O 2 is primarily formed from reduction of O 2 by flavoproteins [2]. Exogenous H 2 O 2 derived from environmental sources is also an issue because H 2 O 2 can penetrate cell membranes. Exogenous sources of H 2 O 2 include the NADPH oxidase- based response of plants and animials to infection, H 2 O 2 -excreting microbes, extracellular oxidation at anoxic/oxic interfaces, and photochemically-driven oxidations [2]. Due to its ability to initiate damage to biomolecules, excessive H 2 O 2 is toxic for living cells. In particular, it can destroy iron-sulfur proteins involved in energy production and biosynthesis [3], and deactivate other enzymes that rely on ferrous iron as a cofactor [4]. The reduction of H 2 O 2 by ferrous iron produces the highly reactive hydroxyl radical (?OH) by the well-known Fenton reaction. The unregulated generation of such a reactive oxidant is known to result in damage to all classes of biological molecules. Therefore, rapid and efficient removal of excess H 2 O 2 is essential for all organisms that are to survive in an aerobic enviornment. In most organisms, H 2 O 2 is scavenged by peroxidases and/or catalases. The latter show a striking ability to disproportionate H 2 O 2 to water concomitant with the generation of dioxygen, according to reaction (1). Enzymes with catalase activity are predominantly found in three gene families: typical (i.e., monofunctional) catalases, catalase-peroxidases (KatGs), and manganese (non-heme) catalases [5]. Among these, the large majority are heme-containing enzymes, and these are widely distributed among prokaryotes and eukaryotes. Manganese catalase is a minor gene group only present in bacteria; they are less proficient catalysts than typical catalases and catalase-peroxidases [5]. Though important, the manganese catalases are not particularly relevant to this discussion and will not be 3 covered further. 1.2 Typical (i.e., monofunctional) catalases: Phylogenetic analyses indicate that there are three main evolutionary clades, which were segregated in an early stage of evolution by way of several gene-duplication events [5, 6]. Clade 1 catalases are small-subunit proteins with heme b as the prosthetic group. They are distributed among bacteria, algae and plants. Clade 2 contains the large-subunit catalases, and these primarily use heme d as prosthetic group and contain an additional ?flavodoxin-like? domain. These are widely distributed among bacteria and fungi. Clade 3 is the most widely distributed; being abundant in archaebacteria, fungi, protists, plants and animals. These are typically small- subunit catalases that contain heme b and NADPH as cofactors. 1.2.1 Structure and active site of monofunctional catalases: The monofunctional catalase from E. coli, hydroperoxidase II (HPII), is a tetrameric large- subunit catalase (Fig. 1.1A) found in clade 2 of the typical catalase family. Available crystal structures indicate that HP II is composed of five different tertiary structural regions or domains, including 1) the amino terminal arm, 2) an antiparallel eight-stranded ?-barrel, 3) an extended wrapping loop, 4) a helical domain, and 5) a C-terminal, flavodoxin-like domain [7]. The core of the protein is similar in structure and sequence to those of small subunit catalases; however the 75-80 residues of the amino-terminal arm and the flavodoxin-like domain are unique to HPII and other clade 2 enzymes (Fig. 1.1B). It has been proposed that both structural elements are dedicated to facilitating the folding and oligomerization process [7, 8]. 4 Although significant differences exist in the overall folding between small- and large- subunit catalases, the active site residues are strictly conserved among all three clades (Fig. 1.1C) [5]. These include Asn 201 and His 128 (HPII numbering) in the distal site, and Tyr 415, which ligates the iron on the heme?s proximal side. Asn 201 is proposed to stabilize the peroxide [9]. The imidazole ring of His 128 is situated 3.5 ? above and parallel to the plane of the heme, allowing for an H-bond with a conserved Ser 167 [7, 10]. The coplanar position of the imidazole ring contrasts with the perpendicular orientation of this group in peroxidases. This is proposed to allow the second oxygen of the peroxide to form an H-bond with the imidazole during catalysis, facilitating O 2 production [5]. In all monofunctional catalases, a tyrosine has been identified as the proximal heme iron ligand in contrast to a histidine in plant peroxidases. The strong anionic character of the ligand is expected to favor the oxidized heme species [11]. Interestingly, a unique feature for E. coli HPII is a covalent bond between the proximal Tyr ligand and an adjacent imidazole ring from His 392 [12]. This has been proposed to enhance resistance to peroxide-dependent deactivation by preventing formation of compound II, an intermediate that is not part of the enzyme?s catalytic cycle [5]. Another feature typical of monofunctional catalases is a long, narrow substrate access channel leading to the deeply buried active site heme group [7]. In contrast, typical peroxidases usually provide more open access to the heme cofactor. In E.coli HPII, two major channels can be defined: one has small diameter and is full of hydrophobic residues, leading H 2 O 2 to the heme distal pocket; the other is a large funnel-shaped opening positioned near residues 590 and 595 and leads to the vicinity of Asn 201 [7]. The shape and size of these channels are essential for selectively directing substrates, particularly H 2 O 2 , to the active site [7, 13]. 5 Figure 1.1: Overall structure (A), one subunit (B) and the active site (C) of the E.coli catalase (HP II) (PDB: 1GGE). 6 1.2.2 The catalytic mechanism of typical catalases: The catalase cycle starts with the oxidation of the heme iron along with the heterolytic cleavage of the O-O bond of hydrogen peroxide. This process produces a ferryl-oxo porphyrin ?-cation radical ([Por . +] Fe IV =O) intermediate known as compound I with the concomitant production of a water molecule (Fig. 1.2) [9]. Kinetics studies [14] indicate that once compound I forms, it is rapidly and efficiently reduced back to the resting (i.e., ferric) enzyme by reacting with a second equivalent of H 2 O 2 , generating H 2 O and O 2 [15]. Two mechanisms have been proposed for O 2 production and the return of the ferric state: 1) a His-mediated mechanism, or 2) a direct mechanism involving radical formation (Fig. 1.2) [16]. The radical mechanism posits a radical intermediate generated upon hydrogen atom transfer from H 2 O 2 to the ferryl heme intermediate. This bears some resemblance to the mechanism proposed for alkane hydroxylation by cytochrome P450 [17]. In contrast, the His-mediated mechanism shows the participation of ionic species resulting from an initial proton abstraction carried out with the assistance of the distal His as the general base. A recent study of the catalase mechanism by QM/MM Car- Parrinello Metadynamics simulations highlight the key role of distal residues acting as an acid/base catalyst [15], lending credence to the His-mediated mechanism. 1.3 Peroxidases: The first part of this section will be dedicated to the structure and function relationships of heme dependent peroxidases from all five different superfamilies/families where the activity is 7 Figure 1.2: Catalytic mechanism of typical catalases. 8 found. Understanding these will provide a structural backdrop for understanding both the diversity and the commonalities of mechanisms and strategies by which heme-dependent peroxidase activity is accomplished. Because catalase-peroxidases (KatGs) fall into the peroxidase-catalase superfamily, the second part will progressively introduce members in this superfamily by discussing their phylogenic relationships and comparing their structures and functions with KatGs in order to understand the importance of the unique two-domain structure observed in KatGs. Together, these sections will lay the groundwork for evaluating the mechanisms by which the C-terminal domain modulates the conformation and performance of an active site that is sequestered in the N-terminal domain ? 30 ? away. They will also highlight strategies for using the stand-alone C-terminal domain as a scaffold for engineering new heme- dependent catalysts. 1.3.1 Systematic Classification of Peroxidases: Peroxidases are capable of reducing H 2 O 2 to water with the concomitant one- or two- electron oxidation of various organic and inorganic substrates according to reactions (2) and (3) [18]. Currently, 15 different Enzyme Commission (EC) numbers from EC 1.11.1.1 to EC 1.11.1.16 have been assigned to peroxidases [19]. Although there are heme and non-heme examples of peroxidases, the most abundant are those containing heme as the prosthetic group. As of January 2012 PeroxiBase (a database dedicated to cataloging genomic and other data on peroxidases), nearly 75% of all genes encoding enzymes with peroxidase activity are the heme- containing variety [20]. Traditionally, heme peroxidases have been divided into ?plant? and ?animal? groups, but the influx of new of genomic data has rendered such schemes obsolete. According to the current scheme, available heme peroxidase gene sequences are grouped 9 phylogenetically into three families (the di-heme peroxidases, the DyP-type peroxidases and the heme-haloperoxidases) and two superfamilies (the peroxidase-cyclooxygenase superfamily and the peroxidase-catalase superfamily) (Fig. 1.3) [21]. The majority of sequences fall into one of the two superfamilies, and the remaining sequences are distributed among the three families. 1.3.2 Di-heme peroxidases family: 1.3.2.1 Structure and function of di-heme peroxidases: Representatives of this family are found predominantly in bacteria with some members found in achaea. A detailed phylogeny has been proposed based on sequence data available with most of the sequences derived from the di-heme cytochrome c peroxidases (DiHCcP) [21]. As the name suggests these enzymes contain two hemes. They are typically periplasmic where they catalyze reduction of hydrogen peroxide to water using single electron donors such as cytochrome c or cupredoxin [22]. Though DiHCcPs bear no sequential or topological resemblance to KatGs, all of them have been shown to be dimeric where each subunit (35-40 kDa) is composed of two domains [23], similar to the two-domain structure of KatGs. In contrast to KatGs where only one domain (the N-terminal domain) is catalytic active, both domains of DiHCcPs participate in catalytic events. There is an electron-transferring heme (E heme) domain and a peroxidatic heme (P heme) domain [23]. Both hemes are of the c-type in that they are covalently linked to the polypeptide chain by way of cysteine side chains [24]. In contrast, the 10 Figure 1.3: Overview of the classification of heme peroxidases (super) families. 11 canonical eukaryotic cytochrome c peroxidase (from the peroxidase-catalase superfamily) contains a single b-type heme in its active site. The P heme is a low-potential center where H 2 O 2 is reduced. The E heme is a high- potential center that serves to transfer electrons from soluble electron-shuttle proteins such as cytochrome c or azurin to the peroxidatic site [25-27]. The P heme is coordinated by two axial histidine ligands, whereas the E heme is ligated by one His and one Met residue [23, 24, 27-31]. Crystal structures of DiHCcP from Pseudomonas aeruginosa reveal that a single Ca 2+ ion is tightly bound at the domain interface of both fully oxidized and mixed-valence forms of the enzyme. This Ca 2+ is absolutely required for catalytic activity [23]. The role of the Ca 2+ ion is unresolved, but it is proposed to maintain the structural integrity of the enzyme and/or modulate electron transfer between the two domains [23, 24, 27-31]. 1.3.2.2 Catalytic cycle and heme function: In most known bacterial DiHCcP enzymes, the inactive resting state can be converted to the active mixed-valence state by reduction of the E heme [32]. A mechanism proposed for the enzyme is shown in Fig. 1.4 [25]. The enzyme is isolated in an inactive state (state I) where both hemes are in a hexacoordinate ferric form. An electron (from ascorbate or the enzyme?s physiological substrates) is transferred to the E heme (step A). E heme reduction triggers a Ca 2+ - dependent conformational change (step B) that results in the removal of the distal histidine from the iron coordination sphere where it is replaced by a water molecule [26, 32, 33]. This produces the active form of the enzyme (state III). State III is then oxidized by H 2 O 2 at the P heme producing a ferryl intermediate and a ferric E heme. This would be roughly analogous to the 12 Figure 1.4: Catalytic cycle of bacterial diheme cytochrome c peroxidase. 13 compound I state of the monoheme peroxidases of eukaryotes (step C). The catalytic cycle is completed by the delivery of two electrons in single-electron steps (steps D and E) from cytochrome c or other substrates. Stage V is electronically equivalent to the resting enzyme, but its P heme is accessible to ligands. Relaxation back to the resting state is likely to be a slow process. It has been proposed that this step is only observed when insufficient reductant is present [32]. 1.3.2.3 Enzyme structure and conformational changes during catalytic cycle: As described earlier, the bacterial DiHCcP is composed of two heme-containing domains [23, 24, 27-31]. Both oxidized and mixed-valence forms of the enzyme share similar topological structures (Fig. 1.5). In the E heme-containing domain, the coordination of the heme remains unchanged by oxidation state. Conversely, the changes of the orientation of heme D-ring propionate group leads to a loss of an H-bond with the Leu 216, and in the mixed-valence state, Leu 216 moves away from the proximal heme site. The A-ring propionate group linked to Ca 2+ ion does not change under activation. In the P heme-containing domain, more dramatic changes occur during activation. The loop carrying His 71 moves toward the dimer interface concomitant with relocation of the aromatic side chain of Trp 73. The movement of this loop allows for the occupation of peroxidatic site by a water molecule, and the introduction of two other conserved residues Gln 104 and Glu 114, which have been suggested to play critical roles in the formation and stabilization of the oxy-ferryl intermediate [23, 27]. 14 E heme domain P heme domain His 201 Met 275 D A Leu 216 Leu 216 Ca 2+ Ca 2+ Ca 2+ Ca 2+ His 71 His 55 His 71 His 55 Arg 97 Arg 97 Glu 114 Glu 114 Trp 73 Trp 73 Gln 104 Gln 104 A B C D Figure 1.5: Structure of di-heme-cytochrome c peroxidase and conformational changes during catalysis. The overall structure of di-heme cytochrome c peroxidase from Pseudomonas aeruginosa (A); the conformational change in the E heme domain (B); the oxidized state of P heme domain (C); the mixed-valence state of P heme domain (D). The spheres represent Ca 2+ . The protein structure colored by green and heme highlighted by red represents the oxidized state (PDB: 1eb7) [27]; the protein structure colored by canyon and heme highlighted by hot pink represents the mixed- valance state (PDB: 2vhd) [27]. 15 1.3.3 DyP-type peroxidase family: Dye-decolorizing peroxidase (DyP) was discovered in the basidiomycete Thanatephorus cucumeris, and it has been shown that this enzyme has the ability to decolorize 18 types of reactive, acidic and dispersive dyes, most of which are xenobiotics [34]. It is a member of a newly discovered and intriguing family of heme peroxidases, the DyP-type peroxidase family. So far, these enzymes have been found only in bacteria and fungi [35]. In this family, four distinct subfamilies (from A to D) can be identified, and it has been surmised that through evolution, subfamilies A and B first diverged from C and D; subsequently, produced the four noted subfamilies [21]. DyP is a member of subfamily D, whereas TyrA, which we will introduce later, belongs to subfamily B. 1.3.3.1 Structures of DyP-type peroxidases: Consistent with its identity as a peroxidase, DyP contains a heme cofactor and requires H 2 O 2 for all reactions. However, DyP has several characteristics that distinguish it from other typical peroxidases. The level of sequence homology with other peroxidases is very low. It is capable of using a wide range of substrates and is tuned to operate at a surprisingly low pH (optimum 3 - 3.2) [36, 37]. Initially, DyP was considered to be a class II member of the peroxidase-catalase superfamily, however, it became increasingly apparent that the primary structure was distinct. Indeed, DyP even lackes the active site signature motif R-X-X-F/W-H that is a hallmark of the peroxidase-catalase superfamily [38]. However, discussion of their structure and function points toward a strategy that might be employed for engineering the C-terminal domain of KatG into a new low pH peroxidase. 16 Figure 1.6: Structures of DyP and DyP-type Peroxidase. A: Overall structure of DyP from Thanatephorus cucumeris Dec 1 (PDB: 2D3Q) [39]; B: Overall structure of TyrA from Shewanella oneidensis (PDB: 2IIZ) [40]; C: The heme and its surrounding residues important for the peroxidase activity of DyP; D: The active site surrounding of TyrA. 17 A striking distinction in active site structure is the absence of the typical distal His, instead, its replacement by an aspartate (Asp 171) [35]. Additionally, DyP contains a unique pair of anti- parallel ?-sheets located above the distal side of the heme (Fig. 1.6 A) [35]. Although sequence homology between DyP and other DyP-type peroxidase is very low, their overall tertiary structures are quite similar. TyrA, a member of type B DyP-type peroxidases, is a two-domain ? + ? protein. Each domain contains a four-stranded, antiparallel ?-sheet sandwiched by ?- helices in a ferredoxin-like fold (Fig. 1.6 B) [40]. At the active site level, three residues: an Asp, an Arg and the proximal His ligand are strictly conserved in this family (Fig 1.6 C and D). The Arg is proposed to assist with the formation of compound I by stabilizing higher oxidation states of heme?s iron center [35]. Asp instead of His as a general acid/base catalyst provides an explanation for the extraordinarily low pH optimum of DyP and DyP-type peroxidases [36, 37]. 1.3.3.2 Catalytic cycle of DyP and Dyp-type peroxidases: A mechanism for the formation of compound I based on the participation of Asp and Arg in the active site has been proposed (Fig. 1.7) [35]. According to this model, a conserved Glu 391 forms an H- bond with proximal ligand His 308, increasing the anionic character of the ligand, serving to stabilize higher heme iron oxidation states. In the resting state of DyP, a sixth ligand to the heme iron is absent, allowing for direct formation of a complex with H 2 O 2 upon its entrance to the active site. Asp 171 acts as a general base to abstract a proton from the peroxide and deliver it to the distal peroxide oxygen atom concomitant with heterolytic cleavage of the O? O bond. Subsequently, one electron is removed from iron, and another electron is transferred from porphyrin to the remaining oxygen, generating an intermediate analogous to compound I observed in other peroxidases. Recently, a modified mechanism for formation of compound I has 18 Figure 1.7: Schematic diagram of the proposed formation of compound I by DyP. Figure 1.8: A modified mechanism of compound I formation by DyP. 19 been proposed [39] (Fig. 1.8). In this model, the carboxylate group of Asp 171 associates with the ligand for heme iron by swinging into the appropriate position [39]. Once the compound I is formed, the carboxylate group can revert to its original position in order to stabilize compound I through an H-bonded network involving an active site water molecule. Interestingly, a compound II (ferryl) has never been detected in these enzymes [35]. To date, the picture of the catalytic mechanism of Dyp (or DyP-type) peroxidases remains incomplete. 1.3.4 Haloperoxidases Family: Haloperoxidase activity has been observed with a variety of heme- or vanadium-dependent enzymes. However, the non-heme (vanadium) haloperoxidases are phylogenetically unrelated to this family. The heme-dependent haloperoxidases are abundant among fungi. They contain protoporphyrin IX as a prosthetic group and catalyze the oxidative transformation of halides and halophenols [41]. A recently constructed phylogenetic tree for this family indicates that heme haloperoxidases form a monophyletic group with frequent gene duplication events [21]. 1.3.4.1 Structure and catalytic mechanisms of haloperoxidases family: In this family, the most intensively investigated member is the chloroperoxidases (CPO) from the ascomycete Caldariomyces fumago. This CPO is a 42 kDa glycoprotein. It catalyzes the H 2 O 2 -dependent chlorination of cyclopentanedione as well as the iodination and bromination of a wide range of substrates [42-46]. 20 Figure 1.9: Structure of CPO. Overall structure of CPO with binding of bromide (A) (PDB: 2CIV) and iodide (B) (PDB: 2CIW) [47], and the active site of CPO (C). Dots represent the halides. 21 Available structures from the protein data bank indicate that CPO has a novel tertiary structure dominated by eight helical segments [48]. In the active site, a cysteine serves as the proximal ligand to the heme iron. Unlike other peroxidases, there is no Arg in the distal cavity, and the catalytic acid/base required to cleave the peroxide O-O bond is Glu rather than His (Fig.1.9). Consequently, CPO is active at low pH with an optimum at 2.8 [42, 47]. This unique active site configuration points to a future strategy by which our stand-alone KatG C-terminal domain (KatG C ) may be engineered. Crystal structures solved in the presence of bromide and iodide have been reported. The bromide/iodide binding sites have been identified. The location of these ions points toward a route by which halides may access the heme center [47]. 1.3.4.2 Catalytic cycle of CPO: The fundamental mechanism of CPO-catalyzed halogenations is presented in Fig 1.10 [41]. It begins with the heterolytic cleavage of a loosely bound H 2 O 2 , resulting in the formation of compound I and H 2 O. Compound I is not stable and is proposed to react with the halide to form a hypothetical ferric hypohalite adduct, termed compound X. Decomposition of compound X is proposed to produce a halonium ion (X + ) along with ferric enzyme. In aqueous solution, the hypohalous acid would be expected to form from X + . Recently, based on substrate-bound structures, a modified mechanism for compound I formation has been proposed [47]. Instead of direct formation of compound I, a ferric-hydroperoxyl species (compound 0) is proposed to be formed first; the Glu 183 then protonates the distal oxygen of compound 0, followed by the formation of compound I along with cleavage of the O-O bond and release of one water molecule (Fig. 1.11). 22 Figure 1.10: Catalytic cycle of haloperoxidases. 23 Figure 1.11: Proposed mechanism of compound I formation by chloroperoxidase. 24 1.3.5 The peroxidase-cyclooxygenase superfamily: The peroxidase-cyclooxygenase superfamily represents one of the two main groups of heme-dependent peroxidases in biology. Peroxidase-cyclooxygenase and peroxidase-catalase superfamilies arose independently; therefore, the structures of both the protein and prosthetic group are distinct between these two groups [18]. The peroxidase-cyclooxygenase superfamily was originally called the mammalian peroxidases, and was later changed to the animal peroxidase family. However, in recent years, even plant, fungal and bacterial representatives were found and assigned to this superfamily necessitating yet another change in nomenclature. Though this superfamily has been subdivided into seven clades, it is still identified by its most famous members (myeloperoxidase [MPO], eosinophil peroxidase [EPO], lactoperoxidase [LPO] and thyroid peroxidase [TPO]) all of which are found in animals. Because this superfamily is not phylogenetically related to the superfamily containing KatG, this section will only consider the intensively investigated representative, myeloperoxidase (MPO). An interesting feature of many representatives from the peroxidase-cyclooxygenase superfamily is novel covalent linkage between the porphyrin cofactor and protein. In this, MPO is an excellent example. 1.3.5.1 Structure of MPO: Myeloperoxidase is abundant in neutrophils but also is expressed in monocytes and certain types of macrophages, where it participates in innate immune defense mechanisms by generating reactive oxidants and diffusible radical species [49]. The mature enzyme exists as a homodimer, where each subunit is dominated by ?-helices with few ?-strands. The subunit is composed of two polypeptide chains resulting from the post-translational excision of a peptide fragment from a single polypeptide precursor [50, 51]. The catalytically essential residues Gln 91, His 95 and 25 Arg 239 occupy the distal heme pocket; in the proximal site, the heme iron is coordinated through the His 336 imidazole resiude. This ligand forms an H-bond with the amide carbonyl oxygen of Asn 421. One Ca 2+ ion has been identified and its coordination is essential for maintaining the distal architecture. A common, but not universal, feature for members of this superfamily is the covalent association of the heme with the protein itself: pyrrole rings A and C of the heme group are methylated, allowing formation of ester linkages with the carboxyl groups of Glu 242 and Asp 94. Additionally, the ?-carbon of the vinyl group on pyrrole ring A forms a covalent bond with the sulfur atom of Met 243 (Fig. 1.12). As one might anticipate, this produces unique spectral properties for the heme. In the ferric state, the Soret band shows an absorption maximum around 428 nm and relatively strong absorption bands in the visible region are also detected, which are responsible for the characteristic green color of the enzyme [52]. This unique post-translational modification gives MPO peculiar redox properties, allowing oxidize Cl - [53]. 1.3.5.2 Enzymatic properties and catalytic mechanism: Under physiological conditions, MPO catalyzes the oxidation of halides (X - ) to the corresponding hypohalite anions (OX - ). To do this the ferric enzyme is oxidized by H 2 O 2 to form compound I. Distal His 95 and Arg 239 are proposed to orient and bind the H 2 O 2 , allowing the heterolytic fission of the O-O bond much as has been observed for other peroxidatic enzymes. In addition to its halogenation activity, MPO is also able to catalyze more typical oxidations of other peroxidatic electron donors. Therefore, once compound I is formed, MPO can catalyze 26 Figure 1.12: Structure of human MPO. Overall structure (A), and active site (B) of human MPO. Different colors highlight the structures from two different polypeptides (PDB:1CXP) [54]. 27 Figure 1.13: Catalytic cycles of MPO. Halogenation cycles: reaction 1 and 2; Peroxidase cycles: reaction 1, 3 and 4 [49]. 28 either halogenation or canonical peroxidation (Fig. 1.13). In the halogenation cycle, compound I is reduced back to the ferric state by the oxidation of halide (X - ) substrates. In the peroxidase cycle, compound I is reduced back to the native state in two sequential one-electron steps using another exogenous reducing substrate (AH 2 ). Presumably, the operating cycle is determined by the availability of substrate. 1.3.6 The peroxidase-catalase superfamily: The first systematic classification of the enzymes of this superfamily was published in 1992 [38] and divided its members into three classes. In contrast to other peroxidase superfamilies, this subdivision has remained unchanged since that time. Class I includes cytochrome c peroxidases, ascorbate peroxidases and catalase-peroxidases (KatG). Class II contains fungal peroxidases, predominantly lignin and manganese peroxidases. Class III includes the secretory plant peroxidases. Though the subdivisions have not changed, the name of the overall superfamily was recently changed from the superfamily of plant, fungal, and bacterial heme peroxidases to peroxidase-catalase superfamily in order to reflect the main enzymatic activities of its members [21]. The Class I members are widely distributed in nature with the exception of animals. Though the physiological roles of the peroxidase-catalases vary considerably, the class I enzymes serve exclusively as H 2 O 2 scavengers [55-57]. The class II peroxidases are found exclusively in fungi and have a major role in degradation of lignin [58, 59]. Class III are mostly plant-secreted glycoproteins with roles in a variety of physiological processes of the plant life 29 cycle such as cell wall metabolism [60], wound healing [61, 62], auxin catabolism [63], H 2 O 2 scavenging, and defense against pathogens [64]. 1.3.6.1 Class III: Plant secretory peroxidases: The most intensively investigated representative of this family is horseradish peroxidase (HRP) from Armoracia rusticana. In this section, the discussion of structure and catalytic mechanism of class III peroxidases will focus on HRP. 1.3.6.1.1 Structure of HRP: Crystal structures reveal that HRP has ten helices assigned A-J which produce a fold that is highly conserved in the Class III family, and several unique features: a long insertion between helices F and G; a short antiparallel ??sheet preceding the C177-C209 disulfide bridge; and a three-residue insertion between the H and I helix [65]. In HRP, all the N-glycosylation sites positioned in loop regions are pointing away from the molecule except Asn 268 which is located right at the end of the I helix, indicating the purpose of glycosylation is to increase solubility of the enzyme [65]. Recent reports indicate that glycan may stabilize the protein and adjust the substrate access [66-68]. In addition, there are four disulfide bonds between cysteine residues 11- 19, 44-49, 97-301 and 177-209. Of those, the disulfide bond in the distal heme-binding domain is proposed to dedicate to stabilizing active site structure [64]. In the active site, the ferric heme b prosthetic group is ligated to the protein though His 170, which also hydrogen bonds to the Asp 247 carboxylate group. Three strictly conserved residues are located at the distal site: His 42 serves as a general base/acid; Arg 38 is essential for 30 the formation and stabilization of compound I; Phe 41 prevents substrate access to the ferryl oxygen of compound I [69]. In addition, Asn 70 is strongly conserved in the peroxidase-catalase superfamily and has been proposed to maintain the basicity of the distal His (His 42) [70]. In both the distal and proximal sites, two hepta-coordinate Ca 2+ ions are required for both structural stability and catalytic functions [69, 71, 72] (Fig. 1.14). Loss of Ca 2+ or disruption of its interactions results in the B-helix bearing distal His shift, allowing the distal His to directly coordinate with the heme iron. Therefore, the Ca 2+ -deficient HRP is inactive, but this is reversible and can be remedied by addition of Ca 2+ [69, 71, 72]. 1.3.6.1.2 Catalytic mechanism of HRP: The mechanism of catalysis of HRP has been extensively investigated [73, 74]. Similar to those of other peroxidases we described previously, the first step in the catalytic cycle is generation of compound I. Once compound I is formed, typically aromatic electron donors reduce compound I to compound II, then compound II is reduced back to the resting state of the enzyme by another equivalence of reducing substrate. Although the formation of these intermediates has been proposed for a number of catalytic heme enzymes, these species are formed only transiently and have been somewhat elusive to identify until recent years. The structures of the HRP catalytic cycle intermediates have been obtained at high resolution by X- ray crystallography, allowing direct and accurate comparison of the bonding interactions in the different intermediates [75]. The catalytic cycle of HRP with the structures of the intermediates is summarized in Fig. 1.15. 31 Figure 1.14: Structure of HRP. Overall structure of horseradish peroxidase (HRP) (PDB: 1h5a) (A) and the active site of HRP (B) [75]. 32 Figure 1.15: Peroxidase cycle of HRP. 33 1.3.6.2 Class II: Manganese, lignin and versatile peroxidases: Extracellular Class II heme peroxidases are currently known only in fungi and have their principle function in lignin degradation [76]. There are three main groups secreted by Basidiomycetes or so-called white rot fungi: manganese peroxidases (MnP), lignin peroxidases (LiP), and versatile peroxidases (VP). 1.3.6.2.1 Structure and function of Class II peroxidases: The degradation of lignin to CO 2 by this small and specialized group of fungi is a major route for the recycling of photosynthetically fixed carbon [77]. Lignin is chemically recalcitrant to catabolism by most organisms because of its complex, heterogeneous and irregular structure. In nature, only the Basidiomycota are able to degrade lignin efficiently. They produce four major groups of enzymes for lignin degradation: LiP, MnP, VP that belong to Class II peroxidase family and a laccase, which is a multi-copper-containing protein. Along with the two other peroxidases, LiP catalyzes the H 2 O 2 -dependent oxidative degradation of lignin, but the mechanism by which this occurs is poorly defined [78, 79]. In addition, LiP displays classical peroxidatic catalytic promiscuity with respect to the electron donor substrate. The enzyme has been shown to oxidize phenolic compounds (especially veratryl alcohol) as well as a broad range of nonphenolic aromatic compounds. Interestingly the reach of LiP in terms of the redox potential of compounds it will oxidize extends as high as 1.4 V [80]. The pH optimum (~3) for the enzyme is strikingly low considering the structure of its active site. Crystal structures for LiP reveal eight major and eight minor ?-helices with a few ? structures in the proximal domain [81]. The overall folding and active site environment of LiP 34 are highly similar to those of HRP (Fig. 1.16A). In addition, LiP contains three glycosylation sites, which have been indicated to protect the C-terminal peptide from proteolysis [77]. Similar to HRP, LiP has four disulfide bridges and two structural calcium ions which are essential for maintaining the structure of the active site [81]. The location of the two Ca 2+ coordination sites are similar to those observed in HRP; however, the disulfide bond that enhances the stability of the distal Ca 2+ binding site in HRP is absent in LiP, making the coordination state and enzyme activity much more sensitive to the loss of the Ca 2+ ion in LiP. Interestingly, a novel post-translational hydroxylation of Trp 171 (C?) has been identified [81]. This is proposed to have an important role in the binding and oxidation of aromatic substrates through long-range electron transfer [82-84]. This Trp residue and the electron- deficient heme are purported to account for the high reduction potentials of the enzyme?s ferryl intermediates [85]. These endow LiP with the unique ability to catalyze oxidative cleavage of C- C bond and ether bonds in non-phenolic aromatic substrates. Manganese peroxidase is an acidic glycoprotein. The overall structure of MnP is similar to that of LiP [86]. Two heptacoordinated structural calciums with similar positions and functions as those observed in LiP [87] (Fig. 1.16B). The active sites of MnP and LiP are virtually identical to one another. MnP does possess an additional disulfide bond in addition to those four that are also seen in LiP. This unique fifth disulfide is proposed to assist in Mn 2+ binding by pushing a C-terminal peptide segment away from the main body of the protein [77]. A bound Mn(II) ion has been identified at the surface of the protein coordinated to the carboxylates of Glu 35, Glu 39 and Asp 179, one of the heme propionates and two water molecules. As described in 35 greater detail below, MnP utilizes Mn 2+ as a direct substrate, the oxidation product of which mediates the oxidation of secondary substutrates. Structurally, versatile peroxidases (VP) resemble LiP more closely than MnP. The available structure shows a 12 helices, 4 disulfide bonds, a heme pocket containing the characteristic proximal and distal triads, 2 structural Ca 2+ sites, and a Mn (II) binding site (Fig. 1.16C) [88]. The solvent accessible Trp 64 is essential in the oxidation of high redox potential aromatic substrates via a long-range electron transfer pathway similar to that observed for LiP [89]. 1.3.6.2.2 Catalytic cycles of Class II peroxidases: LiP has a typical peroxidase catalytic cycle similar to that of horseradish peroxidase (Fig. 1.15). The primary different is in the involvement of long-range electron transfers for LiP. In a sense, MnP utilizes a similar long-range electron transfer strategy as that observed for LiP; however, instead of oxidizing a surface tryptophanyl residue, MnP catalyzes the oxidation of Mn(II) to Mn(III) which in turn may be used to oxidize numerous monomeric phenols including dyes as well as phenolic lignin model compounds [90]. The overall characteristics of the cycles of LiP, VP, and MnP are highly similar to those of HRP (Fig. 1.17). Another difference between LiP and MnP is oxidation of non-phenolic substrates by MnP involves the formation of reactive radicals in the presence of second mediators (e.g., thiols) [91]. The mechanism employed by VP is essentially the same as that of LiP except that VP can also utilize Mn(II) as a mediatory substrate much as MnP does. 36 Figure 1.16: Structure of Class II peroxidases. Overall structures of LiP (PDB: 1LGA)(A) [81], MnP (PDB: 1MNP) (B) [86] and VP (PDB: 2BOQ) (C) [88], and the active sites of LiP (D), MnP (E) and VP (F). 37 Figure 1.17: Catalytic cycle of manganese peroxidase (MnP). 38 1.3.6.3 Class I peroxidases: 1.3.6.3.1 Introduction of Class I peroxidases: Class I peroxidases include intracellular prokaryotic and eukaryotic non-glycosylated heme peroxidases without Ca 2+ [38, 92]. Extensive gene sequence analysis of all three classes indicate that the whole peroxidase-catalase superfamily originated from ancestral Negibacteria, and during evolution the already diversified peroxidase genes were transferred via two ancestral endosymbiotic events to newly formed eukaryotic cells [19, 93]. Class I segregated at an early stage from a common ancestor of the class II and class III peroxidases [94]. However, the evolutionary lineage of CCP and APX from the two-domain KatG remains contested [38, 92]. Ascorbate peroxidases are predominantly observed among green algae and higher plants. Their primary physiological role is to scavenge H 2 O 2 as part of the ascorbate-glutathione cycle [55, 95]. Ascorbate peroxidase plays an important role in root nodules of leguminous plants and chloroplasts in general. Given that O 2 is a known inhibitor of photosynthesis and can deactivate nitrogenase, a peroxide degradation mechanism that avoids generating O 2 is particularly advantageous. The subfamily of CCPs can be divided into two groups: mitochondrial and non- mitochondrial. The enzyme from Saccharomyces cerevisiae is located in the intermembrane space of mitochondria where it decomposes H 2 O 2 generated by the mitochondrial respiratory chain [96]. In addition to acting as a H 2 O 2 scavenger, CCP appears to be integral to the oxidative stress response to methanol utilization in the methylothropic yeast [97]. The enzyme from 39 pathogenic basidiomycetes has been implicated in the bacterium?s resistance against exogenous oxidative stress [98]. 1.3.6.3.2 Structure and Catalytic function of APX and CCP: As with all members of the peroxidase-catalase superfamily, the structures of all class I enzymes are dominated by the ??helix. As a percentage, KatGs exhibit the lowest contribution from ??helix, whereas the APXs have the highest with CCPs falling in between. All class I enzymes carry heme b as a cofactor, and in each case, it is non-covalently held by the protein that carries it. The distal Arg-Trp-His and proximal His-Asp-Trp triads of the active sites are strictly conserved among class I enzymes not only in sequence, but also in their position/orientations within the active site (Fig. 1.18) [2]. There is some evidence that the distal Arg may occupy two positions: one ?out? and one ?in? [99], where only Arg occupying the ?in? position can form H-bond with the iron-linked ferryl oxygen atom. As elsewhere this interaction is proposed to stabilize the ferryl center of compound I [99]. Crystal structures as well as spectroscopic properties of the enzymes indicate several differences between CCP and other peroxidases. The typical compound I (ferryl porphyrin radical) does not accumulate in CCP due to facile electron transfer from an active site, forming instead a ferryl tryptophanyl radical intermediate referred to by some as ?compound ES? [100]. In addition, two fully conserved methionines (230 and 231) serve to stabilize the Trp radical [101]. In contrast, a typical compound I (ferryl porphyrin radical) intermediate is observed upon reaction of APX with H 2 O 2 . In this protein, a potassium ion bound adjacent to the proximal Trp not only stabilizes the conformation of the APX active site [102], but also disfavors Trp oxidation to its corresponding radical intermediate. Beyond this, the catalytic cycles APX and CCP are highly similar to those 40 observed for class II and III peroxidases (Fig. 1.15 and 1.17). One notable feature in both cases is the identity of the preferred electron donor substrate. Reduced cytochrome c and ascorbate are both strikingly poor electron donors for all other peroxidases. In these instances, CCP and APX have clearly adapted to employ the electron donor that is available in the cellular environment of each one. 1.4 Catalase-peroxidases (KatGs): Phylogenetic analyses suggest the duplication event evident in the KatG gene may have occurred in the later phases of evolution, where the individual and distinguishing features of the peroxidase families were already formed [102]. The KatGs are broadly distributed among archaea, bacteria and lower eukaryotes. It appears that their primary purpose across these organisms is to degrade H 2 O 2 . Despite all their similarities to other peroxidase-catalase superfamily members (especially class I enzymes), KatGs are unique in that they use single active site to catalyze two fundamentally different reactions. 1.4.1 The biomedical benefits of understanding KatG structure and function: The vast majority of superfamily members are monomeric, and with the lone exception of KatG, the subunits of these enzymes are composed of a single domain. As a consequence, the novel two-domain composition of KatGs provide a rare opportunity to evaluate mechanisms in which structures peripheral to and distant from an active site influence its architecture and catalytic ability. This aspect of the structure/function relationship is rarely studied and poorly understood. 41 Figure 1.18: Structure of Class I peroxidase. Overall structure of CCP (PDB: 2CYP)(A), APX (PDB: 1OAF) (B) and the active site of CCP (C), APX (D) [103, 104]. 42 There are substantial biomedical benefits to be derived from an understanding of KatG structure and function. First, KatG from Mycobacterium tuberculosis has a central role in the activation of isoniazid (INH), a front-line antitubercular agent. In M. tuberculosis, INH is activated by KatG, leading to the formation of an INH-NAD adduct [105-108]. This adduct inhibits InhA, the enoyl-ACP reductase of the fatty acid synthase type II system, which is the central enzyme complex in mycolic acid biosynthesis [109, 110]. Consequently, inhibition of mycolic acid production interferes with mycobacterial cell wall construction and ultimately results in cell death [109, 110]. It has been estimated that 70% of isoniazid resistant strains carry mutations to the gene for KatG. The mechanisms by which many of the commonly observed mutations compromise isoniazid activation are not known. Clearly, a better understanding of KatG structure and function will be beneficial for resolving these questions. In addition, KatGs have been implicated as virulence factors in organisms like Yersinia pestis [111-113], Legionella pneumophila [114], and enterohemorrhagic E. coli [115, 116]. These organisms produce a periplasmic or extracellular form of KatG. Because a central strategy of antibacterial responses of plants, mammals and other higher eukaryotes relies on the production of copious quantities of reactive oxygen species, particularly H 2 O 2 , the advantage to be gained from a robust extracellular catalase is considerable. The oxidative burst of neutrophils and macrophages rapidly produces high levels of O 2 ?? which disproportionate to H 2 O 2 . In neutrophils, H 2 O 2 is used by myeloperoxidase to oxidize chloride to hypochlorous acid. In macrophages, the rapid generation of ?NO along with O 2 ?? produces peroxynitrite anion. The ability to rapidly degrade H 2 O 2 and other peroxides like peroxynitrite is essential for pathogenic bacteria employ as a defense against the host?s immune response. Therefore, the periplasmic 43 location of these enzymes is an essential strategy that allows access to the target substrate before it has had an opportunity to damage essential cellular systems. The connection between KatG structure and its utility in this role remains to be explored, and a greater knowledge of KatG structure will be necessary in order to exploit this enzyme as a target to hobble bacterial antioxidant defenses, making pathogens like E. coli O157:H7 or Yersinia pestis more vulnerable to the hosts? own natural antibacterial defenses. 1.4.2 Structure of KatGs: The catalase-peroxidases exist as either homodimers or homotetramers [117]. Each subunit is composed of two structurally homologous domains [118]. Neither domain bears any resemblance to monofunctional catalases; rather, both domains strongly resemble monofunctional peroxidases, bearing the same topological structure and similar helical structure around the active sites [119-122]. Structural data and amino acid sequence alignments indicate that there are 10 major helices in each domain, similar to other class I members. Interestingly, the heme-binding consensus sequence is only observed in the N-terminal domain, and hence, this domain contains the only functional active site. In contrast, the C-terminal domain lacks the heme-binding motif, and as a result, does not bind heme or catalyze a reaction. At the active site level, the available crystal structures reveal that the KatGs have the proximal and distal conserved amino acids at almost identical positions as in other class I peroxidases. In particular, both the triads His/Trp/Asp and His/Arg/Trp are strongly conserved and almost in superimposable positions found in other class I peroxidases (Fig. 1.19) [119-122]. In addition, a conserved Asp 135 participates in the extensive KatG-typical hydrogen-bond 44 Figure 1.19: Comparisons of the overall structures and the active sites of monofunctional peroxidases, KatG and monofunctional catalases. APX, (PDB: 1OAF) (A and D, respectively), catalase-peroxidase (PDB: 1SJ2) (B and F, respectively) and HPII from E.coli, (PDB: 1IPH) (C and G, respectively). 45 network at the distal heme pocket. Asn 136 is conserved in all members of the peroxidase- catalase superfamily and supports this network through a hydrogen bond to the distal His. On the basis of site-directed mutagenesis and spectroscopic studies, it has been suggested that the catalase-unique reaction requires this precisely aligned H-bond network [123-125]. Another obvious difference between KatG and monofunctional peroxidase is that there is a pronounced funnel channel to the active site in KatG that is reminiscent of monofunctional catalases. In KatG, the acidic residues are critical for stabilizing the solute matrix and orienting the water dipoles in the channel. Accordingly, exchange of the acidic residues affects catalase but not peroxidase activity [126]. Most KatGs do not contain any disulfide bonds. Extracellular KatGs secreted by certain fungi stand as an exception. Indeed, the recently solved structure for a eukaryotic KatG from Magnaporthe grisea shows two disulfide bridges between the N-terminal domains of two subunits [127]. These cysteines, and presumably the disulfide bonds which result, appear to be fully conserved but only in the group of the secreted fungal KatGs (Fig. 1.20). They are also absent in the other members of class I peroxidases such as CCP and APX. These disulfide bonds significantly increase the structural integrity of the enzyme, making the N-terminal domain more stable than the C-terminal domain [127]. Although KatG belongs to the class I peroxidase family, none of the other class I members (e.g., CCP and APX) has appreciable catalase activity. Given the high similarity of the monofunctional peroxidase and KatG active sites, it is reasonable to suggest that KatG-unique protein structures external to the active site have an important role in fine-tuning the active site 46 Figure 1.20: Two novel disulfide bridges between the N-terminal domains of two subunits in KatG. KatG from M. Grisea KatG 2 (PDB: 3UT2) shown in the overall structure (A) and zoom-in view (B) [127]. 47 for its bifunctional abilities. In support of this hypothesis, sequence alignments [118, 128] and the four available crystal structures of the KatGs [119-122] reveal several structural features that are unique to these enzymes. Two of them are Large Loops 1 and 2, LL1 and LL2, respectively (Fig. 1.21). Relative to the other class I enzymes, each one adds about 35 amino acids to the interhelical loops connecting the D and E helices and F and G helices, respectively. Evaluation of a wide array of substitution and deletion variants targeting the LL2 structure suggests that this loop structure may support hydrogen-bonded networks critical for reactions involving H 2 O 2 and may regulate access of electron donors to the active site [129, 130]. A striking feature of LL1 is that it contains the invariant Tyr (Y226 by E. coli KatG numbering), which participates in a novel Met-Tyr-Trp covalent adduct located on the distal side of the heme active site. This adduct has been observed in the crystal structures [119-121, 131] (Fig. 1.22) as well as by mass spectrometric peptide mapping [132-135]. Elimination of LL1 or substitution of Y226 eliminates catalase activity, but deletion variants are substantially more active as peroxidases, up to an order of magnitude increase to the rate in same cases [128, 136]. In addition to orienting this invariant tyrosine for participation in the essential covalent adduct, LL1 also appears to regulate the access of reducing substrates to the heme edge [128, 136]. 1.4.3 Catalytic cycles of KatGs: KatGs are capable of H 2 O 2 decomposition by catalatic and peroxidatic mechanisms, and they do so using a single active site. In the classical model, both catalase and peroxidase cycle start with heterolytic reduction of H 2 O 2 along with oxidation of ferric enzyme to compound I (Fig. 1.23A). Once compound I is formed, it can be reduced to the resting state either 48 Figure 1.21: Structure of catalase-peroxidase. KatG from M. tuberculosis (PDB: 1SJ2) [131]. 49 Figure 1.22: The covalent adduct in KatG. KatG from M. tuberculosis (PDB: 1SJ2) (Numbering reflects E.coli structure of KatG) [131]. 50 Figure 1.23: Classical model of catalytic cycle of KatG (A), modified model of catalase mechanism of KatG (B). 51 through a catalase or peroxidase cycle. However, this model cannot explain the role of M-Y-W adduct in the catalytic mechanism. In addition, recently, the Goodwin laboratory has reported the stimulation of catalase activity by classical peroxidatic electron donors rather than inhibition, particularly under conditions favorable to peroxidase activity. This is wholly inconsistent with the classical paradigm. One mechanism proposed for the catalatic turnover of KatG that accounts at least for the former observation has been proposed. By this mechanism compound I is converted to compound I* wherein the unique covalent adduct serves as an endogenous electron donor. An additional equivalent of H 2 O 2 then converts compound I* to compound III*. Finally, the adduct radical serves as an electron acceptor returning the enzyme to its ferric state with the concomitant release of O 2 [137-140] (Fig. 1.23B). 1.4.4: The roles of C-terminal domain: As mentioned above, all other members of the peroxidase-catalase superfamily are single domain proteins. Thus, the C-terminal domain, a structure ? 30? from the active site, is unique to KatG. The structure is conserved across all KatGs. Based on its sequential and structural similarity to the N-terminal domain, the C-terminal domain is proposed to have originated from a gene-duplication and fusion event [118, 128, 141]. However, its roles in catalysis are not immediately obvious. Nevertheless, the Goodwin laboratory has observed that KatG without its C-terminal domain (KatG N ) is no longer expressed in a soluble form but is instead found in inclusion bodies. In addition, following denatured purification, refolding and reconstitution, KatG N shows a shifted active site structure, allowing what appears to be the direct coordination of the heme iron by the distal histidine (H106 by Escherichia coli numbering) [142]. Consequently, KatG N has neither catalase nor peroxidase activity. It has been suggested that the 52 C-terminal domain might serve as a platform to direct the proper folding of the N-terminal domain and facilitate dimerization [143]. This is supported by previous observations from the Goodwin laboratory [142, 144]. Introducing the separately expressed and isolated C-terminal domain (KatG C ) results in restructuring KatG N that is evident from a return of high-spin coordination states typical of the wild-type enzyme as well as the restoration of catalase and peroxidase activities [144]. Therefore, the C-terminal domain is an integral component of KatG intimately tied to its structure and function; this is also evident from a recent thermodynamic study of KatGs showing that the N- and C-terminal domains unfold as a single unit rather than independently [145]. However, the mechanism by which the C-terminal domain modulates the active site conformation is still not clear. Inspection of the crystal structures for various KatGs reveals that there are clear interactions between the two domains, which are proposed to contribute to the stability of the overall quaternary structure of the enzyme [119, 120, 122, 146]. As an extension of this hypothesis, it is reasonable to propose that the C-terminal domain, through these interactions, directs conformational adjustments in the adjacent domain including the active site itself. In order to evaluate the individual structures and mechanisms employed by the C-terminal domain, we focused our attention on a conspicuous structural element of this domain, the I?-helix (Chapter 2). Our results show that this helix is a central component of a platform to facilitate folding events and/or conformational adjustment in the N-terminal domain. Upon establishing the critical contributions of this structure, we turned our attention to understanding the mechanism(s) by which this remote structure communicates with and controls the active site (Chapter 3). The C-terminal domain has a unique evolutionary history. It was born in a gene- 53 duplication and fusion event. Either because the maintenance of its own catalytic function was unnecessary, or because a catalytically active C-terminal domain was a hindrance to the overall function of KatG, the specific structures necessary for heme binding and catalytic transformation of H 2 O 2 were eliminated. Nevertheless, the general scaffold for a heme-based catalyst was retained to near perfection. We surmised that this provided an ideal avenue toward the creation of new heme protein-based catalysts. Our efforts and accomplishments in this direction are given in Chapter 4. Finally, our previous observations have shown that the C-terminal domain is essential for both catalase and peroxidase activities of KatG. It is also known that a novel covalent adduct in the N-terminal domain is critical, but only to the catalase activity of the enzyme. More importantly, the establishment of this adduct is both heme and peroxide dependent. This would suggest that the C-terminal domain is also essential for formation of the covalent adduct, but this has never been investigated. By way of a novel procedure to evaluate covalent adduct formation, we have determined that establishment of the covalent adduct cannot occur without the C-terminal domain (Chapter 5). It is noteworthy that although the C-terminal domain accounts for nearly half the structure of this important enzyme, prior to these investigations it remained the least understood of all of KatG?s novel structural features. Together, these studies represent a substantial advance in understanding the function and potential applications of this important structure, and we anticipate that they will contribute not only to answering fundamental questions surrounding protein structure and function, but also to providing insight into the specific biomedical importance of KatG, in particular, its central involvement in Mycobacterial resistance to antibiotics and the virulence of pathogenic bacteria. 54 CHAPTER TWO INTEGRAL ROLE OF THE I?-HELIX IN THE FUNCTION OF THE ?INACTIVE? C- TERMINAL DOMAIN OF KATG 2.1 Introduction: Catalase-peroxidases (KatGs) have been the subject of considerable attention since the discovery of their role in the activation of the frontline anti-tubercular agent isoniazid in Mycobacterium tuberculosis [105-107]. Catalase-peroxidases have also been implicated as virulence factors for some highly virulent pathogens such as Yersinia pestis [111-113], Legionella pneumophila [114]. Nevertheless, the mechanism by which these enzymes may serve as virulence factors have not been illuminated. Clearly, there are important benefits to be derived from understanding the structure and function of KatG. Catalase-peroxidases are distributed among archaea, eubacteria, fungi and protists [93] primarily as hydrogen peroxide scavengers [147]. These enzymes use a single active site to degrade H 2 O 2 by either a catalase or peroxidase mechanism. KatGs are either homodimers or homotetramers [148], and each subunit has two structurally homologous domains [118]. Neither domain bears any resemblance to monofunctional catalases; rather, both domains strongly resemble monofunctional peroxidases, bearing the same topological structure and similar helical structure around active sites [119-122]. Indeed, catalase-peroxidases belong to class I of the 55 peroxidase-catalase superfamily [118]. However, none of the other class I members (e.g., cytochrome c peroxidase [CCP] and ascorbate peroxidase [APX]) has appreciable catalase activity. Sequence alignments [118, 128] and the four available crystal structures of the KatGs [119-122] reveal several structural features that are unique to catalases-peroxidases. Two of them are Large Loop 1 (LL1) and Large Loop 2 (LL2). Relative to the class I enzymes, each one adds about 35 amino acids to the interhelical loops connecting the D and E helices and F and G helices, respectively. Likewise, other peroxidases are single-domain proteins, making the 300- amino acid C-terminal domain another unique feature of KatGs. It may be the persistence of the longer interhelical loops, both of which are essential for catalase but not peroxidase activity [132, 134, 149-151], along with the necessity for a precisely ordered active site to support catalase activity [145] that has compelled the continued presence of the C-terminal domain in KatG but not its most closely related neighbors CcP and APx. The C-terminal domain is ? 30? from the active site. It is conserved across all catalase- peroxidases, and based on its sequential and structural similarity to the N-terminal domain, is proposed to have originated from a gene-duplication and fusion event [118, 128]. However, its roles in catalysis are not immediately obvious. It bears a striking resemblance to monofunctional peroxidases [119-122, 143], but it neither binds heme nor catalyzes any discernable reaction. It is proposed that the class I enzymes were all derived from a two-domain predecessor, suggesting that deletion of one of the domains was a necessary step in the development of CcP and APx [92]. In this, the C-terminal domain has the appearance of a vestigial structure. 56 Contrary to this notion, KatG without its C-terminal domain (KatG N ) is no longer expressed in a soluble form but is instead found in inclusion bodies. In addition, following denatured purification, refolding and reconstitution, KatG N shows a shifted active site structure, allowing what appears to be the direct coordination of the heme iron by the distal histidine (H106 by Escherichia coli numbering) [142]. Consequently, KatG N has neither catalase nor peroxidase activity. It has been suggested that the C-terminal domain might serve as a platform to direct proper folding of the N-terminal domain and facilitate dimerization [143], which is supported by our previous observation. Introducing the separately expressed and isolated C-terminal domain (KatG C ) results in restructuring of KatG N that is evident from a return of high-spin coordination states typical of wild-type enzyme as well as the restoration of catalase and peroxidase activities [144]. That the C-terminal domain is an integral component of KatG intimately tied to its structure and function is also evident from a recent thermodynamic study of KatGs showing that the N- and C-terminal domains unfold as a single unit rather than independently [145]. The I?-helix in the C-terminal domain is highly conserved and appears at the interface between the two domains. It makes contacts with several elements from the N-terminal domain of the adjacent subunit using strongly conserved Leu 690, Arg 691, Glu 695, and Tyr 697. As such, it has all the appearances of a platform by which the C-terminal domain may direct structural adjustments to the N-terminal domain thereby modulating active site conformation. To evaluate this hypothesis, we produced variants substituting alanine for these conserved residues. Single, double, and quadruple alanine substitution variants of intact KatG were expressed, isolated and characterized, and the effects of these substitutions on active site structure, stability and catalytic bifunctionality were evaluated. In addition, the analogous variants were also 57 expressed and isolated for the stand-alone C-terminal domain (KatG C ). The ability of these KatG C variants to restore KatG N to an active conformation was evaluated. Results from both sets of variants demonstrate that the I?-helix is central to the ability of the C-terminal domain to direct structural adjustments in the adjacent N-terminal domain. This supports the hypothesis that the C-terminal domain serves as a platform to direct N-terminal domain active site conformation and bifunctionality. 2.2 Materials and methods: 2.2.1 Materials: Hydrogen peroxide (30%), imidazole, hemin, ampicillin, chloramphenicol, sodium dithionite, phenylmethylsulfonyl fluoride (PMSF), 2,2?-azino-bis(3-ethylbenzthiazoline-6- sulfonate) (ABTS), and guanidine hydrochloride (GuHCl) were purchased from Sigma (St. Louis, MO). Isopropyl-?-d-thiogalactopyranoside (IPTG), urea, mono- and dibasic sodium phosphate, acetic acid, and sodium acetate were obtained from Fisher (Pittsburgh, PA). Bugbuster and benzonase were purchased from Novagen (Madison, WI). All restriction enzymes were purchased from New England Biolabs (Beverly, MA). All oligonucleotide primers were purchased from Invitrogen (Carlsbad, CA). All E. coli strains (BL-21-Gold [DE3] pLysS and XL-1 Blue), Pfu polymerase, and T4 DNA ligase were obtained from Agilent (La Jolla, CA). Nickel-nitrilotriacetic acid (Ni-NTA) resin was purchased from Qiagen (Valencia, CA). Desalting 10DG chromatography columns were purchased from Bio-Rad. All buffers and media were prepared using water purified through a Barnstead EasyPure II system (18.2 M?/cm resistivity). 58 2.2.2 Cloning: All plasmids were prepared using mutagenic coding and non-coding primers (Table 2.1) according to the Round-the-Horn procedure [152]. Amplification products were subjected to blunt-end ligation with T4 DNA ligase and used to transform E. coli (XL-1 Blue) by heat shock according to manufacturer instructions, and transformants were selected on the basis of ampicillin resistance. Plasmids from candidate colonies were isolated and evaluated by diagnostic restriction digest (Table 2.1) and DNA sequence analysis (Davis Sequencing, Davis, CA). Plasmids verified to carry the correct mutations were used to transform E. coli (BL-21- Gold [DE3] pLysS). 2.2.3 Expression and purification: All expression was carried out in E. coli (BL-21-Gold [DE3] pLysS) cells using liquid Luria?Bertani broth supplemented with ampicillin and chloramphenicol with constant agitation. Wild-type KatG was over-expressed in a soluble form as previously described [151]. However, substituting any one of the four strictly conserved residues resulted in accumulation of target protein in inclusion bodies so long as expression was carried out at 37 ?C. Upon adjusting the temperature to 18 ?C, a substantial proportion (~ 60%) of expressed protein was produced in a soluble state for all KatG C variants. Likewise, R691A, E695A, Y697A, and L690A/R691A variants of intact KatG were expressed in a soluble form at appreciable levels (40?60%) at 18 ?C. Interestingly, the L690A variant of intact KatG was only observed in inclusion bodies even at alternative expression temperatures. In all cases, expression was induced with IPTG once cells had reached mid-log phase (OD 600 = 0.4). At 10 h post-induction, cells were harvested by centrifugation, and cell pellets were stored at ? 80 ?C until purification. Expression analysis was 59 Table 2.1: Primer sets for site-directed mutagenesis of I?-helix. Protein: Primer Sequences: Restriction enzymes L690A 5?-Phos-CTGCGTGCGAGTACGC-3? 5?-Phos-TGCGACGAGTAGACAACACAGATCG-3? PmI I R691A 5?-Phos-GCGTGCGAGTTACGCAGT-3? 5?-Phos-GCAGACGAGTAGACAACACAGA-3? Sac I E695A 5?-Phos-GCAGTTACGCAGTAGCGATGCCACG-3? 5?-Phos-AGCACGCACGCAGACGAGTA-3? Pst I Y697A 5?-Phos-GAGTGCAGCAGTAGCGATGCCACG-3? 5?-Phos-AGCACGCACGCAGACGAGTA-3? N/A L690A/R691A 5?-Phos-CGCGTGCGAGTTACGCAGT-3? 5?-Phos-GCGCGACGAGTAGACAACACAG-3? Sac I E695A/Y697A 5?-Phos-GCAGTGCAGCAGTAGCGATGCCACG-3? 5?-Phos-AGCACGCACGCAGACGAGTA-3? Pst I L690/R691A/ E695A/Y697A 5?-Phos-CGTGCTGCAGTGCGCAGTAG-3? 5?-Phos-CGCGCGACGCAGTAGACAACAC-3? Sac I 60 performed using a trichloroacetic acid precipitation technique that we have described elsewhere [153]. Purification of variants expressed at 18 ?C began by resuspending the cell pellets in Bugbuster reagent (Novagen, Madison, WI) in the presence of PMSF (0.1 mM) and benzonase nuclease (250 U). The cell lysate was centrifuged, and the supernatant loaded onto a Ni-NTA column by recirculating the solution with a flow rate of 1 mL/min through the column overnight. Following loading, the column was washed in succession with buffer A (50 mM Tris, pH 8.0), buffer B (50 mM phosphate buffer, pH 7.0, 200 mM NaCl), buffer B supplemented with 2 mM imidazole, and buffer B supplemented with 20 mM imidazole. The target protein was then eluted from the column using buffer B with 50 mM imidazole, and finally, buffer B with 200 mM imidazole. Excess imidazole was then removed by gel filtration using a 10DG column from BioRad. The eluted fractions were collected and analyzed by SDS-PAGE. Protein-containing fractions were combined, concentrated, aliquotted, and stored at ? 80 ?C. Concentrations of purified proteins were estimated according to the method of Gill and von Hippel [154]. Even though no heme precursors had been added to the expression media, a small portion of the purified enzyme already had heme incorporated while inside the cells. The variants KatG N , E695A/Y697A KatG and L690A/R691A/E695A/Y697A KatG were expressed in insoluble inclusion bodies. Purification was carried out as previously described for KatG N [142] with the following modifications. After the centrifugation of the cell lysate, the insoluble pellet was resuspended using 8 M urea in buffer B. Likewise, the initial washing of the loaded Ni-NTA column was carried out using 8 M urea in buffer B supplemented with 10 mM imidazole. Elution of the protein from the Ni-NTA column was accomplished with 8 M urea in 61 buffer B supplemented with 400 mM imidazole. The eluent was then pooled and dialyzed against buffer B (24?30 h; five changes). With these modifications to the purification procedure, we obtained KatG N with a greater capacity for reactivation by KatG C than previously observed [144] as determined by visible absorption, MCD, and EPR spectroscopies as well as catalase and peroxidase activities. 2.2.4 Enzyme reconstitution and absorption spectra: Due to its poor solubility in neutral and acidic solutions, hemin was dissolved in 0.1 M KOH. The concentration of hemin stock solutions was determined by the method of Falk [155]. Hemin (0.75 equivalents) was added to wild-type KatG, intact I?-helix variants, and KatG N . These solutions were allowed to incubate for at least 24 h at 4 ?C. These solutions were then centrifuged to remove insoluble unincorporated heme and other debris. The final concentration of incorporated hemin was determined by the pyridine hemichrome method of Falk [30]. and used to calculate molar absorptivities of the heme absorption spectra of each protein. Spectra for the ferrous states of these proteins were obtained by adding a small amount (< 10 mg) solid dithionite to the ferric form of each. All spectra were obtained at room temperature on a Shimadzu UV-1601 spectrophotometer (Columbia, MD) with a cell path length of 1.0 cm. 2.2.5 Domain mixing and incubation procedures: The concentrations of KatG N , KatG C , and KatG C variants were estimated based on their molar absorptivities as follows: KatG N (?416 = 98 mM ? 1 cm ? 1 ); KatG C , L690A KatG C , R691A KatG C , E695A KatG C , and L690A/R691A KatG C (?280 = 34 mM ? 1 cm ? 1 ); and Y697A KatG C , E695A/Y697A KatG C , and L690A/R691A/E695A/Y697A KatG C (?280 = 33 mM ? 1 cm ? 1 ). 62 Solutions containing 1:1 ratios of KatG N and KatG C or one of its variants were incubated for times ranging from 0 to 96 h in the presence of 50 mM phosphate buffer, pH 7.0, 50 mM NaCl at 4 ?C. Optimal results were obtained with freshly purified and reconstituted KatG N and freshly purified KatG C . 2.2.6 Catalase and peroxidase activity assays: Peroxidase activity was evaluated by monitoring the production of the ABTS radical over time at 417 nm (34.7 mM ? 1 cm ? 1 ) [156]. All assays were carried out at room temperature in 50 mM acetate buffer, pH 5.0. Catalase activity was evaluated by monitoring the decrease in H 2 O 2 concentration with time at 240 nm (39.4 M ? 1 cm ? 1 ) [157]. All assays were carried out at room temperature in 100 mM phosphate buffer, pH 7.0. Initial velocities (v o /[E] T ) were fit to a Michaelis?Menten equation (Eq. (1)) by non-linear regression analysis to determine apparent kinetic parameters k cat , K M , and k cat /K M . !o [E]T = tS + (1) Here, [E] T is determined on the basis of the heme content of the protein(s) present. In all instances reported here, k cat is the asymptotic maximum catalytic output, the apparent K M simply corresponds to the concentration of substrate necessary to produce v o /[E] T equal to ? k cat , and k cat /K M , the apparent second-order rate constant, represents the efficiency with which the substrate is processed and corresponds to the slope of the tangent line to rates obtained at low [S] (i.e., Asp mutation in yeast cytochrome c peroxidase studied by proton NMR spectroscopy, Eur J Biochem / FEBS 224 (1994) 81-87. [189] S.F. DeLauder, J.M. Mauro, T.L. Poulos, J.C. Williams, F.P. Schwarz, Thermodynamics of hydrogen cyanide and hydrogen fluoride binding to cytochrome c peroxidase and its Asn-82-->Asp mutant, Biochem J 302 (1994) 437-442. [190] M. Tanaka, S. Nagano, K. Ishimori, I. Morishima, Hydrogen bond network in the distal site of peroxidases: spectroscopic properties of Asn70 --> Asp horseradish peroxidase mutant, Biochemistry-Us 36 (1997) 9791-9798. [191] R.L. Moore, C.O. Cook, R. Williams, D.C. Goodwin, Substitution of strictly conserved Y111 in catalase-peroxidases: Impact of remote interdomain contacts on active site structure and catalytic performance, J Inorg Biochem. 102 (2008) 1819-1824. 197 [192] J.E. Erman, L.B. Vitello, M.A. Miller, A. Shaw, K.A. Brown, J. Kraut, Histidine 52 is a critical residue for rapid formation of cytochrome c peroxidase compound I, Biochemistry-Us 32 (1993) 9798-9806. [193] N.L. Wengenack, M.P. Jensen, F. Rusnak, M.K. Stern, Mycobacterium tuberculosis KatG is a peroxynitritase, Biochem Biophys Res Commun, 256 (1999) 485-487. [194] T. Sagawa, T. Yomo, S. Tsugawa, Y. Takeda, H. Ihara, E.K. Mitamura, I. Urabe, K. Ohkubo, A novel versatility of catalase I as a dioxygenase for indole-ring-opening dioxygenation, Chem Lett (1999) 339-340. [195] P.C. Loewen (1997) Bacterial catalase. In Oxidative stress and the molecular biology of antioxidant defenses. pp 273-308, Cold Spring Harbor Laboratory Press, Plainview, N.Y [196] E. Santoni, C. Jakopitsch, C. Obinger, G. Smulevich, Comparison between catalase- peroxidase and cytochrome c peroxidase. The role of the hydrogen-bond networks for protein stability and catalysis, Biochemistry-Us 43 (2004) 5792-5802. [197] C. Jakopitsch, A. Ivancich, F. Schmuckenschlager, A. Wanasinghe, G. Poltl, P.G. Furtmuller, F. Ruker, C. Obinger, Influence of the unusual covalent adduct on the kinetics and formation of radical intermediates in synechocystis catalase peroxidase: a stopped-flow and EPR characterization of the MET275, TYR249, and ARG439 variants, J Biol Chem 279 (2004) 46082-46095. [198] S.M. Kapetanaki, X.B. Zhao, S.W. Yu, R.S. Magliozzo, J.P.M. Schelvis, Modification of the active site of Mycobacterium tuberculosis KatG after disruption of the Met-Tyr-Trp cross-linked adduct, J Inorg Biochem 101 (2007) 422-433. [199] C.L. Varnado, Enhancing expression of recombinant hemoproteins: progress toward understanding structure/function and therapeutic application (2006) Dissertation, 198 Auburn University.