recombinant expression, purification and kinetic

44
RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC CHARACTERIZATION OF F 420 -COFACTOR DEPENDENT GLUCOSE-6-PHOSPHATE DEHYDROGENASE FROM MYCOBACTERIUM TUBERCULOSIS by TIJANI OSUMAH Presented to the Faculty of the Graduate School of The University of Texas at Arlington in Partial Fulfillment of the Requirements for the Degree of MASTER OF SCIENCE IN BIOENGINEERING THE UNIVERSITY OF TEXAS AT ARLINGTON May 2013

Upload: others

Post on 27-Jan-2022

10 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC CHARACTERIZATION OF

F420-COFACTOR DEPENDENT GLUCOSE-6-PHOSPHATE DEHYDROGENASE FROM

MYCOBACTERIUM TUBERCULOSIS

by

TIJANI OSUMAH

Presented to the Faculty of the Graduate School of

The University of Texas at Arlington in Partial Fulfillment

of the Requirements

for the Degree of

MASTER OF SCIENCE IN BIOENGINEERING

THE UNIVERSITY OF TEXAS AT ARLINGTON

May 2013

Page 2: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

ii

Acknowledgements

My mother always said to put God first. So I would like to thank God first for all He has

done for me. Without Him, I would not be who or where I am at this moment.

Huge thanks go to my mother, Mrs. Zulietu Osumah, father, Dr. Aruna Osumah and

other mother, Mrs. Awawu Osumah, for their unending love, prayers and support. Also, I would

like to say a big thanks to my many brothers and sisters. I grew up watching all of you. I lived

and learned from your mistakes. I kiss the ground you all walk on. All that I am is because of

you. Thank you so much.

I would also like to extend my unending gratitude to my P.I., Dr. Kayunta Johnson-

Winters. Dr. Kay, you have been like a mother to me. I pray for our long lives, so that I can be

like a son to you.

Last but not least, a loving thanks to my adopted cousin, Adebanke Adetola and to all of

my friends who’ve been with me through it all. I look forward to all the greatness you will all

accomplish. Best wishes, love and respect.

April 18, 2013

Page 3: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

iii

Abstract

RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC CHARACTERIZATION OF F420-

COFACTOR DEPENDENT GLUCOSE-6-PHOSPHATE DEHYDROGENASE FROM

MYCOBACTERIUM TUBERCULOSIS

Tijani Osumah, MS

The University of Texas at Arlington, 2013

Supervising Professor: Kayunta Johnson-Winters

The FGD (F420-dependent Glucose-6-phosphate dehydrogenase) enzyme is an F420

Cofactor (7,8-didemethyl-8-hydroxy-5-deazariboflavin) dependent enzyme found in

Mycobacterium tuberculosis, the causative agent of tuberculosis (TB). TB is still a prominent

cause of illness and death worldwide. Because FGD is not found in humans, makes it a good

target for drug development. By understanding the hydride transfer mechanism of the FGD in

detail, we can aid in the improvement of drug targeting and development for the treatment of TB.

FGD catalyzes the conversion of glucose-6-phosphate to 6-phosophogluconolactone.

This project focuses on the purification and kinetic characterization of recombinant FGD using

steady-state and pre-steady state kinetic methods. A concurrent goal is to probe the functionality

of conserved active site residues that are involved in the hydride transfer reaction. Based upon

crystallographic data, it is believed that Histidine 40 acts as an active site base, abstracting a

proton from the substrate, glucose-6-phosphate, facilitating the hydride transfer from the

substrate to the F420 Cofactor. A separate active site amino acid, Tryptophan 44 is believed to

stabilize an active site intermediate during turnover. We have mutated these conserved

residues, making the following FGD variants, H40A, W44F and W44A. Here, we present

purification methods and steady-state kinetic analyses of both wild type FGD and the H40A

variant.

Page 4: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

iv

Table of Contents

Abstract ......................................................................................................................... iii

List of Illustrations ........................................................................................................... v

List of Tables ................................................................................................................. vi

Chapter 1 Introduction .................................................................................................... 1

1.1 Phylogenetic distribution of the F420 Cofactor ................................................. 5

1.2 Enzymes: Relevance of the cofactor through enzymes .................................14

Chapter 2 F420-Dependent Glucose-6-Phosphate Dehydrogenase (FGD) ......................16

2.1 Expression and Purification of FGD ...............................................................20

2.1.1 Expression of cell extracts ..............................................................20

2.1.2 Purification of FGD .........................................................................20

2.1.3 Site-directed mutagenesis of active site residues in FGD ...............22

2.1.4 Crystallography ..............................................................................24

Chapter 3 Results .........................................................................................................26

3.1 Purification of wtFGD and mutant variants ......................................................26

3.2 Active site mutations .......................................................................................27

3.3 Steady-state kinetics and pH profile ................................................................27

3.4 Conclusions ....................................................................................................30

References ...................................................................................................................33

Biographical Information ................................................................................................38

Page 5: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

v

List of Illustrations

Figure Page 1-1 F420-1 Cofactor and its smaller fragments (FO and F+) ........................................... 3 1-2 Spectra of the oxidized and reduced F420 Cofactor.................................................. 4 1-3 Ionization and resonance structure of F420 Cofactor ................................................ 5 1-4 Phylogenetic tree of the archaea ............................................................................ 7 1-5 Multiple sequence alignment of the FbiC protein..................................................... 10 1-6 Phylogenetic relationship between the PPOx, FDR-A, FDR-B families .................... 13 1-7 Average numbers of putative F420/FMN-binding protein family genes in actinobacterial species ....................................................................................... 14 2-1 Catalyzed reaction of FGD ...................................................................................... 17 2-2 Nitroimidazopyran pro-drug PA-824 ........................................................................ 17 2-3 Crystal structure of FGD ......................................................................................... 19 2-4 Proposed mechanism of FGD ................................................................................. 19 2-5 Schematic of protein purification procedures ........................................................... 22 2-6 Sequence chromatograms showing successful mutation ......................................... 24 2-7 Crystals formed from varying crystallization conditions of the H40A variant ................................................................................ 25 3-1 SDS-PAGE 1 .......................................................................................................... 26 3-2 SDS-PAGE 2 .......................................................................................................... 27 3-3 Steady state kinetic plot .......................................................................................... 29 3-4 pH profile of wtFGD and H40A variant in 50mM K3PO4 at 25°C .............................. 30 3-5 Distance of the histidine amino acid residues in the active site from the modeled substrate .................................................................................... 30

Page 6: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

vi

List of Tables

Table Page 1-1 Distribution of 7,8-didemethyl-8-hydroxy-5-deazariboflavin ..................................... 6

3-1 Summary of concentration and volume parameters of kinetic assays ...................... 28

3-2 Steady-state kinetic assay parameters of wtFGD and

H40A FGD mutant variant. ...................................................................................... 28

Page 7: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

1

CHAPTER 1

Introduction

The availability of an unusual electron-transfer coenzyme, known as the F420 Cofactor

(a 7,8-didemethyl-8-hydroxy-5-deazaflavin transferring agent) (Figure 1-1) has been understood

to be vital for catalysis in certain enzymes. The F420 Cofactor is chemically equivalent to NAD+,

which is exclusively involved in hydride transfer reactions (Equations 1 and 2). However, the

chemical structure is reminiscent of an isoalloxazine chromophore with a side chain comprising

of ribitol, phosphate, and lactate residues as well as a ɣ-linked polyglutamate tail that varies in

length (F420-1 to F420-9), depending upon the species in which the coenzyme is found 5.

(Equation 1)

(Equation 2)

The F420 Cofactor can be oxidized or reduced and is therefore, spectrophotometrically

distinct (Figure 1-2). In a basic solution, the Cofactor is oxidized and bright yellow in color, with

an absorbance maximum between 401-420 nm (Figure 1-2) 6-7

. The oxidized cofactor is an

obligate two-electron acceptor under physiological non-photoreductive conditions 6. Upon

reduction, it loses the 420 nm absorbance and gains a new absorbance at 320 nm 8 with a

much lower extinction coefficient (at pH 8.85, ε420 of F420 is 45,500 M-1

cm-1

and at pH 8.8, ε320

of F420 H2 is 10,800 M-1

cm-1

) 9-10

. Below pH 6.0, the maximum absorbance of the oxidized

Cofactor shifts from 420 nm towards lower wavelengths9-10

. The isosbestic point between pH 4-

10 has been previously determined to be 401 nm. However, as the pH is increased, the

hydroxyl proton at the 8 position (pKa = 6.3) 9 of the F420 Cofactor (Figure 1-3; structure A) is

removed, consequently converting the F420 Cofactor into an anionic form, which has two

resonance structures (Figure 1-3; structure B and C) 10

. These resonance structures attribute for

the shifting of the absorbance wavelength 10

. Also, at a basic pH, another proton is removed

from the 3-NH (pKa = 12.2), resulting in the formation of structure D (Figure 1-3), this also

results in absorbance shifts to longer wavelengths 10

. Similar to the chemical structure, the

Page 8: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

2

fluorescence spectral properties of the F420 Cofactor are also pH and redox sensitive 8, 10

. The

oxidized cofactor has strong fluorescence at 425 nm, while the reduced form does not. FO and

F+ (Figure 1-1), two of several acid-hydrolysis products of the F420

Cofactor, exhibit spectral

properties that are similar to those of the unaltered F420 Cofactor. However, these hydrolysis

products have an extinction coefficient at 420 nm that is approximately 15% less than that

exhibited by the F420 Cofactor. Also, FO and F+ have been observed to possess a shorter

nitrogen-10 side chain than that of the F420 Cofactor. Nonetheless, both FO and F+

compounds

are catalytically active in several F420-dependent enzymatic reactions 9-10

. Additionally, the F420

Cofactor is photosensitive, rapid decomposition occurs when it is exposed to a strong white

light, especially at basic pH 11

.

The redox potential of the F420 Cofactor is between -340 to -350 mV (much lower than

that of flavins) 12

. This lower redox potential has been attributed with several major cellular

implications. For example, instead of mediating the transfer of electrons between NAD+ and

higher potential one- and two-electron acceptors such as flavins, the F420 Cofactor has the

capacity to mediate the reduction of NAD+ with electrons from hydrogen or via formate oxidation

13. In comparison to flavins, which have accessible semiquinone, the absence of the

semiquinone or its formation has no deleterious effects on this role.

Nonetheless, the F420 Cofactor serves as a useful biochemical tool in the studies of

flavoprotein catalysis. It is stable to air and boiling at near neutral pH, degraded by light at high

pH and the side chain is cleaved in a low pH environment. It remains unknown whether, except

in the photolyase, the F420 Cofactor is ever enzyme bound. It appears to behave as a soluble

electron-transfer cofactor, but it has not been thoroughly investigated for its ability to covalently

bind to proteins 6.

Page 9: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

3

Figure 1-1 F420-1 Cofactor and its smaller fragments (FO and F+). The only difference is the length of the side chain

17

NH

NNHO

HO

OH

HO

O

O

O

P

OO

O-

O

HN

5

10

FO

F+

F420-0

COO-

OH

O

F420-1

Page 10: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

4

Figure 1-2. Spectra of the oxidized and reduced F420 Cofactor along with corresponding structures (A. oxidized F420, B. reduced F420 :H2 Cofactor). Data are from Tzeng and other

coworkers 65

-0.1

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

300 400 500 600

Ab

so

rba

nc

e

Wavelength (nm)

A) Oxidized F420 Cofactor

B) Reduced F420 Cofactor

Page 11: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

5

Figure 1-3. Ionization and resonance structure of F420 Cofactor

10: Removing a proton of the 8-

OH group (pKa = 6.3) 9 of the F420 Cofactor (Structure A) converts F420 into an anionic form, which has two resonance forms (Structure B and C)

10

1.1 Phylogenetic Distribution of the F420 Cofactor:

The infrequent presence of the F420 Cofactor and its biosynthesis in certain prokaryotes has

been utilized in comparative phylogenetic analysis and genomics-based investigation of

organisms that employ this cofactor in their metabolic pathways. The F420 Cofactor has been

previously discovered to be distributed sporadically amongst certain prokaryotic organisms, but

observed universally in mycobacteria 14

. However, before this discovery, the F420 Cofactor was

known to play an extensively significant role only in methanogenic bacteria and archaea. While

this distribution is wide-ranging, the F420 Cofactor is abundant only in methanogenic bacteria

where it is believed to function primarily as a redox carrier in energy metabolic pathways.

Although the cofactor is not unique to methanogens, it has become an important indicator for

the identification of methanogens because of its high abundance in these cells and its intense

Page 12: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

6

fluorescence, especially when compared to other species in which this cofactor is found (Table

1)14a, 15

.

Table 1-1 Distribution of 7,8-didemethyl-8-hydroxy-5-deazariboflavin 15a

Organism Content expressed in pmol/mg (dry weight)

Methanogenic archaebacteria

-Methanobacterium thermoautotrophicum 3800

-Methanospirillum hungatei 3700

-Methanobacterium formicium 2400

-Methanosarcina barkeri 190

Nonmethanogenic archaebacteria

-Halobacteria

Strain GN-1 >210

Halococcus salinarium >33

-Acidophilic archaebacteria

Thermoplasma strain 122-1B3 >5

Sulfolobus solfataricus >1

Eubacteria

Streptomyces spp. <20

Mycobacterium tuberculosis 13.5

Nocardia aurantia 8.5

Anacystus nidulans NR

In fact, the F420 Cofactor was found in all methanogens examined at levels varying from

1.2 mg/kg of cell dry weight as in the case of Methanobervibacter ruminatium to approximately

65 mg/kg of cell dry weight as seen in Methanobacterium thermoautotrophicum 16

. All

methanogenic bacteria subsequently discovered have been observed to possess genes that

Page 13: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

7

encode for the synthesis, and ultimately the utilization of the F420 Cofactor in one way or another

16-17. Because of its abundance, this deazaflavin Cofactor apparently plays a critical role in the

physiology of methanogens. Several authors have postulated that the F420 Cofactor is a

functional substitute for ferredoxin16

. This idea has been supported by the absence of ferredoxin

in several methanogens that have been studied. Also, an investigation of Methanosarcina

barkeri found that it possessed an F420 Cofactor with polyglutamate derivatives at an order of

magnitude much lower than in other methanogens, but contained ferredoxin 15a

. Most

biochemical and energetic studies have been performed with a few methanogenic genera:

Methanosarcina, Methanobacterium and Methanococcus. Since these genera are

phylogenetically distantly related (Figure 1-4), the results obtained were hence considered to be

representative for all methanogens6, 15b

.

Figure 1-4. Phylogenetic tree of the archaea indicating the phylogenetic relationship between various methanogenic genera, Archaeoglobus and Pyrococcus

6.

Also, the F420 Cofactor is found in Archaeoglobus fulgidus, where it participates in a

series of two-electron transfer reactions. A. fulgidus is the first sulfur metabolizing organism to

have its genome sequence determined. It possesses all of the enzymes and cofactor required

for methanogenesis and produces a measurable amount of methane during sulfate reduction.

Page 14: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

8

Levels of the F420 Cofactor found in A. fulgidus are similar to methanogens 6, 18

, the discovery of

this cofactor in other organisms suggests that the chemistry of the compound has been useful in

other areas of cellular metabolism.

Although the structure of the F420 was initially determined from methanogenic bacteria19

,

it was present in all Streptomyces species examined, as well as a pool of other related

organisms such as Nocardia6,15a,17,20

. It is also present in one genus of eukarya, and in some

other archaea. Analysis of Streptomyces showed that some F420 is excreted by the organism

during growth; however, it is released in the form of FO. More so, compared to methanogens,

the levels of the F420 Cofactor in Streptomyces was found to be about 10-fold lower. The role of

the F420 Cofactor in most Streptomyces is not known, but in specific Streptomyces, it is vital for

the synthesis of chlorotetracycline, oxytetracycline, and lincolnomycin, as well as DNA light

repair 17, 20-21

. Also, the biosynthesis of riboflavin in Methanobacterium thermoautotrophicum

was studied by Eisenrich et al. (1991) and was found to be identical to that in eubacteria and

fungi 15b, 22

. In the green algae Scenedesmus, the deazaflavin ring of the F420 Cofactor is

required for DNA photolyase function14a, 15a, 21

.

In 2002, the F420 Cofactor became popular as investigation into its function and

mechanism in mycobacteria amassed rapid attention. Purwartini and Daniels have shown that

the F420 Cofactor is involved in the oxidation of glucose-6-phosphate dehydrogenase by and

F420 dependent glucose-6-phosphate dehydrogenase (FGD1, Rv0407 – MTB gene). By doing

so, the F420 Cofactor is biochemically modified to its reduced form 14a

. Phylogenetic investigation

of F420 biosynthesis, using M. bovis as a model, were also studied. Choi et al (2001, 2002)

showed that fbiC gene participates in the earlier steps of F420 biosynthesis between

pyrimidinedione and hydroxyphenyl pyruvate to form FO by encoding for synthase. fbiAB genes

are responsible for the biosynthesis of F420 from the precursor, which involves the addition of a

phospho-lactate group and condensation of glutamate on . M. tuberculosis, M. bovis, M. avium,

Page 15: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

9

M. leprae, Nocardia farcinica, Streptomyces coelicolor, S. avermitilis, Thermobifida fusca and

Rubrobacter xylanophilus all have proteins with high homology for full length fbiC as shown in

multiple amino acid sequence alignment of fbiC using representative organisms (Figure 1-5). In

contrast, Archaeoglobus fulgidus, Methanobacterium thermoautotrophicum, Methanococcus

jannaschii, Halobacterium sp., Synechocystis sp., and Nostoc sp. all have two polypeptides

(located adjacent or non-adjacent) which encode for fbiC23

.

Page 16: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

10

Figure 1-5. Multiple sequence alignment of the FbiC protein in Mycobacterium sp., Norcadia sp. and Streptomyces sp.

23c.

Page 17: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

11

Selengut and Haft (2010) implemented partial phylogenetic profiling (PPP) to discover

protein families that utilize the F420 Cofactor 14b

. This method was efficient in determining these

protein families due to the possibility that the distribution of the F420 Cofactor may not span the

entire profile. By applying hidden Markov models (HMMs) to a set of all 1,451 bacterial and

archaeal genomes available from the National Center for Biotechnology Information (NCBI),

several F420-utilizing enzymes in mycobacteria based on the pattern of the F420 Cofactor

biosynthesis trait were identified. The most prominent family in the distribution was the

Luciferase-like Monooxygenase (LLM) family. A few LLM family proteins have previously been

discovered in archaea to be F420 dependent; one has even been characterized 14b, 24

.

The second most prominent family of enzymes from the partial phylogenetic profiling

was the deazaflavin-dependent nitroreductase (DDN) families. This family was determined to be

limited to only F420-producing species. Like the LLM family, the only characterized member of

this family is F420 dependent. The DDN family is known to include the deazaflavin-dependent

nitroreductase Rv3547, this protein is active in the activation of the promising antimycobacterial

prodrug, PA-824 14, 23a, 23c, 25

.

The third family discovered from this profiling was the pyridoxamine 5’-phosphate

oxidase (PPOX, or PNPOx). At the time of this discovery, there were no known F420-dependent

members of the PPOx family, however, several FMN-dependent enzymes were known. More

recently, two families of F420 dependent reductases (FDR-A and –B) which were previously

uncharacterized have been identified. These enzymes were discovered to be distantly related to

PPOXs (Figure 1-6). However, the PPOx and FDR families are functionally dissimilar.

Structurally and phylogenetically, the FDRs and PPOxs appear to have a common evolutionary

origin, yet these enzyme families have evolved to utilize different Cofactors. Irrespective of their

structural similarities, there are at least two major chemical differences between the two

Cofactors. As a result, catalyzed reaction by the two enzyme families is opposite (oxidation by

Page 18: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

12

the PNPOxs versus reduction by the FDRs). Taylor et al. suggests that this difference in

catalytic activity resulted in the evolutionary expansion of the FDRs in M. smegmatis (28 genes)

compared with the single PPOx gene. Due to the high reducing power of F420, these enzymes

have hence been allowed to catalyze the reduction of a wide variety of compounds, particularly

xenobiotics14b, 26, 27

.

These three (LLM, DDN and PPOx) families account for 32 genes in Mycobacterium

tuberculosis (Figure 1-7) and 123 genes in Mycobacterium smegmatis. Partial phylogenetic

profiling was also able to identify other members of LLM and PPOx as F420-Cofactor related,

however, it was unable to determine how many and which ones specifically amongst the

uncharacterized actually bind to the cofactor 14b

.

Page 19: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

13

Figure 1-6. Phylogenetic relationship between the PPOx, FDR-A, FDR-B families 26

.

Page 20: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

14

Figure 1-7. Average numbers of putative F420/FMN-binding protein family genes in actinobacterial species

30.

1.2 Enzymes: Relevance of Cofactor Through Enzymes

The F420 Cofactor was first found in methanogenic bacteria by Wolfe and his co-workers

31,32. It was initially thought to be a unique cofactor to methanogens

33, but was later discovered

to be present in several other organisms including Streptomyces 5,34-35

, Mycobacteria 36-37

,

Cyanobacteria 38-40

, green algae 41

and halobacteria 10

. Methanogens play a critical role in

carbon cycling because the methanogenic pathway reduces carbon dioxide to produce methane

(Equation 3).

Page 21: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

15

(Equation 3)

As mentioned previously, several F420 Cofactor dependent enzymes have been

identified and observed to require this cofactor for catalysis (Table 2) 8. The most important

component of the F420 Cofactor needed for catalytic activity is the isoalloaxine chromophore 42

(Figures 1 & 2). The consequence of the carbon substitution at the 5-position is the conversion

of the central ring from a pyrazine to a pyrimidine. This change significantly modifies the

thermodynamics and kinetics of the redox processes involved with this Cofactor. Because of

this change, 5-deazaflavins, like the F420 Cofactor, are restricted to the two electron transfer

cycles and only wield half the versatility of flavins. As a result, they are reduced more slowly

when constituted with apoflavoenzymes, but reduced rapidly when the specific oxidant is a two-

electron acceptor 43

.

The alteration of other portions of the F420 Cofactor has also been observed to have a

significant effect on F420 flavoenzyme kinetics. For example, it has been shown that changing

the length of the side chain at nitrogen 10 (as in FO and F+) alters the rates of enzymatic

reactions or the Km for the substrate, but the nature of catalysis remains the same 42

. The F420

reducing hydrogenase in cell extracts of Methanobacterium strain M.o.H has an apparent Km of

25 µM for the cofactor, however, it exhibits a higher Km of 100 µM for F+ and FO compound

33.

More so, for the F420-specific secondary alcohol dehydrogenase, the hydroxyl at Carbon 8 (C-8)

position influences the spectroscopic properties of the F420 Cofactor and the pH activity profiles

of the enzyme. Furthermore, methylation at C-8, C-7, or C-5 reduces the catalytic activity of the

F420 Cofactor, and an alteration of the ring structure inactivates the cofactor 42

.

Page 22: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

16

CHAPTER 2

F420-Dependent Glucose-6-Phosphate Dehydrogenase (FGD)

Unlike most of the previously discussed enzymes, F420 Dependent Glucose-6-

Phosphate Dehydrogenase (FGD) plays no role in methanogenensis or sulfite reduction. FGD

has been discovered in Mycobacteria and Norcadia species, but, its crystal structure has only

been solved from Mycobacterium tuberculosis (Mtb). FGD catalyzes the conversion of glucose-

6-phosphate (G-6-P) to 6-phosphogluconolactone (Figure 2-1, Equation 10). This reaction is a

vital metabolic step in survival of the species in which this enzyme is present 2,24

.

(Equation 10)

Similar to every F420 dependent enzyme, the reduction of the F420 Cofactor is achieved

by the hydride transfer from the substrate molecule (G-6-P) to the C5 position of the cofactor

ring (Figure 2-1). However, the role of the F420 Cofactor to Mtb in general has not been fully

investigated nor understood.

Nonetheless, the F420 Cofactor has been discovered to have some therapeutic

significance. PA-824 (Figure 2-2) is a nitroimidazopyran pro-drug, which was synthesized in the

efforts of combating the tuberculosis pandemic 24

. Tuberculosis remains a major cause of illness

and death worldwide (Center for Disease Control Data, 2011). It is estimated that one third of

the world’s population (approximately 2 billion people) carry this organism and that there are

about 9 million new cases per year occurring at a rate of about one per second. In 2011, there

were 1.4 million TB-related deaths worldwide.

Page 23: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

17

Figure 2-1. Catalyzed reaction of FGD coupled with the activation of anti-tubercular pro-drug PA-824

Figure 2-2. Nitroimidazopyran pro-drug PA-824

Probing the mechanistic interactions between the pro-drug and the F420 Cofactor, it was

discovered that the cofactor is indirectly responsible for the activation of the pro-drug. The

N

N

O

O

O2N

OCF3

Figure 3: Strucure of PA-824 drug

Page 24: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

18

studies showed that the reaction which FGD catalyzes produces reduced F420 to an accessory

protein (Rv3547) which in turn activates PA-824 14a,17,24,43

. The research goal of this project is to

understand the hydride transfer reaction mechanism of FGD in detail. One avenue to study this

reaction is to use steady-state and pre-steady state kinetic methods. Additionally, hydride

tunneling within FGD will be investigated using kinetic isotope effects (KIE). However, this goal

is beyond the scope of this work. More so, site-directed mutagenesis will be used in order to

understand the function of specific residues within the active site of FGD that are believed to

interact with the substrate and intermediates during catalysis.

Before any such studies can be conducted, wild type FGD (wtFGD) must first be

expressed, purified and then characterized. Therefore, the initial emphasis of this project was

the optimization of the expression and purification of wtFGD. A concurrent goal was to mutate a

conserved Histidine residue within the active site of FGD. The crystal structure of FGD from M.

tuberculosis has already been solved using multiwavelength anomalous dispersion methods.

The structure was determined to be an α/β triosephosphateisomerase (TIM) barrel homodimer

(Figure 2-3) with a molecular weight of 78 kDa. Each monomer was determined to weigh

approximately 47 kDa. Based upon this crystal structure, it was postulated that an active site

Histidine (His 40) acts as an active site base which abstracts a proton from the O1 hydroxyl of

glucose-6-phosphate (Figure 2-4). This abstraction facilitates the hydride transfer within FGD.

The reaction continues with the hydride transfer step from the C1 of G-6-P to the C5 of F420.

Finally, the electronic rearrangements in the isoalloxazine system that accompany reduction

also require protonation at N2 in the pyrimidine ring by Glu 109 24

.

To fully understand the functionality of this residue, the H40 amino acid residue was

mutated to an alanine (His40Ala or H40A). This will provide insight into the hydride transfer

reaction mechanism that occurs within FGD as well as the interaction between G-6-P and other

amino acid residues within close proximity. Both projects are consequently described.

Page 25: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

19

Figure 2-3. Crystal structure of FGD, with the F420 Cofactor (green) buried within the active site

24

Figure 2-4. Proposed mechanism of the FGD enzyme

24

Page 26: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

20

2.1 Expression and Purification of FGD

As previously stated, the initial goal of this project is to express and purify recombinant

FGD from Mycobacteria tuberculosis. Because M. tuberculosis is an air borne pathogen, it was

unfavorable to undertake our studies on the FGD enzyme from its native source. Therefore, the

gene has been subcloned into an expression vector, pBluescipt (Stratagene, Santa Clara, CA).

pBluescript is compatible with expression in BL21(DE3) E. coli cells which are non-pathogenic.

A purification protocol has been established previously with recombinant FGD from M.

smegmatis14a

. Because of the high sequence homology between FGD from M. tuberculosis and

FGD from M. smegmatis, it was reasonable to suggest that the purification protocols would be

similar. However, this was not the case. A novel and effective purification protocol for FGD was

developed for the studies outlined in this work.

2.1.1 Expression of cell extracts:

Aliquots from frozen BL21(DE3) E.coli cell stocks were plated on LB agar, containing

50 mg/mL Ampicillin. After incubation for 9-12 hours at 37°C, a single colony from the plates

was used to inoculate 500-1000 mL of Luria-Bertani (LB) broth containing 50 mg/mL Ampicillin.

The cell culture was used to inoculate 10 liters of LB broth containing 50 mg/mL Ampicillin. The

cells were grown at 37°C until an OD600 of 1.00 was attained. Next, IPTG (isopropyl β-D-1-

thiogalactopyranoside) was added to a final concentration of 0.3 mM as an inducer. The cells

were incubated at 37°C for an additional 15 hours for expression of the FGD enzyme.

2.1.2 Purification of FGD:

Unless otherwise stated, all subsequent purification procedures were undertaken at

4°C. Harvested cells were resuspended in 20 mM Tris-HCl buffer (pH 7.0) and lysed via

sonication using a Branson Sonicator (Branson Ultrasonic Corporation, Danbury, CT) three

times at 50% amplitude for 3 to 4 minutes each time, while maintaining the temperature below

10°C. The resulting cell lysate was centrifuged at 12,000 rpm for 30 minutes at 4°C.

Streptomycin sulfate (Figure 2-5) was then added to the supernatant to a final concentration

Page 27: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

21

of 1.5% w/v. The resulting sample was then stirred using a stir-bar for 10 minutes and then

centrifuged at 11,000 × g for 20 minutes.

The resulting supernatant (post-streptomycin sulfate precipitation) was subjected to

45% ammonium sulfate precipitation. The sample was centrifuged at 12,000 × g for 30 minutes,

and the 45% pellet was then re-suspended in 50 mM Tris-HCl (pH 7.0). The sample was then

dialyzed overnight to remove any excess ammonium sulfate. The desalted sample was then

loaded onto a Diethylaminoethyl(DEAE)-Cellulose (5 cm X 30 cm Anion Exchange, Sigma

Aldrich, D3764) column which had been equilibrated with 50mM Tris-HCl (pH 7.0). The protein

was eluted with an increasing NaCl gradient (0 – 500 mM NaCl in 50 mM Tris-HCl, pH 7.0).

Active fractions were confirmed via Sodium Dodecyl Sulfate – Poly-Acrylamide Gel

Electrophoresis (SDS-PAGE). They were consequently pooled, concentrated to approximate 3

mL and loaded onto a 1,6-diaminohexane (EAH) Sepharose Affinity (GE Healthcare Life

Sciences, 17-0569-03) column saturated with the F420 Cofactor. This was followed by a washing

with 300 mL of 20 mM Tris-HCl with 0.5 mM CHAPS, the protein was then eluted with an

increasing gradient from 20 mM Tris-HCl with 0.5 mM CHAPS to 600 mM NaCl – 20 mM Tris-

HCl with 0.5 mM CHAPS.

Active fractions were confirmed via SDS-PAGE. They were consequently pooled,

concentrated and loaded onto a High Sulfite (SO32-

) Cation Exchange column which had been

equilibrated with 50 mM HEPES buffer (pH 7.0). The protein was eluted with a linear gradient of

0 mM NaCl-50 mM HEPES to 600 mM NaCl-50 mM HEPES (pH 7.0).

The purified enzyme was identified with SDS-PAGE and activity assays, isolated and

stored at -80°C until it was ready for use. The enzyme concentration was measured using a

Bradford Protein Assay (Bio-Rad) by observing the wavelength shift of the Coomasie Brilliant

Blue G-250 from 465 nm to 595 nm when protein binding occurs.

Page 28: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

22

Figure 2-5. Schematic of protein purification procedures

2.1.3 Site-directed mutagenesis of active site residues in FGD:

Several variants of the FGD enzyme have been successfully created in order to probe

their functionality. Using Quikchange Site-Directed Mutagenesis (Stratagene, Santa Clara, CA),

three single point mutations within FGD (H40A, W44F and W44A) were created (Figure 2-6).

However, this project focuses on the H40A mutant. As previously mentioned, based upon

crystallographic data, it is believed that Histidine 40 acts as an active site base, abstracting a

proton from the substrate, glucose-6-phosphate, facilitating the hydride transfer from the

substrate to the F420 Cofactor (Figure 2-4) 24

. The proposed mechanism of FGD also postulates

Page 29: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

23

that following the initial proton abstraction from the O1 of G-6-P, the negatively charged reaction

intermediate which would be formed would be stabilized by interacting with Tryptophan 44

(W44) and possibly Glutamic Acid 13 (Glu13) depending on its protonation state. In order to

gain further insight into the hydride transfer mechanism of the FGD enzyme, aromatic and non-

aromatic amino acid substitutions were designed and synthesized (Figure 2-6). The W44

residue was converted to a phenylalanine (W44F) in order to investigate the role of the amino

acid in the catalyzed reaction of the FGD enzyme. The phenylalanine substitute only partially

reduces the aromaticity at the 44 position. However, the conversion of the W44 residue to an

Alanine (W44A) completely removes all aromaticity as the side chain is replaced with a neutral

Alanine residue. Studies with these mutations will perhaps grant further understanding of the

functionality of these residues as well as determining whether the aromaticity has an prominent

role in the stability of the intermediate. While the author was instrumental in the creation of

these variants, these studies are still on-going and beyond the scope of this thesis.

Page 30: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

24

Figure 2-6. Sequence chromatograms showing successful mutation and conversion of designated amino acid residues in the active site of FGD

2.1.4 Crystallography:

G-6-P was added to H40A (10 mg/mL) to a final concentration of 1 mM G-6-P (10 times

the 100 μM Km24

). The search for the initial optimal crystallization conditions was undertaken

using forty-eight crystallization conditions using a PEG/Ion Screen (Hampton Research Inc.,

Aliso Viejo, CA) via the hanging-drop (vapor diffusion) method. The crystals grew best in five of

the forty eight conditions (Figure 2-7) shown below. Each drop consisted of 2 μL of the

enzyme/substrate sample (10 mg/mL) and 2 μL of the reservoir solution. However, these

crystals still have to be analyzed to determine if they are protein crystals.

Page 31: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

25

Figure 2-7. Crystals formed from varying crystallization conditions of the H40A mutant variant. A - 0.2M Sodium fluoride, 20% w/v Polyethylene glycol 3,350; B - 0.2M Ammonium acetate, 20%

w/v Polyethylene glycol 3,350, C - 0.2M Sodium tartrate dibasic dihydrate, 20% w/v Polyethylene glycol 3,350; D - 0.2M Ammonium tartrate dibasic, 20% w/v Polyethylene glycol 3,350; E - 0.2M Potassium sodium tartrate tetrahydrate, 20% w/v Polyethylene glycol 3,350

Page 32: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

26

CHAPTER 3

Results

3.1 Purification of wtFGD and mutant variants:

FGD was purified from E. coli as a recombinant system using ammonium sulfate

fractionation and several chromatographic methods. This method was effective in the

purification of the H40A mutant which is the focus of this work.

Figure 3-1. SDS-PAGE showing purification steps and a highlighted purified wtFGD. M – Marker; 1 – Streptomycin sulfate precipitation; 2 – Ammonium sulfate precipitation; 3 – DE-52

(Anion exchange chromatography); 4 – High S (Cation Exchange Chromatography)

The ammonium sulfate precipitation, although usually very harsh on protein, deemed to

be very effective in the isolation of FGD via salting out. Earlier studies (data not shown) showed

that the 45% ammonium sulfate precipitation (Lane 2) salted out all of the FGD protein while

leaving many others behind. However, the ammonium sulfate has to be removed immediately

after this step in order to maximize enzymatic activity.

The DEAE-Cellulose (DE-52) column was also very effective in increasing separations

between electrophoretic bands as observed via SDS-PAGE (Lane 3), while leaving the over-

expressed protein intact. Because FGD was purified from a recombinant system, the enzyme

was purified apo (without the cofactor). Therefore, the F420 Cofactor was purified from

Page 33: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

27

Mycobacterium smegmatis and intercalated back into the protein. The F420-EAH Sepharose

column was vital as it was a means of merging apo-FGD with the cofactor. This step yielded a

highly purified enzyme as the affinity between FGD and the F420 Cofactor is very high (Figure 3-

1).

3.2 Active site mutations:

The design and development of mutant forms of FGD were successful as confirmed with

sequencing analysis (Sequetech DNA Sequencing Service, Mountain View, CA). The same

purification protocol that was used for wtFGD has been successful in purifying the H40A FGD

variant (Figure 3-2).

Figure 3-2. SDS-PAGE showing successful purification of wtFGD and H40A mutant variant.

Other lanes remain unchanged: Marker; 1 – Streptomycin sulfate precipitation; 2 – Ammonium sulfate precipitation; 3 – DE-52 (Anion exchange chromatography); Lane 4,5 and 6 – High S

(Cation Exchange Chromatography)

3.3 Steady-State Kinetics and pH Profile:

Following successful purification of wtFGD and the H40A FGD variant, steady-state

kinetics experiments were conducted to analyze the effect of the mutation on the catalytic

parameters of the enzyme. FGD activity was assayed by observing the decrease in the

absorbance of the oxidized F420 Cofactor at 420 nm upon reduction. The reagent volume and

Page 34: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

28

concentrations for the assay is summarized in Table 3-1. The reaction kinetics were observed to

follow a Michaelis-Menten model (Table 3-2, Figure 3-3).

Table 3-1. Summary of concentration and volume parameters of kinetic assays

Reagent

Stock Concentration

(μM)

Final Concentration

(μM) Volume (μL)

Holo-FGD 11.4 0.05 8.7

G-6-P 1.0 x 104 Varied 800

K3PO4 50 mM N/A 1190

Table 3-2. Summary of kinetic parameters and comparison of wtFGD and H40A mutant variant

Parameter wtFGD H40A

kcat (1/s) 0.77 ± 0.04 0.89 ± 0.09

Km (μM) 53.8 ± 12.9 11 ± 4

Catalytic Efficiency (1/M·s) 1.43 x 104 6.6 x10

5

Page 35: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

29

Figure 3-3. Steady-state kinetic plot and comparison of the wtFGD (Blue) and H40A (Red) variant

Page 36: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

30

3.4 Conclusion

This is the first study of a substitution of the conserved H40 residue in FGD in which

kinetic properties are reported. The wtFGD had a kcat of 0.77 ± 0.04 s-1

while H40A FGD

exhibited a turnover number of 0.89 ± 0.09s-1

(Table 3-2). Therefore, within error, these values

are similar. However, the Km was observed to decrease significantly for the H40A FGD mutant

(wtFGD Km 53 ± 12; H40A Km 11 ± 4). This means that the H40A variant has a higher affinity for

the substrate as a smaller Km (assuming Km = KD) indicates a tighter binding between the

0

50

100

150

200

250

300

350

3.00 4.00 5.00 6.00 7.00 8.00 9.00 10.00

kcat (1

/s)

pH

H40A

wtFGD

Figure 3-4. pH profile of wtFGD and H40A mutant variant in 50 mM K3PO4 at 25°C

Page 37: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

31

enzyme and substrate. Finally, the H40A variant was observed to exhibit a 45% increase in

catalytic efficiency (Table 3-2). It is possible that the mutation causes a rearrangement in the

enzyme’s conformation resulting in an increase of the mutant’s affinity for the substrate,

consequently increasing the efficiency of the enzyme. The H40A mutant was seen to be as

catalytically active as wtFGD, this was the first indication that this mutant may not serve as the

active site base for the FGD enzyme.

The pH profile of wtFGD, when juxtaposed with the H40A mutant, shows that the H40A

mutant exhibits very interesting dynamics. The wtFGD was observed to have pH optima (pH 4.5

and 8.0) while the H40A mutant was observed to have three (pH 4.5, 5.5 and 7.5). This

dissimilarity provides further insight into the catalytic behavior of the enzyme and possibly the

roles of other amino acids in close proximity to the substrate and cofactor. In general, the H40A

mutant was observed to perform optimally at lower a pH, while the wild type enzyme was

observed to perform optimally at a higher pH. It is plausible that the substitution of a neutral

Histidine at the 40th residue might affect the pKa of the acid/base catalyst in the active site

(E109) indirectly, resulting in a shift of the optimal pH. This information, along with the steady-

state data suggests that H40 is not serving as the active site base within FGD. However, while

the conserved H40 may not serve as an active site base, it is very likely that it serves other

purposes such as assisting in anchoring the substrate or possibly stabilizing an active site

intermediate, a function frequently identified with Tryptophan. This prompted further

investigation into the active site of FGD in order to locate other amino acid residues in close

proximity to the substrate which may serve as a general base to the enzyme.

Page 38: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

32

Figure 3-5. Distance of the histidine amino acid residues in the active site from the modeled substrate, citrate in the active site of the FGD enzyme

24

Catalysis depends critically on the relationship between substrate and cofactor binding,

distance and spatial orientation. Upon further investigation into the active site of FGD, the H40

residue is seen to be further away from the substrate (modeled inhibitor, citrate) with a distance

of 4.59 Å as opposed to another Histidine with closer proximity, H260, which has a distance of

3.01 Å from the substrate (Figure 3-5). H260 is a conserved amino acid residue in the active

site of FGD. A multiple sequence alignment24

of 8 selected F420 Cofactor dependent enzymes

revealed the presence of a Histidine at the 260 position in 4 of the enzymes. Further studies on

H260 will hopefully identify the active site base.

Page 39: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

33

References

1. McCormick, J., and Morton, G. O., Identity of cosynthetic factor I of Streptomyces

aureofaciens and fragment FO from coenzyme F420 of Methanobacterium species. Journal of

the American Chemical Society 1982, 104, 4014-8029.

2. Purwantini, E., Gillis, T. P., and Daniels, L., Presence of F420-dependent glucose-6-

phosphate dehydrogenase in Mycobacterium and Nocardia species, but absence from

Streptomyces and Corynebacterium species and methanogenic Archaea. FEMS Microbiology

Letters 1997, 146, 129–134.

3. Cheesman, P., Toms-Wood, A., and Wolfe, R.S., Isolation and properties of a

fluorescent compound, factor 420 , from Methanobacterium strain M.o.H.

Journal of Bacteriology 1972, 112, 527-31.

4. Zinder, S. H., Methanogenesis: Ecology, Physiology, Biochemistry & Genetics.

Chapman & Hall: New York, 1993.

5. Isabelle, D., Simpson, D., and Daniels, L., Large-scale production of coenzyme F420-5,6

by using Mycobacterium smegmatis. Applied and Environmental Microbiology 2002, 68, 5750-

5755.

6. Howland, J.L., The biochemistry of archaea (Archaebacteria). Biochemical Education

1995, 23, 582.

7. Tzeng, S., Wolfe, R., and Bryant, M., Factor 420-dependent pyridine nucleotide-linked

hydrogenase system of Methanobacterium ruminantium. Journal of Bacteriology 1975, 121,

184-275.

8. DiMarco, A. A., Bobik, T. A., and Wolfe, R. S., Unusual Coenzymes of Methanogenesis.

Annual Review of Biochemistry 1990, 59, 355-94.

9. Eirich, L., Vogels, G., and Wolfe, R., Proposed structure for coenzyme F420 from

Methanobacterium. Biochemistry 1978, 17, 4583-4676.

10. Purwantini, E., Coenzyme F420: Factors Affecting Its Purification from

Methanobacterium Thermoautotrophicum and Its Conversion to F390 and Effect of

Page 40: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

34

Temperature on the Spectral Properties of Coenzyme F420 and Related Compounds. University

of Iowa: 1991.

11. Cheeseman, P., Toms-Wood, A., and Wolfe, R., Isolation and properties of a

fluorescent compound, factor 420 , from Methanobacterium strain M.o.H. Journal of

Bacteriology 1972, 112, 527-558.

12. Jacobson, F., and Walsh, C., Properties of 7,8-didemethyl-8-hydroxy-5-deazaflavins

relevant to redox coenzyme function in methanogen metabolism. Biochemistry 1984, 23 (5),

979-988.

13. Walsh, C. T., Naturally occurring 5-deazaflavin coenzymes: biological redox roles.

Accounts of Chemical Research 1986, 19 (7), 216-221.

14. (a) Purwantini, E., and Daniels, L., Purification of a novel coenzyme F420-dependent

glucose-6-phosphate dehydrogenase from Mycobacterium smegmatis. Journal of Bacteriology

1996, 178 (10), 2861-6; (b) Selengut, J.D., Haft, D.H., Unexpected abundance of coenzyme

F(420)-dependent enzymes in Mycobacterium tuberculosis and other actinobacteria. Journal of

bacteriology 2010, 192, 5788-5798.

15. (a) DiMarco, A. A., Bobik, T. A., and Wolfe, R. S., Unusual coenzymes of

methanogenesis. Annual Review of Biochemistry 1990, 59, 355-94; (b) Ferry, J.,

Methanogenesis: Ecology, Physiology, Biochemistry & Genetics. 1993; (c) Lin, X. L., and White,

R. H., Occurrence of coenzyme F420 and its gamma-monoglutamyl derivative in

nonmethanogenic archaebacteria. Journal of Bacteriology 1986, 168 (1), 444-8.

16. Jones, W. J., Nagle, D.P., and Whitman, W.B., Methanogens and the diversity of

archaebacteria. Microbiological Reviews 1987, 51, 135-177.

17. Bashiri, G., Rehan, A. M., Greenwood, D. R., Dickson, J. M., and Baker, E. N.,

Metabolic engineering of Cofactor F420 production in Mycobacterium smegmatis. PLoS One

2010, 5 (12), e15803.

18. Klenk, H.-p., The complete genome sequence of the hyperthermophilic, sulphate-

reducing archaeon. Nature 1998, 394 (6688), 101.

Page 41: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

35

19. Eirich, L. D., Vogels, G. D., and Wolfe, R. S., Distribution of coenzyme F420 and

properties of its hydrolytic fragments. Journal of Bacteriology 1979, 140 (1), 20-7.

20. Isabelle, D., Simpson, D. R., and Daniels, L., Large-scale production of coenzyme F420-

5,6 by using Mycobacterium smegmatis. Applied and Environmental Microbiology 2002, 68 (11),

5750-5.

21. McCormick, J., and Morton, G. O., Identity of cosynthetic factor I of Streptomyces

aureofaciens and fragment FO from coenzyme F420 of Methanobacterium species. Journal of

the American Chemical Society 1982, 104, 4014-8029.

22. Eisenreich, W. B., Biosynthesis of 5-hydroxybenzimidazolylcobamid (factor III) in

Methanobacterium thermoautotrophicum. The Journal of Biological Chemistry 1991, 266,

23840-9.

23. (a) Choi, K.-p., Bair, T.B., Bae, Y.-min, and Daniels, L., Use of Transposon Tn 5367

Mutagenesis and a Nitroimidazopyran-Based Selection System To Demonstrate a Requirement

for fbiA and fbiB in Coenzyme F420 Biosynthesis by Mycobacterium bovis BCG. Journal of

Bacteriology 2001, 183 (24), 7058-66; (b) Choi, K.-p., Kendrick, N., Daniels, L., Demonstration

that fbiC Is Required by Mycobacterium bovis BCG for Coenzyme F420 and FO Biosynthesis

Demonstration that fbiC Is Required by Mycobacterium bovis BCG for Coenzyme F420 and FO

Biosynthesis. Journal of Bacteriology 2002, 184 (9), 2420–2428; (c) Rao, M. V., A Thesis

Submitted For The Degree Of Master Of Science In. 2009.

24. Bashiri, G., Squire, C. J., Moreland, N. J., and Baker, E. N., Crystal structures of F420-

dependent glucose-6-phosphate dehydrogenase FGD1 involved in the activation of the anti-

tuberculosis drug candidate PA-824 reveal the basis of coenzyme and substrate binding.

Journal of Biological Chemisty 2008, 283 (25), 17531-41.

25. Singh, R., Manjunatha, U., Boshoff, H. I., Ha, Y. H., Niyomrattanakit, P., Ledwidge, R.,

Dowd, C. S., Lee, I. Y., Kim, P., Zhang, L., Kang, S., Keller, T. H., Jiricek, J., and Barry, C. E.,

III, PA-824 Kills Nonreplicating Mycobacterium tuberculosis by Intracellular NO release. Science

2008, 322, 1392-1395.

Page 42: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

36

26. Taylor, M. C., Identification and characterization of two families of F420H2-dependent

reductases from Mycobacteria that catalyse aflatoxin degradation. Molecular Microbiology 2010,

78, 561-75.

27. White, R. H., Biosynthesis of the methanogenic Cofactors. Vitamins and Hormones

2001, 61, 299-337.

28. Grochowski, L. L., Xu, H., and White, R. H., An iron(II) dependent formamide hydrolase

catalyzes the second step in the archaeal biosynthetic pathway to riboflavin and 7,8-didemethyl-

8-hydroxy-5-deazariboflavin. Biochemistry 2009, 48 (19), 4181-8.

29. Graupner, M., Xu, H., and White, R. H., The pyrimidine nucleotide reductase step in

riboflavin and F420 biosynthesis in archaea proceeds by the eukaryotic route to riboflavin.

Journal of Bacteriology 2002, 184 (7), 1952-7.

30. Grochowski L.L., and White, R., Comprehensive Natural Products II. 2010.

31. Graupner, M., and White, R. H., Biosynthesis of the phosphodiester bond in coenzyme

F420 in the methanoarchaea. Biochemistry 2001, 40 (36), 10859-72.

32. Cheeseman, P., Toms-Wood, A., and Wolfe, R. S., Isolation and properties of a

fluorescent compound, Factor 420, from Methanobacterium strain M.o.H. Journal of

Bacteriology 1972, 112 (1), 527-31.

33. Eirich, L., Vogels, G., and Wolfe, R., Distribution of coenzyme F420 and properties of its

hydrolytic fragments. Journal of Bacteriology 1979, 140, 20-27.

34. Eker, A. P. M., Pol, A., Van der Meyden, P., and Vogels, G. D., Purification and

properties of 8-hydroxy-5-deazaflavin derivatives from Streptomyces griseus. FEMS

Microbiology Letters 1980, 8 (3), 161-165.

35. Mayerl, F., Piret, J., Kiener, A., Walsh, C., and Yasui, A., Functional expression of 8-

hydroxy-5-deazaflavin-dependent DNA photolyase from Anacystis nidulans in Streptomyces

coelicolor. Journal of Bacteriology 1990, 172, 6061-6066.

Page 43: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

37

36. Daniels, L., Bakhiet, N., and Harmon, K., Widespread Distribution of a 5-deazaflavin

Cofactor in Actinomyces and Related Bacteria. Systematic and Applied Microbiology 1985, 6

(1), 12-17.

37. Muth, E., Mörschel, E., and Klein, A., Purification and characterization of an 8-hydroxy-

5-deazaflavin-reducing hydrogenase from the archaebacterium Methanococcus voltae.

European Journal of Biochemistry 1987, 169, 571-578.

38. Eker, A., Kooiman, P., Hessels, J., and Yasui, A., DNA photoreactivating enzyme from

the cyanobacterium Anacystis nidulans. The Journal of Biological Chemistry 1990, 265, 8009-

8024.

39. Ma, K., Linder, D., Stetter, K., and Thauer, R., Purification and properties of N5,N10-

methylenetetrahydromethanopterin reductase (coenzyme F420-dependent) from the extreme

thermophile Methanopyrus kandleri. Archives of Microbiology 1991, 155, 593-1193.

40. McCormick, J., and Morton, G. O., Identity of cosynthetic factor I of Streptomyces

aureofaciens and fragment FO from coenzyme F420 of Methanobacterium species. Journal of

the American Chemical Society 1982, 104, 4014-8029.

41. Eker, A., Hessels, J., and Van de Velde, J., Photoreactivating enzyme from the green

alga Scenedesmus acutus. Evidence for the presence of two different flavin chromophores.

Biochemistry 1988, 27, 1758-3523.

42. Aufhammer, S. W., Warkentin, E., Berk, H., Shima, S., Thauer, R. K., and Ermler, U.,

Coenzyme binding in F420-dependent secondary alcohol dehydrogenase, a member of the

bacterial luciferase family. Structure 2004, 12 (3), 361-70.

43. Bashiri, G., Squire, C. J., Baker, E. N., and Moreland, N. J., Expression, purification and

crystallization of native and selenomethionine labeled Mycobacterium tuberculosis FGD1

(Rv0407) using a Mycobacterium smegmatis expression system. Protein Expression and

Purification 2007, 54 (1), 38-44

Page 44: RECOMBINANT EXPRESSION, PURIFICATION AND KINETIC

38

Biographical Information

Born and raised in Lagos, Nigeria, Tijani Osumah moved to Texas, USA on the 8th of

January, 2009 at the age of 16 to start his freshman year at the University of Texas at Arlington.

Since 2010, Tijani has worked under the supervision of Dr. Kayunta Johnson-Winters in

the Department of Chemistry and Biochemistry where he has researched on the recombinant

expression, purification and kinetic characterization of F420-dependent Glucose-6-phosphate

Dehydrogenase (FGD) enzyme from M. tuberculosis.

Tijani has received many awards throughout his career at the University of Texas at

Arlington including a University Scholar Award, John T. Murchison Award for Outstanding

Junior/ Pre-Health Professional student in the Department of Chemistry and Biochemistry and

the Igor Fraiberg Endowed Scholarship in Engineering.

Tijani will begin medical school with the September 2013 class at Ross University

School of Medicine, Portsmouth, Dominica.