A CHEMICAL APPROACH TO DISTINGUISH ATP-DEPENDENT

By

JENNIFER FISHOVITZ

Submitted in partial fulfillment of the requirements

For the degree of Doctor of Philosophy

Dissertation Advisor: Dr. Irene Lee

Department of Chemistry

CASE WESTERN RESERVE UNIVERSITY

January, 2011

CASE WESTERN RESERVE UNIVERSITY

SCHOOL OF GRADUATE STUDIES              Jennifer Fishovitz       PhD      Mary Barkley           Irene Lee   Thomas Gerken   Anthony Berdis   Michael Zagorski        11/19/10                          TABLE OF CONTENTS

Title page 1

Committee Sign-off sheet 2

Table of Contents 3

List of Tables 5

List of Schemes 6

List of Figures 7

Acknowledgments 10

List of Abbreviations 11

Abstract 15

CHAPTER 1: INTRODUCTION 17

CHAPTER 2: COMPARISON OF INHIBITION OF BACTERIAL AND

HUMAN LON ACTIVITY 32

Introduction 33

Materials and Methods 35

Results and Discussion 42

Conclusions 56

CHAPTER 3: CHEMICAL TOOLS TO MONITOR ATP-DEPENDENT PEPTIDASE

ACTIVITY IN ISOLATED MITOCHONDRIA 57

Introduction 58

Materials and Methods 61 3

Results and Discussion 66

Conclusion 78

CHAPTER 4: DESIGN AND CHARACTERIZATION OF ClpXP SPECIFIC

N-TERMINAL BORONIC ACID INHIBITOR 79

Introduction 80

Materials and Methods 84

Results and Discussions 91

Conclusions 99

CHAPTER 5: CONCLUSIONS AND FUTURE DIRECTIONS 100

Introduction 101

Materials and Methods 103

Results and Discussion 106

Current status and future directions 116

BIBLIOGRAPHY 119

4

LIST OF TABLES

CHAPTER 2

2.1: Kinetic parameters for ADP inhibition of peptide cleavage by human Lon 43

2.2: Kinetic parameters of inhibition of human Lon peptidase activity by DBN93

as determined by global fitting of experimental data using DynaFit 52

CHAPTER 5

5.1: -specific peptide substrates 117

5

LIST OF SCHEMES

Scheme 2.1: Noncompetitive inhibition 44

Scheme 2.2: One-and two-step inhibition mechanisms 50

6

LIST OF FIGURES

CHAPTER 1

1.1: Normal and Lon-deficient mitochondria 20

1.2: Cleavage of Insulin B chain by Lon and ClpXP 21

1.3: Domain layout of Lon protease monomer 23

1.4: Structures of Lon and ClpXP and simplified mechanism of proteolysis 24

1.5: Proposed mechanism for peptide hydrolysis by Lon 25

1.6: Continuous fluorescent peptidase assay used to monitor enzyme activity 27

CHAPTER 2

2.1: Structures of peptide-based substrates and inhibitors used in these

experiments 34

2.2: Phosphate does not inhibit the ATPase activity of hLon 47

2.3: Pre-steady state ATP hydrolysis by human Lon 48

2.4: Time-dependent inhibition of human Lon peptidase activity by DBN93 50

2.5: Global fitting of peptide cleavage time courses in the presence of varying

amounts of inhibitor 51

2.6: DBN93 inhibits degradation by human Lon 54

2.7: DBN93 does not significantly inhibit hLon ATPase activity in the presence

of λN protein stimulation 56

7

CHAPTER 3

3.1: Peptide-based substrates are used to monitor activity of human Lon

and human ClpXP 68

3.2: DBN93 does not inhibit human ClpXP peptidase activity 70

3.3: Identification of mitochondrial matrix isolated from HeLa cells

by Western Blot analysis 72

3.4: ATP-dependent peptide cleavage of FRETN 89-98 by isolated mitochondria

is abolished by the addition of Lon specific inhibitor, DBN93 73

3.5: Immunodepletion of Lon from isolated mitochondria abolishes

ATP-dependent peptide cleavage of FRETN 89-98 75

3.6: Chase experiments show CBN93 inhibition of StAR degradation in isolated

mitochondria 77

CHAPTER 4

4.1: Complimentary peptide boronic acid strategies utilizing the P/S and

P’/S’ sites 83

4.2: Purified samples of ClpX and ClpP 87

4.3: N-terminal peptidyl boronic acids synthesized by the Santos group and

screened for ClpXP specific inhibitors 89

4.4: DBN93 inhibits casein degradation by human Lon but not ClpXP 93

4.5: Screening of N-terminal peptidic boronic acids using FRET 89-98

peptidase assay 95 8

4.6: WLS6a IC50 determination by fluorescent peptidase assay 96

4.7: Inhibition of ClpXP degradation of casein by WLS6a 98

4.8: WLS6a does not inhibit ClpXP ATPase activity 99

CHAPTER 5

5.1: Imidazole does not affect the ATPase activity of ClpX 108

5.2: Rate of ATPase activity by ClpX is not significantly affected by the

amount of ClpP 109

5.3: Determination of kinetic parameters for ATP hydrolysis by ClpXP 110

5.4: Screening FRETN 89-98 alanine scan peptides for cleavage by ClpXP,

human Lon, and E. coli Lon 113

5.5: Fluorescently labeled Cleptide is selectively cleaved by ClpXP in the

presence of ATP 114

5.6: Cleavage of Cleptide by ClpXP can be monitored using the fluorescent

peptidase assay 115

9

ACKNOWLEDGMENTS

I would first like to thank my advisor, Dr. Irene Lee, for her patience, support and open door. She was a great teacher and always available when I needed help or a boost of confidence.

I would be lost without the support of past and current lab members: Diana, Jessi and Hilary introduced me to “great days in the Lee lab” and convinced me that I was going to love Lon whether I wanted to or not. Jason Hudak worked as an undergraduate in our lab and spent a lot of time and energy cloning ClpXP for me to use. Many thanks go to Edward Motea for always making me laugh and providing the lab with great (paraphrased) music. Sandra Craig took me under her wing and taught me the intricacies of cell culture, thank you for your help and for assuring me the cells were tougher than I thought they were, at least until they weren’t. James Becker and Kristin synthesized a number of the peptides I used in my studies and Kristin synthesized boronic acid peptide. To the “newbies”, Iteen and Natalie, thanks for making sure I stopped at some point during the day to eat lunch. Now it’s your turn to love Lon like the rest of us.

Special thanks to Dr. Anthony Berdis for numerous instances of help over the years. Collaborations with Dr. Carolyn Suzuki at UMDNJ, Dr. Webster Santos and Ken Knott at Virginia Tech, and Anton Simeonov and Tim Foley at NIH have been invaluable and are much appreciated.

Last, but not least, I want to thank my family and friends, for their unwavering support and encouragement. Now when you ask me if my thesis is done, I can finally answer, “Yes.”

10

LIST OF ABBREVIATIONS

λN Lambda N protein: a λ phage protein that allows E. coli RNA polymerase to transcribe through termination signals in the early operons of the phage

AAA+ ATPases Associated with a variety of cellular Activities

Abu Aminobutyric acid

Abz Anthranilamide

ACN Acetonitrile

ADP Adenosine diphosphate

Aloc Allyloxycarbonyl

ATP Adenosine triphosphate

BME Beta-mercaptoethanol

Boc tert-Butyloxycarbonyl

BSA Bovine Serum Albumin

Bz Benzoic acid amide

Cam Chloramphenicol

Cbz Carboxybenzyl

CBN93 Non-fluorescent C-terminal boronic acid,

Cbz-YRGIT-Abu-B(OH)2

Cleptide Y(NO2)-FAPHMALVPV-K(Abz)

dansyl 5-(dimethylamino)naphthalene-1-sulfonyl chloride

DBN93 fluorescent C-terminal boronic acid,

dansyl-YRGIT-Abu-B(OH)2

DE52 diethylaminoethyl cellulose anion exchange resin used in purification of Lon 11

DMEM Dulbecco's Modified Eagle Medium

DMSO dimethyl sulfoxide

DTT dithiothreitol

E. coli Escherichia coli, a gram negative bacteria

EDTA Ethylenediaminetetraacetic acid eLon E. coli Lon protease

FBS Fetal Bovine Serum

Fmoc Fluorenylmethyloxycarbonyl

FRET Fluorescence Resonance Energy Transfer

FRETN 89-98 Y(NO2)-YRGITCSGRQ-K(Abz)

FRETN 89-98Abu Y(NO2)-YRGIT-Abu-SGRQ-K(Abz)

HATU (2-(7-Aza-1H-benzotriazole-1-yl)-1,1,3,3-tetramethyluronium hexafluorophosphate)

hClpXP Human ClpXP protease

HeLa Immortal cervical cancer cells widely used in scientific research

HEPES (4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid)

hLon Human Lon protease

IPTG Isopropyl-β-D-thio-galactoside

Kburst Burst rate constant for ATP hydrolysis

kcat Vmax/[E]

kcat/Km Substrate specificity constant

Ki Inhibition constant

Km Michaelis constant equal to [substrate] required to reach ½ Vmax

kobs rate/[E]

Kan kanamycin 12

KPi Potassium phosphate

LB Lauria-Bertani medium

Mg(OAc)2 Magnesium actetate

Mito Mitochondria isolated from HeLa cells

MOPS 3-(N-morpholino)propanesulfonic acid

N Hill coefficient

NaCl Sodium chloride

NaPi Sodium phosphate

Ni-NTA agarose Nickel-nitrilotriacetic acid agarose: resin used to purify 6xHis- tagged proteins

NO2 Nitro

P11 Phosphocellulose cation exchange resin used to purify Lon

PCR Polymerase chain reaction

PEI-cellulose Polyethyleneimine-cellulose

pen/strep Penicillin/Streptomycin

Pi Inorganic phosphate

S. Typhimurium Salmonella enterica serovar Typhimurium

SB Super Broth

SDS Sodium Dodecyl Sulfate

SDS-PAGE Sodium Dodecyl Sulfate Poly-Acrylamide Gel Electrophoresis

SSD Sensor and Substrate Discrimination

StAR Steroidogenic Acute Regulatory protein

TBST Tris-buffered saline containing 0.05% Tween 20

TFA trifluoroacetic acid

Tris tris(hydroxymethyl)aminomethane 13

UV ultraviolet v rate

14

A Chemical Approach to Distinguish ATP-dependent Proteases

Abstract

By

JENNIFER FISHOVITZ

Mammalian Lon and ClpXP are two ATP-dependent proteases that are found in the mitochondrial matrix and have been implicated in protecting the mitochondria against damage from oxidative stress. Our lab is interested in understanding the role of

Lon and ClpXP in the regulation of levels of oxidatively damaged proteins. There has been shown to be a transient increase in ATP-dependent proteolysis activity following oxidative stress, but there has been no functional assay to distinguish between the activities of these two . Previously, genetic knock-down studies have been employed with similar proteases, but can lead to changes in cellular metabolism that obscure the accurate detection of substrates. Because Lon and ClpXP have been shown to have different peptide cleavage specificities in degrading protein substrates, I have developed a series of chemical tools which allow for the profiling these enzymes individually at a post-translational level. I now have enzyme-specific peptide substrates and enzyme-specific inhibitors that can be used to inactivate one protease while analyzing the activity of the other in order to identify their respective role in maintenance of mitochondrial function in healthy and oxidatively damaged tissue. These compounds are used in isolated mitochondria to monitor the activity of the individual protease. The information gathered in these studies will lead to the development of potent inhibitors of

15

human Lon and ClpXP, specific peptide substrates, and a method to identify physiologically relevant protein substrates of each enzyme in normal and damaged

mitochondria.

16

CHAPTER 1

INTRODUCTION

17

The mitochondria are known as the powerhouse of mammalian cells because they generate adenosine triphosphate (ATP) which provides chemical energy to carry out a number of processes to maintain the health of the overall cell. ATP is generated by the electron transport chain in the mitochondrial matrix which also produces a large amount of reactive oxygen species (ROS). ROS can cause damage to proteins within the mitochondria and accumulation of ROS and damaged proteins leads to mitochondrial dysfunction and cell death (1, 2).

Mitochondrial dysfunction has been implicated in aging and a number of diseases such as Parkinson’s disease and Alzheimer’s disease (3-7). Therefore, the damaged proteins must be removed, a task that is attributed to proteases within the mitochondria (8, 9).

Lon and ClpXP are two ATP-dependent proteases that are nuclear encoded and contain mitochondrial targeting sequences (10-12) to localize in the mitochondrial matrix (13-15) that have been proposed to remove oxidatively damaged proteins to maintain the integrity of the mitochondria (16-23). As shown in Figure 1.1, in yeast mitochondria that is deficient of Lon, large protein aggregate accumulate and damage the mitochondria compared to normal mitochondria. Recently, it has been shown that ATP-dependent proteolytic activity can be up-regulated by cellular oxidative stress, especially during cardiac ischemia/reperfusion (24, 25). However, this increase in ATP-dependent proteolytic activity has thus far been attributed to Lon without accounting for the contribution of ClpXP. As both Lon and ClpXP are localized in the mitochondrial matrix and have similar functions, it is reasonable to believe ClpXP also functions

18

to degrade oxidatively damaged proteins, most likely with substrate specificity that is different from that of Lon. Indeed, it has already been shown that although

Lon and ClpXP degrade Insulin B chain at two identical sites, each enzyme also cleaves at unique sites (Figure 1.2) (14, 26), indicating that differences in substrate specificity can be exploited to design selective peptide substrates and inhibitors to monitor the activity of the individual enzymes. Herein, I discuss the first quantitative, selective study of Lon activity in the presence of ClpXP from isolated mitochondria and live cells using chemical tools developed in our lab to study the mechanism of the bacterial homolog of Lon.

Structure and function of Lon and ClpXP

Lon and ClpXP are categorized as members of the AAA+ (ATPases associated with a variety of cellular activities) family of proteins. This large family of proteins is involved in DNA replication, transcription, membrane fusion, and proteolysis (27-30). A common feature of this is an ATPase module consisting of the Walker A and Walker B motifs as well as a proposed Sensor and Substrate Discrimination (SSD) domain that interacts with protein substrates

(31). Both Lon and ClpXP share these characteristics. The domain layout of Lon is shown in Figure 1.3, containing an N domain which has been proposed to play a role in substrate recognition (32), the ATPase module for ATP hydrolysis and a

P domain where the Ser-Lys catalytic dyad is located. Like other AAA+ proteins,

Lon and ClpXP oligomerize into ring-like structures, forming a central cavity through which protein substrates are translocated to the proteolytic site for

19

Normal mitochondria Lon-deficient mitochondria

Figure 1.1: Normal and Lon-deficient mitochondria. Adapted from (33).

20

Figure 1.2: Cleavage of Insulin B chain by Lon and ClpXP. Sites of cleavage by each enzyme are indicated by arrows. Although Lon and ClpXP share some cleavage sites, their unique cleavage sites can be exploited to design enzyme- specific peptide substrates and inhibitors.

21

degradation (34), however, their oligomeric structures differ from one another.

While Lon contains its ATPase and proteolytic domains on a single polypeptide

monomer which requires Mg+, but not ATP to oligomerize into a hexamer (35-

37), the functional ClpXP oligomeric form is comprised of a ClpX ATPase and

ClpP proteolytic subunits and requires binding of ATP to complex as two ClpP

heptamers which are flanked on one or both ends by ClpX hexamers (14, 38).

Figure 1.4 shows the structures and simplified mechanism of protein degradation

by Lon and ClpXP. In the presence of ATP, a conformational change takes

places to allow for unfolding and translocation of the protein substrate through

the central cavity to the proteolytic site where it is degraded into small peptide

products. Unfolding and translocation of the protein have been proposed to be

the rate-limiting step of the reaction (39). The proposed mechanism for peptide

bond cleavage by Lon is shown in Figure 1.5, with the Ser acting as a

nucleophile to attack the carbonyl carbon at the scissile site to create a

tetrahedral intermediate, leading to product release. While the coordination of

ATPase and proteolytic activities are not completely clear in the mechanism of

human Lon and ClpXP, the individual activities can be monitored by experiments

designed in our lab that have been previously used to elucidate the mechanism

of bacterial Lon.

22

+ - H3N - -CO2

Figure 1.3: Domain layout of Lon protease monomer. The N domain is

implicated in substrate recognition. The AAA+ module where ATP hydrolysis takes place is conserved in members of the AAA+ family of proteins. The P

domain contains the Ser-Lys catalytic dyad where peptide bond cleavage occurs.

23

Lon ClpXP

Protein substrate

ClpX: ATPase

ATPase ATP subunit domain ClpP: protease

subunit protease ADP domain + Pi

Peptide products

Figure 1.4: Structures of Lon and ClpXP and simplified mechanism of proteolysis. In addition to ATP hydrolysis, there is unfolding and translocation of the protein substrate through the central cavity to the proteolytic site where it is degraded into small peptide products.

24

C-terminal product

Acyl-enzyme N-terminal product intermediate O O-

N CH C N CH C OH H H

H O H O Ser O Ser H H H H2N + Lys H2N Lys

Figure 1.5: Proposed mechanism for peptide hydrolysis by Lon. The active site

Ser of Lon acts as a nucleophile, attacking the carbonyl of the scissile bond. This results in a tetrahedral intermediate and upon collapse, the C-terminal product is released. A water molecule attacks the carbonyl of the acyl-enzyme intermediate,

creating a second tetrahedral intermediate. The collapse of this intermediate

leads to the release of the N-terminal product and regeneration of the active site.

25

Use of previously developed chemical tools to monitor protease activity

Although in the mitochondria Lon and ClpXP target folded proteins, the unfolding and subsequent degradation present a number of complex factors that would have to be taken into account in studying their removal by these proteases. To demonstrate proof of principle and produce an initial mechanism of action, we routinely use unstructured proteins such as the bacterial Lon substrate

λN (26), or small peptides to monitor the ATPase and protease activities of Lon.

Our lab has developed a fluorescent peptide substrate denoted as FRETN 89-98 to be used in a continuous peptidase assay. This 11-mer peptide was derived from the sequence of the λN protein (26) that contains an anthranilamide donor and 3-nitrotyrosine quencher and one cleavage site for Lon (40). Upon cleavage by Lon protease in the presence of ATP hydrolysis (Figure 1.6), the peptide separates into two pieces, and shows an increase in fluorescence as the quencher is separated from the fluorophore. Protease activity is measured by monitoring fluorescence emission over time. The fluorescent trace contains a short lag phase, followed by a linear phase, then a plateau of fluorescence indicating substrate depletion. The slope of the linear phase corresponds to the rate of peptide degradation which can be converted to observed rate constants for comparative studies. This peptide design and assay provides a method to modify the sequence of the peptide and add various fluorophores and quenchers to identify specific substrates of Lon and ClpXP.

26

FRETN 89-98

+ - H N-Y(NO )-RGITC-SGRQ-K(Abz)-CO2 3 2

λex=320nm, λem=420nm

+ATP

Fluorescence -ATP Enzyme, ATP

Time

+H3N-Y(NO2)-RGITC-

+ -SGRQ-K(Abz)-CO2-

Figure 1.6: Continuous fluorescent peptidase assay used to monitor enzyme activity.

27

These studies of E. coli and S. Typhimurium Lon provide a starting point for the study of human Lon. Bacterial and human Lon have long been thought to be similar in their mechanisms as E. coli Lon can substitute for the function of mitochondrial Lon in yeast (41). However, in the following chapters, I describe experiments that compare the activity of human and bacterial Lon homologs and identify differences that illustrate the need for a detailed mechanistic study of human Lon. In Chapter 2, I examine the effect of ADP levels on the peptidase activity of human Lon and compare the resulting inhibition constants to those determined for E. coli Lon. I also look at inhibition of human Lon by the peptidyl boronic acid identified as a potent inhibitor of S. Typhimurium Lon.

Bacterial Lon has been studied by our lab in an effort to elucidate the mechanism of action and determine the rate-limiting step of the reaction. The results from these experiments are used to provide a starting point for the study of human Lon as it has been shown to have overlapping substrate specificity to bacterial Lon (42). FRETN 89-98 has been used to probe the activity of E. coli

Lon by monitoring pre-steady state peptide cleavage via stopped-flow fluorescence spectroscopy and pre-steady state ATP hydrolysis by chemical- quench-flow. In the time course of peptide cleavage, there is a lag prior to steady-state peptide hydrolysis (43, 44) which was attributed to the existence of a step prior to peptide cleavage that is not able to be detected using fluorescence techniques. As it had already been established that peptide cleavage is enhanced by ATP hydrolysis (40, 43) and Lon contains both ATPase and protease activities, the ATP hydrolysis activity on a pre-steady state time scale 28

was monitored under identical conditions using a chemical-acid-quench assay.

During the time frame in which there is a lag in peptide cleavage, a burst in ADP production was seen (44), suggesting that ATP hydrolysis occurs prior to peptide cleavage and the rate limiting step occurs after ATP hydrolysis, possibly the release of ADP from the active site of Lon. Steady state inhibition constants for

ADP were determined using the fluorescent peptidase assay. ADP was shown to inhibit peptide cleavage of FRETN 89-98 with inhibition constants in the low micromolar range, indicating that ADP binds tightly to the active site of Lon and

ADP release could be the rate-limiting step in the mechanism of E. coli Lon.

Although it is implicated that human Lon protects the mitochondria from damage, it has been demonstrated that Lon deficient S. Typhimurium showed protection against virulence (45). This implicates S. Typhimurium Lon as a possible therapeutic target. As Lon belongs to the family that includes the 26S proteasome, a starting point for identification of potential inhibitors of Lon activity

was those compounds that have already been identified as inhibitors of the

proteasome. Our lab has screened a number of proteasome inhibitors against S.

Typhimurium Lon. Using the fluorescent peptide substrate FRETN 89-98 and a

continuous spectroscopic assay, IC50 values for seven proteasome inhibitors

were determined (42). Of these compounds, two were identified as potent Lon

inhibitors, MG132 and MG262, though both are ~2000-fold more potent against

the proteasome. These two compounds have an identical peptide sequence,

differing only in the functional moiety located at the C-terminus. MG132 contains

an aldehyde group while MG262 contained a boronate moiety. The IC50 of

29

MG262 is ~40-fold less than that of MG132, suggesting that the boronate moiety is key in the inhibition of Lon. Other compounds containing the boronate moiety were tested for inhibition of Lon, but had much higher IC50 values. The fact that

the boronic acid moiety alone is not enough to inhibit Lon in at least the low-

micromolar range gives credence to the need for a peptide moiety that binds to

the enzyme in addition to the existence of boronic acid to interact with the active site. To this end, a peptidyl boronic acid compound, DBN93 (dansyl-YRGIT-Abu-

B(OH)2, was synthesized that contains the sequence of a product of peptide

hydrolysis by Lon and a boronic acid moiety. The inhibition of S. Typhimurium

Lon by this compound has been studied in detail and identify it as a potent

inhibitor of proteolytic activity (46).

As the activity of Lon differs among the bacterial and human homologs, it

is reasonable to assume that human Lon and ClpXP are also different in their

substrate specificity and mechanism of action although they are both found in the

mitochondrial matrix and have similar functions. In Chapter 3, I discuss the

identification of an enzyme-specific peptide substrate and the use of the peptidyl

boronic acid inhibitor to selectively monitor the activity of Lon in isolated

mitochondria and live cells. The use of isolated mitochondria to monitor ATP-

dependent proteolytic activity has been well established (47), but here I describe the first quantitative detection of Lon activity in mitochondria and method to inactivate Lon in the cell and monitor the effects of oxidative stress. Chapter 4 described the development and characterization of a ClpXP specific boronic acid inhibitor that can be used in the future to inactivate ClpXP to monitor changes in

30

mitochondrial integrity and identify protein targets. Finally, Chapter 5 describes a

number of preliminary studies that build on the results of the previous chapters to develop a peptide substrate that would be used to specifically monitor ClpXP activity as well as future directions of the project.

31

CHAPTER 2

COMPARISON OF INHIBITION OF BACTERIAL AND HUMAN LON

PROTEASE ACTIVITY

32

Lon preferentially degrades damaged or misfolded proteins at its

proteolytic site while ATP is bound and hydrolyzed into ADP at its ATPase site.

The ATPase function of the protease is assumed to be modulated by ATP/ADP

levels. An excess of ADP would inhibit continued hydrolysis of ATP, thereby

inhibiting protein degradation. The effect of ADP on proteolytic activity of E. coli

Lon has been studied by our lab, producing inhibition constants in the low

micromolar range (43). Although bacterial and human Lon have similiarities,

including the steady-state kinetic parameters of ATPase activity(42), our lab has

discovered differences in their protease activity (42). To further uncover the

mechanism of action by human Lon in an effort to exploit the differences between the analogs for drug design and protein targets, I have tested the effect of ADP on human Lon peptidase activity using the fluorescent assay developed by Lee and Berdis (40) and our lead peptide substrate, FRETN 89-98 (Figure 2.1).

From the peptide sequence , our lab has developed a peptidyl boronic acid (DBN93, Figure 2.1) to target the proteolytic active site of Lon, which has been shown to inhibit the bacterial enzyme homolog (46). Because the peptide from which this inhibitor was derived is a substrate of both bacterial and human

Lon, we predict that it would also be a potent inhibitor of the human enzyme. In this chapter, I discuss the inhibition of human Lon by DBN93 and the determination of kinetic parameters from global fitting of experimental data. I compare these results to those published for bacterial Lon and discuss the implications for future inhibitor design.

33

Figure 2.1: Structures of peptide-based substrates and inhibitors used in these experiments. FRETN 89-98 is derived from the sequence of the endogenous

Lon substrate, λN protein. FRETN 89-98Abu contains an Aminobutyric acid substitution for Cys. The peptidyl boronic acid inhibitor, DBN93, was derived from a product of FRETN 89-98 peptide hydrolysis by Lon with a boronic acid moiety on the carboxyl end to interact with the active site of the protease.

34

Materials and Methods

Materials. Tris, IPTG, chromatography media, DTT, Mg(OAc)2, trypsin, kanamycin, chloramphenicol, ATP, dansyl chloride, DMSO, Tween 20, and all other materials were purchased from Fisher, Sigma, and Amresco.

General methods. All reactions conditions are listed as final concentrations.

Enzyme concentrations are reported as monomer concentration as quantified by

Bradford Assay (48) or absorbance at 280 nm using the molar extinction coefficient (49). Peptides were quantified by extinction coefficient at A280. In order to correct for the inner filter effect, at concentrations of FRETN 89-98 above 100

µM, the peptide used is a mixture of 10% FRETN 89-98 and 90% a non- fluorescent analogue containing a tyrosine in place of the nitro-tyrosine and benzoic acid on the lysine in place of the anthranilamide. All reactions were run at 37 ºC unless otherwise stated.

Expression and purification of human Lon protease. Human Lon was expressed and purified as previously described (42) with the following modifications. Human Lon expressed in Rosetta (DE3) cells were grown at 37 ºC in superbroth (SB, per L: 30 g tryptone, 20 g yeast extract, 10 g MOPS pH 7.5) containing 30 µg/mL kanamycin and 34 µg/mL chloramphenicol until they reached an OD600 of 1.0 at which they were induced with 1 mM IPTG for 1 hr at

37 ºC. After induction, cells were harvested at 3000 x g at 4 ºC. Pelleted cells

35

were combined and resuspended in 50 mM KPi lysis buffer (all buffers contain 5 mM BME, 20% glycerol, and 0.01% Tween 20 unless otherwise stated) and lysed in a Dounce homogenizer three times. For complete lysis, cells were sonicated for 5 min in 15 sec pulses at 100 V. Cell lysate was cleared by centrifugation at 20000 x g for 2 hr at 4 ºC. Cleared lysate was immediately loaded onto a P11 column equilibrated in lysis buffer and the flow through was collected. The column was then washed with 0.1 M KPi wash buffer until protein

was no longer coming off the column. Finally, Lon was eluted with a linear

gradient of 0.1 M KPi to 0.5 M KPi buffer, collected in 20 mL fractions. Fractions

were tested for protein content by Bradford dye and positive fractions were

analyzed by SDS-PAGE. Fractions containing Lon were combined and diluted to

a final KPi concentration of 110 mM then loaded onto a DE52 column

equilibrated in 110 mM KPi buffer. Flow through of the load was collected and the protein was eluted with 120 mM KPi buffer. The load was concentrated using

Amicon YM-30 MWCO membrane. Load and elution fractions were analyzed by

SDS-PAGE and Lon positive fractions were combined and concentrated to ~6

mL. Protein was loaded onto a Sepharose S-300 gel filtration column equilibrated

in hLon storage buffer (50 mM HEPES, 75 mM KPi pH 7, 5 mM DTT, 1 mM

Mg(OAc)2, 150 mM NaCl, 20% glycerol, 0.01% Tween 20) and eluted with buffer.

Fractions were analyzed by SDS-PAGE and Lon positive fractions were

combined, concentrated, quantified, aliquoted, and stored at -80 ºC.

36

Peptide Synthesis. Synthesis of FRETN 89-98, FRETN 89-98Abu, and DBN93

(Figure 2.1) were performed as previously described (43, 46).

ADP inhibition of human Lon peptidase activity. Reactions containing 50 mM

HEPES (pH 8.0), 5 mM Mg(OAc)2, 2 mM DTT, 300 nM hLon and varying amounts of FRETN 89-98 and ADP were initiated by the addition of 50 µM ATP.

Peptide cleavage was monitored at 420 nm (excitation = 320 nm) on a

FluoroMax-3 fluorometer (Horiba Group) at 37 °C. The rate of peptide cleavage was determined by the slope of a line tangent to the linear phase of the time course and normalized by the rate of complete peptide cleavage by trypsin.

Observed rate constants were determined by dividing by the concentration of enzyme. Kinetic parameters were determined by global fitting of the data using the program EnzFitter (BioSoft) and equation 2.1 for non-competitive inhibition

(50).

n kcat  S kobs  Equation 2.1  I  n  I  K' 1   S 1   Kis   Kii 

Where kobs is the observed rate constant for peptide cleavage, kcat is the maximum rate constant, S is peptide substrate concentration, n is the Hill coefficient, and K’ is the observed Michaelis constant for the peptide substrate. I is the inhibitor concentration, and Kis and Kii are the dissociation constants for

37

binding of inhibitor to free enzyme and enzyme:substrate complex, respectively.

K’ is converted to the true Michaelis constant (Km) using Equation 2.2.

log K' log K  Equation 2.2 m n

Effect of phosphate on steady-state ATPase activity. Reactions containing

50 mM HEPES (pH 8), 5 mM Mg(OAc)2, 2 mM DTT, 150 nM hLon in the absence

and presence of 1 mM sodium phosphate (NaPi, pH 7.2) were initiated with 1 mM

[α-32P]ATP and incubated at 37 °C. Aliquots were quenched at various time points (0-15 min) in 0.5 N formic acid and 3 µL was spotted on a PEI-cellulose

TLC plate and developed in 0.3 M KPi (pH 3.4). The amount of ADP produced

was determined from using equation 2.3

ADPDLU ADP  *ATP i Equation 2.3 ATPDLU  ADPDLU

Where [ADP] is amount of ADP produced, DLU is density light units quantified and [ATP]i is the initial concentration of ATP.

Chemical quench ATPase assay. Pre-steady state time courses of ATP hydrolysis were performed on a KinTek rapid-chemical-quench-flow instrument.

All solutions contained 50 mM HEPES (pH 8), 5 mM Mg(OAc)2, and 2 mM DTT.

A 15 µL solution of human Lon (6 µM final) and casein (10 µM final) was rapidly

mixed with a solution of ATP (500 µM final) containing [α-32P] ATP at 37 ºC for 38

varying times (0-3 s) then quenched with 0.5 N formic acid and extracted with

phenol/chloroform/isoamyl alcohol (25:24:1, pH 6.7). 3 µL of the aqueous layer were spotted on PEI-cellulose and the amount of ADP produced was determined from equation 2.3 as described above.

To determine rate constants, the average amount of ADP produced (0-400 msec) from at least three trials is plotted against time and fit to equation 2.4

Y  A  (ek burst t )  C Equation 2.4

Where Y is [ADP] in micromolar, A is the burst amplitude, kburst is the burst rate constant in per seconds, C is the end point in micromolar, and t is time in

seconds.

Inhibition of human Lon peptidase activity by DBN93. Fluorescent peptidase

assays were performed on a Fluoro-Max 3 spectrophotometer (Horiba group). All

reactions contained 50 mM Tris (pH 8.1), 5 mM Mg(OAc)2, 150 mM NaCl, 2 mM

DTT, 1 mM peptide and 150 nM human Lon. After 1 min at 37 °C, 1 mM ATP was added. After an additional 90 sec, varying amounts of DBN93 was added and fluorescent emission at 420 nm was monitored for 2400 sec with excitation

at 320 nm. A mixture of 10% fluorescent/90% non-fluorescent peptide (Figure

2.1) was used to avoid complications arising from the inner filter effect. All experiments were performed at least in duplicate.

39

Determination of kinetic parameters of inhibition of human Lon by DBN93.

Kinetic parameters for the inhibition of human Lon by boronic acid peptide

DBN93 were derived using methods previously used for S. Typhimurium Lon (46,

51). For reactions containing no inhibitor, steady state velocities were determined

from the linear phase of the reaction time course using Kaleidagraph (Synergy).

For reactions containing inhibitor, the steady state velocities were determined by

fitting the reaction time course using Prism 4 (GraphPad) with equation 2.5

vi  vss kint er t P  vsst  1 e Equation 2.5 kint er

where P is the amount of peptide cleaved, νss is the final steady-state rate, t is

time, νi is the initial rate, and kinter is the rate constant for the interconversion of

νss and νi.

Apparent Ki values were determined by fitting ν/ν0 vs [I] to equations 2.6

and 2.7,

2 v E  I  K app  E  I  K app   4E I i  1 i i Equation 2.6 v0 2E

2 v E  I  K *app  E  I  K *app   4E I ss  1 i i Equation 2.7 v0 2E

app *app where E is enzyme concentration, I is inhibitor concentration, Ki and Ki are the apparent inhibition constants for initial and steady-state phases, respectively.

.

40

* Apparent Ki and Ki values from experimental data were converted to true

* Ki and Ki values by equation 2.8

 [S]  app   Equation 2.8 Ki  Ki 1   Km 

where S is substrate concentration and Km is the reported Km value for the

peptide (42).

Inhibition of human Lon protein degradation by DBN93. Reactions containing

50 mM Tris (pH 8.1), 15 mM Mg(OAc)2, 5 mM DTT, 10 µM λN or StAR protein,

and 1 µM hLon were initiated with 5 mM ATP and incubated at 37 °C in the

absence and presence of 10 µM DBN93. At t=0, 10, 20, 30, 40, 60 min reaction

aliquots were quenched with SDS-PAGE loading dye. Each time point was

loaded onto a 12.5% SDS-PAGE gel and stained by Coomassie Brilliant Blue.

Effect of DBN93 on radioactive steady-state ATPase assay. Steady-state

ATPase activity was monitored as described above. Briefly, reactions containing

50 mM HEPES (pH 8), 5 mM Mg(OAc)2, 2 mM DTT, 150 nM hLon, and 10 μM λN protein in the absence and presence of 2 μM DBN93 were initiated with 1mM [α-

32P]ATP. The amount of ATP hydrolyzed was determined using equation 2.3 as

described above.

41

Results and Discussion

ADP inhibition of peptide cleavage by human Lon. Steady-state peptidase

time courses were run in the presence of Km level ATP (42), varying amounts of

peptide substrate and varying amounts of ADP. The rate of each time course

was quantified by the slope of a line tangent to the linear phase of the time

course. The resulting rate constant data was analyzed using the global fitting

program EnzFitter (BioSoft) and fitting of the data to Equation 2.1 for non-

competitive inhibition yielded kinetic parameters shown in Table 2.1. The values

of Kis (~ 1100 µM) and Kii (~3500 µM) refer to the dissociation constants of

inhibitor binding to free enzyme and enzyme:substrate complex, respectively

(Scheme 2.1). When compared to the parameters determined for bacterial Lon

at Km,ATP: Kis = 1 µM, Kii = 7 µM (43), it can be seen that ADP binds ~500-1000- fold less tightly to human Lon than it does to bacterial Lon, making its inhibitory effect on peptidase activity weaker. This result suggests that while ADP release may be rate-limiting in the mechanism of bacterial Lon (43, 44), it is more likely

human Lon has a different rate-limiting step for enzymatic turnover which occurs

after chemistry, a distinction between the mechanisms that must be explored

further in the future.

Effect of phosphate on steady-state ATPase activity. As Lon catalyzes the

hydrolysis of ATP to yield ADP and inorganic phosphate, the detection of burst

kinetics could be attributed to the slow product Pi release step. To determine if

42

Parameter Human Lon

-1 kcat (sec ) 6.0 ± 0.1

Km (µM) 1000 ± 100

n 1.33 ± 0.01

Kii (µM) 3500 ± 600

Kis (µM) 1000 ± 200

Table 2.1: Kinetic parameters for ADP inhibition of peptide cleavage by human

Lon.

43

Scheme 2.1: Noncompetitive inhibition. Inhibitor binding to free enzyme and enzyme:substrate complex is described by Kis and Kii, respectively.

44

this is the case, I compared the rates of ATPase activity in the absence and

presence of 1 mM NaPi as described in Materials and Methods. As shown in

Figure 2.2, the rate of ATP hydrolysis in the presence of 1mM NaPi is not

significantly inhibited, suggesting the phosphate release is not the rate-limiting step in the mechanism. Combined with the fact that ADP binds very weakly to

human Lon, these results indicate that the rate limiting step occurs after chemistry but are not associated with release of either product.

Chemical quench ATPase assay. Radioactive chemical-quench experiments were run to examine ATP hydrolysis by human Lon in the pre-steady state

region. This method allows for rapid mixing and acid-quenching to monitor the

reaction on a millisecond time scale. The presence of a burst in the ATPase time

course (Figure 2.3a) provides further evidence that the rate limiting step is after

ATP hydrolysis, suggesting ADP release or possibly a conformational change. As

I have shown that ADP does not bind as tightly to human Lon as it does to

bacterial Lon, substrate translocation is most likely the rate limiting step. The full

time course is triphasic, showing a burst in ADP production followed by a slow

transition phase prior to a linear phase indicating steady-state ATP turnover. As

this unusual time course cannot be fit to the classic burst equation, the initial

burst phase (0 – 400 msec) is fit to a single exponential equation (Equation 2.4)

to determine the burst amplitude (Figure 2.3b). This value was determined to be

3.2 ± 0.3 µM, which corresponds to the concentration of active enzyme in the

experiment. As the amount of Lon monomer I used was assumed to be 6 µM,

45

Figure 2.2: Phosphate does not inhibit the ATPase activity of hLon. Rates of

ATP hydrolysis were determined in the absence and presence of 1 mM sodium phosphate. As the rate in the presence of phosphate is not significantly decreased, this confirms that product release is most likely not the rate-limiting step of the mechanism of Lon activity.

46

(a)

(b)

Figure 2.3: Pre-steady state ATP hydrolysis by human Lon. (a) The full time course is triphasic: a burst in ADP production followed by a slow transition phase then steady-state turnover and does not fit to a traditional burst equation. (b) The burst phase between 0 and 400 msec is fit to Equation 2.4 to yield a burst

-1 amplitude of 3.2 ± 0.3 µM and kburst of 10 ± 3 sec . The presence of a burst indicates that a step after ATP hydrolysis is rate-limiting.

47

this indicates that I am seeing only a half-burst amplitude, a trend that has also

been seen in the pre-steady state activity of E. coli Lon, indicating that not all of

the ATPase sites in the Lon complex are active under these reaction conditions.

Further experiments will be done in the future to probe the active site

concentration and the similarities in ATPase activity between E. coli and human

Lon.

Site-directed inhibition of human Lon by DBN93. To further evaluate the

potency of DBN93 towards the inhibition of hLon, the kinetic parameters of

inhibition were determined using the method previously developed for bacterial

Lon using FRETN 89-98 as the reporter probe (46). As with S. Typhimurium

Lon, the peptidase activity of hLon was inhibited by DBN93 in a time-dependent

manner, illustrated by the two distinct rates of peptide cleavage in the

representative time course shown in Figure 2.4a: vi for the initial rate and vss for the final rate before substrate depletion. The rate constant for the interconversion of the two rates, kinter, can also be quantified (51). The vi and vss rates decrease

to different degrees with the increase in [DBN93] (Figure 2.4b), suggesting a two-

step mechanism of inhibition. Analysis of this data yielded the following inhibition

* constants: Ki = 1.35 ± 0.19 µM and Ki = 0.014 ± 0.001 µM. To support the two-

step mechanism, the experimental time courses were globally fitted to both a

one- and a two-step inhibition mechanism (Scheme 2.2) using the nonlinear

fitting program Dynafit (Biokin) (52) (Figure 2.5). The experimental data fit best

to the two-step mechanism and the resulting kinetic parameters presented in 48

k inter v ss

v i

(b)

Figure 2.4: Time-dependent inhibition of human Lon peptidase activity by

DBN93. (a) Representative time course of peptide cleavage by human Lon in the

presence of inhibitor. The time course contains an initial linear phase (vi) anda

steady-state linear phase (vss) with kinter as the rate constant of the

interconversion between the two phases. (b) The rates of these phases (vi ■; vss▲) can be determined by fitting the data to equations 2.6 and 2.7 to yield

* inhibition constants Ki and Ki as described in Materials and Methods.

49

Scheme 2.2: One-and two-step inhibition mechanisms which were fit to the reaction time courses of varying [DBN93] using the global fitting program,

DynaFit.

0.50µM DBN93

0.75µM 1.00µM 1.25µM 1.50µM 1.75µM 2.00µM

Figure 2.5: Global fitting of peptide cleavage time courses in the presence of varying amounts of inhibitor. Black lines indicate experimental data, gray lines are the best fit of the two-step inhibition mechanism shown in Scheme 2.2. The kinetic parameters determined from this fit are outlined in Table 2.2.

50

Table 2.2 agree closely with those obtained by determining the kinetic

parameters individually through analyzing the plots shown in Figure 2.4b as done

previously for S. Typhimurium Lon (46). Through mechanistic characterization, it

is discerned that DBN93 inhibits hLon via a two-step mechanism with an overall

* Ki of 21 nM. This value is similar to that determined for S. Typhimurium Lon.

However, it is interesting to note that the Ki values, which reflect binding of the inhibitor to the enzyme, of hLon and bacterial Lon exhibit a similar difference as the previously reported Km values of FRETN 89-98 in the two enzymes (~4-6-fold higher for hLon). The Km of FRETN 89-98 is 300 µM and 1300 µM for bacterial

and human Lon, respectively (42). Taken together, these results suggest that

even a peptidyl sequence that weakly interacts with the proteolytic active site of

Lon is sufficient to confer inhibition by the boronic acid functionality.

Inhibition of human Lon protein degradation by DBN93. To evaluate the

efficacy of DBN93 inhibiting the proteolytic activity of hLon, the ATP-dependent

degradation of two protein substrates, λN and Steroidogenic Acute Regulatory

(StAR) protein, an endogenous substrate of hLon (53, 54), were examined in the

presence and absence of the inhibitor. As the peptide sequence in DBN93 is

found in the Lon-specific substrate FRETN 89-98Abu and λN protein, we

predicted that the protease activity of hLon would be inhibited. As shown in

Figure 2.6, Lon degrades both λN (a) and StAR (b) within 20 min in the presence

of ATP. However, in the presence of 10 µM DBN93, Lon-mediated degradation of

51

S. Typhimurium Parameter Human Lon Lon†

5 -1 -1 k3 (x10 M s ) 4.3 ± 0.2 1.0 ± 0.3

-1 k4 (s ) 0.060 ± 0.002 0.022 ± 0.007

Ki (μM) 1.37 0.216

-1 k5 (s ) 0.0090 ± 0.0004 0.0032 ± 0.0003

-1 k6 (s ) 0.00014 ± 0.00001 0.00028 ± 0.00002

Ki* (μM) 0.021 0.017

Table 2.2: Kinetic parameters of inhibition of human Lon peptidase activity by

DBN93 as determined by global fitting of experimental data using DynaFit. The resultant values are compared to those previously determined for bacterial Lon.

Reported errors are s.e.m.

†Data previously published (46)

52

(a) -DBN93 +DBN93

Time (min) 0 10 20 30 40 60 0 10 20 30 40 60 hLon -

λN -

-DBN93 +DBN93 (b)

Time (min) 0 10 20 30 40 60 0 10 20 30 40 60 hLon -

StAR-

Figure 2.6: DBN93 inhibits protein degradation by human Lon. Protein substrates λN (a) and StAR (b) were digested with human Lon in the presence of

ATP at 37 °C. Complete degradation of the proteins was inhibited by the addition

of 10 µM DBN93.

53

λN and StAR is not observed even after 60 min. While the inhibition of λN

degradation is expected as the peptide sequence of DBN93 is derived from one

of the products, it is not found in StAR. However, the interaction of YRGIT-Abu in

DBN93 with Lon likely directs the electrophilic boronic acid functionality to react with the proteolytic site Ser.

DBN93 does not affect Lon ATPase activity. The ATPase activity of

mitochondrial Lon has been proposed to possess chaperone-like functions

independent of its proteolytic activity (22). Since DBN93 inhibits hLon in the

presence of ATP, it is conceivable that DBN93 inhibits only the proteolytic, but

not the ATPase functionality. As shown previously, the ATPase activity of Lon is

stimulated by peptide or protein substrate (40, 55, 56). Using a radiolabeled

assay, we demonstrated that the λN protein-stimulated ATPase activity of hLon

was only slightly affected by DBN93 (Figure 2.7). Therefore, as a chemical

genetic tool, this inhibitor should allow for the specific evaluation of the proteolytic

activity. While genetic techniques deplete the cell of both ATPase and proteolytic

activities, the chemical inhibitor DBN93 will only deactivate the proteolytic site in

a dose-dependent manner, thereby permitting the direct evaluation of the

physiological functions of Lon as a protease in cell culture. To the best of our knowledge, the chemical inhibitor described here is the only way to confer this kind of selectivity in deactivating the proteolytic activity of Lon.

54

Figure 2.7: DBN93 does not significantly inhibit hLon ATPase activity in the presence of λN protein stimulation. Reactions measuring the amount of radiolabeled ATP hydrolyzed by human Lon in the absence and presence of

DBN93 were run. The addition of inhibitor does not affect the ATPase activity of human Lon, indicating that DBN93 interacts with the proteolytic site to inhibit activity.

55

Conclusion. In this chapter, I have discussed the differences in inhibition of bacterial and human Lon by ADP and the peptidyl boronic acid, DBN93. As shown by the data, ADP binds with much lower affinity to human Lon than it does to bacterial Lon. ADP release has been proposed to be the rate-limiting step in the mechanism of bacterial Lon, but the low affinity of ADP for human Lon suggests that there is a different rate-limiting step. Pre-steady state ATP hydrolysis experiments discussed here show a burst in ADP production, suggesting that the slow step in the mechanism is after ATP hydrolysis.

When comparing the effect of the peptidyl boronic acid inhibitor, DBN93, on bacterial and human Lon, I showed that the overall potency of the compound

* (Ki ) was similar between the two, in the low nanomolar range. The discovery that overall potency of the inhibitor (Ki*) is not necessarily reflected in the binding affinity of the inhibitor for the enzyme (Ki), indicates that even a peptide sequence that binds with weak affinity can inhibit protease activity. This information will be useful in the design of inhibitors that will be specific for bacterial Lon, but do not affect the activity of human Lon.

56

CHAPTER 3

CHEMICAL TOOLS TO MONITOR ATP-DEPENDENT PEPTIDASE ACTIVITY

IN ISOLATED MITOCHONDRIA

57

ATP-dependent proteases are energy-powered proteolytic machines that

coordinately link ATP hydrolysis with peptide bond cleavage (57, 58) and govern

a protein quality control system dedicated to the elimination of oxidatively

damaged, misfolded and unassembled proteins (55, 59, 60). In mammalian

mitochondria, human Lon (hLon) and ClpXP (hClpXP) are the only soluble ATP- dependent proteases in the matrix (61) and their respective in vivo functions

have yet to be clearly defined. In biochemical experiments using purified

proteins, Lon and ClpXP are apparently promiscuous and degrade some of the

same proteins; however, these proteases exhibit distinct peptide cleavage sites

specificities within a protein substrate (12, 14, 26). Knowledge of the respective

peptide cleavage specificities of Lon and ClpXP can be used to design peptidyl

substrates and inhibitors, which distinguish these proteases and permit the

interrogation of their respective physiological functions. An additional benefit of

such chemical biological tools is the ability to rapidly and/or transiently down-

regulate enzymatic activity without altering protease expression or manipulating

mitochondrial or cellular metabolism. The development of specific chemical tools

is also necessary to probe the functions of Lon and ClpXP upon transient

changes in the metabolism in mitochondria, such as oxidative stress. In an

animal model of cardiac ischemia, an increase in oxidatively modified

mitochondrial matrix proteins following reperfusion is correlated with an increase

in ATP-dependent proteolytic activity, which is then followed by a decrease in

oxidized proteins (62). Similarly, treatment of isolated mitochondria with a H2O2

generating system results in the stimulation of energy-dependent proteolytic

58

activity (62, 63). These data suggest that ATP-dependent proteases play a crucial role in the turnover of oxidatively modified proteins in the mitochondrial matrix. Of the two ATP-dependent proteases found in mitochondrial matrix, Lon has been specifically implicated in targeting oxidatively damaged proteins, as results have shown that it degrades mildly oxidized aconitase (64, 65). The evaluation of ATP-dependent protease activity in mammalian mitochondria, cultured cells or tissues has been performed without rigorously controlling for the contributions of ClpXP or the other membrane-associated ATP-dependent proteases. The exploitation of cleavage site specificity to design enzyme- selective substrates will therefore aid in determining the functions of these two matrix proteases individually and in protein mixture.

Our previous studies on Lon have employed a fluorogenic peptide

substrate, Y(NO2)-RGITCSGRQK(Abz), FRETN 89-98 (Figure 2.1), which

contains residues 89-98 of the endogenous Lon substrate, λN protein (26), to

monitor peptidase activity. This peptide also contains an anthranilamide/nitro-

tyrosine fluorescence donor/quencher pair flanking the cleavage site at Cys-Ser.

In the presence of ATP, the rate of peptide cleavage can be quantified by

monitoring an increase in fluorescence emission over time (40). From the

sequence of this peptide, we have developed a peptidyl boronic acid (DBN93,

Figure 2.1) to target the proteolytic active site of Lon, which has been shown to

inhibit the bacterial enzyme homolog (46). Because the peptide from which this

inhibitor was derived is a substrate of both bacterial and human Lon, we

predicted that it would be a potent inhibitor of the human enzyme. The results

59

from Chapter 2 show that DBN93 is indeed a potent inhibitor of human Lon, with an overall inhibition constant similar to that for bacterial Lon. As hLon and hClpXP have been shown to degrade the same protein at different cleavage sites, it is suggested that they have different substrate specificities. Though I showed in the previous chapter that a lower binding affinity for the peptidyl moiety does not necessarily affect the overall potency of an inhibitor, it is reasonable to assume that at least minimal binding of the peptide is necessary for inhibition. Identification of a peptide sequence that does not bind to hClpXP can lead to the design of a boronic acid inhibitor that will specifically inhibit hLon and not hClpXP, thereby lending service as a selective chemical probe to interrogate the physiological functions of Lon. This approach provides the tools needed to elucidate the mechanism by which hLon is regulated in mitochondria, especially in response to stress, and is general enough that it could be easily adapted to monitor mitochondrial ClpXP or the membrane-bound ATP- dependent protease activities in future studies.

In this chapter, I discuss the use of the fluorescently labeled peptide substrate FRETN 89-98 to monitor ATP-dependent activity in mitochondria isolated from HeLa cells. HeLa are an immortal cell line derived from cervical cancer cells that are regularly used in scientific research (66). The results discussed here suggest that Lon and ClpXP are able to be individually monitored even in a mixture of proteins such as isolated mitochondria using the chemical tools that have been developed in our lab.

60

Materials and Methods

Materials. DMEM media, FBS was purchased from USA Scientific, penicillin/streptomycin, Amphotericin B, gentamycin, 0.25% trypsin/EDTA. All other materials were purchased from Fisher, Sigma or Amresco.

Fluorescent peptidase assay to identify selective peptide substrate and inhibitor. Reactions containing 50 mM HEPES (pH 8.0), 5 mM Mg(OAc)2, 2 mM

DTT, 200 nM purified hLon or hClpXP, 100 µM FRETN 89-98 or FRETN 89-

98Abu, and 500 µM ATP were run on a FluoroMax-3 fluorometer at 37 °C as described in Chapter 2. For Lon reactions, HEPES, Mg(OAc)2, DTT, peptide and

Lon were incubated for 1 min at 37 °C prior to initiation with ATP. For ClpXP reactions, HEPES, Mg(OAc)2, DTT, ClpX, ClpP, and ATP were incubated for 1 min at 37ºC prior to initiation with peptide. For reactions containing inhibitor,

DBN93 was added after an additional 90 sec incubation at 37ºC following the initiation with ATP or peptide to allow the reaction to go past the initial lag phase.

Growing HeLa cells. HeLa cells were grown to passage 6 in DMEM/F12 media containing 10% FBS and 1% penicillin/streptomycin with Amphotericin B in 150 mm tissue culture plates. Plates were incubated at 37 °C at 5% CO2 until 100% confluent, then trypsinized and harvested.

61

Isolation of mitochondria from HeLa cells. Intact mitochondria was isolated

from 120 x 106 HeLa cells using the Mitochondria Isolation Kit for Mammalian

Cells from Pierce (Rockford, IL). Per the manufacturer recommendations, 20 x

106 HeLa cells were resuspended in 800 µL ice cold Reagent A and EDTA (pH 8)

was added to a final concentration of 0.2 mM. The cells were vortexed at medium

speed for five seconds and incubated on ice for exactly 2 min before being

transferred to a chilled 2 mL Dounce homogenizer and lysed on ice with 60 strokes as recommended by the manufacturer. The homogenized cells were then transferred to a polycarbonate centrifuge tube and 800 µL of Reagent C and 0.2

mM final EDTA (pH 8) were added. The homogenizer was rinsed with 200 µL

Reagent A and the rinse was added to the centrifuge tube which was inverted

three times to mix. This procedure was repeated until all 120 x 106 cells were

lysed. The mixture was then centrifuged at 700 x g for 10 min at 4 ºC to pellet cell

debris. The supernatant was transferred to a new centrifuge tube and further

centrifuged at 3000 x g for 15 min at 4 ºC to isolate mitochondria from cytosol

and lysozomal and peroxisomal contaminants. The supernatant was removed

and the pellet containing mitochondria was resuspended in 500 µL Reagent C.

The suspension was divided in two equal parts and centrifuged at 12000 x g for 5

min at 4 ºC. The supernatant was removed and the pellet in tube 1 was

resuspended in 50 µL 50 mM KPi (pH 7), 20% glycerol, 0.1% Tween 20. The

mitochondria was then put through a CentriSpin-10 column (Princeton

Separations) equilibrated in the same buffer to remove small molecular weight

contaminants and stored at 4 °C. The amount of protein present in the isolated

62

mitochondria was quantified using the Bradford assay using BSA as a standard

(48). The second tube of isolated mitochondria was immunodepleted of Lon as

detailed below.

Western Blot analysis of isolated mitochondria. The presence and relative

amount of the respective protease complex components in mitochondria isolated

from 120 x 106 HeLa cells suspended in 100 μl of 50 mM KPi, 20% glycerol,

0.1% Tween 20 buffer were extrapolated by Western Blot Analysis. Known amounts of purified hLon, hClpP, or hClpX as reference standards were compared to a sample of partially purified isolated mitochondrial protein mixture.

The amount of each enzyme present in the mitochondrial sample (mito) was approximated and divided by the volume of sample loaded on the gel to yield ng protein/μl mito. This concentration was then converted to number of copies/cell.

This analysis suggests the proteins exist in approximately the following concentrations: 5000 copies of Lon/cell, 5000 copies of ClpX/cell, and 2000 copies of ClpP/cell.

Immunodepletion of human Lon from isolated mitochondria. Antiserum raised against hLon was purified by incubating production bleed with nitrocellulose containing purified hLon protein and eluting the purified antibody from the nitrocellulose using 100 mM glycine (pH 2.5). Mitochondria isolated from

HeLa cells as described above were resuspended in 50 µL of 25 mM Tris pH 7.5,

100 mM NaCl, 0.1% Tween 20 immunoprecipitation buffer and incubated

63

overnight at 4 °C with 100 µL of the purified anti-hLon. This mixture was then

added to Protein A Agarose (Pierce) that was previously blocked with 1% BSA in

TBST and rotated at 4 °C for two hours. After incubation, the mixture was

centrifuged at 2500 x g for 2 min to pellet the Protein A agarose bound to

antibody and hLon. The supernatant containing isolated mitochondria lysate

depleted of Lon was removed and subjected to a CentriSpin-10 column

equilibrated in 50 mM KPi, 20% glycerol, 0.1% Tween 20. The amount of protein

was quantified by a Bradford assay, using BSA as a standard.

Monitoring specific activity from isolated HeLa mitochondria. Reactions

containing 50 mM HEPES, 5 mM Mg(OAc)2, 2 mM DTT, 1 mM FRETN 89-98 and

1 mM ATP were incubated at 37 °C for 1 min. 5 µg of the purified isolated

mitochondria as quantified by Bradford assay was added and cleavage of

peptide was observed by monitoring the fluorescent emission at 420 nm (λex=320 nm) for 20 min in the absence and presence of 1 mM ATP and 10 µM inhibitor

DBN93.

Inhibition of StAR degradation in COS-1 cells. The experiment testing the stability of StAR in cells treated with DBN93 was done by the Suzuki Lab at

University of Medicine and Dentistry of New Jersey using a method modified from Granot et al (53), on Day 1, two 6 cm dishes of COS-1 cells (1.2 x 106) were

transfected with pCMV5 StAR-his plasmid and two with the pCMV5 control

vector. For each dish of cells, 3 mg of plasmid and 7.5 ml Lipofectamine 2000

64

were diluted in 250 ml OPTI-MEM medium and then added to the cells incubated

in DMEM medium. After 5 hrs, FBS was added to 10%. On Day 2, the two

dishes of cells were split into seven 6 cm dishes. On Day 3, the cells were

incubated in methionine-free DMEM for 30 min prior to labeling for 60 min with

methionine- and cysteine- free DMEM containing [35S]methionine

(100mCi/ml;Trans35S-Label, ICN). The non-fluorescent boronic acid inhibitor,

CBN93 (Figure 3.1) was added to a final concentration of 20 µM 15 min before

the onset of the chase, as well as during the chase. Cells were chased in

complete DMEM medium (10% FBS, penicillin/streptomycin) and harvested at

the time points indicated and then stored at -80 °C. Cellular protein was extracted in lysis buffer (50 mM Tris, pH 7.4, 300 mM NaCl, 0.5% Triton X-100)

containing a protease inhibitor cocktail (Roche Applied Science) on ice for 30

min. The lysates were centrifuged at 13,000 rpm for 30 min. The cell extracts

were pre-cleared with 20 ml Protein G for 2-4 hrs at 4 °C on a rocker, and the extracts were incubated with 10-15 ml anti-his tag antibody (Santa Cruz

Biotechnology SC-8036) overnight at 4 °C on a rocker. The next day, 40 ml of

Protein G agarose was added and incubated for 2 h at 4 °C. The Protein G agarose complexes were washed in ice-cold wash buffer three times, and once with ice cold PBS. Immunoprecipitated radiolabeled protein(s) present in the pulse and chase extracts, were separated electrophoretically on 11% SDS-PAGE and visualized by autoradiography or PhosphorImaging.

65

Results and Discussion

Fluorescent peptide reporter of ATP-dependent proteolysis. The identification of different cleavage profiles of oxidized insulin B by Lon and ClpXP

reported by Maurizi and others (14, 26) (Figure 1.2) lends the possibility of

generating specific peptide substrate reporters for the proteases. As such, the

peptide reporter FRETN 89-98 (Figure 2.1), which has been shown to be

degraded by bacterial and human Lon (40) will be at least a preferred, if not

specific, substrate of hLon, but not hClpXP. Indeed, under identical conditions,

the rate of hLon meditated ATP-dependent FRETN89-98 cleavage was 5-fold

faster than that of hClpXP (Figure 3.1a). By contrast, the FRETN 89-98Abu

(Figure 2.1), where the Cys at the cleavage site of FRETN 89-98 is replaced with

the non-natural amino acid aminobutyric acid (Abu), was only cleaved by hLon

and not hClpXP (Figure 3.1b). These results demonstrate that specific probes

for monitoring ATP-dependent protease activity can be generated by exploiting

the differences in their peptide cleavage specificities, a task that is not easily

accomplished with protein substrates as reporters. Furthermore, the peptide

FRETN 89-98 could be used to monitor the activities of purified hLon or hClpXP

individually in vitro or the total contribution of ATP-dependent protease activity in

a mitochondrial matrix sample whereas FRETN 89-98Abu can be used to

selectively monitor the activity of hLon in the presence of hClpXP.

66

(a) FRETN 89-98

Lon + ATP

ClpXP + ATP

Lon - ATP

(b) FRETN 89-98Abu

Lon + ATP

ClpXP + ATP

Figure 3.1: Peptide-based substrates are used to monitor activity of human Lon and human ClpXP. (a) FRETN 89-98 is cleaved by both human Lon (black) and human ClpXP (dark gray) in the presence of ATP with different efficacies. The time course in the absence of ATP is shown in light gray. (b) FRETN 89-98Abu is a specific substrate for human Lon (black) and is not cleaved by human ClpXP

(dark gray).

67

Generation of selective inhibitor, DBN93. The preferential cleavage of FRETN

89-98Abu peptide by hLon suggests that the hydrolyzed peptide product

YRGIT(Abu) would be bound to the proteolytic site of hLon but not hClpP.

Therefore, derivatizing the carboxyl terminal of YRGIT(Abu) with a boronic acid moiety, which acts as an electrophile to sequester the proteolytic site Ser in Lon, should generate a mechanism-based inhibitor that blocks proteolysis. This chemical tool will aid in evaluating the contribution of hLon in protein quality control in mammalian mitochondria. The peptidyl boronic acid dansyl-YRGIT-

Abu-B(OH)2, abbreviated as DBN93 (Figure 2.1), contains part of the sequence derived from a hydrolyzed peptide product generated from λN, as well as FRETN

89-98, degradation by Lon (46). The amino terminal of DBN93 is also derivatized with a dansyl group to offer the opportunity to probe its mechanism of interaction with Lon if needed (67).

The specificity of DBN93 towards hLon is confirmed by the observation that the cleavage of FRETN 89-98 by purified hClpXP was not affected by up to

4 µM DBN93 (Figure 3.2), a condition that would have completely inhibited Lon

(46). This result highlights the usefulness of exploiting the peptide sequence of substrates and inhibitors to develop specific compounds which can be used to determine the physiological functions of ATP-dependent proteases. Despite the promiscuity of protein substrate specificity and lack of defined peptide cleavage specificity in Lon, a proteolytic site-specific inhibitor can be generated from the hydrolyzed peptide product.

68

Figure 3.2: DBN93 does not inhibit human ClpXP peptidase activity. ClpXP degradation of FRETN89-98 is not inhibited by DBN93. 300 nM ClpP and 900 nM

ClpX were preincubated with 1 mM ATP at 37 °C for 1 min prior to the addition of

1 mM FRETN89-98. After 90 s, 0 µM (black), 1 µM (light gray) or 4 µM (dark gray) DBN93 was added and the reaction was monitored for 1500s.

69

Inhibition of mitochondrial protein mixture. To assess the ability of our peptide reporter to detect ATP-dependent protease activity, and the efficacy of

DBN93 to selectively inhibit hLon in biological samples containing hClpXP, we evaluated the cleavage of FRETN 89-98 in mitochondrial matrix proteins isolated from HeLa cell cultures. These proteins were isolated using a method that presumably removed mitochondrial membranes and were thus free of membrane bound ATP-dependent protease contamination. The presence of hLon and hClpXP in the isolated protein mixture was confirmed by Western Blot analysis using antibodies against each enzyme (Figure 3.3). The lower molecular weight of ClpP in the mitochondrial matrix protein mixture might be due to autoprocessing of ClpP as reported previously (68) or the fact that the purified protein standard contains a 6x His-tag, which increases the molecular weight.

To monitor peptidase activity, 5 µg of the isolated matrix protein mixture was incubated with 100 µM FRETN 89-98 with and without the addition of 1 mM

ATP. As shown in Figure 3.4, an increase in fluorescence emission signal attributed to FRETN 89-98 cleavage over time is detected in the 1 mM ATP time course. By comparison, there is little change in fluorescence intensity over the same time period when no additional ATP is present. This observation confirms that ATP-dependent cleavage of FRETN 89-98 is detectable in mitochondrial protein mixture isolated from cell culture. When the Lon inhibitor, DBN93, was added to the FRETN 89-98 cleavage reaction containing 1 mM ATP, the peptidase signal was reverted to the background level, where ATP was omitted

(Figure 3.4). This result suggests that FRETN 89-98 was detecting only hLon 70

25ng 10ng 5ng 1ng Mito (a) α-hLon

25ng 10ng 5ng 1ng Mito

(b)

α-ClpX 25ng 10ng 5ng 1ng 0.5ng Mito (c)

α-ClpP

Figure 3.3: Identification of mitochondrial matrix proteins isolated from HeLa cells. Western blots of mitochondria isolated from HeLa cells visualized with antibodies against known amounts of (a) human Lon, (b) ClpX, or (c) ClpP as indicated to approximate amount of enzyme in the isolated mitochondria. The amounts above the blots indicate the amount of purified protein ran as a reference to approximate the amount of each protease in the mito sample.

71

+ ATP

- ATP

+ ATP + DBN93

Figure 3.4: ATP-dependent peptide cleavage of FRETN 89-98 by isolated mitochondria is abolished by the addition of Lon specific inhibitor, DBN93.

Cleavage of 100 µM FRETN 89-98 by 5 µg of the purified mitochondrial matrix protein mixture was monitored in the absence (black) and presence (dark gray)

of 1 mM ATP. The addition of 10 μM DBN93 inhibited ATP-dependent peptidase activity (light gray) to intrinsic levels.

72

activity in the protein mixture. As discovered earlier, purified hClpXP is at least

5-fold less efficient in mediating the ATP-dependent cleavage of FRETN 89-98

(Figure 3.1a); therefore, the lack of detectable ClpXP activity in the isolated HeLa mitochondrial matrix proteins could be attributed to the presence of a relatively low concentration of functional hClpXP complex. This possibility is tentatively supported by the estimation (from Western Blot analysis) that there are ~ 5000 copies of Lon, 5000 copies of ClpX, and 2000 copies of ClpP per cell, yielding a theoretical concentration of hLon exceeding that of hClpXP by ~ 2.5-fold

(5000/2000).

To confirm that I am able to detect only Lon activity from isolated mitochondria using FRETN 89-98, I immunodepleted the isolated mitochondria of human Lon and tested the sample for ATP-dependent activity. As shown in

Figure 3.5, in the presence of ATP (black), the rate of peptide cleavage is similar to that in the absence of ATP (dark gray), indicating that Lon activity is no longer able to be detected. The addition of DBN93 to the reaction did not decrease the rate of peptide cleavage as it did in the mitochondria that still contained Lon.

Taken together, the peptidase assay accurately measures the respective

ATP-dependent protease activity based on their specific concentrations in the sample. Based on this observation, I propose that the combined usage of specific fluorescent peptide substrates and the boronic acid inhibitor demonstrates a viable chemical strategy to detect hLon and/or hClpXP activity in biological samples. As the deactivation of mitochondrial Lon does not require any genetic

73

Figure 3.5: Immunodepletion of Lon from isolated mitochondria abolishes ATP- dependent peptide cleavage of FRETN 89-98. In the presence of ATP (black), there is an increase in the rate of peptide cleavage by mitochondria containing

Lon over that in the absence of ATP (dark gray). Upon addition of 10 µM DBN93

(light gray), ATP-dependent peptide cleavage was brought back down to the

background rate in the mitochondria containing Lon. In mitochondria

immunodepleted of Lon (IP), there is no increase in peptide cleavage in the

presence of ATP (white crosshatch) over that in the absence of ATP (gray

crosshatch). Error bars indicate the standard error from least three trials.

74

manipulation, the chemical approach described herein should be applicable to the profiling of Lon and/or ClpXP activities in mitochondrial protein matrix isolated from primarily cell cultures and animal tissues that are not easily amendable to genetic manipulation.

Inhibition of StAR protein degradation in vivo. To assess the cell permeability of the boronic acid inhibitor, we tested whether treatment of cells with CBN93 will downregulate Lon activity and lead to increased StAR stability using a previously employed method (53). The relative amount of StAR present in the presence or absence of CBN93 was determined by immunoprecipitation of StAR, SDS-PAGE and autoradiography. As shown in Figure 3.6, higher levels of 35S-Met labeled

StAR were detected at 2, 4, and 8 hours in the cells treated with 20 µM CBN93.

This result supports the notion that the peptidyl boronic acid inhibitor CBN93 exhibits cell permeability in cultured cells. As the non-fluorescent analog,

CBN93, inhibits the degradation of the endogenous substrate StAR in cell culture, this inhibitor should be able to enter isolated mitochondria to target Lon, thereby allowing the identification of substrates of this enzyme by proteomic techniques, as well as physiological functions by monitoring alterations in mitochondrial function within minutes to hours. Since the active-site directed approach used for generating specific reagents for monitoring Lon exploits the interaction between the enzyme and peptide substrate, we propose that the same strategy can be employed to develop specific chemical reagents to study other ATP-dependent proteases which may be indistinguishable by their protein

75

- CBN93 + CBN93

Chase (h) 0 2 4 8 0 2 4 8 StAR

Figure 3.6: Chase experiments show CBN93 inhibition of StAR degradation in

isolated mitochondria. Degradation of 35S Met-labeled StAR protein by isolated mitochondria was monitored in the absence and presence of 20 μM CBN93. The

amount of 35S- labeled StAR remaining in the mitochondria was quantified by

radiography.

76

substrates but not peptide cleavage specificities. One caveat is that in addition to

mitochondrial Lon, StAR is also degraded by the proteasome which is found in

the cytosol, but not the matrix of mitochondria. It is possible that the increased levels of StAR observed in the cell culture study reflect a combined CBN93 block of the proteasome and Lon. This ambiguity further reiterates the non-selective nature of protein substrate, and the need to develop specific peptide substrates as activity probes for ATP-dependent proteases.

77

Conclusion. In this chapter, I have shown that the fluorescent peptide substrate

for Lon and ClpXP (FRETN 89-98) can be used to monitor ATP-dependent

activity of mitochondria isolated from HeLa cells. As shown by Western Blot

analysis, the amount of ClpXP present in the mitochondria is approximately 2.5-

fold less than the amount of Lon. As ClpXP is about 5-fold less efficient at

cleaving FRETN 89-98, it is understandable that I would not be able to detect

peptide cleavage by ClpXP under the conditions of the experiment. To confirm that I was able to only detect Lon activity, I immunodepleted the mitochondria of

Lon and tested for peptidase activity. The rate of peptide cleavage was nearly identical in the absence and presence of ATP, confirming that the ATP- dependent activity that I was able to detect was due to Lon activity.

78

CHAPTER 4

DESIGN AND CHARACTERIZATION OF ClpXP SPECIFIC N-TERMINAL

BORONIC ACID INHIBITOR

79

As discussed in Chapter 3, ClpXP cleaves FRETN 89-98 with about 5-fold

less efficiency than human Lon (Figure 3.1a). Upon substitution of Cys with Abu,

the peptide is no longer cleaved by ClpXP, making it a specific substrate to

monitor Lon activity. A hydrolyzed peptide product of FRETN 89-98Abu was

used to design a potent inhibitor of bacterial and human Lon (DBN93), with an

overall inhibition constant of 21 nM, which is similar to that which was found for

bacterial Lon (17 nM) (46). When tested against ClpXP, it was found that up to 4

µM DBN93, there was no inhibition of peptidase activity (Figure 3.2). As such,

DBN93 has been identified as a chemical tool that can monitor Lon activity in mitochondria isolated from HeLa cells. The success with the design of DBN93 as a Lon-specific inhibitor suggests that a peptidyl moiety that binds tightly to ClpXP along with a boronate moiety to interact with the active site could produce an inhibitor that is specific to ClpXP, allowing for its selective monitoring even in the presence of other proteases. To this end, in collaboration with the Santos Lab at

Virginia Tech, a series of peptidyl boronic acid compounds were synthesized and tested against both human Lon and ClpXP to identify promising inhibitor candidates.

Peptidic α-amino boronic acids are excellent inhibitors of proteases because of the ability of Lewis basic residues located in the enzyme active site to bind to the empty p-orbital of boron forming a stable “ate” complex, which mimics the tetrahedral intermediate formed during amide bond hydrolysis (69-73). When the protease substrate is known, this strategy becomes attractive because the P-

80

site residues can readily be incorporated to provide a good starting point for inhibitor development (Figure 4.1). It is well-known that substrate selectivity within the enzyme active site relies heavily on both the P/S- and P’/S’-site complementarity. However, current methods incorporating the boronic acid moiety only take advantage of the binding pocket on P/S side of the scissile amide bond (C-terminal boronic acid). In this chapter, I report the characterization of a series of N-terminal peptidic boronic acids that were synthesized to harvest the potential of residues on the opposite side (P’-sites). In collaboration with the Santos group, we reason that the P’-sites may provide unique selectivity for a number of biologically important proteases. In the design of N-terminal peptidic boronic acids, the carboxyl group is replaced with a boronic acid moiety and the nitrogen derived from the scissile amide bond is changed to carbon because the B-N bond is labile. The terminal boronic acid residue can be conveniently modified to incorporate both natural and unnatural amino acid functional groups. To evaluate the utility of this approach, I compared the inhibition profiles of a series of N-terminal peptidic boronic acids toward human

Lon and ClpXP. As previously discussed, both of these enzymes are present in mitochondrial matrix (17), therefore, the observed up-regulation in ATP- dependent proteolytic activity upon oxidative stress (62, 63) could be attributed to one or both enzymes. The contribution of the respective proteases in protecting mitochondria from oxidative stress is not fully known. Therefore, the development of a selective inhibitor against the respective protease will provide

81

S1' S2

P2 O P1' O H H N N N N H H O P1 O P2'

S2' S1 protease

HO OH P1' H Tetrahedral peptide N peptide N Intermediate H O P O 1

OH OH P1' H peptide N B B peptide OH HO O O P1 C-terminal peptidic boronic acid N-terminal peptidic boronic acid

Prior Work Current Work

Figure 4.1: Complimentary peptide boronic acid strategies utilizing the P/S and

P’/S’ sites.

82

valuable chemical tools to interrogate the physiological functions of the two

mitochondrial proteases. Furthermore, several physiological protein substrates

have been identified for human Lon but not for human ClpXP. The availability of

selective inhibitors should benefit the identification of physiological substrates of

hClpXP when used in conjunction with proteomic studies.

This chapter describes the purification of recombinant ClpX and ClpP and

confirms the inability of C-terminal peptidyl boronic acid, DBN93, to inhibit ClpXP

protease activity. I also discuss the screening of the series of N-terminal boronic

acid compounds that were synthesized by the Santos Lab at Virginia Tech in

order to identify a specific ClpXP inhibitor. A boronic acid compound was indeed

identified as an inhibitor of ClpXP peptidase activity and I determined its IC50 value. I also confirmed specificity of this inhibitor by testing its effect on protein degradation by hLon and ClpXP. The results gathered from these experiments provide a starting point for development of a potent, ClpXP specific inhibitor that can be used to monitor ClpXP activity in vitro and later in vivo.

83

Materials and Methods

Materials. Oligonucleotides were purchased from IDT. Ni-NTA agarose was purchased from Qiagen. All other reagents and materials were purchased from

Fisher, Sigma, and Amresco.

Cloning of recombinant ClpX and ClpP. Using a modified method of Kang et al

(14), the ClpP was amplified from genomic DNA by Jason Hudak using

oligonucleotides oJH075 (5’-GACCCGGCATATGCCGCTCATTCCCATCG-3’)

and oJH114 (5’-GGCCCAGCTCGAGGGTGCTAGCTG-3’). The ClpX gene was

amplified from genomic DNA using oligonucletoides oJH073 (5’-

CACCAGCACATATGGCCTCAAAAGATG G-3’) and oJH115 (5’-

GCAATATGACCTCGAGGCTGTTTGCAG-3’). The resultant PCR products were cloned into the NdeI and XhoI sites of pET24c and transformed into XL1Blue cell line for sequencing confirmation. The ClpX plasmid contained a mutation E373K which was changed back to wild-type using Quick-Change kit from Qiagen with primers oJH125 (5’-

CATCAGCAGGAGGAAAAATGAAAAGTATCTTGGATTTGGAACACC -3’) and oJH126 (5’- GGTGTTCCAAATCCAAGATACTTTTCATTTTTCCTCCTGCTGATG

-3’). Following sequence confirmation, the plasmids ClpP and ClpX were transformed into Rosetta (DE3).

84

Purification of recombinant ClpX and ClpP. Recombinant wild type His-tagged

human ClpX and human ClpP were overexpressed in Rosetta (DE3) using the

plasmids ClpX and ClpP, respectively. Cells were grown in SB containing 30

μg/mL kanamycin and 34 μg/mL chloramphenicol at 37 °C to OD600 ~0.7.

Cultures were then induced with 0.5 mM IPTG for 21 hours at 25 °C. Cells were

harvested and pellets were resuspended in loading buffer containing 50 mM Tris

(pH 8.1), 150 mM NaCl, 10 mM imidazole, 15% glycerol and 0.05% Tween 20,

then lysed by homogenization and sonication. Lysate was applied to Ni-NTA

agarose column equilibrated in loading buffer. The column was washed with 5 column volumes of loading buffer, with flow-through collected. The column was then washed with 8 volumes of wash buffer containing 50 mM Tris (pH 8.1), 500 mM NaCl, 10 mM imidazole, 15% glycerol and 0.05% Tween 20, with the flow- through collected. The column was eluted with 5 column volumes elution buffer containing 50 mM Tris (pH 8.1), 300 mM NaCl, 200 mM imidazole, 15% glycerol and 0.05% Tween 20, with 1 mL fractions collected. Collected fractions were analyzed on 12.5% SDS-PAGE gels stained by Coomassie. Fractions containing purified protein were pooled and concentrations were determined by Bradford assay (48), using BSA as a standard. Protein was then aliquoted and stored at

-80 °C Purified aliquots of ClpX and ClpP are shown in Figure 4.2 after SDS-

PAGE analysis, with molecular weights of ~64 kDa and ~25 kDa, respectively.

85

MW (kDa)

170 135 100 72 ClpX

55

40

33

24 ClpP

17

Figure 4.2: Purified samples of ClpX and ClpP.

86

Steady state peptidase activity of ClpXP. ClpXP peptidase activity was monitored using a previously designed assay which utilizes FRETN 89-98 (40) as described in Chapter 2. Reactions containing 50 mM HEPES (pH 8.1), 5 mM

Mg(OAc)2, 2 mM DTT, 900 nM ClpX and 300 nM ClpP were incubated in the

absence and presence of 1 mM ATP for 1 min at 37 °C. Reactions were initiated

by the addition of 500 μM FRETN 89-98 and fluorescence emission was

monitored for 1200 sec. In all assays, a mixed substrate was used (90% non-

fluorescent 10% fluorescent as described in Chapter 2) to avoid the inner filter effect (43).

Human ClpXP steady-state inhibition of protein degradation. Reactions containing 50 mM Tris (pH 8.1), 15 mM Mg(OAc)2, 5 mM DTT, 2 µM hClpX, 2 µM

ClpP, and 5 mM ATP were incubated at 37 °C for 1 min prior to the addition of

10µM casein in the absence and presence of 10 μM DBN93. Reaction aliquots were quenched in SDS loading dye at t=0, 30, 60, 90 min. Each time point was loaded onto a 12.5% SDS-PAGE gel and stained by Coomassie Brilliant Blue.

Synthesis of N-terminal boronic acid compounds. N-terminal boronic acid compounds (Figure 4.3) were synthesized by the Santos lab at Virginia Tech as published (74).

Screening of N-terminal boronic acid compounds as specific inhibitors of

ClpXP. To screen the effect of various compounds on the peptidase activity of

87

Figure 4.3: N-terminal peptidyl boronic acids synthesized by the Santos group and screened for ClpXP specific inhibitors. The peptide sequence GRQAY was derived from the sequence of the second product of FRETN 89-98 peptide hydrolysis by Lon (see Figure 1.6).

88

Lon and ClpXP, reactions containing 50 mM HEPES (pH 8.1), 5 mM Mg(OAc)2, 2 mM DTT, 300 nM human Lon or human ClpXP, and 20 μM various inhibitors were incubated for 30 min at room temperature. For the Lon assays, the reactions were then incubated for 1 min at 37 °C prior to initiation by the addition of 500 µM FRETN 89-98 and 1 mM ATP. For ClpXP assays, after incubation at room temperature for 30 min, 1 mM ATP was added and the reaction was further incubated at 37 °C for 1 min. The reaction was then initiated by the addition of

500 µM FRETN 89-98 and fluorescent emission was monitored for 1200 sec at

420 nm with excitation at 320 nm. Observed rate constants were determined as previously described and were normalized by dividing by the kobs for the reaction

in the absence of modulator. All experiments were performed in triplicate.

Compounds that inhibited ClpXP, but not human Lon were further tested at a concentration of 1 mM using the same fluorescent assay.

IC50 determination of lead N-terminal boronic acid compound, WLS6a.

Reactions contained 50 mM HEPES (pH 8.1), 5 mM Mg(OAc)2, 2 mM DTT, and

300 nM ClpXP. Varying amounts of inhibitor were added and the reaction was

incubated at room temperature for 30 min. The reaction was incubated with 1

mM ATP for 1 min at 37°C, then initiated with the addition of 500 μM FRETN 89-

98. Observed rate constants were normalized as described above and plotted against concentration of WLS6a and fit to Equation 4.1 (75) using Prism 4

(GraphPad)

89

v 1 i  Equation 4.1 v [I] 0 1 IC50

where [I] is concentration of inhibitor and IC50 is the concentration of inhibitor at which cleavage is inhibited by 50%.

WLS6a inhibition of protein degradation. Reactions containing 50 mM HEPES

(pH 8.1), 15 mM Mg(OAc)2, 5 mM DTT, 2μM ClpXP, and 1 mM WLS6a were incubated at 37 °C for 1 min prior to the addition of 5 mM ATP. Following further incubation for 1 min at 37 °C, reactions were initiated with 10 μM casein.

Reaction aliquots were quenched in SDS loading dye at t=0, 5, 15, 30, 60, 120 minutes. Each time point was loaded onto a 12.5% SDS-PAGE gel and stained by Coomassie Brilliant Blue.

Effect of WLS6a on steady-state ATPase activity of ClpXP. Reactions containing 50 mM HEPES (pH 8), 5 mM Mg(OAc)2, 2 mM DTT, 300 nM ClpXP, in the absence and presence of 1 mM WLS6a were initiated with 1 mM [α-32P] ATP and incubated at 37 ºC. Reaction aliquots were quenched with 0.5 N formic acid at various time points (0-15 min) and 3 µL were spotted on PEI-cellulose TLC plates. The plates were developed in 0.3 M KPi (pH 3.4) and the amount of ADP

produced was quantified by PhosphorImaging as described in Chapter 2.

90

Results and Discussion

Degradation of casein by ClpXP. To confirm the inability of DBN93 to inhibit the protease activity of ClpXP, the degradation of α-casein, a known protein substrate of both Lon and ClpXP (12, 14), was examined in the presence and absence of the inhibitor. As I have shown that the peptide from which the sequence of DBN93 is derived is not a substrate of ClpXP and that the peptidase activity of ClpXP was not affected by up to 4 µM DBN93, the degradation of protein substrate should also not be inhibited. As shown in Figure 4.4, under identical experimental conditions, DBN93 inhibited α-casein degradation by hLon

(Figure 4.4a), but not by hClpXP (Figure 4.4b), as expected. This result confirms that the peptide sequence of substrates and inhibitors can be exploited develop specific compounds which can be used to determine the physiological functions of ATP-dependent proteases.

Screening of N-terminal boronic acid compounds as specific inhibitors of

ClpXP. As illustrated in the previous chapter, hClpXP shows the expected ATP- dependent cleavage of the fluorescent peptide, illustrated by an increase in fluorescence emission which can be monitored in the absence and presence of modulators. Using a FRET assay previously developed for hLon (76), I adapted the same protocol to determine inhibitors for hClpXP using FRETN 89-98 as the substrate. These potential inhibitors peptidyl N-terminal boronic acid compounds were synthesized by the Santos group at Virginia Tech. The peptide sequence

91

-DBN93 +DBN93 (a) Time (min) 0 10 20 30 0 10 20 30

hLon -

casein-

-DBN93 (b) +DBN93 Time (min) 0 30 60 120 0 30 60 120

ClpX -

casein- ClpP-

Figure 4.4: DBN93 inhibits casein degradation by human Lon but not ClpXP. (a)

1 μM hLon or (b) 2 μM ClpX and 2 μM ClpP (b) were reacted at 37 °C with 10 μM casein and 5 mM ATP in the absence and presence of 10 μM DBN93. Reaction aliquots were quenched at specified time points in SDS-PAGE loading dye and run on a 12.5% gel, then stained by Coomassie.

92

was derived from the carboxy product of peptide hydrolysis by Lon (Figure 1.5).

Initial screening of N-terminal peptidic boronic acids WLS6a-6g (Figure 4.3) at 20

µM indicated that diasteromers 6g1 and 6g2 inhibited both hLon and hClpXP but

6b showed no inhibition of either enzyme (Figure 4.5a). In contrast, WLS6a, 6e and 6f inhibited hClpXP but not Lon. Encouraged by these results, I then tested these compounds at a higher concentration (1 mM) and discovered that the proteolytic activity of hClpXP was almost completely inhibited (>90%) by N- terminal boronic peptide WLS6a in a time-dependent manner while hLon was unaffected (Figure 4.5b). 6e and 6f also inhibited hClpXP, but to a lesser extent.

IC50 determination of lead compound, WLS6a. After identifying compound

WLS6a as the most effective inhibitor from the series, I performed experiments

under steady-state conditions to estimate its potency. Fitting the kobs values of

peptide cleavage against varying amounts of WLS6a to equation 4.1 yielded an

IC50 of 29 ± 9 µM (Figure 4.6). Detailed kinetic assays to elucidate the

mechanism of inhibition as well as true inhibition constants for WLS6a are

currently underway using an optimized peptidase assay.

WLS6a inhibition of casein degradation by ClpXP. To confirm that WLS6a is

a selective inhibitor of hClpXP, I performed a secondary assay and investigated

whether compound WLS6a inhibited hClpXP and hLon mediated protein

degradation using SDS-PAGE analysis. I used casein as the substrate as it is

known to be a protein substrate for both hLon and hClpXP (77, 78). As

93

(a)

(b) B

Figure 4.5: Screening of N-terminal peptidic boronic acids using FRET 89-98 peptidase assay. (a) 20 μM or (b) 1 mM putative inhibitor was incubated with

300 nM hLon (black) or hClpXP (gray) at rt for 30 min prior to addition of ATP and FRETN 89-98 substrate. All values are the average of three experiments.

94

Figure 4.6: WLS6a IC50 determination by fluorescent peptidase assay. Rates of

FRETN 89-98 cleavage by ClpXP in the presence of varying amounts of WLS6a were determined using the fluorescent peptidase assay. Rates were normalized by dividing by the rate in the absence of inhibitor and plotted against WLS6a concentration. Data were fit to equation 4.1 as described in Materials and

Methods and an IC50 of 29 ± 9uM was determined.

95

expected, casein was completely degraded by hLon in the absence of inhibitor while minimal inhibition is observed in the presence of inhibitor as a small amount of casein persisted after 15 min of incubation (Figure 4.7a). It is clear, however, that casein is completely hydrolyzed after 30 min. In the presence of hClpXP, proteolytic cleavage of casein was complete within 2 h in the absence of inhibitor but almost fully intact during the same time course of the experiment upon incubation with inhibitor WLS6a (Figure 4.7b). These results are in complete agreement with the FRET assay and suggest that WLS6a is a selective inhibitor for hClpXP over hLon, even at a concentration of 1 mM. To the best of our knowledge, compound WLS6a is the first reported selective inhibitor for hClpXP. Although the exact mechanism of inhibition is currently unknown, this discovery should facilitate the understanding of role of hClpXP and hLon in the mitochondria.

WLS6a does not inhibit the ATPase activity of ClpXP. To confirm that the boronic acid compound WLS6a was inhibiting only the protease activity of ClpXP,

ATPase actvity assays were done in the presence of 1 mM WLS6a. As expected, even at a high concentration of WLS6a, there is no inhibition of ClpX ATPase activity (Figure 4.8).

96

A - WLS6a + WLS6a (a) Tim e (m in ) 0 5 15 30 0 5 15 30

Human Lon

Casein product

B (b) - WLS6a

Time (min) 0 5 15 30 60 120

ClpX

Casein ClpP product

+ WLS6a

ClpX

Casein ClpP product

Figure 4.7: Inhibition of ClpXP degradation of casein by WLS6a. Reactions

containing 1 µM hLon (a) or hClpXP (b) were incubated for 1 min at 37 °C in the absence and presence of 1 mM WLS6a. 5 mM ATP was added and the ClpXP reactions were further incubated for 1 min at 37 °C before initiation with 10 µM casein. Aliquots were quenched at specified time points and run on a 12.5%

SDS-PAGE gel.

97

Figure 4.8: WLS6a does not inhibit ClpXP ATPase activity. Reactions containing 300 nM ClpXP in the absence (black) and presence (gray) of 1mM

WLS6a and casein were initiated with [α-32P] ATP. The addition of ClpXP inhibitor does not affect the ATPase activity of ClpXP, confirming that the inhibitor interacts with the protease site to inhibit protease activity.

98

Conclusions. In this chapter, I have shown that the C-terminal boronic acid inhibitor, DBN93, which I showed to be a potent inhibitor of human Lon activity, does not affect the protease activity of ClpXP. In an effort to identify a boronic acid inhibitor that would be specific for ClpXP and not Lon, a series of N-terminal peptidyl boronic acids were synthesized by the Santos Lab at Virginia Tech. I screened these compounds and identified three potential ClpXP inhibitors. Upon further studies at higher concentration of inhibitor, I identified a single compound,

WLS6a which does indeed inhibit only ClpXP protease activity and not the activity of human Lon. I then monitored ClpXP cleavage of FRETN 89-98 in the presence of varying amounts of WLS6a to determine an IC50 value of ~29 µM.

More in-depth kinetic studies of the inhibition of ClpXP activity by this compound are necessary to determine true inhibition constants. The ideal inhibitor would have an inhibition constant in the low nanomolar range, but as WLS6a does not inhibit human Lon even at higher concentrations (1 mM), it will be a useful tool in selectively monitoring the activity of ClpXP in the future as DBN93 was used to detect Lon activity in isolated mitochondria (Chapter 3). The peptide sequence of the N-terminal boronic acid compounds can also be varied to possibly identify a more potent inhibitor.

99

CHAPTER 5

CONCLUSIONS AND FUTURE DIRECTIONS

100

In Chapters 2-4, I have discussed experiments that exploit the differences

in three ATP-dependent proteases. In previous studies from our lab, it has been

shown that ADP inhibits the peptidase activity of E. coli Lon (43) and proposed

that the step after ATP hydrolysis is rate limiting, which could be attributed to

ADP release (44, 55). In Chapter 2, I looked at the inhibition of human Lon

peptidase activity by ADP and determined inhibition constants for comparison to

those published for E. coli Lon. As described in the previous chapters, I have identified a peptide substrate that is cleaved by human Lon, but not human

ClpXP (FRETN 89-98Abu), a peptidyl boronic acid inhibitor that inhibits only human Lon and not ClpXP protease activity (DBN93), and a boronic acid inhibitor that only inhibits ClpXP and not human Lon protease activity (WLS6a). I have also shown in Chapter 3 that DBN93 can be used to abolish ATP-dependent peptidase activity of mitochondria isolated from HeLa cells. These results suggest that by identifying a peptide substrate that binds with high affinity to

ClpXP and a potent inhibitor of ClpXP will allow for the selective monitoring of

ClpXP activity in isolated mitochondria and live cells. The identification of

WLS6a as a ClpXP-specific inhibitor as described in Chapter 4 provides a starting point for the development of a more potent inhibitor that will help elucidate the mechanism of ClpXP protease activity, but there is still a need for a specific peptide substrate of ClpXP.

To this point, I am able to use experiments that have been previously developed for the study of Lon mechanism to begin to understand the activity of

ClpXP. This chapter describes the preliminary experiments I have done to

101

identify and characterize specific peptide substrates of ClpXP which can later be used to develop potential specific inhibitors for use studies of isolated mitochondria and live cells. Because the ClpXP complex contains individual

ATPase and proteolytic components, I must look at both activities. I begin by describing the determination of kinetic parameters for the ATPase activity of ClpX using the steady-state radioactive ATPase assay. As I showed in Chapter 3, Lon and ClpXP have different affinities for our lead peptide substrate, FRETN 89-98, so I began experiments to determine the kcat and Km for its cleavage by ClpXP to compare to those published for human Lon. I also introduce experiments that lead to the design of ClpXP specific substrates and inhibitors.

102

Materials and Methods

Radioactive steady-state ATPase activity assay. Steady-state ATPase

activity was monitored as previously described in Chapter 2. To test the effect of

imidazole concentration on ATPase activity by ClpXP, reactions containing 50

mM HEPES (pH 8), 5 mM Mg(OAc)2, 2 mM DTT, 100 nM ClpX, with and without

the addition of 40 mM imidazole (pH 8) were initiated with 1 mM [α-32P]ATP. The

reaction was incubated at 37 °C and aliquots were quenched in 0.5 N formic acid

at various time points (0-15 min). The rate of ATP hydrolysis was determined as

described in Chapter 2.

To examine the effect of ClpP on ClpX ATPase activity, reactions

containing 300 nM ClpX and varying amounts of ClpP were initiated with 1 mM α-

32P-ATP and aliquots were quenched in 0.5 N formic acid at various time points

(0-15 min).The rates of ATP hydrolyzsis were determined as previously

described.

For the determination of kcat and Km values of ClpX ATPase activity, reactions containing 300 nM ClpX and 100nM ClpP were initiated with varying amounts of [α-32P]ATP and incubated at 37 °C. Reaction aliquots were removed

at various time points (0-15 min) and quenched in 0.5 N formic acid and the

amount of ATP hydrolysis was determined as described above.

Steady state ATP-dependent FRETN 89-98 peptidase assay. Reactions

containing 50 mM HEPES (pH 8.0), 5 mM Mg(OAc)2, 2 mM DTT, 900 nM ClpX, 103

300nM ClpP, and ATP were incubated at 37 °C for 1 min prior to addition of

fluorescently labeled peptide. An increase in fluorescence emission at 420 nm

(λex= 320 nm) was monitored as described in Chapter 2.

For the determination of kcat and Km for peptide cleavage of FRETN 89-98, varying amounts of peptide (10-1000 µM) were added to initiate the reaction in the presence of 1 mM ATP. The amount of peptide hydrolyzed was measure by determining the maximum fluorescence generated per micromolar peptide after complete digestion with trypsin. The steady-state rate of the reaction was determined from the slope of a line tangent to the linear portion of the time course was converted to an observed rate constant, kobs, by dividing by the

concentration of enzyme used in the assay. Upon completion of experimental

data from time courses with higher concentrations of peptide, kinetic parameters

will be determined by fitting the kobs data versus concentration of peptide to equation 5.2

k * S n k  cat Equation 5.2 obs n n K m  S

Where kobs is the observed rate constant in per second, kcat is the maximal rate

constant in per second, S is the concentration of peptide in micromolar, n is the

Hill coefficient, and Km is the concentration of substrate at half maximal rate in

micromolar.

To identify peptides of interest from a library of peptides containing alanine

substitutions for each residue of FRETN 89-98 (79), continuous fluorescent 104

peptidase activity assays were done as described above with 50 µM peptide and

1 mM ATP in the presence of 900 nM ClpX and 300 nM ClpP or 300 nM hLon.

To compare the peptides in the library, the rate of cleavage of each alanine

substituted peptide was normalized by dividing by the rate of cleavage of the lead

peptide FRETN 89-98.

ATP-dependent cleavage of fluorescently labeled Cleptide. Reactions

containing 50 mM HEPES (pH 8), 5 mM Mg(OAc)2, 2 mM DTT, 900 nM ClpX,

300 nM ClpP and 1 mM ATP were initiated with varying amounts of fluorescently labeled Cleptide (14) and peptide cleavage was monitored for 900 sec at 37 °C.

The amount of peptide hydrolyzed was measure by determining the maximum fluorescence generated per micromolar peptide after complete digestion with chymotrypsin. The steady-state rate of peptide cleavage was determined by the slope of a line tangent to the linear phase of the time course which was converted to kobs by dividing by the enzyme concentration used in the assay.

105

Results and Discussion

Determination of kinetic parameters for ClpX ATPase activity. As the storage

buffer of ClpXP contains 200 mM imidazole, I tested what effect the

concentration of imidazole had on ClpXP ATPase activity. As shown in Figure

5.1, the addition of 40 mM imidazole had no influence on the rate of ATP

hydrolysis by ClpX.

As the ATPase component, ClpX, complexes with the proteolytic

component, ClpP, to form the functional protease, I wanted to know what effect

the presence of ClpP has on the ATPase activity of ClpX. To determine this, I

varied the amount of ClpP in the reaction and compared the resulting rates of

ATP hydrolysis. As shown in Figure 5.2, the amount of ClpP in the reaction has no effect on the rate of ATP hydrolysis.

For the determination of kinetic parameters kcat and Km for ATP hydrolysis

by ClpXP, I used the radioactive steady-state ATPase assay. Rates of ClpXP

ATPase activity were determined as described in Materials and Methods, and

converted to kobs values by dividing by the concentration of ClpX. These observed rate constants were plotted against initial ATP concentration and the data was fit to the Hill equation as data assumed a sigmoidal function (Figure

-1 5.3). Fitting of this data provided a kcat of 2 ± 0.2 sec , a Km of 330 ± 40 µM, and

a Hill coefficient of 1.5 ± 0.2. A Hill coefficient greater than 1 suggests positive

; binding of ATP to one active site of the ClpX complex affects the

binding and hydrolysis of ATP of the other active sites.

106

Figure 5.1: Imidazole does not affect the ATPase activity of ClpX. The rate of

ATP hydrolysis by ClpX was determined in the presence of an additional 40 mM imidazole, which did not affect

107

Figure 5.2: Rate of ATPase activity by ClpX is not significantly affected by the amount of ClpP. The rates of radiolabeled-ATP hydrolysis by 300 nM ClpX in the presence of varying amounts of ClpP were determined by PhosphorImaging as described in Materials and Methods.

108

Figure 5.3: Determination of kinetic parameters for ATP hydrolysis by ClpXP.

The observed rate constants from time courses in the presence of varying amounts of ATP were fit to the Hill equation to yield kcat and Km values of 2 ± 0.1 sec-1 and 330 ± 40 µM, respectively.

109

ClpXP cleavage of FRETN 89-98. Observed rate constants for cleavage of

FRETN 89-98 by ClpXP were determined by dividing the rate of the time course

by the concentration of ClpP used in the assay as previously described. At this

point, the Vmax does not seem to be reached, therefore, reactions with higher

concentrations of FRETN 89-98 will have to be done in order to determine a kcat

and Km for cleavage by ClpXP. These kinetic parameters can be compared to

those for human Lon. As I have already shown that under certain experimental

conditions, ClpXP cleaves FRETN 89-98 less efficiently than Lon, I expect the

specificity constant, defined as kcat/Km, for ClpXP to be less than that of human

Lon, which has been published to be 4.5 x 103 M-1sec-1 (42).

Alanine scan peptides with ClpXP. A library of peptides have been synthesized

in our lab by substituting each residue of our lead peptide 89-98FR with alanine.

This library has been used to identify the residues important for eLon activity (79)

and led to the suggestion that I91, C93 and R96 are necessary for peptide

recognition by Lon. As a peptide with alanine substitutions at each of these

residues was not cleaved by bacterial Lon, it provided a possible peptide that

would be specifically cleaved by ClpXP and could be used in mixtures of ClpXP

and Lon, such as in mitochondrial extract, to selectively monitor ClpXP activity.

Each peptide was synthesized with the Abz/nitrotyrosine fluorescent

donor/quencher pair for use with our continuous steady state peptidase assay.

Steady state time courses were run with both hLon and ClpXP in the presence of

110

ATP and the rates were normalized to that of FRETN 89-98 (Figure 5.4). This

study identified peptides that were cleaved at a higher efficiency by ClpXP than

by both bacterial and human Lon and at a higher efficiency than FRETN 89-98

(R89A, G90A, G95A, R96A). As was predicted, the peptide with the triple alanine

substitution which was not cleaved by bacterial Lon was also not cleaved by

human Lon, but was cleaved by ClpXP. We have named this peptide hClp001 as

it is the first ClpXP-specific peptide substrate identified in our lab. Although this

peptide is cleaved with less efficiency than FRETN 89-98, it does provide a

specific peptide substrate that can be used to monitor the activity of ClpXP in

isolated mitochondria in the future.

Cleavage of fluorescently labeled Cleptide. The sequence of Cleptide

(FAPHMALVPV, Figure 5.5) has been identified by Maurizi and others (14) as a substrate of ClpXP. We have labeled this sequence with a Lys(Abz) group on the C-terminus and a (3-NO2)Tyr on the N-terminal for use in our continuous steady-state peptidase assay. As shown in Figure 5.5, this peptide is indeed cleaved by ClpXP, but not by human Lon in the presence of ATP. This result provides us with another ClpXP-specific peptide substrate whose sequence can be exploited to design specific inhibitors for the enzyme. As we are interested in the kinetic parameters of cleavage of this peptide by ClpXP, I have begun experiments to determine kcat and Km values. As shown in Figure 5.6, I have not

yet tested a concentration of peptide where Vmax has been reached, therefore, a

non-fluorescent analogue of the Cleptide will have to be synthesized and used in

111

Figure 5.4: Screening FRETN 89-98 alanine scan peptides for cleavage by

ClpXP (black), human Lon (gray) and E. coli Lon (white) (79). The peptide with alanine substitutions at I91, C93, and R96 is cleaved by ClpXP, but not human or bacterial Lon, though less efficiently than FRETN 89-98.

112

Y(NO )-FAPHMALVPV-K(Abz) 2

ClpXP + ATP

hLon + ATP

ClpXP - ATP

Figure 5.5: Fluorescently labeled Cleptide is selectively cleaved by ClpXP in the presence of ATP. 50 µM peptide was cleaved by 300 nM enzyme in the presence

(ClpXP: black, hLon: light gray) and absence (dark gray) of 1 mM ATP.

113

Figure 5.6: Cleavage of Cleptide by ClpXP can be monitored using the

fluorescent peptidase assay. kobs values were determined by taking the slope of a line tangent to the linear phase of the cleavage time course in per sec and converted to µM/sec by normalizing with the fluorescence upon complete digestion with chymotrypsin then dividing by enzyme concentration. A complete kinetic workup will be done to determine kcat and Km values for Cleptide cleavage.

114

combination with the fluorescent peptide to monitor cleavage at higher concentrations of peptide while avoiding the inner filter effect.

115

Current status and future directions. To this point, I have begun experiments to decipher the mechanism of ClpXP ATPase and proteolytic activities. In addition to the peptide substrates that have been previously used in our lab to study Lon mechanism (FRETN 89-98 and FRETN 89-98Abu, Table 5.1), I have identified two ClpXP specific peptides (hClp001 and Cleptide, Table 5.1), and begun the steady-state characterization of cleavage of the Cleptide by ClpXP.

The results of the casein digest analyzed by mass spectrometry will allow for the design of additional specific peptide substrates, as well as inhibitors that can be used in isolated mitochondria and live cells as I have used DBN93 to inhibit Lon as described in Chapter 3. The use of chemical tools such as peptide substrates and inhibitors should contribute to the identification of physiologically relevant protein substrates of each enzyme and the mechanisms by which they degrade these proteins to protect the integrity of the mitochondria and overall health of the cell.

The experiments utilizing the fluorescent peptide substrates and inhibitors to monitor peptidase activity in isolated mitochondria and live cell survival are the first of their kind to be done in our lab. It was an interesting change of pace to be able to take the results of experiments done on purified proteins in a cuvette and apply them to physiologically relevant models. One drawback to using isolated

mitochondria is that so much protein is needed to get a good signal on the

fluorometer that a large amount of HeLa cells have to be grown to isolate enough

mitochondria. The design of a peptide substrate that is more efficiently cleaved

by human Lon should give a better signal and require less mitochondria. As of

116

Cleaved Cleaved Peptide Name Sequence by by hLon? hClpXP?

FRETN 89-98 Y(NO2)-RGITCSGRQ-K(Abz) Yes Yes

FRETN 89- Y(NO )-RGIT-Abu-SGRQ-K(Abz) Yes No 98Abu 2

hClp001 Y(NO2)-RGATASGAQ-K(Abz) No Yes

Cleptide* Y(NO2)-FAPHMALVPV-K(Abz) No Yes

Table 5.1: Enzyme-specific peptide substrates.

117

right now, ClpXP peptidase activity is not able to be detected from the isolated mitochondria, partly due to the lower amount of ClpXP in the mitochondria compared to Lon and the absence of a peptide that is cleaved efficiently enough by ClpXP to be able to see fluorescent signal. Hopefully using the fluorescently labeled Cleptide will allow for detection of ClpXP activity.

In summary, the results I have presented in the previous chapters provide key information towards the development of a mechanism for proteolysis by human Lon and ClpXP. The chemical tools that have been identified and characterized can be used in proteomic studies to identify physiologically relevant protein substrates of each enzyme. Discovery of what proteins Lon and

ClpXP function to remove after oxidative damage due to aging or disease will certainly advance our knowledge of not only the individual enzymes, but the cause of the disease in general, which would lead to more effective drug design to prevent or treat cell damage.

118

BIBLIOGRAPHY

1. Friguet, B., Bulteau, A.-L., and Petropoulos, I. (2008) Mitochondrial protein

quality control: Implications in ageing, Biotechnology Journal 3, 757-764.

2. Bulteau, A. L., Szweda, L. I., and Friguet, B. (2006) Mitochondrial protein

oxidation and degradation in response to oxidative stress and aging, Exp

Gerontol 41, 653-657.

3. Shastry, B. S. (2003) Neurodegenerative disorders of protein aggregation,

Neurochem Int 43, 1-7.

4. Jenner, P. (2003) Oxidative stress in Parkinson's disease, Annals of

Neurology 53, S26-S38.

5. Bowling, A. C., and Beal, M. F. (1995) Bioenergetic and oxidative stress in

neurodegenerative diseases, Life Sciences 56, 1151-1171.

6. Schapira, A. (1996) Oxidative stress and mitochondrial dysfunction in

neurodegeneration., Current Opinion in Neurology 9, 260-264.

7. Beckman, K. B., and Ames, B. N. (1998) The Free Radical Theory of

Aging Matures, Physiol. Rev. 78, 547-581.

8. Gottesman, S., and Maurizi, M. R. (1992) Regulation by proteolysis:

energy-dependent proteases and their targets, Microbiol. Rev. 56, 592-

621.

9. Rosen, R., Biran, D., Gur, E., Becher, D., Hecker, M., and Ron, E. (2002)

Protein aggregation in Escherichia coli: role of proteases., FEMS Microbiol

Lett 207, 9-12.

119

10. Corydon, T. J., Wilsbech, M., Jespersgaard, C., Andresen, B. S., Borglum,

A. D., Pedersen, S., Bolund, L., Gregersen, N., and Bross, P. (2000)

Human and mouse mitochondrial orthologs of bacterial ClpX, Mamm

Genome 11, 899-905.

11. de Sagarra, M. R., Mayo, I., Marco, S., Rodriguez-Vilarino, S., Oliva, J.,

Carrascosa, J. L., and Casta n, J. G. (1999) Mitochondrial localization and

oligomeric structure of HClpP, the human homologue of E. coli ClpP, J

Mol Biol 292, 819-825.

12. Wang, N., Gottesman, S., Willingham, M. C., Gottesman, M. M., and

Maurizi, M. R. (1993) A human mitochondrial ATP-dependent protease

that is highly homologous to bacterial Lon protease, Proc. Natl. Acad. Sci.

USA 90, 11247-11251.

13. Kang, S. G., Dimitrova, M. N., Ortega, J., Ginsburg, A., and Maurizi, M. R.

(2005) Human mitochondrial ClpP is a stable heptamer that assembles

into a tetradecamer in the presence of ClpX, J Biol Chem 280, 35424-

35432.

14. Kang, S. G., Ortega, J., Singh, S. K., Wang, N., Huang, N. N., Steven, A.

C., and Maurizi, M. R. (2002) Functional proteolytic complexes of the

human mitochondrial ATP-dependent protease, hClpXP, J Biol Chem 277,

21095-21102.

15. van Dyck, L., and Langer, T. (1999) ATP-dependent proteases controlling

mitochondrial function in the yeast Saccharomyces cerevisiae., Cell Mol.

Life Sci. 56, 825-842.

120

16. Davies, K. J., and Lin, S. W. (1988) Oxidatively denatured proteins are

degraded by an ATP-independent proteolytic pathway in Escherichia coli,

Free Radic Biol Med 5, 225-236.

17. Käser, M., and Langer, T. (2000) Protein degradtion in mitochondria, Cell

Dev. Biol 11, 181-190.

18. van Dyck, L., Pearce, D. A., and Sherman, F. (1994) PIM1 encodes a

mitochondrial ATP-dependent protease that is required for mitochondrial

function in the yeast Saccharomyces cerevisiae, J. Biol. Chem. 269, 238-

242.

19. Wagner, I., Arlt, H., van Dyck, L., Langer, T., and Neupert, W. (1994)

Molecular chaperones cooperate with PIM1 protease in the degradation of

misfolded proteins in mitochondria, EMBO J. 13, 5135-5145.

20. Langer, T., and Neupert, W. (1996) Regulated protein degradation in

mitochondria, Experentia 52, 1019-1157.

21. Rep, M., and Grivell, L. A. (1996) The role of protein degradation in

mitochondrial function and biogenesis, Curr. Genet. 30, 367-380.

22. van Dijl, J. M., Kutejova, E., Suda, K., Perecko, D., Schatz, G., and

Suzuki, C. K. (1998) The ATPase and protease domains of yeast

mitochondrial Lon: roles in proteolysis and respiration-dependent growth,

Proc Natl Acad Sci U S A 95, 10584-10589.

23. van Dyck, L., Neupert, W., and Langer, T. (1998) The ATP-dependent

PIM1 protease is required for the expression of intron-containing in

mitochondria, Genes Dev. 12, 1515-1524.

121

24. Bota, D. A., and Davies, K. J. (2001) Protein degradation in mitochondria:

implications for oxidative stress, aging and disease: a novel etiological

classification of mitochondrial proteolytic disorders, Mitochondrion 1, 33-

49.

25. Bulteau, A. L., Lundberg, K. C., Ikeda-Saito, M., Isaya, G., and Szweda, L.

I. (2005) Reversible redox-dependent modulation of mitochondrial

aconitase and proteolytic activity during in vivo cardiac

ischemia/reperfusion, Proc. Natl. Acad. Sci. U S A 102, 5987-5991.

26. Maurizi, M. R. (1987) Degradation in vitro of bacteriophage lambda N

protein by Lon protease from Escherichia coli., J. Biol. Chem. 262, 2696-

2703.

27. Hanson, P. I., and Whiteheart, S. W. (2005) AAA+ proteins: have engine,

will work., Nature Rev. Mol. Cell. Biol. 6, 519-529.

28. He, T. C., Jiang, N., Zhuang, H., and Wojchowski, D. M. (1995)

Erythropoietin-induced recruitment of Shc via a receptor phosphotyrosine-

independent, Jak2-associated pathway, J. Biol. Chem. 270, 11055-11061.

29. Neuwald, A. F., Aravind, L., Spouge, J. L., and Koonin, E. V. (1999) AAA+:

A Class of Chaperone-Like ATPases Associated with the Assembly,

Operation, and Disassembly of Protein Complexes, Genome Res. 9, 27-

43.

30. Ogura, T., and Wilkinson, A. J. (2001) AAA+ superfamily ATPases:

common structure--diverse function, Genes Cells 6, 575-597.

122

31. Smith, C. K., Baker, T. A., and Sauer, R. T. (1999) Lon and Clp family

proteases and chaperones share homologous substrate- recognition

domains, Proc Natl Acad Sci U S A 96, 6678-6682.

32. Li, M., Rasulova, F., Melnikov, E. E., Rotanova, T. V., Gustchina, A.,

Maurizi, M. R., and Wlodawer, A. (2005) Crystal structure of the N-

terminal domain of E. coli Lon protease, Protein Sci 14, 2895-2900.

33. Suzuki, C. K., Suda, K., Wang, N., and Schatz, G. (1994) Requirement for

the yeast gene LON in intramitochondrial proteolysis and maintenance of

respiration, Science 264, 273-276 and 891.

34. Lupas, A., Flanagan, J. M., Tamura, T., and Baumeister, W. (1997) Self-

compartmentalizing proteases, Trends Biochem Sci 22, 399-404.

35. Park, S. C., Jia, B., Yang, J. K., Van, D. L., Shao, Y. G., Han, S. W., Jeon,

Y. J., Chung, C. H., and Cheong, G. W. (2006) Oligomeric structure of the

ATP-dependent protease La (Lon) of Escherichia coli, Mol Cells 21, 129-

134.

36. Rudyak, S. G., Brenowitz, M., and Shrader, T. E. (2001) Mg2+-linked

oligomerization modulates the catalytic activity of the Lon (La) protease

from Mycobacterium smegmatis, Biochemistry 40, 9317-9323.

37. Rotanova, T. V., Botos, I., Melnikov, E. E., Rasulova, F., Gustchina, A.,

Maurizi, M. R., and Wlodawer, A. (2006) Slicing a protease: structural

features of the ATP-dependent Lon proteases gleaned from investigations

of isolated domains, Protein Sci 15, 1815-1828.

123

38. Gottesman, S., Clark, W. P., de Crecy-Lagard, V., and Maurizi, M. R.

(1993) ClpX, an alternative subunit for the ATP-dependent Clp protease of

Escherichia coli. Sequence and in vivo activities., J. Biol. Chem. 268,

22618-226226.

39. Kenniston, J. A., Baker, T. A., Fernandez, J. M., and Sauer, R. T. (2003)

Linkage between ATP consumption and mechanical unfolding during the

protein processing reactions of an AAA+ degradation machine, Cell 114,

511-520.

40. Lee, I., and Berdis, A. J. (2001) Adenosine triphosphate-dependent

degradation of a fluorescent lambda N substrate mimic by Lon protease.,

Anal. Biochem. 291, 74-83.

41. Teichmann, U., van Dyck, L., Guiard, B., Fischer, H., Glockshuber, R.,

Neupert, W., and Langer, T. (1996) Substitution of PIM1 protease in

Mitochondria by Escherichia coli Lon protease, J. Biol. Chem. 271, 10137-

10142.

42. Frase, H., Hudak, J., and Lee, I. (2006) Identification of the proteasome

inhibitor MG262 as a potent ATP-dependent inhibitor of the Salmonella

enterica serovar Typhimurium Lon protease, Biochemistry 45, 8264-8274.

43. Thomas-Wohlever, J., and Lee, I. (2002) Kinetic characterization of the

peptidase activity of Escherichia coli Lon reveals the mechanistic

similarities in ATP-dependent hydrolysis of peptide and protein substrates,

Biochemistry 41, 9418-9425.

124

44. Vineyard, D., Patterson-Ward, J., Berdis, A. J., and Lee, I. (2005)

Monitoring the timing of ATP hydrolysis with activation of peptide cleavage

in Escherichia coli Lon by transient kinetics, Biochemistry 44, 1671-1682.

45. Matsui, H., Suzuki, M., Isshiki, Y., Kodama, C., Eguchi, M., Kikuchi, Y.,

Motokawa, K., Takaya, A., Tomoyasu, T., and Yamamoto, T. (2003) Oral

immunization with ATP-dependent protease-deficient mutants protects

mice against subsequent oral challenge with virulent Salmonella enterica

serovar typhimurium, Infect Immun 71, 30-39.

46. Frase, H., and Lee, I. (2007) Peptidyl boronates inhibit Salmonella

enterica serovar Typhimurium Lon protease by a competitive ATP-

dependent mechanism, Biochemistry 46, 6647-6657.

47. Bulteau, A. L., Ikeda-Saito, M., and Szweda, L. I. (2003) Redox-dependent

modulation of aconitase activity in intact mitochondria, Biochemistry 42,

14846-14855.

48. Bradford, M. M. (1976) A rapid and senstive method for the quantitation of

microgram quantities of protein utilizing the principle of protein-dye

binding., Anal Biochem 72, 248-254.

49. Gill, S. C., and von Hippel, P. H. (1989) Calculation of Protein Extinction

Coefficients from Amino Acid Sequence Data, Analytical Biochemistry

182, 319-326.

50. Cleland, W. W. (1979) Statistical analysis of enzyme kinetic data, Methods

Enzymol 63, 103-138.

125

51. Copeland, R. A. (2005) Evaluation of Enzyme Inhibitors in Drug Discovery:

A Guide for Medicinal Chemists and Pharmacologists, John Wiley & Sons,

Hoboken, NJ.

52. Kuzmic, P. (1996) Program DYNAFIT for the analysis of enzyme kinetic

data: application to HIV proteinase, Anal Biochem 237, 260-273.

53. Granot, Z., Kobiler, O., Melamed-Book, N., Eimerl, S., Bahat, A., Lu, B.,

Braun, S., Maurizi, M. R., Suzuki, C. K., Oppenheim, A. B., and Orly, J.

(2007) Turnover of mitochondrial steroidogenic acute regulatory (StAR)

protein by Lon protease: the unexpected effect of proteasome inhibitors,

Mol Endocrinol 21, 2164-2177.

54. Ondrovicova, G., Liu, T., Singh, K., Tian, B., Li, H., Gakh, O., Perecko, D.,

Janata, J., Granot, Z., Orly, J., Kutejova, E., and Suzuki, C. K. (2005)

Cleavage Site Selection within a Folded Substrate by the ATP-dependent

Lon Protease, J. Biol. Chem. 280, 25103-25110.

55. Goldberg, A. L., Moerschell, R. P., Chung, C. H., and Maurizi, M. R.

(1994) ATP-dependent protease La (Lon) from Escherichia coli, Meth.

Enzymol. 244, 350-375.

56. Waxman, L., and Goldberg, A. L. (1986) Selectivity of intracellular

proteolysis: protein substrates activate the ATP-dependent protease (La),

Science 232, 500-503.

57. Goldberg, A. L., and Waxman, L. (1985) The role of ATP hydrolysis in the

breakdown of proteins and peptides by protease La from Escherichia coli,

J Biol Chem 260, 12029-12034.

126

58. Menon, A. S., Waxman, L., and Goldberg, A. L. (1987) The energy utilized

in protein breakdown by the ATP-dependent protease (La) from

Escherichia coli, J Biol Chem 262, 722-726.

59. Gur, E., and Sauer, R. T. (2008) Recognition of misfolded proteins by Lon,

a AAA(+) protease, Genes & Development 22, 2267-2277.

60. Maurizi, M. R. (1992) Proteases and protein degradation in Escherichia

coli, Experentia 48, 178-201.

61. Tatsuta, T., and Langer, T. (2009) AAA proteases in mitochondria: diverse

functions of membrane-bound proteolytic machines, Research in

Microbiology 160, 711-717.

62. Bulteau, A. L., Lundberg, K. C., Ikeda-Saito, M., Isaya, G., and Szweda, L.

I. (2005) Reversible redox-dependent modulation of mitochondrial

aconitase and proteolytic activity during in vivo cardiac

ischemia/reperfusion, Proc Natl Acad Sci U S A 102, 5987-5991.

63. Sweetlove, L. J., Heazlewood, J. L., Herald, V., Holtzapffel, R., Day, D. A.,

Leaver, C. J., and Millar, A. H. (2002) The impact of oxidative stress on

Arabidopsis mitochondria, Plant J 32, 891-904.

64. Bota, D. A., and Davies, K. J. (2002) Lon protease preferentially degrades

oxidized mitochondrial aconitase by an ATP-stimulated mechanism, Nat

Cell Biol 4, 674-680.

65. Bota, D. A., Van Remmen, H., and Davies, K. J. (2002) Modulation of Lon

protease activity and aconitase turnover during aging and oxidative stress,

FEBS Lett 532, 103-106.

127

66. Masters, J. R. (2002) HeLa cells 50 years on: the good, the bad and the

ugly, Nat Rev Cancer 2, 315-319.

67. Patterson-Ward, J., Huang, J., and Lee, I. (2007) Detection and

characterization of two ATP-dependent conformational changes in

proteolytically inactive Escherichia coli lon mutants by stopped flow kinetic

techniques, Biochemistry 46, 13593-13605.

68. Maurizi, M. R., Clark, W. P., Kim, S.-H., and Gottesman, S. (1990) Clp P

Represents a Unique Family of Serine Proteases, J Biol Chem 265,

12546-12552.

69. Matthews, D. A., Alden, R. A., Birktoft, J. J., Freer, S. T., and Kraut, J.

(1975) X-ray crystallographic study of boronic acid adducts with subtilisin

BPN' (Novo). A model for the catalytic transition state, J. Biol. Chem. 250,

7120-7126.

70. Matteson, D. S., Sadhu, K. M., and Lienhard, G. E. (1981) R-1-Acetamido-

2-phenylethaneboronic acid. A specific transition-state analog for

chymotrypsin J. Am. Chem. Soc 103, 5241-5242.

71. Kinder, D. H., and Katzenellenbogen, J. A. (1985) Acylamido boronic

acids and difluoroborane analogs of amino acids: potent inhibitors of

chymotrypsin and elastase, J. Med. Chem. 28, 1917-1925.

72. Shenvi, A. B. (2002) a-Aminoboronic acid derivatives: effective inhibitors

of , Biochemistry 25, 1286-1291.

128

73. Kettner, C. A., and Shenvi, A. B. (1984) Inhibition of the serine proteases

leukocyte elastase, pancreatic elastase, cathepsin G, and chymotrypsin

by peptide boronic acids, J. Biol. Chem. 259, 15106-15114.

74. Knott, K., Fishovitz, J., Thorpe, S. B., Lee, I., and Santos, W. L. (2010) N-

Terminal peptidic boronic acids selectively inhibit human ClpXP, Organic

& Biomolecular Chemistry 8, 3451-3456.

75. Copeland, R. A. (2000) Enzymes: A Practical Introduction to Structure,

Mechanism, and Data Analysis, 2nd ed., John Wiley & Sons, Inc., New

York.

76. Frase, H., Hudak, J., and Lee, I. (2006) Identification of the proteasome

inhibitor MG262 as a potent ATP-dependent inhibitor of Salmonella

enterica serovar Typhimurium Lon protease, Biochemistry 45, 8264-8274.

77. Kang, S. G., Ortega, J., Singh, S. K., Wang, N., Huang, N. N., Steven, A.

C., and Maurizi, M. R. (2002) Functional proteolytic complexes of the

human mitochondrial ATP-dependent protease, hClpXP, J. Biol. Chem.

277, 21095-21102.

78. Wang, N., Gottesman, S., Willingham, M. C., Gottesman, M. M., and

Maurizi, M. R. (1993) A human mitochondrial ATP-dependent protease

that is highly homologous to bacterial Lon protease, Proc. Natl. Acad. Sci.

USA 90, 11247-11251.

79. Patterson-Ward, J., Tedesco, J., Hudak, J., Fishovitz, J., Becker, J.,

Frase, H., McNamara, K., and Lee, I. (2009) Utilization of sunthetic

peptides to evaluate the importance of substrate interaction at the

129

proteolytic site of Escherichia coli Lon protease, Biochim Biophys Acta

1794, 1355-1363.

130