A Study of Peptide Peptide Interaction by Matrix-Assisted Laser Desorption/Ionization

Similar documents
Biological Mass Spectrometry. April 30, 2014

Molecular Biology. general transfer: occurs normally in cells. special transfer: occurs only in the laboratory in specific conditions.

Properties of amino acids in proteins

Objective: You will be able to explain how the subcomponents of

This exam consists of two parts. Part I is multiple choice. Each of these 25 questions is worth 2 points.

Matrix Assisted Laser Desorption Ionization Time-of-flight Mass Spectrometry

PAPER No. : 16, Bioorganic and biophysical chemistry MODULE No. : 22, Mechanism of enzyme catalyst reaction (I) Chymotrypsin

Copyright 2008 Pearson Education, Inc., publishing as Pearson Benjamin Cummings

Biological systems interact, and these systems and their interactions possess complex properties. STOP at enduring understanding 4A

PHAR3316 Pharmacy biochemistry Exam #2 Fall 2010 KEY

CHAPTER 21: Amino Acids, Proteins, & Enzymes. General, Organic, & Biological Chemistry Janice Gorzynski Smith

Chemical Nature of the Amino Acids. Table of a-amino Acids Found in Proteins

CS612 - Algorithms in Bioinformatics

Bioinformatics for molecular biology

Proteins are sometimes only produced in one cell type or cell compartment (brain has 15,000 expressed proteins, gut has 2,000).

TECHNICAL BULLETIN. R 2 GlcNAcβ1 4GlcNAcβ1 Asn

AP Bio. Protiens Chapter 5 1

Gentilucci, Amino Acids, Peptides, and Proteins. Peptides and proteins are polymers of amino acids linked together by amide bonds CH 3

PROTEINS. Amino acids are the building blocks of proteins. Acid L-form * * Lecture 6 Macromolecules #2 O = N -C -C-O.

Introduction to proteins and protein structure

The Structure and Function of Large Biological Molecules Part 4: Proteins Chapter 5

Characterization of Noncovalent Complexes of Polyelectrolytes by Mass Spectrometry

Biomolecules: amino acids

2. Ionization Sources 3. Mass Analyzers 4. Tandem Mass Spectrometry

Chemical Mechanism of Enzymes

PROTEINS. Building blocks, structure and function. Aim: You will have a clear picture of protein construction and their general properties

Amino acids. (Foundation Block) Dr. Essa Sabi

MALDI-TOF. Introduction. Schematic and Theory of MALDI

Charges on amino acids and proteins. ph 1. ph 7. Acidic side chains: glutamate and aspartate

Chemistry 121 Winter 17

Ionization Methods. Neutral species Charged species. Removal/addition of electron(s) Removal/addition of proton(s)

Draw how two amino acids form the peptide bond. Draw in the space below:

ESI and MALDI Mass Spectrometry. S. Sankararaman Department of Chemistry Indian Institute of Technology Madras Chennai

Methionine (Met or M)

Previous Class. Today. Detection of enzymatic intermediates: Protein tyrosine phosphatase mechanism. Protein Kinase Catalytic Properties

Short polymer. Dehydration removes a water molecule, forming a new bond. Longer polymer (a) Dehydration reaction in the synthesis of a polymer

Biomolecular Mass Spectrometry

Introduction to Peptide Sequencing

! Proteins are involved functionally in almost everything: " Receptor Proteins - Respond to external stimuli. " Storage Proteins - Storing amino acids

Lecture 3: 8/24. CHAPTER 3 Amino Acids

WHY IS THIS IMPORTANT?

Amino Acids and Proteins Hamad Ali Yaseen, PhD MLS Department, FAHS, HSC, KU Biochemistry 210 Chapter 22

Biochemistry - I. Prof. S. Dasgupta Department of Chemistry Indian Institute of Technology, Kharagpur Lecture 1 Amino Acids I

From Atoms to Cells: Fundamental Building Blocks. Models of atoms. A chemical connection

1-To know what is protein 2-To identify Types of protein 3- To Know amino acids 4- To be differentiate between essential and nonessential amino acids

Amino Acids. Review I: Protein Structure. Amino Acids: Structures. Amino Acids (contd.) Rajan Munshi

The Basics: A general review of molecular biology:

Organic Molecules: Proteins

Macromolecules of Life -3 Amino Acids & Proteins

Comparison of mass spectrometers performances

Robert J. Cotter Department of Pharmacology and Molecular Sciences, Johns Hopkins University School of Medicine, Baltimore, Maryland, USA

Chapter 3: Amino Acids and Peptides

2. Which of the following amino acids is most likely to be found on the outer surface of a properly folded protein?

Human Biochemistry Option B

For questions 1-4, match the carbohydrate with its size/functional group name:

Levels of Protein Structure:

AP Biology. Proteins. Proteins. Proteins. Amino acids H C OH H R. Effect of different R groups: Polar amino acids polar or charged & hydrophilic

Secondary Structure North 72nd Street, Wauwatosa, WI Phone: (414) Fax: (414) dmoleculardesigns.com

Additional problems: 1. Match and label the conjugate acid and base pairs in the following reactions. Which one of these systems is a good buffer?

Mechanisms of Enzymes

Proteins. Proteins. Proteins. Proteins. Effect of different R groups: Nonpolar amino acids. Amino acids H C OH H R. Multipurpose molecules.

Electronic Supplementary Information

Sequence Identification And Spatial Distribution of Rat Brain Tryptic Peptides Using MALDI Mass Spectrometric Imaging

Lecture 3. Tandem MS & Protein Sequencing

Understand how protein is formed by amino acids

METHODS AND REVIEWS MATRIX ASSISTED LASER DESORPTION IONIZATION TIME-OF-FLIGHT MASS SPECTROMETRY

Proteins. Amino acids, structure and function. The Nobel Prize in Chemistry 2012 Robert J. Lefkowitz Brian K. Kobilka

[ Care and Use Manual ]

paper and beads don t fall off. Then, place the beads in the following order on the pipe cleaner:

The Structure and Function of Macromolecules

Structure of -amino acids. Stereoisomers of -amino acids. All amino acids in proteins are L-amino acids, except for glycine, which is achiral.

MBB 694:407, 115:511. Please use BLOCK CAPITAL letters like this --- A, B, C, D, E. Not lowercase!

1. Describe the difference between covalent and ionic bonds. What are the electrons doing?

Biological Molecules B Lipids, Proteins and Enzymes. Triglycerides. Glycerol

9/6/2011. Amino Acids. C α. Nonpolar, aliphatic R groups

UNIVERSITY OF GUELPH CHEM 4540 ENZYMOLOGY Winter 2005 Quiz #2: March 24, 2005, 11:30 12:50 Instructor: Prof R. Merrill ANSWERS

Details of Organic Chem! Date. Carbon & The Molecular Diversity of Life & The Structure & Function of Macromolecules

BIO 311C Spring Lecture 15 Friday 26 Feb. 1

Lecture 15. Membrane Proteins I

Agilent Protein In-Gel Tryptic Digestion Kit

BCHS 3304/ Exam I. September 26,

For questions 1-4, match the carbohydrate with its size/functional group name:

Page 8/6: The cell. Where to start: Proteins (control a cell) (start/end products)

Protein Modeling Event

Reactions and amino acids structure & properties

Solving practical problems. Maria Kuhtinskaja

Chymotrypsin Lecture. Aims: to understand (1) the catalytic strategies used by enzymes and (2) the mechanism of chymotrypsin

Double charge of 33kD peak A1 A2 B1 B2 M2+ M/z. ABRF Proteomics Research Group - Qualitative Proteomics Study Identifier Number 14146

Amino Acids. Lecture 4: Margaret A. Daugherty. Fall Swiss-prot database: How many proteins? From where?

Acid/Base chemistry. NESA Biochemistry Fall 2001 Review problems for the first exam. Complete the following sentences

Arginine side chain interactions and the role of arginine as a mobile charge carrier in voltage sensitive ion channels. Supplementary Information

Head. Tail. Carboxyl group. group. group. air water. Hydrocarbon chain. lecture 5-sa Seth Copen Goldstein 2.

Proteins. AP Biology. Proteins. Proteins. Proteins. Effect of different R groups: Nonpolar amino acids. Amino acids H C OH H R. Structure.

Biochemistry Prof. S. Dasgupta Department of Chemistry Indian Institute of Technology Kharagpur. Lecture -02 Amino Acids II

9/16/15. Properties of Water. Benefits of Water. More properties of water

Secondary Structure. by hydrogen bonds

Supporting information

Globular proteins Proteins globular fibrous

Biological Mass spectrometry in Protein Chemistry

Phenylketonuria (PKU) Structure of Phenylalanine Hydroxylase. Biol 405 Molecular Medicine

Transcription:

A Study of Peptide Peptide Interaction by Matrix-Assisted Laser Desorption/Ionization Amina S. Woods and Marilyn A. Huestis Chemistry and Drug Metabolism, NIDA Intramural Research Program, NIDA, NIH, Baltimore, Maryland, USA Matrix-assisted laser desorption/ionization (MALDI) mass spectrometry was used to study peptide peptide interaction. The interaction was seen when 6-aza-2-thiothymine was used as a matrix (ph 5.4), but was disrupted with a more acidic matrix, -cyano-4-hydroxycinnamic acid (ph 2.0). In the present study, we show that dynorphin, an opioid peptide, and five of its fragments that contain two adjacent basic residues (Arg 6 -Arg 7 ), all interact noncovalently with peptides that contain two to five adjacent acidic residues (Asp or Glu). Two other nonrelated peptides containing two (Arg 6 -Arg 7 ) or three (Arg 1 -Lys 2 -Arg 3 ) adjacent basic amino acid residues were studied and exhibited the same behavior. However, peptides containing adjacent Lys or His did not form noncovalent complexes with acidic peptides. The noncovalent bonding was sufficiently stable that digestion with trypsin only cleaved Arg and Lys residues that were not involved in hydrogen bonding with the acidic residues. In an equimolar mixture of dynorphin, dynorphin fragments (containing the motif RR), and an acidic peptide (minigastrin), the acidic peptide preferentially complexed with dynorphin. If the concentration of minigastrin was increased 10 fold, noncovalent interaction was seen with dynorphin and all its fragments containing the motif RR. In the absence of dynorphin, minigastrin formed noncovalent complexes with all dynorphin fragments. These findings suggest that conformation, equilibrium, and concentration do play a role in the occurrence of peptide peptide interaction. Observations from this study include: (1) ionic bonds were not disrupted by enzymatic digests, (2) conformation and concentration influenced complex formation, and (3) the complex did not form with fragments of dynorphin or unrelated peptides that did not contain the motifs RR or RKR, nor with a fragment of dynorphin where Arg 7 was mutated to a phenylalanine residue. These findings strongly suggest that peptide peptide interaction does occur, and can be studied by MALDI if near physiologic ph is maintained. (J Am Soc Mass Spectrom 2001, 12, 88 96) 2001 American Society for Mass Spectrometry Mass spectrometry is becoming the tool of choice to study biopolymers. Ganem et al. [1, 2] were the first to demonstrate that noncovalent compounds could be studied by electrospray ionization (ESI) mass spectrometry and Loo wrote an extensive review on the subject [3]. Although ESI is considered to be the method of choice for studying noncovalent interaction, matrix-assisted laser desorption/ionization (MALDI) provides an alternative simpler approach. In 1995, Woods et al. [4] showed that noncovalent interaction (zinc finger, enzyme substrate complexes) could be studied by MALDI if matrix and sample conditions were optimized to prevent hydrogen bonding disruption. Several laboratories have used MALDI to observe noncovalent interaction. Lecchi and Pannell [5] showed that a less acidic matrix [6-aza-2- thiothymine (ATT)] in the presence of ammonium citrate allowed the detection of intact double stranded Address reprint requests to Dr. Amina S. Woods, Chemistry and Drug Metabolism, Intramural Research Program, NIDA, NIH, 5500 Nathan Shock Drive, Baltimore, MD 21224. E-mail: awoods@intra.nida.nih.gov DNA. Biemann s and Vertes groups [6, 7] demonstrated the attachment between highly acidic DNA strands and basic peptides. Farmer and Caprioli also studied protein protein and peptide protein interaction, as well as publishing an excellent review on the subject [8]. Hillenkemp s group claimed that noncovalent interaction can only be seen by MALDI if you gather first shots of crystals that have not been previously irradiated [9]. Further work by Lin, Cotter, and Woods detected noncovalent interaction of single and double stranded DNA with peptides [10, 11]. The purpose of this study is to explore the interaction between basic peptides, such as dynorphin and its fragments [12], and acidic peptides such as minigastrin [13], and to identify the structural elements that drive the interaction. Dynorphin and its fragments are of great interest as they are alleged to act at aspartate and glutamate receptors in the nervous system [14]. We describe the conditions that govern the formation of noncovalent complexes between peptides. Another of our goals is to set up new techniques that can facilitate the study of such complexes. We will discuss the 2001 American Society for Mass Spectrometry. Published by Elsevier Science Inc. Received June 2, 2000 1044-0305/01/$20.00 Revised August 23, 2000 PII S1044-0305(00)00197-5 Accepted August 24, 2000

J Am Soc Mass Spectrom 2001, 12, 88 96 STUDY OF PEPTIDE PEPTIDE INTERACTION BY MALDI 89 Table 1. Peptide list Sequence MW IP a BB b Name YGGFLRRIRPKLKWNDQ 2147.5 11.36 1990 dynorphin YGGFLRRIRPKLK 1604.0 12.16 3260 dynorphin 1 13 YGGFLRRIRP 1234.5 12.16 2530 dynorphin 1 10 YGGFLRRIR 1137.4 12.16 2360 dynorphin 1 9 YGGFLRRI 981.2 11.05 3050 dynorphin 1 8 YGGFLRR 868.2 11.05 1600 dynorphin 1 7 YGGFLRF 859.0 11.05 3810 [phe 7 ]dynorphin 1 7 IRPKLKWNDQ 1297.5 9.99 390 dynorphin 8 17 RKRARKE 943.1 11.72 4110 basic peptide 1 QRRQRKSRRTI 1484.7 12.60 5110 basic peptide 2 KKGE 460.5... 2240 basic peptide 3 SYSMEHFRWGKPVGKKR 2093.4 10.45 690 basic peptide 4 ADAQHATPPKKKRKVEDPKDF 2406.7 9.40 6330 basic peptide 5 LEEEEEAYGWMDF-NH2 1646.7 2.95 1880 minigastrin YLRKDDDDDY 1317.3 3.71 310 acidic peptide 1 GYLRKDDDDY 1259.3 3.86 110 acidic peptide 2 KGYLRKDDDY 1272.4 7.01 260 acidic peptide 3 PKGYLRKDDY 1254.4 9.69 1040 acidic peptide 4 a IP: Isoelectric point. b BB: Bull and Breeze hydrophobicity index [20]. stability of the interaction, and the use of peptide footprinting to determine the site of interaction, as well as the role conformation and equilibrium play in such interaction. We will also show the crucial role of sample preparation, specifically matrix ph, plays in conserving ionic bonding and allowing detection by MALDI. Materials and Methods (a) Peptides: Dynorphin (D), dynorphin 1 7 (DF 1 7), dynorphin 1 8 (DF 1 8), dynorphin 1 9 (DF 1 9), dynorphin 1 10 (DF 1 10), dynorphin 1 13 (DF 1 13), minigastrin (MG), and basic peptides 1, 2, 3, 4, and 5 (BP 1, BP 2, BP 3, BP 4, BP 5) were purchased from Sigma (St. Louis, MO), dynorphin fragment 8 17 (DF 8 17) and [phe 7 ]dynorphin 1 7 (DF [phe 7 ] 1 7) from American Peptide (Sunnyvale, CA) and acidic peptides 1, 2, 3, and 4 (AP1, AP2, AP3, AP4) were synthesized by the core facility of the Biological Chemistry department at the Johns Hopkins School of Medicine. All peptides were dissolved in milli-q water to a concentration of 10 pmol/ L (Table 1). The dynorphin acidic peptide mixtures were made by adding equal amounts of the working solutions of each of the following to an acidic peptide: dynorphin, dynorphin 1 7, dynorphin 1 8, dynorphin 1 9, dynorphin 1 10, and dynorphin 1 13. Another mixture containing dynorphin and its fragments (the concentration of each of the components was 1 pmol/ L) and minigastrin (10 pmol/ L). (b) Enzyme: Sequencing grade, modified trypsin, which cleaves at the carboxyl terminus of both Arg and Lys, was purchased from Roche Molecular Biochemicals (Indianapolis, IN). The enzyme was diluted to a concentration of 0.01 g/ L, so that very short incubation times could be used. (c) Matrix: 6-aza-2-thiothymine (ATT) and -cyano- 4-hydroxycinnamic acid (CHCA) were purchased from Aldrich (Milwaukee, WI). All matrices were prepared fresh daily as a saturated solution in 50% ethanol. (d) Tryptic digests: Test samples: 2 L peptide 4 L 25 mm ammonium bicarbonate ph 8.0 1 L 0.01 g/ L trypsin, 0.3 L aliquots of the digest were deposited on the sample plate, and the reaction was stopped at 1, 3, 5, and 10 min by addition of 0.3 L matrix. Blank samples were simultaneously studied, water was substituted for peptide. (e) Sample preparation: 0.3 L peptide 0.3 L matrix (ATT or CHCA). (f) Instrument: Mass spectra were acquired in linear mode in both positive and negative ion mode on a DE-PRO MALDI from PE-Biosystems (Framingham, MA), equipped with a nitrogen laser (337 nm) and an extraction voltage of 20 kv. All spectra were the average of 50 shots. Results and Discussion Peptide Peptide Interaction To test if peptide peptide noncovalent interaction could be studied by MALDI, mixtures of equal amounts of a solution of minigastrin LEEEEEAYGWMDF-NH 2 and solutions of various peptides were made. Aliquots (0.3 L) from each mixture were deposited on the sample plate, followed by the addition of 0.3 L of ATT and spectra were acquired. Only the mixture of minigastrin, and the two peptides that contained the motifs RKR or RR formed a complex. complexes were formed with peptides containing the motif KK. The positive ion mode spectrum of the mixture of minigastrin and RKRLARKE (basic peptide 1) gave three molecular ions (MH ) at m/z 944.1, 1647.7 (the two peptides), and

90 WOODS AND HUESTIS J Am Soc Mass Spectrom 2001, 12, 88 96 2590.8, the noncovalent complex (943.1 1646.7 1) (data not shown). The negative ion mode spectrum exhibited the following (M H) at m/z 942.2, 1645.7, and 2588.6 u (data not shown). The positive ion mode spectrum of the mixture of minigastrin and QRRQRK- SRRTI also gave three MH, m/z 1485.7, 1647.7, and 3132.4 (1484.7 1646.7 1) (data not shown). We then acquired spectra of both mixtures using CHCA as a matrix and only detected the MH of the free peptides (data not shown). The absence of peptide complexes with the CHCA matrix confirms the disruption of noncovalent peptide complexes in acidic mediums. To confirm peptide peptide interaction, between peptides containing two or more acidic residues (Asp or Glu) and peptides containing the motif RR, we studied the behavior and dynamics of a set of solutions of dynorphin, a 17 amino acid opioid peptide and solutions of five dynorphin fragments that contain the motif RR (residues 6 and 7), dynorphin 1 7, dynorphin 1 8, dynorphin 1 9, dynorphin 1 10, dynorphin 1 13, and two dynorphin fragments that do not contain the RR motif, [phe 7 ]dynorphin 1 7 and dynorphin 8 17 when mixed with various acidic peptides. Equimolar solutions of minigastrin, dynorphin and minigastrin and each of the dynorphin fragments were made, and spectra were acquired using ATT or CHCA for matrix. When using ATT as matrix, noncovalent complexes were seen with all fragments except dynorphin 8 17 and [phe 7 ]dynorphin 1 7, which do not contain the RR motif (Table 2). Equally good spectra were observed in both positive and negative ion modes, as seen in Figure 1a,b which show the spectra of dynorphin fragment 1 13 and minigastrin mixture. However the positive ion spectrum shows a higher rate of in source decay of minigastrin when compared to the negative mode spectrum. In Figure 1a, the b series (from b 4 to b 12 ) can be seen in the lower mass range. The peak at m/z 1788.9 is minigastrin ATT, and the peak at m/z 1745.3 is probably an adduct of ATT and a fragment of minigastrin that lost CO 2. When CHCA (ph 2.0) was used, the noncovalent interaction was not seen, as no complexes were formed with any compound due to the electrostatic changes that occur at acidic ph. Solutions of the four peptides containing 2, 3, 4, and 5 Asp were made (Table 2 only shows data obtained with acidic peptides 1 and 2). Equimolar solutions of each acidic peptide with dynorphin and every dynorphin fragment were made and spectra acquired. Again, when using ATT, noncovalent complexes with dynorphin and all dynorphin fragments containing the RR motif (Table 2) were observed. complexes were detected when CHCA was used as a matrix. However if CHCA s ph was raised (ph 3.0) as previously done [4] by addition of ammonium citrate or bicarbonate complexes can be seen (data not shown). Equilibrium To find out if dynorphin and its fragments showed equal avidity for minigastrin, an equimolar solution of Table 2. MH of mixtures of D MG, DF MG, D AP1, DF AP1, D AP2, DF AP2 Mixture Dynorphin minigastrin Dynorphin 1 7 minigastrin Dynorphin 1 8 minigastrin Dynorphin 1 9 minigastrin Dynorphin 1 10 minigastrin Dynorphin 1 13 minigastrin Dynorphin 8 17 minigastrin [Phe 7 ]dynorphin 1 7 minigastrin Basic peptide 1 minigastrin Basic peptide 2 minigastrin Basic peptide 3 minigastrin Basic peptide 4 minigastrin Basic peptide 5 minigastrin Dynorphin acidic peptide 1 Dynorphin 1 7 acidic peptide 1 Dynorphin 1 8 acidic peptide 1 Dynorphin 1 9 acidic peptide 1 Dynorphin 1 10 acidic peptide 1 Dynorphin 1 13 acidic peptide 1 Dynorphin 8 17 acidic peptide 1 [Phe 7 ]dynorphin 1 7 acidic peptide 1 Dynorphin acidic peptide 2 Dynorphin 1 7 acidic peptide 2 Dynorphin 1 8 acidic peptide 2 Dynorphin 1 9 acidic peptide 2 Dynorphin 1 10 acidic peptide 2 Dynorphin 1 13 acidic peptide 2 Dynorphin 8 17 acidic peptide 2 [Phe 7 ]dynorphin 1 7 acidic peptide 2 Complex formation dynorphin, its fragments containing the RR motif, and minigastrin was made. The only complex detected was between dynorphin and minigastrin at m/z 3795.0, although we see the MH of each of the dynorphin fragments (Figure 2a). The preference for dynorphin demonstrated by acidic peptides could be due to the tryptophan residue at position 14. Arg and Trp are the only two residues that are strict proton donors [15]. However, if the concentration of minigastrin in the mixture was 10 fold relative to dynorphin and its fragments, then noncovalent complexes of minigastrin and dynorphin and all its fragments were formed, despite the fact that the [M H] of dynorphin minigastrin had a high relative intensity (Figure 2b). This second finding suggests that equilibrium might be involved. A side effect of the 10 fold increase in minigastrin relative to dynorphin and its fragments was the formation of a prominent minigastrin dimer at m/z 3293.1 that partially suppressed the dynorphin fragment 1 to 13 minigastrin complex at m/z 3249.3. The peak is clearly seen if that area of the spectrum is enlarged (Figure 2b). To test the role of Trp 14 or possibly the dynorphin conformation, we made an equimolar mixture of minigastrin and the five dynorphin fragments containing the RR motif (the largest is made up of residues 1 13). In the absence of dynorphin, the

J Am Soc Mass Spectrom 2001, 12, 88 96 STUDY OF PEPTIDE PEPTIDE INTERACTION BY MALDI 91 Figure 1. (a) Positive ion mode spectrum of an equimolar mixture of minigastrin (1647.7 u) and dynorphin fragment 1 13 (1604.7 u). The MH of the noncovalent complex is seen at m/z 3251.7. te the in source decay of minigastrin. (b) Negative ion mode spectrum of an equimolar mixture of minigastrin (1645.8 u) and dynorphin fragment 1 13 (1602.9 u). The [M H] of the noncovalent complex is seen at m/z 3249.3.

92 WOODS AND HUESTIS J Am Soc Mass Spectrom 2001, 12, 88 96 Figure 2. (a) Negative ion mode spectrum of an equimolar mixture of minigastrin (1645.7 u) and dynorphin (2146.2 u) and dynorphin fragments 1 7 (867.4 u), 1 8 (980.9 u), 1 9 (1136.6 u), 1 10 (1233.6 u), and 1 13 (1603.1 u). The only noncovalent complex seen is that of MG D (3792.9 u). (b) Negative ion mode spectrum of a mixture where the concentration of minigastrin (1646.0 u) is 10 fold that of the equimolar mixture of dynorphin (2146.4 u) and dynorphin fragments (DF) 1 7 (867.5 u), 1 8 (980.5 u), 1 9 (1137.0 u), 1 10 (1233.5 u), and 1 13 (1602.8 u). When the concentration of MG is raised, noncovalent complexes are seen with dynorphin and all its fragments. MG D (3793.0 u), MG DF 1 7 (2513.9 u), MG DF 1 8 (2626.9 u), MG DF 1 9 (2783.1 u), MG DF 1 10 (2880.2 u), and MG DF 1 13 (3249.1 u).

J Am Soc Mass Spectrom 2001, 12, 88 96 STUDY OF PEPTIDE PEPTIDE INTERACTION BY MALDI 93 Figure 3. Negative ion mode spectrum of an equimolar mixture of minigastrin (1645.8 u) and dynorphin fragments (DF) 1 7 (867.1 u), 1 8 (980.4 u), 1 9 (1136.7 u), 1 10 (1233.8 u), and 1 13 (1603.0 u). In the absence of dynorphin, noncovalent complexes of MG and all DF are formed. MG DF 1 7 (2513.6 u), MG DF 1 8 (2626.2 u), MG DF 1 9 (2782.6 u), MG DF 1 10 (2880.2 u), and MG DF 1 13 (3248.9). spectra showed complexes of minigastrin and every one of the fragments (Figure 3). Mixtures where the dynorphin fragments were 2, 5, 10, 20, and 50 fold more dilute than minigastrin were also tested. Complexes were seen in all cases; however, the intensities of the complexes peaks were significantly decreased for the 20 and 50 fold dilutions (data not shown). The avidity demonstrated by dynorphin could also be due to its secondary structure. Dynorphin is the only molecule that forms -helix (59%), -turns (6%) and has only 35% random coils, whereas all dynorphin fragments and minigastrin are made of random coils. The -helix structure allows the positively charged residues to be maximally exposed and situated so as to interact with other molecules in solution [16]. When minigastrin is complexed with dynorphin, modeling shows reduction of its random coils. Also, when charge versus ph is plotted for dynorphin, a maximal charge of 6 is seen at ph 0 3, and 4.5 from 3 12 producing an almost horizontal line. This could be explained by the isoelectric point (IP) of dynorphin (IP 11.4), as the solubility of a peptide increases at ph values that are furthest away from its IP. The Bull and Breeze (BB) hydrophobicity index [17] seemed to indicate that the hydrophobicity or hydrophilicity of the interacting peptides did not play a role in complex formation, as it did not enhance or detract the process (Table 1). Basic peptides 1, 2, and 4 are very hydrophilic, whereas dynorphin and dynorphin fragments show a range of hydrophobicity from 110 to 3260. Peptide Footprinting A widely used technique to locate promoter sites on DNA is done by adding RNA polymerase to DNA and then digesting the DNA RNA polymerase complex with a deoxyribonuclease. The polymerases remain bound to the promoters and protect them from digestion. The string of bases that remains intact pinpoints the interaction site. Hence, the idea of digesting the peptide complex with trypsin, to see if the residues involved in the noncovalent interaction would be shielded from the enzyme s reach. We first digested the equimolar mixture of basic peptide 1 and minigastrin. The positive ion mode spectrum of the 5 min digest showed complex peaks at m/z 2590.6 (minigastrin RKRARKE H ) and 2333.5 (minigastrin RKRAR H ). The 10 min digest only shows one complex at m/z 2106.0 (minigastrin RKR H ) (Figure 4a,b). The minigastrin dynorphin mixture was also digested with trypsin under the same conditions. The reaction progressed much faster. A complex of a dynorphin fragment and minigastrin was observed as early as 1 and 3 min, at m/z 2515.6 (1646.7 868.0 1, [minigastrin YGGFLRR H ]). The dynorphin fragment

94 WOODS AND HUESTIS J Am Soc Mass Spectrom 2001, 12, 88 96 Figure 4. (a) Positive ion mode spectrum of the 5 min incubation time of a tryptic digest of an equimolar mixture of minigastrin (1647.9 u) and basic peptide 1 [RKRARKE] (944.0 u). The MH at 686.0 u is the tryptic fragment [RKRAR]. The MH at 2590.6 u is the noncovalent complex of minigastrin and basic peptide 1 and the MH at 2333.6 u is the noncovalent complex of minigastrin and the tryptic fragment [RKRAR]. (b) Positive ion mode spectrum of the 10 min incubation time of a tryptic digest of an equimolar mixture of minigastrin (1647.7 u) and basic peptide 1 [RKRARKE]. The MH at 2106.0 u is the noncovalent complex of minigastrin and basic peptide 1 tryptic fragment [RKR].

J Am Soc Mass Spectrom 2001, 12, 88 96 STUDY OF PEPTIDE PEPTIDE INTERACTION BY MALDI 95 Figure 5. Positive ion mode spectrum of the 3 min incubation time of a tryptic digest of an equimolar mixture of minigastrin (1647.7 u) and dynorphin. The only noncovalent complex seen is that of MG and the D tryptic fragment 1 7 (869.0). a 6,c 6,a 7,b 11, and a 13 are D ISD fragments, while C 8,Y 10,Y 11, and Z 12 are MG fragments. 1 7, MH 869.0 u was the most prominent fragment obtained from the digest (Figure 5). In both the cases of basic peptide 1 and dynorphin, the interactions were not disrupted by trypsin. Cleavage did occur at the Arg residue at the carboxyl terminus of the basic peptides in the complex. However, Arg 1 and Lys 2 in the case of basic peptide 1, and Arg 6 in the case of dynorphin, were not accessible for cleavage. Dynorphin alone was digested under the same conditions, a peak at m/z 713 (YGGFLR) was seen among the tryptic fragments, showing that cleavage does occur if the site is not shielded. Peptide peptide interaction could be explained through the physical and chemical properties of the side chains of the residues involved. Arg and Trp are the only residues whose side chain can serve as hydrogen donors, whereas Asp and Glu can both receive and donate hydrogen. The carboxyl groups of Glu and Asp differ marginally in their physical and chemical properties. The carboxyl groups of Asp and Glu side chains ionize with intrinsic pka values of 3.9 and 4.3, respectively; these residues are ionized and very polar under physiological conditions. The Arg side chain consists of three nonpolar methylene groups and a strongly basic -guanidium group, which is dissociated over the entire ph range (pka is 12.0) [15]. It is likely that a proton from the Arg side chain is sought after by the delocalized lone pair of electrons on the Asp or Glu side chain carboxyl group. These salt bridge interactions also consist of some degree of hydrogen bonding in addition to the electrostatic interaction. To test the possible role of coulombic repulsion, we synthesized acidic peptides with an Arg and a Lys residue adjacent to the Asp residues (acidic peptides 1, 2, 3, and 4). The presence of these residues did not prevent noncovalent interaction between the Arg residues on dynorphin and dynorphin fragments and the Asp residues on AP1, AP2, AP3, and AP4 from occurring. Thus, we can assume that the ionic bonds were stable enough to withstand disruption due to nearby weak coulombic repulsion forces. Spectra obtained in both positive and negative modes were basically equally good. However each mode had its advantage and disadvantage. Positive mode usually showed minimal adduct formation with ATT, but more ISD fragmentation. Negative mode on the other hand showed more adduct formation with ATT, but less ISD fragmentation. The ISD of peptides usually increases with laser power. On our DE Pro where the maximum laser power is 4600, a laser power of 2600 was used for most of the samples prepared with ATT matrix. However if conditions were optimized, a laser power of 2100 or less could be used resulting in less saturation and fragmentation of the free peptides. In conclusion, the chemistry of basic residues such as arginine, and acidic residues such as aspartic and glutamic acid, seem to support the likelihood that electrostatic attraction of salt bridges and hydrogen

96 WOODS AND HUESTIS J Am Soc Mass Spectrom 2001, 12, 88 96 bonding between residues is at the base of the observed peptide peptide interaction. It is obvious that a ph range close to physiological ph is needed. Our results seem to indicate that structure plays an important role, as dynorphin (the only peptide with -helix structure) preferentially interacted with minigastrin. The relative intensity of the complexes varied from as little as 3% to as much as 70% (data not shown). However, a 10 fold increase in the concentration of minigastrin relative to that of dynorphin and its fragments led to complex formation with all the fragments in addition to dynorphin, probably due to a shift in complex formation and dissociation. This finding was further supported by the fact that removal of dynorphin from the mixture resulted in all its fragments forming complexes with minigastrin. The dynamics of noncovalent complex formation between dynorphin, its fragments, and minigastrin also suggest that equilibrium may be involved. The observation of noncovalent complex stability was further supported by the peptide footprinting experiment, which demonstrated that the interaction was stable enough to avoid disruption by enzymatic digestion. Acknowledgments We would like to thank Dr. Martha Vestling from the University of Wisconsin for critically reading the manuscript and Dr. Lisa Marzilli from the Johns Hopkins School of Medicine for helpful discussions as well as her help in preparing figures and editing the manuscript. References 1. Ganem, B.; Li, Y.; Henion, J. J. Am. Chem. Soc. 1991, 113, 6294 6296. 2. Ganem, B.; Li, Y.; Henion, J. J. Am. Chem. Soc. 1991, 113, 7818 7819. 3. Loo, J. A. Mass Spectrom. Rev. 1997, 16, 1 17. 4. Lecchi, P.; Pannell, L. K. J. Am. Soc. Mass Spectrom. 1995, 6, 972 975. 5. Woods, A. S.; Buchsbaum, J. C.; Worrall, T. A.; Cotter, R. J.; Berg, J. M. Anal. Chem. 1995, 67, 4462 4465. 6. Juhasz, P.; Biemann, K. PNAS 1994, 91, 4333 4337. 7. Tang, X.; Callahan, J. H.; Zhou, P.; Vertes, A. Anal. Chem. 1995, 67, 4542 4548. 8. Farmer, T. B.; Caprioli, R. M. J. Mass Spectrom. 1998, 33, 697 704. 9. Cohen, L. R. H.; Stupart, K.; Hillenkamp, F. J. Am. Soc. Mass Spectrom. 1995, 8, 1046 1052. 10. Lin, S.; Cotter, R. J.; Woods, A. S. Proteins Structure Function Genetics 1998, 2, 12 21. 11. Lin, S.; Long, S.; Ramirez, S. M.; Cotter, R. J.; Woods, A. S. Anal. Chem. 2000, 72, 2635 2640. 12. Goldstein, A.; Tachibana, S.; Lowney, L. I.; Hunkapiller, M.; Hood, L. PNAS 1979, 76, 6666 6670. 13. Gregory, R. A.; Tracy, H. J. Gut 1974, 15, 683 687. 14. Caudel, R. M.; Dubner, R. Neuropeptides 1998, 32, 87 95. 15. Creighton, T. E. Proteins Structure and Molecular Properties, 2nd ed.; W. H. Freeman: New York, 1993; Chap 1. 16. Pole Bio-Informatique Lyonnais: http://pbil.ibcp.fr.cgi-bin/ secpred_sopma.pl. 17. Bull, H. B.; Breese, K. Arch. Biochem. Biophys. 1974, 161, 665 670.