12.12.2012 Views

Thermodynamic Analysis of Interaction of eIF4E Protein with mRNA 5

Thermodynamic Analysis of Interaction of eIF4E Protein with mRNA 5

Thermodynamic Analysis of Interaction of eIF4E Protein with mRNA 5

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>Thermodynamic</strong> <strong>Analysis</strong> <strong>of</strong> <strong>Interaction</strong> <strong>of</strong> <strong>eIF4E</strong> <strong>Protein</strong><br />

<strong>with</strong> <strong>mRNA</strong> 5' Cap Structure and 4E-BP1 or eIF4G<br />

<strong>Protein</strong> Fragments<br />

Anna NiedŸwiecka<br />

Praca doktorska<br />

pod kierunkiem dr. hab. Ryszarda Stolarskiego, pr<strong>of</strong>. UW<br />

Zak³ad Bi<strong>of</strong>izyki, Instytut Fizyki Doœwiadczalnej, Wydzia³ Fizyki<br />

Uniwersytet Warszawski<br />

Warszawa 2003


Recenzenci: pr<strong>of</strong>. dr hab. Kazimierz Wierzchowski<br />

dr hab. Maciej Geller<br />

II


Sk‡adam serdeczne podziŒkowania<br />

dr. hab. Edwardowi Dar¿ynkiewiczowi, pr<strong>of</strong>. UW, ktry swoim<br />

wysi‡kiem organizatorskim umo¿liwi‡ mi prowadzenie badaæ,<br />

oraz za za wsparcie w rozmaitych sytuacjach,<br />

dr. hab. Ryszardowi Stolarskiemu, pr<strong>of</strong>. UW, za wszechstronn„<br />

pomoc, cierpliwo i ¿yczliwe nastawienie do rozwijaj„cej siŒ<br />

tematyki pracy,<br />

pr<strong>of</strong>. dr. Stephenowi K. Burley’owi z Uniwersytetu Rockefellera<br />

w Nowym Jorku za inspiracjŒ,<br />

dr hab. Agnieszce Bzowskiej, dr Beacie Wielgus-Kutrowskiej,<br />

mgr Joannie fluberek, dr. hab. Janowi Antosiewiczowi, pr<strong>of</strong>.<br />

UW, p. dr hab. Ewie Kulikowskiej, p. Dorocie Haber oraz innym<br />

pracownikom Zak‡adu Bi<strong>of</strong>izyki za cenne dyskusje i<br />

zapewnienie mi‡ej atmosfery,<br />

dr. hab. Januszowi StŒpiæskiemu, dr Marzenie Jankowskiej-<br />

Anyszce i mgr Lidii Chlebickiej za syntezy chemiczne,<br />

biosyntezŒ bia‡ka oraz wiele po¿ytecznych wskazwek,<br />

mgr. Jzefowi Greupnerowi i mgr Urszuli Wonikowskiej-Bezak<br />

oraz lek. med. Tomaszowi Zawadzkiemu za skierowanie moich<br />

zainteresowaæ ku fizyce i biologii molekularnej,<br />

moim Rodzicom za wszelk„ pomoc.<br />

Praca by‡a finansowana przez Komitet Badaæ Naukowych (3 PO4A 071 22 - grant<br />

promotorski, 6 PO4A 055 17, 6 P04A 034 09), Human Frontier Science Program<br />

(RG0303/1998-M) i II. Amerykaæsko-Polski Fundusz im. Marii Sk‡odowskiej-Curie<br />

(MEN/NSF-98-337, MEN/NSF-96-257).<br />

III


Table <strong>of</strong> Contents<br />

ABBREVIATIONS ................................................................................................................VIII<br />

PREFACE................................................................................................................................XI<br />

1. INTRODUCTION ............................................................................................................. 1<br />

1.1. Brief Description <strong>of</strong> Biological Aspects................................................................................ 1<br />

1.1.1. Central Dogma <strong>of</strong> Molecular Biology............................................................................. 1<br />

1.1.2. <strong>Protein</strong> Functions.............................................................................................................. 2<br />

1.1.3. Translation Initiation........................................................................................................ 2<br />

1.1.4. Eukaryotic Initiation Factor 4E ....................................................................................... 5<br />

1.1.5. Translational Control ....................................................................................................... 9<br />

1.1.5.1. Regulation <strong>of</strong> eIF4F Complex Formation ................................................................ 10<br />

1.1.6. Clinical Aspects <strong>of</strong> the Cellular <strong>eIF4E</strong> Activity Control ............................................. 11<br />

1.1.6.1. Relation <strong>of</strong> <strong>eIF4E</strong> Activity to Immunology and Malignancy.................................. 11<br />

1.1.6.2. Relation <strong>of</strong> 4E-BP1 and eIF4G Activities to Human Health and Disease.............. 11<br />

1.1.6.2.1. Phosphorylation <strong>of</strong> 4E-BP1................................................................................. 11<br />

1.1.6.2.2. Dephosphorylation <strong>of</strong> 4E-BP1 and Cleavage <strong>of</strong> eIF4G..................................... 12<br />

1.1.6.3. Anticancer Gene Therapy and Pharmacotherapy..................................................... 12<br />

1.2. Some Remarks about <strong>Thermodynamic</strong> Approach to Molecular <strong>Interaction</strong>s in<br />

Solution................................................................................................................................................. 13<br />

1.2.1. <strong>Thermodynamic</strong> and Apparent Association Constant.................................................. 13<br />

1.2.2. Gibbs Free Energy.......................................................................................................... 14<br />

1.2.3. <strong>Thermodynamic</strong> Parameters <strong>of</strong> Macromolecular <strong>Interaction</strong>s in Solution.................. 15<br />

1.2.4. Noncovalent <strong>Interaction</strong>s in Aqueous Solutions........................................................... 17<br />

1.2.5. Parsing <strong>of</strong> Gibbs Free Energy........................................................................................ 18<br />

1.2.6. <strong>Thermodynamic</strong>s <strong>of</strong> Coupled Processes ....................................................................... 18<br />

2. AIM OF THE WORK...................................................................................................... 20<br />

3. MATERIALS AND METHODS ..................................................................................... 22<br />

3.1. Chemical and Biochemical Syntheses................................................................................. 22<br />

3.1.1. Cap Analogues ............................................................................................................... 22<br />

3.1.2. Peptides........................................................................................................................... 22<br />

3.1.3. <strong>Protein</strong> Biosynthesis....................................................................................................... 24<br />

3.2. Preparation <strong>of</strong> <strong>Protein</strong> and Peptide Samples to Spectroscopic Measurements ........... 24<br />

3.3. Absorption and Emission Spectroscopy............................................................................. 25<br />

3.3.1. Fluorescence Quenching................................................................................................ 26<br />

3.3.1.1. Static Quenching........................................................................................................ 26<br />

3.3.1.2. Dynamic Quenching.................................................................................................. 27<br />

3.3.1.3. Resonance Energy Transfer ...................................................................................... 27<br />

3.3.1.3.1. Resonance Energy Transfer in <strong>eIF4E</strong> ................................................................. 28<br />

3.3.2. Tryptophan Fluorescence in <strong>Protein</strong>s............................................................................ 28<br />

3.3.3. Quenching <strong>of</strong> <strong>eIF4E</strong> Fluorescence by Cap Analogues................................................. 29<br />

V


3.4. Experimental Conditions <strong>of</strong> Spectroscopic Measurements.............................................30<br />

3.4.1. Titration Assay................................................................................................................31<br />

3.4.2. Fluorescence Data Corrections ......................................................................................31<br />

3.4.2.1. Inner Filter Effect......................................................................................................31<br />

3.4.2.2. Dilution.......................................................................................................................32<br />

3.4.2.3. Fluorescence Drift in Time........................................................................................33<br />

3.5. Fluorescence Data <strong>Analysis</strong>..................................................................................................33<br />

3.5.1. Simultaneous Determination <strong>of</strong> Association Constants and <strong>Protein</strong> Activity.............33<br />

3.5.2. The Gibbs Free Energy <strong>of</strong> Binding................................................................................35<br />

3.5.3. Osmotic Stress and Electrostatic <strong>Interaction</strong>s...............................................................35<br />

3.5.3.1. Davies-Stockes-Robinson Electrostatic Screening Approach .................................35<br />

3.5.3.2. Wyman Linkage <strong>Analysis</strong> .........................................................................................36<br />

3.5.4. Protonation Equilibria ....................................................................................................37<br />

3.5.5. <strong>Thermodynamic</strong>s ............................................................................................................38<br />

3.5.5.1. Coupling Between Binding and Conformational Transition <strong>of</strong> Ligand..................38<br />

3.5.5.2. Enthalpy-Entropy Compensation in Congener Series – General Model 199 .............39<br />

3.6. Isothermal Titration Calorimetry.......................................................................................40<br />

3.6.1. Calorimetric Measurements for m 7 GpppG....................................................................41<br />

3.6.1.1. Modified "Single Injection" Experiment ..................................................................41<br />

3.6.2. Calorimetric Data Treatment .........................................................................................42<br />

3.6.2.1. Buffer Ionization Heats..............................................................................................42<br />

3.7. NMR Spectroscopy................................................................................................................42<br />

3.7.1. NMR Spectra Recording................................................................................................43<br />

3.7.2. NMR Data <strong>Analysis</strong> .......................................................................................................44<br />

3.7.2.1. Concentration Dependence........................................................................................44<br />

3.7.2.2. Temperature Dependence..........................................................................................44<br />

3.8. Dynamic Light Scattering ....................................................................................................45<br />

3.9. Statistical <strong>Analysis</strong>.................................................................................................................46<br />

3.10. Additional Information − Crystallography...................................................................47<br />

4. RESULTS AND DISCUSSION..................................................................................... 48<br />

4.1. Methodological Aspects <strong>of</strong> Studies <strong>of</strong> Non-Enzymatic <strong>Protein</strong>.......................................48<br />

4.1.1. Solution <strong>of</strong> Spectroscopic Difficulties...........................................................................49<br />

4.1.2. Activity <strong>of</strong> Non-Enzymatic <strong>Protein</strong>...............................................................................51<br />

4.1.3. Incorrectness <strong>of</strong> Data Linearization in Case <strong>of</strong> Strong <strong>Interaction</strong>s ............................58<br />

4.2. Molecular Mechanism <strong>of</strong> Recognition <strong>of</strong> <strong>mRNA</strong> 5' Cap Structure by <strong>eIF4E</strong> Cap-<br />

Binding <strong>Protein</strong> ...................................................................................................................................60<br />

4.2.1. Affinity <strong>of</strong> Cap Analogues for <strong>eIF4E</strong>............................................................................60<br />

4.2.1.1. Equilibrium Binding Constants.................................................................................60<br />

4.2.1.2. Binding Affinity vs Inhibitory Properties <strong>of</strong> Cap Analogues ..................................62<br />

4.2.2. Parsing the Free Energy <strong>of</strong> <strong>eIF4E</strong> Binding to <strong>mRNA</strong> 5' Cap.......................................64<br />

4.2.2.1. Energetic Cost <strong>of</strong> 7-Substituent Alteration...............................................................66<br />

4.2.2.2. General Separation <strong>of</strong> the Binding Free Energy.......................................................68<br />

4.2.2.3. Relation <strong>of</strong> Stacking - Hydrogen Bonding Cooperativity to Other Cap-Binding<br />

Molecules ....................................................................................................................................69<br />

4.2.2.4. Energetic Cost <strong>of</strong> Methylation <strong>of</strong> the N(2)-Amino Group.......................................70<br />

VI


4.2.2.5. Energetic Cost <strong>of</strong> the Second Nucleoside Addition and Modification ................... 70<br />

4.2.2.6. Contributions <strong>of</strong> Phosphate Groups and Trapped Water Molecules....................... 72<br />

4.2.3. <strong>Interaction</strong>s in Context <strong>of</strong> Environment........................................................................ 73<br />

4.2.3.1. <strong>Interaction</strong> <strong>of</strong> <strong>eIF4E</strong>-Cap in the Presence <strong>of</strong> Electrolyte ......................................... 74<br />

4.2.3.1.1. Davies-Stockes-Robinson Electrostatic Screening Approach........................... 74<br />

4.2.3.1.1.1. Electrostatic Effects ...................................................................................... 74<br />

4.2.3.1.1.2. Two-Step Mechanism <strong>of</strong> the Cap-<strong>eIF4E</strong> Association ............................. 77<br />

4.2.3.1.1.3. Relation to the "Clamping Cycle" ............................................................ 78<br />

4.2.3.1.1.4. Osmotic Stress............................................................................................... 78<br />

4.2.3.1.2. Wyman Linkage <strong>Analysis</strong> ................................................................................... 79<br />

4.2.3.1.2.1. Electrolyte Effect........................................................................................... 79<br />

4.2.3.1.2.2. Osmotic Stress............................................................................................... 80<br />

4.2.3.1.3. Comparison <strong>of</strong> the Two Approaches .................................................................. 80<br />

4.2.3.1.4. Reasons for KCl-Dependence <strong>of</strong> Internal Rearrangement Rate Constants....... 80<br />

4.2.3.1.5. Discussion <strong>of</strong> Hydration Effects ......................................................................... 81<br />

4.2.3.1.6. Conformational Change <strong>of</strong> <strong>eIF4E</strong> upon Cap Binding........................................ 82<br />

4.2.3.1.6.1. Deaggregation <strong>of</strong> <strong>eIF4E</strong> Induced by Cap Binding ...................................... 82<br />

4.2.3.2. Ionic Equilibria in <strong>eIF4E</strong>-Cap Complex................................................................... 83<br />

4.2.4. Conclusions .................................................................................................................... 85<br />

4.3. Molecular <strong>Interaction</strong>s <strong>with</strong>in Ternary Complexes Involving <strong>eIF4E</strong>, a Cap Analogue,<br />

and an <strong>eIF4E</strong>-Binding Motif from eIF4G or 4E-BP1 <strong>Protein</strong>s.................................................... 86<br />

4.3.1. <strong>eIF4E</strong> Fluorescence Quenching Pattern upon Peptide Binding................................... 87<br />

4.3.2. Cooperativity in Ternary Peptide-<strong>eIF4E</strong>-Cap Complexes ........................................... 90<br />

4.3.2.1. Cooperativity <strong>of</strong> Cap and eIF4G Peptides Binding.................................................. 91<br />

4.3.2.2. Cooperativity <strong>of</strong> Cap and 4E-BP1 Peptides Binding............................................... 93<br />

4.3.3. Regulation <strong>of</strong> 4E-BP1 Activity by Phosphorylation .................................................... 94<br />

4.3.4. Concluding Remarks...................................................................................................... 96<br />

4.4. <strong>Thermodynamic</strong>s <strong>of</strong> <strong>mRNA</strong> 5' Cap Binding by <strong>eIF4E</strong> .................................................... 97<br />

4.4.1. Van't H<strong>of</strong>f <strong>Analysis</strong> <strong>of</strong> Binding <strong>of</strong> <strong>eIF4E</strong> to Cap Analogues....................................... 97<br />

4.4.2. Direct Calorimetric Measurements <strong>of</strong> Binding Enthalpy........................................... 106<br />

4.4.3. Protonation Equilibrium <strong>of</strong> m 7 GpppG ........................................................................ 108<br />

4.4.4. Conformational Equilibrium <strong>of</strong> Cap Analogues......................................................... 110<br />

4.4.5. Could the Positive Values <strong>of</strong> ΔCp° Result from <strong>Protein</strong> Deaggregation ? ................ 113<br />

4.4.6. Enthalpy-Entropy Compensation for Individual Cap Analogues .............................. 113<br />

4.4.7. Biological Implications <strong>of</strong> Non-Linear van't H<strong>of</strong>f <strong>Thermodynamic</strong>s........................ 116<br />

4.4.8. Isothermal Enthalpy-Entropy Compensation in Congener Series ............................. 117<br />

4.4.8.1. Molecular Interpretation <strong>of</strong> Isothermal Enthalpy-Entropy Compensation ........... 119<br />

4.4.9. Inquiries about Origins <strong>of</strong> the Unusual Positive Heat Capacity Change................... 121<br />

4.4.9.1. Changes <strong>of</strong> Solvent Accessible Surfaces Upon Binding ....................................... 122<br />

4.4.9.2. Electrostatic Contributions to Heat Capacity Changes.......................................... 123<br />

4.4.9.3. Tryptophan Stacking <strong>with</strong> Cationic 7-Methylguanosine Moiety .......................... 124<br />

4.4.9.4. Coupling <strong>of</strong> Other Intermolecular Equilibria to <strong>eIF4E</strong> – Cap Binding................. 127<br />

4.4.10. Linear Correlation between ΔCp° and ΔG°................................................................. 128<br />

4.4.11. Discussion <strong>with</strong> Other Authors.................................................................................... 130<br />

4.4.12. Conclusions .................................................................................................................. 131<br />

5. SUMMARY................................................................................................................... 133<br />

6. REFERENCES............................................................................................................. 135<br />

VII


Abbreviations<br />

used throughout the text:<br />

4E-BP 4E binding protein;<br />

A absorbance;<br />

Å angström, 10 -10 m;<br />

bz 7 GTP 7-benzylguanosine 5’-triphosphate;<br />

CPK colours scheme developed by Corey, Pauling and Koltun: carbon - light grey;<br />

oxygen – red; hydrogen – white; nitrogen - light blue; phosphorus –<br />

orange.<br />

cx complex;<br />

DLS Dynamic Light Scattering;<br />

DNA deoxyribonucleic acid;<br />

DTT dithiothreitol;<br />

EDTA disodium ethylenediaminetetraacetate;<br />

eIF eukaryotic initiation factor;<br />

ESI MS Electrospray Ionization Mass Spectrometry;<br />

et 7 GTP 7-ethylguanosine 5’-triphosphate;<br />

F observed fluorescence;<br />

F0<br />

calculated initial fluorescence;<br />

GDP guanosine 5’-diphosphate;<br />

GMP guanosine 5’-monophosphate;<br />

GpppG P 1 -guanosine-5’ P 3 -guanosine-5’ triphosphate;<br />

GTP guanosine 5’-triphosphate;<br />

Hepes N-[2-hydroxyethyl]piperazine-N'-[2-ethanesulfonic acid];<br />

HPLC High Performance Liquid Chromatography;<br />

I ionic strength;<br />

ITC Isothermal Titration Calorimetry;<br />

k Boltzmann constant, 1.38 ⋅ 10 -23 J⋅K -1 ;<br />

K observed association constant for stacking, determined for individual<br />

protons from NMR signals;<br />

Kas<br />

Kas<br />

observed association constant;<br />

(micro)<br />

Kas<br />

observed microscopic association constant;<br />

pH-ind<br />

calculated, hypothetical association constant for binding <strong>of</strong> two<br />

species in the ionic states optimal for binding;<br />

Kt<br />

thermodynamic association constant;<br />

kDa kilodalton, 1000 atomic mass units;<br />

1K self-stacking/unstacking equlibrium constant;<br />

L ligand;<br />

m2 2,7 GTP N 2 ,7-dimethylguanosine 5’-triphosphate;<br />

m3 2,2,7 GTP N 2 ,N 2 ,7-trimethylguanosine 5’-triphosphate;<br />

m 7 G 7-methylguanosine;<br />

m 7 GDP 7-methylguanosine 5’-diphosphate;<br />

m 7 GMP 7-methylguanosine 5’-monophosphate;<br />

m 7 Gpppm 2’O G P 1 -7-methylguanosine-5’ P 3 -2'-O-methylguanosine-5’ triphosphate;<br />

m 7 GpppA P 1 -7-methylguanosine-5’ P 3 -adenosine-5’ triphosphate;<br />

m 7 GpppC P 1 -7-methylguanosine-5’ P 3 -cytosine-5’ triphosphate;<br />

VIII


m 7 GpppG P 1 -7-methylguanosine-5’ P 3 -guanosine-5’ triphosphate;<br />

m 7 Gpppm 7 G P 1 -7-methylguanosine-5’ P 3 -7-methylguanosine-5’ triphosphate;<br />

m 7 GppppG P 1 -7-methylguanosine-5’ P 3 -guanosine-5’ tetraphosphate;<br />

m 7 Gppppm 7 G P 1 -7-methylguanosine-5’ P 3 -7-methylguanosine-5’ tetraphosphate;<br />

m 7 GTP 7-methylguanosine 5’-triphosphate;<br />

M molar concentration (1 M = 1 mol⋅dm -3 );<br />

MALDI-TOF MS Matrix-assisted Laser Desorption/Ionization Time-Of-Flight Mass<br />

Spectrometry;<br />

<strong>mRNA</strong> messenger ribonucleic acid;<br />

NMR Nuclear Magnetic Resonance;<br />

NMWL nominal molecular weigh limit;<br />

P probability that distribution <strong>of</strong> fitting residuals is random;<br />

p, PO3 - ; phosphate group;<br />

p-Cl-bz 7 GTP 7-(p-chlorobenzyl)-guanosine 5’-triphosphate;<br />

Pact<br />

active protein;<br />

PDB <strong>Protein</strong> Data Bank (http://rcsb.icm.edu.pl/);<br />

pH -log10 <strong>of</strong> H + activity approximated by concentration;<br />

pHopt<br />

pKa<br />

optimal pH value;<br />

-log10 <strong>of</strong> H + acidic dissociation constant;<br />

pKL<br />

-log10 from the effective acidic dissociation constant <strong>of</strong> a ligand;<br />

pKP<br />

-log10 from the effective acidic dissociation constant <strong>of</strong> a protein<br />

residue;<br />

P(ν1,ν2) probability that improvement <strong>of</strong> the fit <strong>with</strong> the smaller number <strong>of</strong><br />

degrees <strong>of</strong> freedom (ν2) in comparison <strong>with</strong> that <strong>with</strong> the greater<br />

number (ν2) is random;<br />

R universal gas constant, 8.3143 J⋅mol -1 ⋅K -1 ;<br />

r 2<br />

linear correlation coefficient;<br />

R 2<br />

goodness <strong>of</strong> fit;<br />

RNA ribonucleic acid;<br />

SDS PAGE sodium dodecyl sulphate polyacrylamide gel electrophoresis;<br />

t temperature in °C;<br />

T absolute temperature in K;<br />

TH<br />

temperature at which ΔH° = 0;<br />

TS<br />

temperature at which ΔS° = 0;<br />

[x] molar concentration <strong>of</strong> x expressed in M (1 M = 1 mol⋅dm -3 );<br />

° denotes standard state; the hypothetical state <strong>of</strong> solute at the standard<br />

concentration (1 mol⋅dm -3 ) and at arbitrarily chosen temperature <strong>of</strong><br />

293 K (20 °C), and exhibiting infinitely dilute solution behaviour; ∗<br />

ε molar extinction coefficient;<br />

δ chemical shift;<br />

ΔCp° standard molar heat capacity change at constant pressure;<br />

ΔCp°sst<br />

contribution to standard molar heat capacity change, resulting from<br />

thermodynamic coupling <strong>with</strong> self-stacking;<br />

Δδ change <strong>of</strong> chemical shift (Δδ = δ(mixture) − δ(free));<br />

ΔG° standard molar Gibbs free energy change (observed);<br />

∗ according to IUPAC Compendium <strong>of</strong> Chemical Terminology, 2 nd Edition (1997)<br />

IX


ΔG°el<br />

contribution <strong>of</strong> electrolyte effect to standard molar Gibbs free energy<br />

<strong>of</strong> complex stabilization;<br />

ΔG° pH-ind<br />

standard molar Gibbs free energy <strong>of</strong> binding <strong>of</strong> two species in the<br />

ionic states that are optimal for binding;<br />

ΔG°L<br />

standard molar Gibbs free energy change <strong>of</strong> partial protonation <strong>of</strong> a<br />

ligand at a given pH;<br />

ΔG°P<br />

standard molar Gibbs free energy change <strong>of</strong> partial deprotonation <strong>of</strong> a<br />

protein residue at a given pH;<br />

ΔH° standard molar enthalpy change;<br />

ΔH°0<br />

intrinsic standard molar enthalpy change;<br />

ΔH°cal<br />

calorimetric standard molar enthalpy change;<br />

ΔH°ion<br />

standard molar enthalpy <strong>of</strong> buffer ionization at given pH;<br />

ΔH°sst<br />

contribution to standard molar enthalpy change, resulting from<br />

thermodynamic coupling <strong>with</strong> self-stacking;<br />

ΔH°vH<br />

standard molar van't H<strong>of</strong>f enthalpy change;<br />

1ΔH° standard molar enthalpy change <strong>of</strong> self-stacking;<br />

ΔN number <strong>of</strong> exchanged water molecules;<br />

ΔS° standard molar entropy change;<br />

ΔS°0<br />

intrinsic standard molar entropy change;<br />

ΔS°el<br />

contribution <strong>of</strong> electrolyte effect to standard molar entropy change;<br />

ΔS°sst<br />

contribution to standard molar entropy change, resulting from<br />

thermodynamic coupling <strong>with</strong> selfsatcking;<br />

1ΔS° standard molar entropy change <strong>of</strong> self-stacking;<br />

λem<br />

emission wavelength;<br />

excitation wavelength.<br />

λex<br />

X


Preface<br />

The thesis concerns the rules <strong>of</strong> intermolecular recognition <strong>of</strong> <strong>mRNA</strong> 5' terminus by<br />

eukaryotic translation initiation <strong>eIF4E</strong> factor and thermodynamic origins <strong>of</strong> the complex<br />

stability. Molecular spectroscopy and isothermal calorimetry were used to address the<br />

mechanistic bases <strong>of</strong> the process <strong>of</strong> association. Conclusions resulting from the<br />

experimental data were discussed in context <strong>of</strong> biological processes in the living cell and <strong>of</strong><br />

general thermodynamic features <strong>of</strong> protein-ligand interactions.<br />

The work is composed <strong>of</strong> the following chapters:<br />

1. INTRODUCTION which contains a concise and far from exhaustive survey <strong>of</strong> the<br />

present knowledge about significance <strong>of</strong> the studied objects in biology and medicine,<br />

and points to some selected thermodynamic aspects <strong>of</strong> macromolecular interactions,<br />

that are fundamental but not always explicitly accented in literature;<br />

2. AIM OF THE WORK defines the context and the particular goals <strong>of</strong> the thesis;<br />

3. MATERIALS AND METHODS – provide necessary information about chemicals,<br />

syntheses, applied equipment, experimental protocols, and procedures <strong>of</strong> data<br />

analysing;<br />

4. RESULT AND DISCUSSION are merged so that the results can be immediately<br />

interpreted in light <strong>of</strong> other phenomena or literature data, <strong>with</strong>out perpetual referring to<br />

different chapters. The chapter addresses four main problems:<br />

4.1. Methodological aspects <strong>of</strong> studies <strong>of</strong> non-enzymatic proteins;<br />

4.2. Molecular mechanism <strong>of</strong> recognition <strong>of</strong> <strong>mRNA</strong> 5' cap structure by <strong>eIF4E</strong>;<br />

4.3. Molecular interactions <strong>with</strong>in ternary complexes that consist on <strong>eIF4E</strong>, a cap<br />

analogue, and additionally involve an <strong>eIF4E</strong>-binding motif from either the<br />

initiation factor 4G (eIF4G) or the translation inhibitor (4E-BP1). The former is<br />

represented by two is<strong>of</strong>orms (eIF4GI and eIF4GII), and the latter is studied as the<br />

unphosphorylated, monophosphorylated, and diphosphorylated species.<br />

4.4. <strong>Thermodynamic</strong>s <strong>of</strong> <strong>mRNA</strong> 5' cap binding to <strong>eIF4E</strong> and its possible consequences<br />

for biological activity.<br />

5. SUMMARY recollects the main conclusions drawn from the studies.<br />

The results presented in this work were partially published, 1-7 some <strong>of</strong> them are in<br />

press (Chapter 4.4.9.3.) or in preparation for submission (data from 4.1. and 4.4.).<br />

XI


1. Introduction<br />

1.1. Brief Description <strong>of</strong> Biological Aspects<br />

Life <strong>of</strong> organisms depends on delicate balance <strong>of</strong> many biochemical processes. Such<br />

balance needs sensitive and efficient regulatory mechanisms to coordinate growth,<br />

development and reproduction. The most fundamental mechanisms are related to gene<br />

expression.<br />

1.1.1. Central Dogma <strong>of</strong> Molecular Biology<br />

The Central Dogma was first proposed by F. Crick in 1958, and states that genetic<br />

information flow is unidirectional, from DNA to protein via an RNA intermediate. 8 In<br />

order to be expressed, genetic information encoded in DNA has first to be transcribed into<br />

a complementary copy <strong>of</strong> RNA. There are some exceptions from this rule, i. e. reverse<br />

transcriptases <strong>of</strong> retroviruses synthesize DNA from an RNA template, and RNA viruses<br />

such as poliovirus or influenza virus apply RNA replication. 9<br />

DNA replication<br />

DNA<br />

Reverse transcription<br />

Transcription<br />

(RNA synthesis)<br />

Scheme 1-1. Central Dogma <strong>of</strong> Molecular Biology.<br />

RNA transcripts are either used directly as transfer RNAs (tRNAs), ribosomal RNAs<br />

(rRNAs), and small nuclear RNAs (snRNAs) or, as messenger RNAs (<strong>mRNA</strong>s), are<br />

translated into proteins. The <strong>mRNA</strong> transcripts are cotranscriptionally processed by<br />

capping (linkage <strong>of</strong> 7-methylguanosine by a 5' - 5' triphosphate bridge to the first<br />

transcribed nucleoside), splicing (intron excision), and polyadenylation <strong>of</strong> the 3'<br />

terminus. 10 The mature <strong>mRNA</strong> is transported into the cytoplasm (except for mitochondrial<br />

or chloroplast protein <strong>mRNA</strong>s). In eukaryotes, transcripts contain one sequence encoding a<br />

1<br />

<strong>mRNA</strong> PROTEIN<br />

RNA<br />

replication<br />

Translation initiation<br />

Translation<br />

(protein synthesis)


polypeptide, but some sequences can yield more than one polypeptide as a consequence <strong>of</strong><br />

alternative transcriptional starting sites and/or alternative splicing. Translation occurs on<br />

ribosomes. The ribosome is made up <strong>of</strong> two nonidentical subunits, the large (60S) and<br />

small one (40S). Each <strong>of</strong> them contains one or more rRNA molecules and 20-50 different<br />

ribosomal proteins. Several ribosomes may simultaneously translate the same <strong>mRNA</strong><br />

molecule. <strong>Protein</strong>s are synthesized starting <strong>with</strong> their N termini, corresponding to<br />

translation <strong>of</strong> the <strong>mRNA</strong> in the 5' to 3' direction. Translation may be divided into three<br />

stages: initiation, elongation, and termination. Some proteins can fold to their native<br />

conformation during elongation <strong>of</strong> the peptide chain, i. e. before the synthesis is<br />

completed, or it is only after that, <strong>of</strong>ten <strong>with</strong> help <strong>of</strong> other proteins, molecular chaperones.<br />

<strong>Protein</strong>s can undergo many post-translational covalent modifications, e. g. proteolysis,<br />

protein splicing, phosphorylation, and disulphide bond formation, yielding finally the<br />

biologically active form. 11<br />

1.1.2. <strong>Protein</strong> Functions<br />

<strong>Protein</strong>s participate in almost every biological process. Exemplary functions, which<br />

proteins are responsible for, are as follows: structure building, metabolism, transport<br />

throughout body, storage <strong>of</strong> ions, membrane transport, muscle contraction, osmotic<br />

regulation, immunological defence, cell recognition, generation and transduction <strong>of</strong><br />

nervous signals, and control functions. 9 Control factors are even by orders <strong>of</strong> magnitude<br />

less abundant than structural proteins or enzymes, but they perform superior functions in<br />

the living cell, regulating abundancies and activities <strong>of</strong> the remaining proteins. <strong>Protein</strong><br />

control factors act as hormones and fulfil control on gene expression, cell growth,<br />

proliferation, and differentiation.<br />

One <strong>of</strong> the eukaryotic regulatory proteins is <strong>eIF4E</strong>, initiation translation factor 4E,<br />

the key protein in translational control <strong>of</strong> biosynthesis. 12<br />

1.1.3. Translation Initiation<br />

Translation initiation is a complex (Scheme 1-2), rate-limiting step <strong>of</strong> gene<br />

expression. 13 During initiation, the correct starting site on <strong>mRNA</strong> is identified and binding<br />

<strong>of</strong> the ribosome occurs. All <strong>mRNA</strong>s and snRNAs that are transcribed in the nucleus<br />

possess a 5'-terminal "cap", m 7 GpppN, in which 7-methylguanosine is linked by a 5’-5’-<br />

triphosphate bridge to the first nucleoside 14 (Scheme 1-3). The cap structure participates in<br />

2


Scheme 1-2. Eukaryotic translation initiation pathway. Initiation factors (eIF) are shown as colourcoded<br />

labelled circles, and appear as rectangulars when first implicated in the pathway. 60S, 40S -<br />

large and small ribosomal subunits, respectively; Met – initiator methionyl-tRNA; m7G AUG -<br />

<strong>mRNA</strong> <strong>with</strong> the 7-methylguanosine cap and the starting codon (AUG). <strong>eIF4E</strong> is responsible for<br />

recognition <strong>of</strong> the <strong>mRNA</strong> cap and mediates joining <strong>of</strong> the remaining eIFs, which finally leads to<br />

ribosome recruitment and positioning. Scheme adopted from Hershey and Merrick. 15<br />

3


the splicing <strong>of</strong> <strong>mRNA</strong> precursors, 16,17 is necessary for RNA nuclear export 18-20 and for<br />

optimal <strong>mRNA</strong> translation, 14,21 and affects <strong>mRNA</strong> stability. 19 After nuclear export to the<br />

cytosol, the cap <strong>of</strong> snRNAs is further methylated at the amino group <strong>of</strong> the guanosine<br />

moiety, forming m GpppN<br />

7 , 2 , 2<br />

22<br />

This trimethylated structure is responsible for import <strong>of</strong><br />

3<br />

the snRN-protein spliceosomal complexes (snRNPs) back into the nucleus, 23,24 where<br />

snRNPs take part in pre-<strong>mRNA</strong> splicing. 25 A variety <strong>of</strong> <strong>mRNA</strong> cap structures can be<br />

present in the cytoplasm, and most <strong>of</strong> them are additionally methylated at 2'O <strong>of</strong> the second<br />

ribose moiety: m 7 Gpppm2 6,2'O A, m 7 Gpppm 2'O G, m 7 Gpppm 2'O A, m 7 Gpppm 2'O C, m 7 GpppA,<br />

and m 7 GpppG. 26 The original <strong>mRNA</strong> cap can be replaced by the trimethylguanosine cap<br />

structure as a result <strong>of</strong> trans-splicing which occurs in primitive animals, 27,28 e. g. in the C.<br />

elegans nematode about 70% <strong>of</strong> the <strong>mRNA</strong> population contains GpppN<br />

4<br />

m 7 , 2 , 2<br />

3<br />

Regular translation initiation is cap-dependent, though there are also two alternative<br />

ways to recruit the 40S ribosome subunit to the <strong>mRNA</strong>, mediated by an internal ribosome<br />

entry site (IRES) or by the poly(A) tail. 30 The cap function in translation initiation is<br />

mediated by the eukaryotic initiation factor 4E (<strong>eIF4E</strong>), 12 which is the smallest subunit <strong>of</strong><br />

the heterotrimeric eIF4F initiation complex. 31 eIF-4F appears to be also required for the<br />

cap-independent translation <strong>of</strong> naturally uncapped <strong>mRNA</strong>s. 32 In higher eukaryotes eIF4F<br />

consists <strong>of</strong> <strong>eIF4E</strong>, eIF4A, an RNA-dependent ATPase and RNA helicase, and eIF4G, a<br />

multifunctional, adaptor protein. 30,31,33 Two is<strong>of</strong>orms <strong>of</strong> eIF4G, namely eIF4GI (171<br />

kDa) 34 and eIF4GII (176 kDa) 35 are present in mammals. These scaffolding molecules act<br />

as a molecular bridge between <strong>eIF4E</strong> and eIF4A. 33,35 The eIF4Gs interact <strong>with</strong> the<br />

ribosome-bound multiprotein factor eIF3, to recruit the 40S ribosomal subunit. Additional<br />

interaction <strong>of</strong> the eIF4Gs <strong>with</strong> the poly(A)-binding protein (PABP) allows <strong>mRNA</strong><br />

circularization 30,31 and ribosome recycling. 36<br />

* H<br />

N<br />

H 2<br />

O<br />

N 6<br />

1<br />

2<br />

3<br />

N<br />

CH 3<br />

N<br />

7<br />

+<br />

9<br />

N<br />

8<br />

H<br />

H<br />

H<br />

H<br />

OH<br />

HO<br />

O<br />

CH 2<br />

H<br />

O O O<br />

O P O P O P O<br />

O - O - O -<br />

Scheme 1-3. Chemical structure <strong>of</strong> eukaryotic <strong>mRNA</strong> 5' terminus called "cap". Proton that partially<br />

dissociates at physiological pH is marked <strong>with</strong> asterisk.<br />

CH 2<br />

H<br />

H<br />

O<br />

O<br />

H<br />

Base<br />

H<br />

. 29<br />

OH O-H or -CH3 <strong>mRNA</strong> chain


1.1.4. Eukaryotic Initiation Factor 4E<br />

The eukaryotic initiation factor 4E (<strong>eIF4E</strong>) is a 25 kDa protein (217 amino acids)<br />

which plays a pivotal role in translational control. <strong>eIF4E</strong> is a cytoplasmic protein, and acts<br />

as a subunit <strong>of</strong> the eIF4F initiation complex. However, a fraction <strong>of</strong> <strong>eIF4E</strong> localizes also in<br />

the nucleus 37 and participates in nucleocytoplasmic transport <strong>of</strong> specific, growth regulatory<br />

<strong>mRNA</strong>. 38 Since <strong>eIF4E</strong> colocalizes in nuclear speckles <strong>with</strong> splicing factors, it is also<br />

supposed to be involved in splicing and 3' <strong>mRNA</strong> processing. 39 All these additional<br />

functions <strong>of</strong> <strong>eIF4E</strong> appear to be dependent on its intrinsic ability to recognize the <strong>mRNA</strong> 5'<br />

cap. Thus, this single biochemical activity <strong>of</strong> <strong>eIF4E</strong> is versatile enough to play roles in<br />

divergent processes in different cellular compartments. The shuttling protein, <strong>eIF4E</strong>-<br />

transporter (4E-T), that mediates the nuclear import <strong>of</strong> <strong>eIF4E</strong> via the α/β importin pathway<br />

by a piggy-back mechanism was discovered recently. 40<br />

<strong>eIF4E</strong> is the least abundant protein initiation factor in the cell (0.8 ⋅ 10 6 molecules<br />

per 1 HeLa cell, i. e. only 0.021 % <strong>of</strong> cell protein), 41 and its accessibility for formation <strong>of</strong><br />

the initiation eIF4F heterocomplex is supposed to regulate the overall efficiency <strong>of</strong><br />

ribosome recruitment. 31,42 Expression <strong>of</strong> <strong>eIF4E</strong> is regulated at the transcription level. 43 On<br />

the other hand, <strong>eIF4E</strong> is a very instable protein; its instability index calculated on the basis<br />

<strong>of</strong> the sequence 44 is 52.17 for human full length <strong>eIF4E</strong> and 47.41 for murine protein<br />

(residues 28-217) (Table 1-1). This biochemical property may account for the low<br />

abundance <strong>of</strong> <strong>eIF4E</strong> in the cell. <strong>Interaction</strong> between <strong>eIF4E</strong> and eIF4G upon eIF4F complex<br />

formation is inhibited by 4E-binding proteins (mammalian 4E-BP1, 4E-BP2, 4E-BP3, 31<br />

and yeast p20 45 ).<br />

Table 1-1. Comparison <strong>of</strong> protein stabilities resulting from frequency <strong>of</strong> occurrence <strong>of</strong> particular<br />

dipeptides in the amino acid sequences. 44<br />

<strong>Protein</strong> (Organism) Instability Index<br />

<strong>eIF4E</strong> (residues 1-217; human) 52.17<br />

<strong>eIF4E</strong> (residues 28-217; mouse) 47.41<br />

stability limit 40<br />

BPTI (bovine) 32.60<br />

S100 (human) 25.22<br />

PNP (E. coli) 23.13<br />

lysozyme (mouse) 17.76<br />

5


Figure 1-1. Sequence alignments <strong>of</strong> <strong>eIF4E</strong> from mammalia. The relative accessibility <strong>of</strong> each<br />

protein residue was extracted from DSSP 46 and PHD 47 files and shown as colour boxes: white –<br />

buried; cyan – intermediate accessible; blue – accessible; blue <strong>with</strong> red borders – exposed into<br />

solvent; red – unstructured lysine side chains; gap – unstructured main chain. The secondary<br />

structural elements were assigned from the X-ray structures <strong>of</strong> human <strong>eIF4E</strong> in complex <strong>with</strong><br />

m 7 GpppA (1IPB.pdb 48 , top) and murine <strong>eIF4E</strong> <strong>with</strong> m 7 GpppG (1L8B 1 , bottom); α - alpha helices;<br />

η - 310 helices; β - beta strands; TT - strict beta turns.<br />

Table 1-2. Characteristics <strong>of</strong> <strong>eIF4E</strong> sequence conservation among organisms<br />

Group <strong>of</strong> Organisms (<strong>Protein</strong> Is<strong>of</strong>orm) Identity (%) Similarity (%)<br />

Mammalia 98 99<br />

Vertebrates 88 95<br />

Animals a<br />

19 49<br />

Plants (<strong>eIF4E</strong>-1) 55 69<br />

Plants (<strong>eIF4E</strong>-2) 52 69<br />

Plants b<br />

26 53<br />

Fungi c<br />

28 50<br />

All eukaryotes (<strong>eIF4E</strong>-1) 10 35<br />

All eukaryotes d<br />

5 23<br />

Human vs Yeast 29 44<br />

a including only <strong>eIF4E</strong>-1 from Caenorhabditis elegans<br />

b including both <strong>eIF4E</strong>-1 and <strong>eIF4E</strong>-2 is<strong>of</strong>orms<br />

b including only <strong>eIF4E</strong>-1 from Schizosaccharomyces pombe<br />

c including all is<strong>of</strong>orms <strong>of</strong> <strong>eIF4E</strong> (33 sequences)<br />

6


Figure 1-2 (in the previous page). Sequence alignments <strong>of</strong> <strong>eIF4E</strong> (or <strong>eIF4E</strong>-1 if more is<strong>of</strong>orms<br />

exist) from various organisms: human 49 , mouse 50 , rabbit 51 , Xenopus laevis 52 , Aplysia californica 53 ,<br />

Drosophila melanogaster 54 , Caenorhabditis elegans 55 , maize 56 , rice 57 , wheat 58 , Arabidopsis<br />

thaliana 59 , Candida glabrata 60 , Saccharomyces cerevisiae 61 , Schizosaccharomyces pombe 62 . The<br />

sequences are separated in groups corresponding to the animals (1), plants (2) and fungi kingdoms<br />

(3). The alignments were done by the ClustalW program 63 and are colour coded according to the<br />

following rules: red box, white character - strict identity; red character - similarity in a group (><br />

70% <strong>of</strong> residues are similar according to physico-chemical properties); blue frame - similarity<br />

across groups; orange box - differences between conserved groups (> 50% <strong>of</strong> residues are<br />

conserved <strong>with</strong>in a group but not conserved from one group to the other). Risler matrix was used<br />

for similarity and difference score calculations 64 . A consensus line indicates as follows: uppercase<br />

– identity; lowercase - consensus level > 0.5; ! is I or V; $ is L or M; % is F or Y; # is N, D, Q or E.<br />

<strong>eIF4E</strong> activity is also regulated by phosphorylation at Ser209 in response to<br />

treatment <strong>of</strong> cells <strong>with</strong> growth factors, hormones, and mitogens. 65-67 Phosphorylation<br />

seems to positively correlate <strong>with</strong> protein synthesis, cell growth and proliferation, 66,68,69 but<br />

the explanation how it could happen at the molecular level remains unclear. 70-74<br />

<strong>eIF4E</strong> is phylogenetically highly conserved (Table 1-2, Fig. 1-1, 1-2) and unusually<br />

rich in tryptophans. Each <strong>of</strong> its eight tryptophans is absolutely conserved among all known<br />

sequences, from fungi to human. Three-dimensional structures <strong>of</strong> murine <strong>eIF4E</strong> bound to<br />

7-methylGDP 75 and 7-methylGpppG, 1 and recently also <strong>of</strong> human <strong>eIF4E</strong> in complex <strong>with</strong><br />

7-methylGTP and 7-methylGpppA 48 were solved by X-ray crystallography. The structure<br />

<strong>of</strong> their yeast homologue bound to 7-methylGDP was established by solution NMR<br />

spectroscopy. 76 Each protein consists <strong>of</strong> one α/β domain, and is shaped like a cupped hand,<br />

<strong>with</strong> the cap analogue located in a narrow cap-binding slot on the concave surface <strong>of</strong> the<br />

protein (Fig. 1-3). Recognition <strong>of</strong> the 7-methylguanine moiety by the human and murine<br />

proteins is mediated by base sandwich-stacking between Trp56 and Trp102, formation <strong>of</strong><br />

Watson-Crick-like hydrogen bonds <strong>with</strong> a side chain carboxylate <strong>of</strong> a conserved Glu103<br />

and a backbone NH <strong>of</strong> Trp102, and a van der Waals contact <strong>of</strong> the N(7)-methyl group <strong>with</strong><br />

Trp166. The phosphate chain forms salt bridges and direct or water-mediated hydrogen<br />

bonds <strong>with</strong> the NH <strong>of</strong> Trp102 and Trp166 indole rings, and side chains <strong>of</strong> Arg112, Lys162,<br />

Arg157, and Lys 206. As regards the second nucleoside, adenosine is stabilized by a direct<br />

hydrogen bond between N 6 -amino group and the backbone carbonyl <strong>of</strong> Thr205. The<br />

second guanosine in invisible in the crystal structure since it is not capable <strong>of</strong> forming this<br />

hydrogen bond. At the same time, the protein loop containing Thr205 is disordered.<br />

8


Figure 1-3. Global fold <strong>of</strong> <strong>eIF4E</strong> bound to a fragment <strong>of</strong> cap, 7-methylguanosine 5’-triphosphate<br />

(shown as sticks in CPK colours). 1 Secondary structure elements are colour coded: yellow - β<br />

strands, magenta - α helices, white - loops and turns. Amino acid belonging to the cap-binding site<br />

are shown as balls and sticks: blue – Lys206, Arg112, Lys162, Arg157 (from the left); purple –<br />

absolutely conserved tryptophans 102, 166, 56 (from the left), red – glutamic acid 103.<br />

1.1.5. Translational Control<br />

To the middle nineties <strong>of</strong> the XX-th century transcription was thought to be the main<br />

subject to a multitude <strong>of</strong> controls. 9 However, the cell applies also a lot <strong>of</strong> regulatory<br />

mechanism at the level <strong>of</strong> translation. Translational control provides the cell <strong>with</strong> benefits<br />

more important than the energetic cost paid for the reaching that step <strong>of</strong> gene expression. 77<br />

The most conspicuous advantage <strong>of</strong> translational control is immediacy. Direct and rapid<br />

intervention gives the cell no time to adjust coupled biochemical reactions and the protein<br />

synthesis is immediately stopped. On the other hand, most translational controls consist in<br />

reversible modifications <strong>of</strong> translation factors, e. g. phosphorylation. The reversibility is<br />

fast and economical in energetic terms, especially for energy-deprived cells. The changes<br />

in transcription rates are considerably greater in magnitude than the changes in translation<br />

9


ates. Thus, while transcription is responsible for the coarse control, translation provides a<br />

means for fine regulation. Moreover, it enables a spatial control <strong>of</strong> protein synthesis, which<br />

plays a key role in effecting synaptic plasticity in the hippocampus. 78 The local protein<br />

synthesis at a stimulated synapse leads to the synapse-specific growth. This, in turn, is<br />

necessary to long-term potentiation that underlies processes <strong>of</strong> learning and memory<br />

encoding. 79,80 Translation gives a possibility to control <strong>of</strong> expression <strong>of</strong> a single gene or <strong>of</strong><br />

whole classes <strong>of</strong> <strong>mRNA</strong>s. For instance, host-cell shut<strong>of</strong>f induced by heat shock 81 or by<br />

picornaviruses, 82 as well as retroviral gene expression (such as human immunodeficiency<br />

virus type 1, HIV-1 83 ) are controlled at the translational level. Translational controls are<br />

common for regulation <strong>of</strong> gene expression during development (e. g. in oocytes), and in<br />

other systems lacking transcriptional control. Finally, translational control has proven to be<br />

an important factor in human cancer. 84-87<br />

1.1.5.1. Regulation <strong>of</strong> eIF4F Complex Formation<br />

The accessibility <strong>of</strong> <strong>eIF4E</strong> in mammals is regulated by interactions <strong>with</strong> small<br />

translational repressors, <strong>eIF4E</strong>-binding proteins, 4E-BP1, 4E-BP2 and 4E-BP3, which<br />

prevent productive interactions between <strong>eIF4E</strong> and eIF4G. 88,89 They also inhibit<br />

phosphorylation <strong>of</strong> <strong>eIF4E</strong> at Ser 209. 68<br />

Sequence analysis <strong>of</strong> the 4E-BPs and eIF4Gs suggests that these otherwise unrelated<br />

protein families have converged on the same <strong>eIF4E</strong> binding strategy. 45,89 The <strong>eIF4E</strong><br />

binding site shared between the 4E-BPs and the eIF4Gs is BXXYDRXFLΦ, where B is a<br />

conserved basic residue, X is variable, Φ is a conserved hydrophobic residue, and invariant<br />

residues are shown in boldface. 34,35,67,90 Crystal structures <strong>of</strong> two ternary complexes <strong>of</strong><br />

<strong>eIF4E</strong> <strong>with</strong> 7-methyl-GDP and peptides encompassing the <strong>eIF4E</strong> binding sites, derived<br />

from eIF4GII or 4E-BP1, revealed that both peptides recognize a phylogenetically<br />

invariant, partially hydrophobic, partially acidic surface <strong>of</strong> the convex dorsum <strong>of</strong> <strong>eIF4E</strong> in<br />

the vicinity <strong>of</strong> Trp73, and share a common mode <strong>of</strong> interaction. 91<br />

Some <strong>of</strong> the hypophosphorylated forms <strong>of</strong> 4E-BPs bind to <strong>eIF4E</strong> and these<br />

interactions are highly regulated in cells. 82 Hyperphosphorylation abrogates the interaction<br />

<strong>with</strong> <strong>eIF4E</strong>, which enables recruitment <strong>of</strong> eIF4G, allowing eIF4F complex formation and<br />

translation initiation to proceed. 67,92 The phosphorylation mechanism has been extensively<br />

investigated for 4E-BP1. The protein possesses at least six phosphorylation sites: Thr37,<br />

Thr46, Ser65, Thr70, Ser83 and Ser112 (numbering for the human protein). The ordered,<br />

10


hierarchical model <strong>of</strong> phosphate addition to endogenous 4E-BP1, in which phosphorylation<br />

<strong>of</strong> Thr37 and Thr46 is followed by phosphorylation <strong>of</strong> Thr70 and finally <strong>of</strong> Ser65, 3,92 is<br />

generally accepted, although some points remain controversial. 93-95 At issue is the<br />

influence <strong>of</strong> the single phosphorylation <strong>of</strong> Ser65 or combined Ser65/Thr70<br />

phosphorylation on the 4E-BP1 binding to <strong>eIF4E</strong>. 96<br />

1.1.6. Clinical Aspects <strong>of</strong> the Cellular <strong>eIF4E</strong> Activity Control<br />

1.1.6.1. Relation <strong>of</strong> <strong>eIF4E</strong> Activity to Immunology and Malignancy<br />

Since initiation is a rate-limiting step for translation, it is not surprising that <strong>eIF4E</strong><br />

activity is frequently related to disorder <strong>of</strong> cellular proliferation, growth and<br />

differentiation. 87 The <strong>eIF4E</strong> gene is a proto-oncogene. Overexpression <strong>of</strong> <strong>eIF4E</strong> is caused<br />

by gene amplification. 85 When present in excess <strong>of</strong> the normal physiological amount,<br />

<strong>eIF4E</strong> can stimulate global protein synthesis. 97 Elevated levels <strong>of</strong> <strong>eIF4E</strong> were found in<br />

many cancers, and were shown to stimulate division and cause malignant transformation <strong>of</strong><br />

both normal and immortal cells. 87 Immunological detection <strong>of</strong> <strong>eIF4E</strong> was even reported as<br />

a useful diagnostic tool to identify malignant cells for removal during surgery <strong>of</strong> head and<br />

neck cancer. 98 Affinity <strong>of</strong> <strong>eIF4E</strong> for <strong>mRNA</strong> 5' cap is strongly reduced by the nuclear<br />

promyelocytic leukaemia protein (PML) which mediates suppression <strong>of</strong> oncogenic<br />

transformation and <strong>of</strong> growth. 99<br />

On the other hand, enhanced <strong>eIF4E</strong> activity and its phosphorylation at Ser209 are<br />

engaged in signalling pathways necessary for human peripheral blood T lymphocytes to<br />

progress from the resting state to proliferation and to attain immune competency. 66<br />

1.1.6.2. Relation <strong>of</strong> 4E-BP1 and eIF4G Activities to Human Health and Disease<br />

1.1.6.2.1. Phosphorylation <strong>of</strong> 4E-BP1<br />

4E-BP1 is also known as PHAS-I, phosphorylated heat- and acid-stable protein<br />

regulated by insulin, since insulin causes 4E-BP1 to become hyperphosphorylated and<br />

dissociate from eIF-4E, thus stimulating the overall rate <strong>of</strong> translation and promoting cell<br />

growth. 67<br />

Activation <strong>of</strong> human blood T lymphocytes needs the release <strong>of</strong> <strong>eIF4E</strong> by<br />

hyperphosphorylation <strong>of</strong> 4E-BP1 to enable synthesis <strong>of</strong> proteins that are important for<br />

proliferation. 66<br />

Dissociation <strong>of</strong> 4E-BP1 from <strong>eIF4E</strong> is involved in tumour cell survival. In particular,<br />

11


the α6β4 integrin stimulates the phosphorylation and inactivation <strong>of</strong> 4E-BP1, thus<br />

preventing inhibition <strong>of</strong> <strong>eIF4E</strong>, and finally enhancing translation <strong>of</strong> vascular endothelial<br />

growth factor (VEGF) in human breast carcinoma cells. 100<br />

1.1.6.2.2. Dephosphorylation <strong>of</strong> 4E-BP1 and Cleavage <strong>of</strong> eIF4G<br />

In contrast, heat shock 81 and infection <strong>of</strong> cells <strong>with</strong> picornaviruses, such as poliovirus<br />

and encephalomyocarditis virus, 82 render 4E-BP1 dephosphorylated. Picornaviruses use a<br />

cap-independent mechanism for the translation <strong>of</strong> their RNA, and specific inhibition <strong>of</strong><br />

cap-dependent translation by activation <strong>of</strong> 4E-BP1 is thought to be the major cause <strong>of</strong> a<br />

shut<strong>of</strong>f <strong>of</strong> host protein synthesis in encephalomyocarditis virus-infected cells. Poliovirus<br />

acts doubly, and its second strategy involves also translational control at the initiation<br />

level, since the infection results in the cleavage <strong>of</strong> the eIF4GII protein. 101 The eIF4GII<br />

protein is also a target <strong>of</strong> the rhino and foot-and-mouth-disease viral proteases, 102 while the<br />

HIV-1 viral protease cleaves the eIF4GI protein. 83<br />

Rapid cleavage <strong>of</strong> eIF4GI by caspase-3 occurs also during induced apoptosis. 103,104<br />

1.1.6.3. Anticancer Gene Therapy and Pharmacotherapy<br />

4E-BP1 belongs to a class <strong>of</strong> multiphosphorylated proteins that are dephosphorylated<br />

rapidly after rapamycin treatment. 105 Rapamycin represents a novel family <strong>of</strong> agents <strong>of</strong><br />

antibiotic, antifungal, and, which is the most important, immunosuppressory and<br />

antitumour activity. 106 Rapamycin acts successfully as a cytostatic or induces apoptosis in<br />

malignant cells through inhibition <strong>of</strong> a function <strong>of</strong> mTOR (mammalian target <strong>of</strong><br />

rapamycin) which is a huge, highly conserved protein (~280 kDa). mTOR is a central<br />

controller <strong>of</strong> cell growth and proliferation, influencing an unusually abundant and diverse<br />

set <strong>of</strong> reactions. 107 It acts both as a protein-kinase and protein-phosphatase regulator, and<br />

among others it modulates 4E-BP1 phosphorylation. The mTOR − 4E-BP1 signalling<br />

pathway is important not only for malignancy but also for proliferation in T lymphocytes. 66<br />

The Food and Drug Administration has approved rapamycin in 1999, and the European<br />

Commission in 2000. However, recent data indicate that genetic mutations or<br />

compensatory changes in tumour cells render resistance to rapamycin. 108,109<br />

The important role for translation initiation in the regulation <strong>of</strong> the cell cycle makes<br />

that other components <strong>of</strong> the initiation machinery already begin to be considered attractive<br />

targets for therapeutic intervention. Attention is focused at the activity <strong>of</strong> 4E-BP1, as a<br />

12


crucial mTOR-controlled downstream effector molecule, and at <strong>eIF4E</strong>. The most recent<br />

publications (2002) report that cellular activity <strong>of</strong> <strong>eIF4E</strong> is already now exploited to<br />

selectively kill breast cancer cells during clinical gene therapy. 110 However, there are still<br />

no chemotherapeutic inhibitors <strong>of</strong> the <strong>eIF4E</strong> activity. Thorough biophysical studies <strong>of</strong><br />

functional complexes involving <strong>eIF4E</strong> can make up for this delay and provide a specific<br />

inhibitory agent based on the rationally modified cap structure.<br />

1.2. Some Remarks about <strong>Thermodynamic</strong> Approach to<br />

Molecular <strong>Interaction</strong>s in Solution<br />

Mechanistic insights into molecular bases <strong>of</strong> biological activity do not flow<br />

naturally from the description <strong>of</strong> the crystal structure. The stationary structure does not<br />

reveal all exchange processes that need completing so that the final state, detectable by<br />

crystallography, can come into being. In general, proteins and ligands structures are not<br />

rigid. In particular, non-enzymatic regulatory protein factors are <strong>of</strong>ten very flexible and<br />

experience large and even global conformational changes upon binding to their targets, in<br />

contrast to most <strong>of</strong> enzymes. Such features are characteristic for e. g. transcription factors<br />

that interact <strong>with</strong> DNA. 111 In addition to conformational transitions, formation <strong>of</strong> the<br />

functional macromolecular complexes is accompanied by changes <strong>of</strong> interactions <strong>with</strong> the<br />

environment, first <strong>of</strong> all <strong>with</strong>in the first-layer solvation shell. Hence, the static structural<br />

view needs to be supplemented by thermodynamic description in order to render the<br />

biological functioning on the molecular level comprehensible.<br />

Binding studies <strong>of</strong> specific complexes involving proteins, nucleic acids and small<br />

ligands have proven to be useful to investigate the stabilization energy and the influence <strong>of</strong><br />

solvent on the complex formation, e. g. electrolytic effect, 112-116 water exchange between<br />

the macromolecules and the bulk solvent, 117-121 and linked protonation equilibria. 122-124<br />

1.2.1. <strong>Thermodynamic</strong> and Apparent Association Constant<br />

A thermodynamic equilibrium constant (Kt) which is a function <strong>of</strong> temperature (T)<br />

and pressure (p) only, is expressed in terms <strong>of</strong> equilibrium molar activities (ai): 125<br />

∏ α<br />

νi<br />

t = a i<br />

i=<br />

1<br />

K , Eq. 1-1<br />

where νi are stoichiometric coefficients <strong>of</strong> the reaction and summation ranges by all<br />

species taking part in the reaction. A logarithmic form <strong>of</strong> this relationship:<br />

13


log( K<br />

t<br />

∑ i<br />

i 1<br />

α<br />

=<br />

) = ν ⋅ log( a )<br />

Eq. 1-2<br />

i<br />

makes it possible to analyse easily the dependence <strong>of</strong> Kt on the presence <strong>of</strong> individual<br />

participants <strong>of</strong> the reaction: macromolecules, cations, anions, protons, and water.<br />

An experimentally observed equilibrium association constant (Kas) is defined for<br />

the process converting chosen reactants (e. g. protein (P) and ligand (L)) to products (e. g.<br />

complex (cx)) at a specified set <strong>of</strong> solution conditions, i. e. when activities <strong>of</strong> remaining<br />

reactants (e. g. ions, protons, water molecules) are fixed, e. g.: 118<br />

K<br />

as<br />

[ cx]<br />

= Eq. 1-3<br />

[ P]<br />

⋅[<br />

L]<br />

0<br />

0<br />

This apparent constant is formulated in terms <strong>of</strong> concentrations <strong>of</strong> only a protein, a ligand<br />

and their complex, instead <strong>of</strong> thermodynamic activities <strong>of</strong> all interacting species that could<br />

putatively participate in the reaction. Kas contains involved, hidden information <strong>of</strong> both the<br />

true thermodynamic association constant (Kt) and the changes <strong>of</strong> the activity coefficients<br />

as a function <strong>of</strong> the solution conditions. Therefore, the observed equilibrium binding<br />

constant depends on the environmental variables, i.e. pH, ionic strength and osmolality <strong>of</strong><br />

the solution. 118,119,126 Hence, Kas is a relative quantity. To probe the molecular and<br />

thermodynamic origins <strong>of</strong> stability and specificity <strong>of</strong> <strong>eIF4E</strong> interactions <strong>with</strong> cap in<br />

solution the binding studies need to be enriched by measurements at different experimental<br />

conditions. This is a starting point to analyse possible intermolecular processes that can<br />

accompany binding <strong>of</strong> cap analogues by <strong>eIF4E</strong>: protonation or ion and water exchange.<br />

1.2.2. Gibbs Free Energy<br />

Chemical potential (μi) <strong>of</strong> interacting species at given conditions consists on the<br />

standard potential (μi°, index "°" refers to the pseudostandard state at concentrations <strong>of</strong> 1<br />

mol/dm 3 , i.e. unit molarity 127 ) and a contribution (μi m ) which depends on the presence <strong>of</strong><br />

other solution components (xi): 125<br />

o<br />

m<br />

μ = μ T,<br />

p)<br />

+ μ ( T,<br />

p,<br />

x ,..., x )<br />

Eq. 1-4<br />

i i ( i 1 α−1<br />

The activity <strong>of</strong> species is defined as:<br />

⎛ m<br />

⎞<br />

⎜<br />

μi<br />

( T,<br />

p,<br />

x1,...,<br />

x α−1<br />

)<br />

a<br />

≡<br />

⎟<br />

i ( T,<br />

p,<br />

x1,...,<br />

x α−1<br />

) exp<br />

, Eq. 1-5<br />

⎜<br />

⎟<br />

⎝<br />

RT<br />

⎠<br />

m<br />

so: i ≡ RT ln a i<br />

μ . Eq. 1-6<br />

14


Definition <strong>of</strong> the molar Gibbs free energy change (ΔG):<br />

∑ α<br />

i<br />

i=<br />

1<br />

ΔG<br />

≡ μ ν<br />

Eq. 1-7<br />

at equilibrium, where:<br />

m<br />

Δ ∑ i<br />

i 1<br />

α<br />

=<br />

i<br />

G = ΔG°<br />

+ μ νi<br />

= 0<br />

leads to the relationship for the standard molar Gibbs free energy change (ΔG°):<br />

o<br />

t<br />

15<br />

Eq. 1-8<br />

Δ G = −RT<br />

ln K . Eq. 1-9<br />

Stability <strong>of</strong> a noncovalent complex is determined by the difference in noncovalent<br />

interactions involving both the macromolecular recognition surfaces and solvent<br />

components in the complex and in the uncomplexed state. Endeavour to interpret the<br />

apparent Gibbs free energy <strong>of</strong> binding (ΔG°app) defined by the apparent association<br />

constant (Kas) and not by the thermodynamic association constant must take into account<br />

the fact <strong>of</strong> the participation <strong>of</strong> water, electrolyte ions and other solution components in the<br />

interactions <strong>of</strong> a protein <strong>with</strong> a ligand, as well as the other macromolecular equilibria.<br />

For shortening, the apparent Gibbs free energy change (ΔG°app) will be simply<br />

referred to here as "the Gibbs free energy change" but one should bear in mind that this<br />

"ΔG°" will be related to the observed and not to the thermodynamic association constant.<br />

1.2.3. <strong>Thermodynamic</strong> Parameters <strong>of</strong> Macromolecular <strong>Interaction</strong>s in<br />

Solution<br />

The Gibbs free energy change in an isothermal-isobaric process is related to changes<br />

in standard molar enthalpy (ΔH°) and entropy (ΔS°): 128<br />

ΔG ° = ΔH°<br />

− TΔS°<br />

. Eq. 1-10<br />

At constant but different temperatures the process is described by different pairs <strong>of</strong><br />

the ΔH° and ΔS° values, since they can vary <strong>with</strong> temperature according to the standard<br />

molar heat capacity change at constant pressure ( Δ ) and to two critical temperatures that<br />

are characteristic for the analysed process:<br />

o o<br />

H p H<br />

Δ = ΔC<br />

( T − T ) ,<br />

o<br />

C p<br />

o<br />

Δ H = 0 at TH; Eq. 1-11<br />

o o ⎛ T ⎞<br />

ΔS<br />

= ΔC<br />

p ln<br />

⎜<br />

⎟ , ΔS° = 0 at TS; Eq. 1-12<br />

⎝ TS<br />


⎛ ∂H°<br />

⎞<br />

C p = ⎜ ⎟<br />

⎝ ∂T<br />

⎠ p<br />

Δ o<br />

The parameters <strong>of</strong> interest can be determined from the van't H<strong>of</strong>f equation:<br />

ln K<br />

as<br />

o p<br />

ΔC<br />

⎡TH<br />

⎛ TS<br />

⎞ ⎤<br />

= ⎢ − ln⎜<br />

⎟ −1⎥<br />

R ⎣ T ⎝ T ⎠ ⎦<br />

16<br />

Eq. 1-13<br />

Eq. 1-14<br />

The isobaric heat capacity in solution is related to the fluctuations in internal energy<br />

(E) <strong>of</strong> the system ( X denotes an ensemble average <strong>of</strong> a quantity X):<br />

B<br />

( ) 2<br />

E E<br />

1<br />

C p ≈ C v = −<br />

Eq. 1-15<br />

2<br />

k T<br />

In most cases, the thermodynamic properties <strong>of</strong> the system expressed by ΔCp° are<br />

dominated by changes in water structure induced by macromolecular association. The<br />

increased energy fluctuations are related to lack <strong>of</strong> possibility <strong>of</strong> formation <strong>of</strong> hydrogen<br />

bonds by water molecules forced to contact <strong>with</strong> nonpolar (nonpolarizable) or aromatic<br />

(polarizable) solvent accessible groups <strong>of</strong> macromolecules (so called "hydrophobic<br />

effect"). The nature <strong>of</strong> aromatic rings shows that the terms "hydrophobic" and "nonpolar,<br />

apolar" must not be used equivalently, since the aromatic systems can carry induced charge<br />

distribution, being then partly polar, yet still hydrophobic objects. 129 Consequently, their<br />

contribution to the heat capacity changes, related to changes in the solvent accessible<br />

surface area (SASA), differs from that <strong>of</strong> aliphatic groups 130 (Table 1-3). The fluctuations<br />

<strong>of</strong> water energy are silenced when hydrogen bonds are maximally satisfied, i. e. when<br />

polar (uncharged) or charged molecular surface is water exposed. The second main origin<br />

<strong>of</strong> energy fluctuation in the system is related to macromolecular internal s<strong>of</strong>t vibrations. 131<br />

Table 1-3. Increments to the standard molar heat capacity resulting from exposure <strong>of</strong> 1 Å <strong>of</strong><br />

aliphatic, aromatic, and polar molecular surface to water (ΔCp°/SASA); data from different groups<br />

<strong>of</strong> authors.<br />

ΔCp° transfer /SASA (J⋅mol -1 ⋅K -1 ⋅Å -2 ) Authors (date) ref<br />

Surface: aliphatic aromatic polar<br />

+1.88 ± 0.08 a<br />

−1.09 ± 0.13 Murphy et al. (1991, 1992) 132,133<br />

+1.34 ± 0.17 a<br />

−0.59 ± 0.17 Spolar et al. (1992) 134<br />

+2.14 +1.55 −1.27 Makhatadze & Privalov (1995) 130<br />

a average values for aliphatic and aromatic surfaces


1.2.4. Noncovalent <strong>Interaction</strong>s in Aqueous Solutions<br />

"Energy <strong>of</strong> hydrogen bonds ranges from −12.5 to −38 kJ/mol" is the habitually<br />

repeated statement in the context <strong>of</strong> protein stability and protein-ligand interactions, both<br />

in many biochemical textbooks 135-137 as well as during academic lectures. However, this is<br />

true only if the "energy" means the internal energy (ΔE°) calculated usually in vacuum,<br />

and not the enthalpy (ΔH°) and Gibbs free energy (ΔG°). Binding between<br />

macromolecules in aqueous solution is an exchange, entropy-related process involving<br />

both water and solutes, since it needs disruption <strong>of</strong> pre-existing interactions <strong>of</strong> individual<br />

groups <strong>of</strong> each macromolecule <strong>with</strong> water and ions, and replacement <strong>of</strong> these interactions<br />

by contacts between complementary groups on each macromolecule. Hence, the stability<br />

<strong>of</strong> noncovalent complexes is totally different than in vacuum, and is highly dependent on<br />

the details <strong>of</strong> the solution environment. The real Gibbs free energy <strong>of</strong> the hydrogen bond<br />

formed in the water accessible region <strong>of</strong> the macromolecule dissolved in aqueous solution<br />

is only from −3 to −8 kJ/mol, 138-143 which may appear surprising but is a well documented<br />

rule. This is due mainly to entropic effects related to the presence <strong>of</strong> surrounding polar<br />

water molecules that compete for the hydrogen bond formation <strong>with</strong> the residues <strong>of</strong> the<br />

macromolecule. The strong decrease in free energy <strong>of</strong> a single amide hydrogen bond<br />

formation in water in comparison <strong>with</strong> that in apolar solvent (CCl4), from –35.31 ± 0.46 to<br />

−1.42 ± 1.2 kJ/mol, has been also demonstrated theoretically. 144 Similar entropic effects<br />

due to the presence <strong>of</strong> counterions should be taken into account when analysing the Gibbs<br />

free energy <strong>of</strong> salt bridges (Coulombic interactions <strong>with</strong> some contribution <strong>of</strong> hydrogen<br />

bonding) that usually link unlikely ionised residues.<br />

The second factor decreasing ΔG° <strong>of</strong> hydrogen bonds and salt bridges in<br />

macromolecules is the entropic penalty for ordering, e. g. conformational entropy <strong>of</strong><br />

protein side chains packing. Consequently, for macromolecules and their complexes<br />

interacting in buffered aqueous solutions (dielectric constant ε ~ 80, ionic strength I ~ 100-<br />

150 mM), the free energy <strong>of</strong> stabilization <strong>of</strong> a hydrogen bond is the same as that <strong>of</strong> a salt<br />

bridge, van der Waals interaction (induction and dispersion forces), and a hydrophobic<br />

contact. 145<br />

Stacking <strong>of</strong> aromatic rings (van der Waals and hydrophobic interactions) can attain<br />

free energy from 0 to −42 kJ/mol, depending on interatomic distances, but the energy is<br />

usually close to that <strong>of</strong> hydrogen bonds in proteins, 4 - 12.5 kJ/mol. 128,146,147 A nontypical<br />

17


ut biologically important example <strong>of</strong> stacking is cation - π stacking which consists in<br />

interaction between a charge and an induced quadrupole moment <strong>of</strong> charge distribution at<br />

the aromatic ring. 129,148,149<br />

Each type <strong>of</strong> noncovalent bonds has a typical range <strong>of</strong> interatomic distances. The<br />

optimal configuration <strong>of</strong> hydrogen bonds and salt bridges is linear, <strong>with</strong> the atoms/ions<br />

distant by 2.6 – 3.1 Å. The optimal distance for the van der Waals interaction is usually by<br />

0.3 - 0.5 Å greater than the sum <strong>of</strong> the van der Waals atomic radii, and this is ~ 4 Å for the<br />

C – C pair.<br />

1.2.5. Parsing <strong>of</strong> Gibbs Free Energy<br />

At issue is widely utilized parsing <strong>of</strong> ΔG° into contributions from individual,<br />

specific, non-covalent interactions, 120,137,139,150,151 i. e. hydrogen bonding, salt bridges, van<br />

der Waals contacts etc., from stating it to be meaningful 152 to proving its unreliability. 153<br />

The argument against parsing is based on the fact that even if the energy <strong>of</strong> the system can<br />

be approximated as linear combination <strong>of</strong> terms, it is not possible to write such general<br />

expression for the total free energy. The ΔG° value can be approximately factorized only<br />

when the resulting interactions are not correlated, e. g. operate on different coordinates.<br />

The free energy parsing gives reliable physical insight into macromolecular processes as<br />

long as the experimentally or theoretically obtained data are interpreted <strong>with</strong> care. 137,152,154<br />

1.2.6. <strong>Thermodynamic</strong>s <strong>of</strong> Coupled Processes<br />

<strong>Thermodynamic</strong>ally coupled processes that accompany binding <strong>of</strong> a ligand to a<br />

protein influence the thermodynamic parameters <strong>of</strong> binding to be determined. 122 In<br />

particular, the existence <strong>of</strong> many <strong>of</strong> protein conformational microstates can significantly<br />

shift the binding equilibrium in comparison <strong>with</strong> that for the binding to the single state.<br />

Additionally, conformational rearrangement <strong>of</strong> the ligand as well as protonation, ionic and<br />

hydration equilibria <strong>with</strong>in the interacting molecules can modulate the resultant<br />

thermodynamic features <strong>of</strong> the complex formation.<br />

A0<br />

1K, 1ΔH°<br />

(+B) K0, ΔH°0 K1, ΔH°1 (+B)<br />

BA0<br />

18<br />

A1<br />

A1B


For instance, if the molecule (A) exists in equilibrium <strong>of</strong> only two states (A0, A1) that<br />

is described by the constant (1K) and the transition enthalpy (1ΔH°), and can bind the<br />

second molecule (B) <strong>with</strong> different affinities and binding enthalpies <strong>of</strong> K0, K1 and ΔH°0,<br />

ΔH°1, respectively, then the observed parameters are as follows: 122<br />

( ΔH°<br />

+ ΔH°<br />

− ΔH°<br />

)<br />

1ΔH°⋅1<br />

K γ⋅1<br />

K 1<br />

1 0<br />

Δ H°<br />

= ΔH°<br />

0 − +<br />

; Eq. 1-16<br />

1+<br />

K<br />

1 + γ⋅<br />

K<br />

1<br />

( 1+<br />

K)<br />

1<br />

19<br />

( 1ΔH°<br />

+ ΔH°<br />

1 − ΔH°<br />

0 )<br />

T(<br />

1+<br />

γ⋅<br />

K)<br />

1ΔH°⋅1<br />

K<br />

γ⋅1<br />

K<br />

Δ S°<br />

= ΔS°<br />

0 − R ln(<br />

1+<br />

1K)<br />

− + R ln(<br />

1+<br />

γ⋅1<br />

K)<br />

+<br />

; Eq. 1-17<br />

T<br />

p<br />

p 0<br />

2 ( 1ΔH°<br />

) ⋅1<br />

2<br />

RT ( 1+<br />

K)<br />

1<br />

2<br />

1<br />

( )<br />

( ) 2<br />

1ΔH°<br />

+ ΔH°<br />

1 − ΔH°<br />

0<br />

2<br />

RT 1+<br />

γ⋅<br />

K<br />

K γ⋅1<br />

K<br />

Δ C°<br />

= ΔC°<br />

−<br />

+<br />

; Eq. 1.18<br />

where γ = K1/K0 (K1 < K0, γ [0,1]), and terms indexed "0" denote intrinsic parameters<br />

for binding <strong>of</strong> B to A in the A0 state. This case is called "nonmandatory coupling" since<br />

both A states can participate in the binding to some extent. In contrast, "mandatory<br />

coupling" occurs when only one state <strong>of</strong> A is capable <strong>of</strong> binding B, then K1 = 0 and γ = 0.<br />

In this case, the equations simplify significantly.<br />

Due to the two-state transition <strong>of</strong> A, the resultant binding free energy is always<br />

shifted toward less negative values:<br />

⎛ 1+<br />

⎞<br />

⎜ 1K<br />

ΔG°<br />

= ΔG°<br />

0 + RT ln<br />

⎟ . Eq. 1-19<br />

⎝1<br />

+ γ⋅1<br />

K ⎠<br />

It is interesting that macromolecular equilibria that have no intrinsic heat capacity<br />

changes resulting from surface-related phenomena and conformational stiffening (silencing<br />

<strong>of</strong> internal s<strong>of</strong>t modes) can yield a non-zero heat capacity change due only to the presence<br />

<strong>of</strong> coupled processes, e. g. due to ligand-induced shift in the equilibrium <strong>of</strong> protein states.<br />

The heat capacity change resulting from mandatory coupling is always negative, while the<br />

nonmandatory coupling can yield both negative and positive contribution, depending on<br />

the particular values <strong>of</strong> equilibrium constants and enthalpy changes.<br />

Hence, the overall heat capacity change <strong>of</strong> association that is accompanied by<br />

coupled processes can be written as a sum <strong>of</strong> the following terms:<br />

ΔC<br />

o p<br />

= ΔC<br />

+ ΔC<br />

o transfer<br />

p aliphatic<br />

o transfer<br />

+ ΔCp<br />

aromatic<br />

o<br />

p mandatory _ coupling<br />

+ Δ<br />

+ ΔC<br />

C<br />

o transfer<br />

p polar<br />

o<br />

p nonmandatory<br />

_ coupling<br />

1<br />

+ ΔC<br />

2<br />

o<br />

pint<br />

ernal _ s<strong>of</strong>t _ vibrations<br />

1<br />

+<br />

Eq. 1-20


2. Aim <strong>of</strong> The Work<br />

While the structure determinations provided important information on the mode <strong>of</strong><br />

cap binding, the consistent energetic description <strong>of</strong> the cap-<strong>eIF4E</strong> interaction was still<br />

lacking. In order to build a biophysical basis <strong>of</strong> the cap-dependent translation initiation and<br />

explain some discrepancies among structural, biochemical and biological<br />

observations, 75,155-157 measurements <strong>of</strong> molecular interactions in solution were necessary.<br />

The <strong>eIF4E</strong> protein possesses eight tryptophans which are conserved both in number<br />

and location, 21 and binding <strong>of</strong> cap results in quenching <strong>of</strong> the intrinsic protein<br />

fluorescence. 5,6,156,158-164 However, previous fluorescence measurements <strong>of</strong> the interaction<br />

between cap analogues and <strong>eIF4E</strong> purified by m 7 G-affinity chromatography yielded<br />

puzzling data. The reported association constants, 5,156,161-163 in the range <strong>of</strong> 10 5 -10 6 M -1 , did<br />

not reflect the differing inhibitory potency <strong>of</strong> cap analogues observed in vitro, 157 and<br />

contrasted <strong>with</strong> the structural data. 75,91 The crystal structure suggested that the binding<br />

constant for the <strong>eIF4E</strong>-m 7 GDP complex should be a few orders <strong>of</strong> magnitude higher, since<br />

there was a clearly visible hole in the middle <strong>of</strong> the six-membered ring <strong>of</strong> 7-<br />

methylguanosine in the electron density omit map. Such a high resolution <strong>with</strong>in the<br />

protein binding centre is usually observed by crystallographers for the complexes that have<br />

binding constants much higher than 10 6 . 165 These divergences were a prompt to perform<br />

systematic measurements to reveal the true affinity <strong>of</strong> cap analogues for <strong>eIF4E</strong>.<br />

The work was focused on the process <strong>of</strong> association <strong>of</strong> recombinant untagged murine<br />

<strong>eIF4E</strong> (residues 28-217) and full length human <strong>eIF4E</strong> <strong>with</strong> a large series <strong>of</strong> mono- and<br />

dinucleotide cap analogues, in order to attain several particular goals leading together to a<br />

complete decription <strong>of</strong> the molecular mechanism <strong>of</strong> specific recognition <strong>of</strong> the <strong>mRNA</strong> 5'<br />

structure by <strong>eIF4E</strong>. The particular goals were as follows:<br />

1. determination <strong>of</strong> association constants that would reflect the structure-activity<br />

relationship;<br />

2. parsing <strong>of</strong> the free energy <strong>of</strong> the binding into contributions <strong>of</strong> individual structural<br />

elements in the cap-binding centre,<br />

3. elucidation <strong>of</strong> the nature <strong>of</strong> the stacking-hydrogen bonding cooperativity,<br />

4. searching for conformational changes and water exchange between the complex and<br />

bulk solution,<br />

20


5. revealing the role for ionic equilibria and electrostatic interactions in cap-<strong>eIF4E</strong><br />

recognition,<br />

6. analysis <strong>of</strong> the specific binding between <strong>eIF4E</strong> and <strong>mRNA</strong> 5' cap in terms <strong>of</strong> exact<br />

quantitative thermodynamic parameters determined independently from the van't<br />

H<strong>of</strong>f equation and isothermal titration calorimetry (ITC).<br />

Such quantitative data are <strong>of</strong> primary importance for rational design <strong>of</strong> new cap-<br />

analogues <strong>of</strong> potential therapeutic activity since the high <strong>eIF4E</strong> cellular level is relevant to<br />

malignancy and apoptosis 13 .<br />

The fluorescence affinity measurements have been extended to include ternary<br />

complexes which consisted <strong>of</strong> <strong>eIF4E</strong>, a cap-analogue, and synthetic peptides corresponding<br />

to the <strong>eIF4E</strong> recognition motifs from the mammalian proteins eIF4GI, eIF4GII, 4E-BP1,<br />

and 4E-BP1 peptides monophosphorylated at Ser65 and diphosphorylated at Ser65/Thr70.<br />

The particular questions asked in the work were related to two issues:<br />

1. what is the influence <strong>of</strong> phosphorylation at Ser65 and Thr70 on regulation <strong>of</strong> 4E-BP1<br />

binding to <strong>eIF4E</strong>,<br />

2. how to reconcile a lot <strong>of</strong> contradictory biological and biochemical experimental data<br />

reported hitherto, regarding the putative cooperation between two <strong>eIF4E</strong> binding<br />

sites. 166-168<br />

21


3. Materials and Methods<br />

3.1. Chemical and Biochemical Syntheses<br />

3.1.1. Cap Analogues<br />

Syntheses and purification <strong>of</strong> cap analogues (Scheme 3-1) as sodium salts were<br />

performed by Dr. Dr. Edward Dar¿ynkiewicz, Janusz Stêpiñski and Marzena Jankowska-<br />

Anyszka as described previously 169-173 . Their concentrations were obtained from weighed<br />

amounts (± 5 %) and checked spectrophotometrically. 157 The cap analogues in solution at<br />

elevated pH (> 8) undergo opening <strong>of</strong> the five-membered ring <strong>of</strong> 7-substituted guanine,<br />

followed by hydrolysis <strong>of</strong> the glycosidic bond. 174 As checked by NMR, 175 the cap was very<br />

stable at pH 7.2, and stable enough to perform fast experiments at higher pH.<br />

3.1.2. Peptides<br />

Peptides corresponding to residues 569-580 <strong>of</strong> mammalian eIF4GI:<br />

KKRYDREFLLGF, 34 621-637 <strong>of</strong> mammalian eIF4GII: KKQYDREFLLDFQFMPA, 35 and<br />

51-67 <strong>of</strong> mammalian 4E-BP1: RIIYDRKFLMECRNSPV 67 were synthesized by Dr. Joseph<br />

Marcotrigiano (The Rockefeller University, N.Y., U.S.A.) by Boc protocols for solid phase<br />

peptide synthesis and cleaved using a standard HF procedure. 176 Syntheses <strong>of</strong> a<br />

mammalian phosphopeptide P-Ser65 51-67 4E-BP1 and a diphosphopeptide P-<br />

Ser65/Thr70 51-75 4E-BP1 RIIYDRKFLMECRNSpPVTKTpPPKDL 67 were performed<br />

by Dr. Aleksandra Wys³ouch-Cieszyñska (IBB PAS, Warszawa) on a Wang resin<br />

(Novabiochem) using Fmoc protocols for solid phase phosphopeptide synthesis. 1 Peptides<br />

were purified to homogeneity by semi-preparative HPLC (> 95 % purity), and<br />

characterized by MALDI-TOF or ESI mass spectrometry (569-580 eIF4GI peptide,<br />

predicted mass 1697.0 Da, measured mass 1698.0 Da; 621-637 eIF4GII peptide, predicted<br />

mass 2177.8 Da, measured mass 2178.0 Da; 51-67 4E-BP1 peptide, predicted mass 2141.8<br />

measured mass 2141; phosphopeptide, P-Ser65 51-67 4E-BP1, predicted mass 2220.1,<br />

measured mass 2221.0; diphosphopeptide, P-Ser65/Thr70 51-75 4E-BP1, predicted mass<br />

3180.55, measured mass 3181.0).<br />

Concentrations <strong>of</strong> the five investigated peptides were determined by acid gas-phase<br />

hydrolysis and three independent repeats <strong>of</strong> amino acid analysis, by Dr. Pawe³ Mak<br />

(Jagiellonian University, Kraków).<br />

22


Scheme 3-1. Structures <strong>of</strong> the 16 methylated cap-analogues. Φ denotes the phenyl ring. Protons<br />

which partially dissociate at pH 7.2, are marked <strong>with</strong> asterisk (pKa N(1)-H ~7.24-7.54, depending on<br />

R2, R3, and n 177 ; pKa phosph ~6.1-6.5, depending on n 178 ).<br />

m 7 GTP: R1 = CH3, R2 = R3 = H, n = 3;<br />

m 7 GDP: R1 = CH3, R2 = R3 = H, n = 2;<br />

m 7 GMP: R1 = CH3, R2 = R3 = H, n = 1;<br />

m 7 G: R1 = CH3, R2 = R3 = H, n = 0;<br />

et 7 GTP: R1 = CH2-CH3, R2 = R3 = H, n = 3;<br />

bz 7 GTP R1 = CH2-Φ, R2 = R3 = H, n = 3;<br />

p-Cl-bz 7 GTP: R1 = CH2-Φ-Cl, R2 = R3 = H, n = 3;<br />

m 7 , 2<br />

2 GTP : R1 = CH3, R2 = CH3, R3 = H, n = 3;<br />

m 7 , 2 , 2<br />

R2<br />

R2<br />

* H<br />

N<br />

R3<br />

*<br />

H<br />

N<br />

R3<br />

O<br />

N 6<br />

1<br />

2<br />

3<br />

N<br />

N<br />

O<br />

N<br />

R1<br />

N<br />

7<br />

+<br />

9<br />

N<br />

8<br />

H<br />

H<br />

H<br />

H<br />

R1<br />

N<br />

+<br />

N<br />

H<br />

OH<br />

HO<br />

3 GTP : R1 = CH3, R2 = R3 = CH3, n = 3;<br />

O<br />

H<br />

H H<br />

OH<br />

HO<br />

O<br />

m 7 Gpppm 2’O G: R1 = CH3, R2 = R3 = H, n = 3, R4 = OCH3, B = G<br />

m 7 GpppG: R1 = CH3, R2 = R3 = H, n = 3, R4 = OH, B = G<br />

m 7 GpppA: R1 = CH3, R2 = R3 = H, n = 3, R4 = OH, B = A<br />

m 7 GpppC R1 = CH3, R2 = R3 = H, n = 3, R4 = OH, B = C<br />

m 7 Gpppm 7 G: R1 = CH3, R2 = R3 = H, n = 3, R4 = OH, B = m 7 G<br />

m 7 GppppG: R1 = CH3, R2 = R3 = H, n = 4, R4 = OH, B = G<br />

m 7 Gppppm 7 G: R1 = CH3, R2 = R3 = H, n = 4, R4 = OH, B = m 7 G<br />

23<br />

CH 2<br />

H<br />

CH 2<br />

H<br />

O<br />

O (P O) n<br />

O -<br />

O<br />

O (P O) n CH 2<br />

O -<br />

H<br />

H *<br />

H<br />

OH<br />

O<br />

H<br />

B<br />

H<br />

R4


A dodecapeptide DGIEPMWEDEKN was kindly provided by Dr. Laslo Balaspiri<br />

(A. Szent-Gyorgyi Medical University, Szeged, Hungary).<br />

All other chemicals were analytical grade, purchased from Sigma-Aldrich, Merck,<br />

Carl Roth (Germany) or Fluka (U.S.A.).<br />

3.1.3. <strong>Protein</strong> Biosynthesis<br />

Human eukaryotic initiation factor <strong>eIF4E</strong> (residues 1-217) and murine <strong>eIF4E</strong><br />

(residues 28-217 and residues 33-217) was expressed by Mrs. Lidia Chlebicka <strong>with</strong> a help<br />

<strong>of</strong> Dr. Micha³ Dadlez (IBB PAS, Warszawa) in Escherichia coli (strain<br />

BL21(DE3)pLys). 179,180 The bacterial cells were transformed by a pET11d plasmid,<br />

containing the cloned coding region for <strong>eIF4E</strong>, and the T4 promoter. Induction <strong>of</strong> T7<br />

polymerase in liquid culture <strong>of</strong> bacteria on Luria-Bertani broth, <strong>with</strong> ampicillin and<br />

chloramphenicol, was initiated by addition <strong>of</strong> isopropyl-β-D-galactopyranoside. Bacterial<br />

pellets were lysed by sonication.<br />

The human protein was initially isolated from the soluble fraction by means <strong>of</strong><br />

affinity chromatography 181 and purified on an ion exchange MonoQ column. Next, the<br />

human and the murine proteins were purified from inclusion bodies pellets and folded by<br />

one step dialysis from 6 M guanidine hydrochloride, followed by ion-exchange<br />

chromatography on a MonoQ (human) or MonoS (murine) column, thus avoiding contact<br />

<strong>with</strong> cap analogues at any stage <strong>of</strong> purification. The protein purity, checked by a routine<br />

SDS PAGE, was > 95 %.<br />

3.2. Preparation <strong>of</strong> <strong>Protein</strong> and Peptide Samples to<br />

Spectroscopic Measurements<br />

The protein solutions were buffer exchanged if necessary <strong>with</strong> use <strong>of</strong> Ultrafree-15 ml<br />

filters <strong>with</strong> Biomax 5 kDa NMWL membrane (Millipore Co., U.S.A.). Immediately before<br />

the spectroscopic measurements, the protein sample was filtered through Millipore<br />

Ultrafree-0.5 ml Biomax 100 kDa NMWL. Total concentration <strong>of</strong> <strong>eIF4E</strong> was determined<br />

from absorbance ( 280 53900 = ε cm-1M -1 ).<br />

Solutions <strong>of</strong> 4E-BP1 peptides were strictly controlled for lack <strong>of</strong> disulphide dimer<br />

formation. Dimers occurred in the 4E-BP1 solution because <strong>of</strong> lack <strong>of</strong> any additives in the<br />

stock water solutions, which was necessary for accurate determination <strong>of</strong> the peptide<br />

24


concentrations by amino acid analysis. Peptide dimers were broken if incubated at very<br />

high concentration <strong>of</strong> the reducing agent (13 mM DTT) at pH 7.5 for two weeks. This<br />

secured satisfactory monomerization, as rigorously checked on analytical HPLC (Waters<br />

Inc.) and ESI mass spectroscopy (Q-T<strong>of</strong> 2 spectrometer, Micromass Inc.). Then, the<br />

affinity measurements could be performed properly <strong>with</strong> use <strong>of</strong> more dilute peptide<br />

solution at 1 mM DTT. The problem concerned the phosphorylated 4E-BP1 peptides to<br />

lesser extent, since the presence <strong>of</strong> the negatively charged phosphate groups at Ser65 in the<br />

vicinity <strong>of</strong> Cys62 can partially prevent the dimerization.<br />

3.3. Absorption and Emission Spectroscopy<br />

Emission <strong>of</strong> light from a fluorophore occurs from electronically excited states. 182<br />

Return to the ground electronic state (S0) from excited singlet states (Sn) is attained by a<br />

rapid (after 10 -10 - 10 -8 s), spin-allowed emission <strong>of</strong> a photon (fluorescence), while<br />

transition from excited tripled states (phosphorescence) is forbidden and the lifetimes are<br />

longer (10 -3 – 10 1 s). Phosphorescence is not visible in aqueous solutions at temperatures<br />

applied during this work and will not be considered.<br />

Light absorption in UV-Vis wave region occurs typically from the lowest vibrational<br />

level (<strong>of</strong> S0) to some higher vibrational level <strong>of</strong> S1 or S2 in a time <strong>of</strong> about 10 -15 s. In<br />

solutions, molecules rapidly relax to the lowest vibrational level <strong>of</strong> S1 via internal<br />

conversion in a time ≤ 10 -12 s. Fluorescence return to S0 occurs to one <strong>of</strong> its vibrational<br />

levels. Then the molecule quickly reaches thermal equilibrium (10 -12 s). The rapid<br />

irradiative decay to the lowest vibrational level <strong>of</strong> S1 results in the Stokes' shift between the<br />

absorption and emission spectra. The shift is further enhanced by possible decay to higher<br />

vibrational levels <strong>of</strong> S0, reorientation <strong>of</strong> fluorophore in respect to solvent molecules,<br />

excited-state reactions, complex formation, and energy transfer.<br />

When incident light (intensity I0) is absorbed by a diluted chromophore solution<br />

(concentration [ch], extinction coefficient ε) on the optical length (l), then the absorbance<br />

(A) and the intensity <strong>of</strong> transmitted light (I1) can be reliably expressed by the linear Beer-<br />

Lambert law: 183<br />

light:<br />

I0<br />

A = log = ε ⋅[<br />

ch]<br />

⋅ l . Eq. 3-1<br />

I<br />

1<br />

The rate <strong>of</strong> emission <strong>of</strong> fluorescence (F) is related to the intensity <strong>of</strong> the absorbed<br />

25


F = (I0 − I1) ⋅ q, Eq. 3-2<br />

where q is the quantum yield <strong>of</strong> fluorescence. For weakly absorbing solutions for which<br />

the optical density is very small, the equation can be simplified to:<br />

F = 2.3⋅I0⋅ε⋅[ch]⋅l⋅q. Eq. 3-3<br />

Fluorescence intensities are proportional to the fluorophore concentration over a very<br />

limited range <strong>of</strong> optical densities, usually 2 orders <strong>of</strong> magnitude lower than those which<br />

satisfy the Beer-Lambert law for absorption. If the conditions ensuring linearity are<br />

satisfied, the observed fluorescence changes can be directly interpreted in terms <strong>of</strong><br />

changing concentrations <strong>of</strong> the species present in a studied sample. Otherwise suitable<br />

corrections for the inner filter effect are necessary 183 (see below, 3.4.2.1., p. 31).<br />

3.3.1. Fluorescence Quenching<br />

Fluorescence intensity can be decreased by: 182<br />

1. dynamic quenching:<br />

a) irradiative return do S0 by a diffusive collisions <strong>of</strong> the excited fluorophore <strong>with</strong><br />

other molecules,<br />

b) resonance energy transfer (RET) <strong>of</strong> the excited-state energy from an excited donor<br />

to an acceptor <strong>with</strong>out the appearance <strong>of</strong> a photon, as a result <strong>of</strong> long-range<br />

dipole-dipole interactions,<br />

2. static quenching by formation <strong>of</strong> nonfluorescent complexes <strong>of</strong> fluorophore <strong>with</strong> a<br />

quencher in the fluorophore ground state,<br />

3. inner filter effect, i. e. attenuation <strong>of</strong> the excitation beam and/or reabsorption <strong>of</strong> the<br />

emitted beam by a layer <strong>of</strong> an optically dense solution.<br />

3.3.1.1. Static Quenching<br />

Static quenching results from formation <strong>of</strong> a dark complex that returns to the ground<br />

state immediately after absorption <strong>with</strong>out emission <strong>of</strong> a photon. A fraction <strong>of</strong> the<br />

fluorophores is removed from observation and, contrary to dynamic quenching, the<br />

lifetime <strong>of</strong> the uncomplexed fraction is unchanged. The quenching constant is the<br />

association constant <strong>of</strong> the dark complex: 182<br />

[ cx]<br />

K s = , Eq. 3-4<br />

[ f ] ⋅[<br />

qu]<br />

0<br />

0<br />

26


where [cx], [f]0 and [qu]0 are the equilibrium concentrations <strong>of</strong> the complex, free<br />

fluorophore and free quencher, respectively. The fluorescence changes analogically as<br />

upon the dynamic quenching (below):<br />

F0<br />

= 1+<br />

K s[<br />

qu]<br />

0 , Eq. 3-5<br />

F<br />

but the equilibrium concentration the free quencher must be analysed here.<br />

3.3.1.2. Dynamic Quenching<br />

Dynamic quenching by a quencher at a total concentration [qu] is a rate process<br />

which depopulates the excited state, leading to shortening <strong>of</strong> fluorescence lifetimes (τ0).<br />

The resultant lifetime in the presence <strong>of</strong> the quencher (τ) is shorter by a decay rate (kq)<br />

which is characteristic to the given quencher and environment:<br />

τ0<br />

= 1+ k qτ0<br />

[ qu]<br />

. Eq. 3-6<br />

τ<br />

An equivalent decrease in fluorescence intensity is:<br />

0 F<br />

F<br />

τ<br />

=<br />

τ<br />

0<br />

. Eq. 3-7<br />

The Stern-Volmer quenching constant is defined as:<br />

KSV = kq ⋅ τ0, Eq. 3-8<br />

so that the Stern-Volmer equation is expressed as:<br />

F0<br />

= 1+<br />

K SV[<br />

qu]<br />

. Eq. 3-9<br />

F<br />

3.3.1.3. Resonance Energy Transfer<br />

The rate <strong>of</strong> RET (kT) is related to a factor describing the relative space orientation <strong>of</strong><br />

the transition dipoles <strong>of</strong> the donor and acceptor (κ 2 ), to the quantum yield <strong>of</strong> the donor (qd),<br />

to the lifetime <strong>of</strong> the donor in the absence <strong>of</strong> the acceptor (τd), to the reciprocal <strong>of</strong> sixth<br />

power <strong>of</strong> their distance (r), and to the integral <strong>of</strong> spectral overlap between the donor<br />

emission and the acceptor absorption (J(λ)):<br />

k<br />

T<br />

q<br />

~<br />

τ<br />

d<br />

d<br />

⋅ κ<br />

⋅ r<br />

2<br />

6<br />

J(<br />

λ)<br />

. Eq. 3-10<br />

27


The distance at which the transfer rate is equal to the decay rate <strong>of</strong> the donor in the<br />

absence <strong>of</strong> acceptor (by the radiative and nonradiative ways) is called the Förster distance<br />

(R0), usually ~ 30 – 60 Å.<br />

The orientation factor (κ 2 ) is given by:<br />

2<br />

κ = cos θ − 3cos<br />

θ ⋅ cos θ , Eq. 3-11<br />

T<br />

d<br />

a<br />

where θT is the angle between the donor and acceptor dipoles, θd and θa are the angles<br />

between these dipoles and the vector joining them. The κ 2 value can range from 0 for the<br />

perpendicular orientation <strong>of</strong> the dipoles to 4 for the collinear, parallel orientation.<br />

3.3.1.3.1. Resonance Energy Transfer in <strong>eIF4E</strong><br />

The <strong>eIF4E</strong> molecule contains 6 tyrosines and 8 tryptophans. 50 The diameter <strong>of</strong> the<br />

molecule is ~ 40 Å, i. e. in the typical range <strong>of</strong> the Förster distances. Hence, the tyrosine<br />

(donor) fluorescence can be transferred to tryptophans (acceptor), 182 and the shape <strong>of</strong> the<br />

emission spectrum excited at 280 nm is dominated by the final tryptophan fluorescence<br />

(Fig. 4-1, p. 49). It has been checked by comparison <strong>of</strong> <strong>eIF4E</strong> emission spectra registered<br />

for different excitation wavelengths (λex) that the contribution <strong>of</strong> the tyrosine emission in<br />

the resultant fluorescence intensity at λex = 280 nm is not more than 10 %. Putative<br />

conformational changes can modulate both tyrosine and tryptophan emission but<br />

eventually the spectral information about the changes is provided by the tryptophan<br />

emission.<br />

One issue is the interaction <strong>of</strong> <strong>eIF4E</strong> <strong>with</strong> the tyrosine containing peptides 4E-BP1,<br />

4E-BP1 P-Ser65, 4E-BP1 P-Ser65/Thr70, eIF4GI, and eIF4GII. The peptide tyrosine is<br />

distant by only 10 Å from Trp73 <strong>of</strong> <strong>eIF4E</strong> in the bound state. 91 However, the mutual<br />

orientation <strong>of</strong> the two aromatic rings is almost perpendicular (Fig. 4-22, 23, p. 86-87). The<br />

tyrosine is stabilized in this plane by a close phenylalanine ring and an –OH hydrogen<br />

bond, and the tryptophan is tightly bound by a network <strong>of</strong> intra- and intermolecular<br />

interactions. These constraints suggest that in spite <strong>of</strong> proximity <strong>of</strong> the donor and acceptor<br />

the RET is most probably inefficient.<br />

3.3.2. Tryptophan Fluorescence in <strong>Protein</strong>s<br />

Tryptophan fluorescence varies widely in quantum yield and in the wavelength <strong>of</strong><br />

maximum intensity among proteins and solvents. 182 The 280-nm absorption band <strong>of</strong> Trp is<br />

28


the result <strong>of</strong> transition to two excited states, called 1 La and 1 Lb, <strong>of</strong> considerably different<br />

properties. 184 The 1 La state has a large dipole moment, and thus the high sensitivity to<br />

environment. Transition from 1 La usually yields more fluorescence but 1 Lb can be<br />

sometimes also the emitting state and then they are hardly distinguishable in proteins. The<br />

transitions have different structures, intensities and relative spectral positions, but the<br />

details are blurred by inhomogeneous broadening in proteins due mainly to the variation <strong>of</strong><br />

the relative orientation <strong>of</strong> the local electric field felt by tryptophans. This solvent-solute<br />

interaction accounts for the large Stokes' shift in water and much less inside the protein<br />

hydrophobic core. The large red shift is related to the cooperative effect <strong>of</strong> many water<br />

molecules, ~ 100 cm -1 /1 water molecule <strong>with</strong>in 6 Å. 184 Additionally, protein tryptophans<br />

can be dynamically quenched by neighbouring charged amino acids. 182 Hence, the<br />

spectroscopic properties (quantum yield, accurate shape <strong>of</strong> the spectrum e.t.c.) <strong>of</strong> <strong>eIF4E</strong><br />

were not the main focus <strong>of</strong> the work to avoid too far-reaching, uncertain conclusions.<br />

Instead, the fluorescence intensity was used to monitor quantitatively the changes<br />

experienced by <strong>eIF4E</strong>.<br />

3.3.3. Quenching <strong>of</strong> <strong>eIF4E</strong> Fluorescence by Cap Analogues<br />

The intrinsic protein fluorescence quenching observed upon formation <strong>of</strong> the<br />

complex <strong>of</strong> <strong>eIF4E</strong> <strong>with</strong> a cap analogue can result from two factors: direct quenching <strong>of</strong> the<br />

tryptophans 102, 56 and 166 in the cap-binding site, and changes <strong>of</strong> local environment <strong>of</strong><br />

other, more distant fluorescent amino acids as a result <strong>of</strong> possible conformational changes<br />

<strong>of</strong> the protein upon the binding. The former is thought to be static quenching, since:<br />

1. The Förster resonance energy transfer in unlikely due to several reasons:<br />

a) Spectral overlapping <strong>of</strong> the cap analogue emission and the protein absorption<br />

spectra, and reversely (Fig. 4-1, p. 49), the protein emission and the cap analogue<br />

absorption spectra, is negligible.<br />

b) Neither any batochromically shifted emission band nor fluorescence<br />

enhancement <strong>of</strong> cap analogues were observed. The emission <strong>of</strong> the free cap in the<br />

presence <strong>of</strong> the protein obtained by fitting <strong>of</strong> the full equation (Eq. 3-26, p. 34)<br />

was the same as the emission in the absence <strong>of</strong> the protein;<br />

c) Brownian dynamic calculations involving electrostatic effects together <strong>with</strong> the<br />

stopped-flow binding studies 4 suggested that quenching can occur only when the<br />

ligand is apart from its final position known from the crystal structure 75 not more<br />

29


than a few angströms. It would probably still diminish this distance if polarization<br />

effects were taken into account in the calculations. 185<br />

2. Collisional quenching <strong>of</strong> the protein tryptophans seems to be unlikely at such low<br />

concentration <strong>of</strong> the protein, 50 nM – 1 μM, and <strong>of</strong> the cap analogue, not exceeding<br />

5 μM for the most specific analogue and 150 μM for the least specific one.<br />

However, the unambiguous assessment could be only provided by time-resolved<br />

measurements.<br />

Fluorescence changes <strong>of</strong> tyrosines, and the tryptophans which do not have direct<br />

contact <strong>with</strong> cap are related to dynamic quenching by surrounding charged amino acids, 182<br />

changes <strong>of</strong> mutual distances and orientations in Trp-Tyr pairs, and changes <strong>of</strong> local<br />

polar/hydrophobic neighbourhood. The quenching has a dynamic character but reflects the<br />

changes that the protein undergoes upon cap binding and thus depends not on the total<br />

quencher concentration but on the equilibrium concentration. Hence, it is described by the<br />

same equation as static quenching.<br />

3.4. Experimental Conditions <strong>of</strong> Spectroscopic Measurements<br />

Absorption and fluorescence spectra were recorded on Lambda 20 UV/VIS and LS-<br />

50B instruments (Perkin-Elmer Co., Norwalk, CT., USA), in a quartz semi-micro cuvette<br />

(Hellma, Germany) <strong>with</strong> optical lengths 4 mm and 10 mm for absorption and emission,<br />

respectively. The titration experiments were performed in a standard buffer: 50mM<br />

Hepes/KOH pH 7.20, 100mM KCl, 1mM dithiothreitol (DTT) and 0.5mM disodium<br />

ethylenediaminetetraacetate (EDTA), except for the experiments at variable pH and ionic<br />

strength. pH (± 0.01 unit) was measured independently at each temperature and ionic<br />

strength (Beckman Φ300 pH-meter, Germany). Solutions were filtered through 0.22 μm<br />

pore size. <strong>Protein</strong> samples were s<strong>of</strong>tly degassed prior to titration. The cuvette was<br />

thermostated and the temperature was controlled <strong>with</strong> a thermocouple inside it (± 0.2 °C).<br />

For the <strong>eIF4E</strong>-cap association the excitation wavelength <strong>of</strong> 280nm (slit 2.5nm, auto<br />

cut-<strong>of</strong>f filter), and the emission wavelength <strong>of</strong> 335, 336 or 337nm (slit 2.5 to 4nm, 290 nm<br />

cut-<strong>of</strong>f filter) were applied, <strong>with</strong> an automatic correction for the photomultiplier sensitivity.<br />

For the <strong>eIF4E</strong>-peptide binding studies, the excitation wavelength <strong>of</strong> 290nm and the<br />

emission wavelength <strong>of</strong> 350 or 355 nm were used. These conditions ensured observation <strong>of</strong><br />

the protein tryptophan emission, only. The fluorescence intensity was monitored by the<br />

30


egistration <strong>of</strong> the whole spectrum (310 nm to 400 nm) and during continuous, time-<br />

synchronized titration at a single wavelength, <strong>with</strong> the integration time <strong>of</strong> 30 seconds and<br />

the gap <strong>of</strong> 30 seconds for adding the ligand, <strong>with</strong> slow but sufficient magnetic stirring to<br />

ensure mixing and keeping the temperature constant in the whole volume <strong>of</strong> the cuvette.<br />

During the gap, the UV xenon flash lamp was switched <strong>of</strong>f to avoid photobleaching the<br />

photosensitive protein sample. The cuvette has not been touched during the whole titration<br />

experiment to ensure constant geometry for the optical measurements.<br />

3.4.1. Titration Assay<br />

Titrations were performed for <strong>eIF4E</strong> at several concentrations (50 nM to 1 μM), in<br />

steady-state conditions provided by preincubation <strong>of</strong> <strong>eIF4E</strong> in the buffer <strong>of</strong> the appropriate<br />

pH and ionic strength for the given experiment. 1 μl aliquots <strong>of</strong> increasing concentrations<br />

(1 μM to 5 mM) <strong>of</strong> a ligand were injected manually to 1400 μl <strong>of</strong> <strong>eIF4E</strong> solution. Each<br />

titration consisted <strong>of</strong> 28 to 45 data points <strong>with</strong> a suitable number <strong>with</strong>in the range, at which<br />

the total ligand concentration ([L]) was close to the concentration <strong>of</strong> active <strong>eIF4E</strong> ([Pact])<br />

(Fig. 4-7, p. 57). The curvature <strong>of</strong> the fitted function (Eq. 3-26, p. 34) in this range,<br />

∂<br />

2<br />

F<br />

∂[<br />

L]<br />

2<br />

[L] = [Pact<br />

]<br />

~<br />

K<br />

P<br />

as<br />

act<br />

, Eq. 3-12<br />

mostly influences the accuracy <strong>of</strong> fitting <strong>of</strong> [Pact] and Kas. For determination <strong>of</strong> the active<br />

protein fraction in the samples studied by stopped-flow fluorimetry and ITC, the cap<br />

analogue <strong>of</strong> the highest association constant for <strong>eIF4E</strong> was used (7-methylGTP), since this<br />

Kas ensures the optimal quotient <strong>of</strong> Kas/[Pact].<br />

3.4.2. Fluorescence Data Corrections<br />

3.4.2.1. Inner Filter Effect<br />

The fluorescence intensities were corrected for the inner filter effect 183 (Fig. 3-1).<br />

The cuvette had two thicker walls and thus the optical pathlength for absorption was<br />

significantly shorter than that for emission (4 mm × 10 mm), which minimized both the<br />

inner filter effect and a recapture <strong>of</strong> the excitation beam from a little mirror behind the<br />

thick cuvette wall. A following polynomial dependence <strong>of</strong> the fluorescence intensity (F) on<br />

the absorbance (A):<br />

F(<br />

A)<br />

2<br />

3<br />

4<br />

= 1−<br />

2.<br />

348⋅<br />

A + 7.<br />

56 ⋅ A − 20.<br />

2 ⋅ A + 21.<br />

5 ⋅ A<br />

Eq. 3-13<br />

F(<br />

0)<br />

31


proved to be better than the correction proposed for the typical 10 mm × 10 mm cuvette 186<br />

(P < 0.0001). The experimental data (Fobs) were divided by the above factor (F(A)/F(0))<br />

determined individually for each titration, yielding the data-set corrected for absorption<br />

(FcorrA):<br />

F(<br />

A)<br />

= F<br />

Eq. 3-14<br />

F(<br />

0)<br />

FcorrA obs<br />

The inner filter effect was negligible for the specific cap analogues but could change Kas<br />

values ~ 2-fold for the weakly interacting and strongly absorbing cap analogues (Fig. 3-1).<br />

F(A)/F(0)<br />

Residuals<br />

1.0<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.02<br />

0<br />

-0.02<br />

Figure 3-1. (left) Inner filter effect measured for tryptophan solution in the 4 mm × 10 mm cuvette,<br />

titrated by GMP in 50mM Hepes/KOH buffer, pH 7.20, 100mM KCl. A comparison <strong>of</strong> a<br />

polinomial function (solid line, black residuals) <strong>with</strong> the analytical function proposed for the 10<br />

mm × 10 mm cuvette 186 (broken line, open residuals). The maximal GMP concentration is < 100<br />

μM, which rules out a possibility <strong>of</strong> dynamic quenching <strong>of</strong> tryptophan. (right) Fluorescence<br />

titration <strong>of</strong> <strong>eIF4E</strong> <strong>with</strong> GDP. Data uncorrected () and corrected () for the inner filter effect.<br />

3.4.2.2. Dilution<br />

0.0 0.1 0.2 0.3 0.4<br />

Absorbance<br />

0.0 0.1 0.2 0.3 0.4<br />

The final dilution was always ≤ 3.2%. Suitable data correction was applied for these<br />

titrations for which the final dilution was ≥ 2%. FcorrV is the corrected fluorescence, Fobs is<br />

the measured fluorescence, V is the actual sample volume at each point <strong>of</strong> the titration, and<br />

V0 is the initial sample volume:<br />

V<br />

FcorrV = Fobs<br />

⋅ . Eq. 3-15<br />

V<br />

0<br />

Fluorescence (a. u.)<br />

32<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0 20 40 60 80 100 120<br />

GDP (μM)


3.4.2.3. Fluorescence Drift in Time<br />

In some cases at higher temperatures (~ 40 °C) the protein fluorescence decreased<br />

somewhat in time due most probably to partial thermal denaturation <strong>of</strong> <strong>eIF4E</strong>. A correction<br />

was needed to cancel numerically the fluorescence changes that did not result from ligand<br />

binding. Thanks to synchronization <strong>of</strong> each injection during titration, a correction for a<br />

fluorescence decrease in the time <strong>of</strong> the titration experiment could be applied. Monitoring<br />

<strong>of</strong> the fluorescence signal was continued after completing the titration by the next 10 - 20<br />

minutes in the same way as during the titration. The exponential curve was fitted to the<br />

time-dependent data at the constant ligand concentration:<br />

F ( t)<br />

= a ⋅ exp( −b<br />

⋅ t)<br />

+ c . Eq. 3-16<br />

The titration data were then corrected (FcorrT) by the exponential function:<br />

FcorrT obs<br />

= F ⋅ exp( b ⋅ t)<br />

. Eq. 3-17<br />

The latter two corrections had only slight influence on the resultant Kas values<br />

(<strong>with</strong>in 0.3 standard deviation) but improved the goodness <strong>of</strong> fit (R 2 and P value, see 3.9.,<br />

p. 46)<br />

3.5. Fluorescence Data <strong>Analysis</strong><br />

3.5.1. Simultaneous Determination <strong>of</strong> Association Constants and <strong>Protein</strong><br />

Activity<br />

As a result <strong>of</strong> binding <strong>of</strong> a ligand at a total ligand concentration <strong>of</strong> [L] to an active<br />

protein at a total concentration <strong>of</strong> [Pact], there are three species in the solution: the free<br />

ligand at the equilibrium concentration <strong>of</strong> [L]0, the free active protein at the equilibrium<br />

concentration <strong>of</strong> [Pact]0, and their complex at the equilibrium concentration <strong>of</strong> [cx]:<br />

[L] = [L]0 + [cx], Eq. 3-18<br />

[Pact] = [Pact]0 + [cx]. Eq. 3-19<br />

Equilibrium among the concentrations <strong>of</strong> free species and the complex is ruled by the<br />

association constant Kas that is defined in the same manner as the static quenching constant<br />

(Eq. 3-4):<br />

[ cx]<br />

K as = . Eq. 3-20<br />

[ L]<br />

[ P ]<br />

0<br />

act<br />

0<br />

Solution <strong>of</strong> the three above equations yields a square equation <strong>with</strong> the positive root for<br />

[cx]:<br />

33


[ cx]<br />

( K ([ L]<br />

−[<br />

P ]) + 1)<br />

2<br />

[ L]<br />

+ [ Pact<br />

] 1−<br />

as<br />

act + 4Kas<br />

⋅[<br />

Pact<br />

]<br />

= +<br />

Eq. 3-21<br />

2<br />

2Kas<br />

The total protein concentration [P] is the sum <strong>of</strong> the active fraction [Pact] and the<br />

inactive [Pinact] fraction that does interact <strong>with</strong> the ligand:<br />

[P] = [Pact] + [Pinact]. Eq. 3-22<br />

The initial fluorescence intensity (F(0)) <strong>of</strong> the pure protein is equal:<br />

F(0) = [Pact]⋅φPact_free + [Pinact]⋅φPinact. Eq. 3-23<br />

The observed fluorescence intensity (F) after addition <strong>of</strong> the fluorescent ligand <strong>of</strong> the<br />

fluorescence efficiency φlig-free is equal:<br />

F = [Pact]0⋅φPact_free + [cx]⋅φcx + [L]0⋅φlig_free + [Pinact]⋅φPinact. Eq. 3-24<br />

No assumptions regarding the fluorescence efficiency <strong>of</strong> the inactive protein are necessary.<br />

Simple substitution <strong>of</strong> F(0) and [cx], and definition <strong>of</strong> the difference between the<br />

fluorescence efficiencies <strong>of</strong> the apo-protein and the complex as:<br />

Δφ = φ − − φ<br />

Eq. 3-25<br />

Pact<br />

free<br />

cx<br />

yield the full equation which describes the fluorescence intensity as a function <strong>of</strong> the total<br />

ligand concentration <strong>with</strong>in the course <strong>of</strong> titration:<br />

= − ⋅ Δφ<br />

+ φ + ⋅ φ , Eq. 3-26<br />

F F(<br />

0)<br />

[ cx]<br />

( lig −free)<br />

[ L]<br />

lig −free<br />

This theoretical curve was fitted to the experimental data points. The parameters to be<br />

extracted from the fit were as follows: Kas, [Pact], Δφ, φlig-free, F(0). The latter two<br />

parameters were independently verified experimentally. The accordance between the non-<br />

linearly fitted φlig-free value and the value determined from linear regression to the<br />

fluorescence data for the free ligand at increasing concentrations in the absence <strong>of</strong> the<br />

protein was 4 %. For the experiments at the most elevated temperature, the parameter [Pact]<br />

had a meaning <strong>of</strong> an average value in the course <strong>of</strong> the titration.<br />

eliminated.<br />

Total quenching upon binding was calculated as:<br />

Q = F(<br />

0)<br />

− F(<br />

∞)<br />

= [ Pact<br />

] ⋅ Δφ<br />

. Eq. 3-27<br />

For the peptides, φlig-free was fixed as zero, since the direct tyrosine fluorescence was<br />

The final Kas were calculated as a weighted average <strong>of</strong> 3 to 12 independent titration<br />

series, except for m 7 GTP, for which more than 30 titration experiments were performed, in<br />

order to make statistical analysis <strong>of</strong> the method and as control for each protein batch. The<br />

34


Eadie-H<strong>of</strong>stee representation 187,188 was not used, for its basic assumption <strong>of</strong> a great excess<br />

<strong>of</strong> the ligand in relation to the protein is not satisfied here.<br />

3.5.2. The Gibbs Free Energy <strong>of</strong> Binding<br />

The Gibbs free energy <strong>of</strong> binding <strong>of</strong> cap analogues to <strong>eIF4E</strong> was calculated from Eq.<br />

1-9. The binding constants (Kas) were used for unsymmetrical dinucleotide and<br />

mononucleotide cap analogues, and the microscopic binding constants (Kas (micro) = 0.5 ⋅<br />

Kas) in case <strong>of</strong> the symmetrical cap analogues (m 7 Gpppm 7 G, m 7 Gppppm 7 G, GpppG)<br />

because <strong>of</strong> entropic effects.<br />

3.5.3. Osmotic Stress and Electrostatic <strong>Interaction</strong>s<br />

Increase <strong>of</strong> the KCl concentration at constant pH and temperature is accompanied by<br />

an increasing osmolality (2ϕ[KCl]) and a decreasing activity coefficient <strong>of</strong> water (aw):<br />

2ϕ⋅<br />

[ KCl]<br />

log( a w ) = − , Eq. 3-28<br />

ln10⋅<br />

Ω<br />

where: ϕ = 0 . 91±<br />

0.<br />

01,<br />

an osmotic coefficient <strong>of</strong> KCl approximately constant from 0 to<br />

0.5 M KCl; 189 Ω the number <strong>of</strong> moles <strong>of</strong> free solvent water in 1 kg <strong>of</strong> the solution. The<br />

latter is not assumed to be constant, but the presence <strong>of</strong> other buffer components <strong>of</strong> the<br />

molecular masses μi at the concentrations ci, and the hydration number <strong>of</strong> KCl ( ν ≈ 5),<br />

190<br />

are taken into account:<br />

1000 − μ KCl[<br />

KCl]<br />

− ∑μ<br />

ic<br />

i<br />

Ω =<br />

− ν ⋅[<br />

KCl]<br />

Eq. 3-29<br />

μ<br />

H O<br />

2<br />

Keeping constant Ω = 55.<br />

5 yields overestimation <strong>of</strong> the results concerning the number <strong>of</strong><br />

the water molecules that are exchanged between <strong>eIF4E</strong>-cap complex and the bulk solvent<br />

(ΔN) by ~ 18% (see next page).<br />

3.5.3.1. Davies-Stockes-Robinson Electrostatic Screening Approach<br />

Binding <strong>of</strong> two ions <strong>of</strong> opposite charges (protein z1 and ligand z2) is screened by the<br />

ionic atmosphere <strong>of</strong> the excess salt and other ionized components <strong>of</strong> the solution. The<br />

influence <strong>of</strong> ionic strength and osmotic stress on the activity coefficients <strong>of</strong> the interacting<br />

species leads to the expression for log(Kas) in function <strong>of</strong> [KCl]: 191,192<br />

35


2Az1z<br />

2 I<br />

log( K as ) = log( K as ( 0))<br />

+ + ΔN<br />

log( a w ) , Eq. 3-30<br />

1+<br />

a B I<br />

where ionic strength:<br />

j<br />

2<br />

I = [ KCl]<br />

+ ∑c<br />

iz i , Eq. 3-31<br />

zi is a charge <strong>of</strong> the i-th species, aj is a sum <strong>of</strong> radii <strong>of</strong> interacting species, and ΔN is a<br />

number <strong>of</strong> water molecules taken up (or released, if ΔN < 0) to the macromolecular<br />

surfaces upon complex formation. The temperature-dependent coefficients at 20°C are:<br />

A = 0.<br />

50585,<br />

B = 0.<br />

32789.<br />

192<br />

3.5.3.2. Wyman Linkage <strong>Analysis</strong><br />

Hydration effects as well as ionic interactions are considered stoichiometrically: 193<br />

0<br />

log( K as ) = log( K as ) − c ⋅ log( a KCl)<br />

+ ΔN<br />

⋅ log( a w ) , Eq. 3-32<br />

where c is the total number <strong>of</strong> all ions released from macromolecular surfaces upon<br />

binding. <strong>Thermodynamic</strong> activity <strong>of</strong> KCl (aKCl) is used. Assuming aKCl = [KCl] leads to<br />

underestimation <strong>of</strong> the Kas value extrapolated to 1 M KCl by ~ tw<strong>of</strong>old. Kas 0 corresponds to<br />

the value <strong>of</strong> Kas, when the thermodynamic effects <strong>of</strong> ion release (log([KCl]) = 0 at 1 M<br />

KCl) and <strong>of</strong> the hydration (log(aw) = 0 at 0 M KCl) cancel each other out. Kas 0 occurs at the<br />

KCl concentration <strong>of</strong> approximately 0.3 M. The ion release does not influence the resultant<br />

ionic strength, since the protein and the ligand are present at nano- to micromolar<br />

concentrations, while [KCl] ranges from 0.05 M to 1 M, and the buffer contribution to I at<br />

pH 7.2 is ~ 0.017 M.<br />

The contribution <strong>of</strong> the electrolyte effect (ΔG°el) to the stability <strong>of</strong> the cap-<strong>eIF4E</strong><br />

complex can be calculated from the derivative SKas: 118<br />

as:<br />

SK<br />

as<br />

⎛ ∂ log K as ⎞<br />

≡ ⎜ ⎟<br />

⎝ ∂ log[ KCl]<br />

⎠<br />

T,<br />

p<br />

2φ[<br />

KCl]<br />

= −c<br />

− ΔN<br />

Ω<br />

36<br />

Eq. 3-33<br />

Δ ° = −SK<br />

RT ln[ KCl]<br />

, Eq. 3-34<br />

G el as<br />

and the entropy <strong>of</strong> the electrolyte effect was calculated as:<br />

ΔG°<br />

el<br />

Δ S°<br />

el = − . Eq. 3-35<br />

T


3.5.4. Protonation Equilibria<br />

When the binding event is accompanied by protonation and deprotonation <strong>of</strong> two<br />

specifically interacting residues at constant temperature, ionic strength and osmotic stress,<br />

these processes interfere to produce "bell-like" shaped dependence <strong>of</strong> the binding constant<br />

on pH. In case <strong>of</strong> 7-methylGTP, the population was considered as a mixture <strong>of</strong> cationic<br />

(Lc) and zwitterionic (Lzw) forms:<br />

= [ L ] [ L ] . Eq. 2-36<br />

[ L]<br />

c zw +<br />

The protein was assumed to exist in equilibrium <strong>of</strong> two states, one <strong>with</strong> Glu103 protonated<br />

(Glu) and the other <strong>with</strong> Glu103 deprotonated (Glu), leading to:<br />

−<br />

[ Pact<br />

] = [ Pact<br />

( Glu)]<br />

+ [ Pact<br />

( Glu )] . Eq. 3-37<br />

When the cationic form <strong>of</strong> cap binds to <strong>eIF4E</strong> <strong>with</strong> deprotonated Glu103, the pH-<br />

independent association constant, for the species in their appropriate ionic states, can be<br />

expressed as:<br />

where:<br />

and<br />

−<br />

pH−ind<br />

[ LcPact<br />

( Glu<br />

K as<br />

−<br />

[ Lc<br />

] 0[<br />

Pact<br />

( Glu<br />

[ P<br />

act<br />

c<br />

( Glu<br />

)]<br />

= Eq. 3-38<br />

)]<br />

−<br />

c<br />

c<br />

act<br />

−<br />

0<br />

)] = [ L P ( Glu )] + [ P ( Glu )]<br />

Eq. 3-39<br />

[ L ] = [ L P ( Glu )] + [ L<br />

act<br />

−<br />

c<br />

]<br />

0<br />

act<br />

37<br />

−<br />

0<br />

. Eq. 3-40<br />

The populations <strong>of</strong> both the free protonated ligand ([Lc]0) and the free deprotonated protein<br />

([Pact(Glu − )]0) that actually do not form the complex, are in ionic equilibrium defined by<br />

the effective acidic dissociation constants KL and KP, the former for the ligand in the<br />

presence <strong>of</strong> the protein, and the latter for Glu103 in the presence <strong>of</strong> the cap analogue:<br />

and<br />

K<br />

K<br />

+<br />

[ Lzw<br />

][ H ]<br />

L = Eq. 3-41<br />

[ Lc<br />

] 0<br />

P<br />

[ Pact<br />

( Glu )] 0[<br />

H ]<br />

= Eq. 3-42<br />

[ P ( Glu)]<br />

act<br />

−<br />

The experimentally observed association constant is:<br />

K<br />

+<br />

[ LcPact<br />

( Glu )]<br />

= Eq. 3-43<br />

−<br />

−<br />

([ L]<br />

− [ L P ( Glu )])([ P ( Glu )] −[<br />

L P ( Glu )])<br />

as −<br />

c act<br />

act<br />

c act<br />


and after transformation may be expressed as a function <strong>of</strong> pH:<br />

log( K<br />

as<br />

pH−ind<br />

as<br />

pH−pK<br />

L<br />

P<br />

) = log( K ) − log( 1+<br />

10 ) − log( 1+<br />

10 ) , Eq. 3-44<br />

38<br />

pK −pH<br />

where log(Kas pH-ind ), pKL and pKP are fitted parameters. The optimal pH for binding is:<br />

pH opt L P<br />

= ( pK + pK ) / 2 . Eq. 3-45<br />

From the pH-dependence <strong>of</strong> log(Kas), the observed free energy change for the <strong>eIF4E</strong>-<br />

cap association at a given pH can be regarded as:<br />

pH−ind<br />

ΔG<br />

° ( pH)<br />

= ΔG°<br />

+ ΔG°<br />

+ ΔG°<br />

L<br />

P<br />

, Eq. 3-46<br />

where ΔG°L and ΔG°P denote energies <strong>of</strong> bringing <strong>of</strong> the interacting ligand and the protein<br />

residue at a given pH to the appropriate ionic states, for which the binding energy would<br />

equal ΔG° pH-ind :<br />

pH−ind pH−ind<br />

as<br />

Δ G°<br />

= −RT<br />

ln( K ) ; Eq. 3-47<br />

pH−pK<br />

L<br />

Δ G°<br />

= RT ln( 1+<br />

10 ) ; Eq. 3-48<br />

L<br />

pK<br />

pH<br />

Δ G°<br />

= RT ln( 1+<br />

10 ) . Eq. 3-49<br />

P<br />

P −<br />

3.5.5. <strong>Thermodynamic</strong>s<br />

The temperature dependence <strong>of</strong> Kas was analysed according to the van't H<strong>of</strong>f isobaric<br />

equation (Eq. 1-14). 194 The molar heat capacity change ( Δ C ) and the characteristic<br />

temperatures at which<br />

the fitting.<br />

195 196 o<br />

Δ G ,<br />

o<br />

o<br />

Δ S = 0 (TS) or H vH<br />

Eq. 1-9, 1-12, and 1-11, respectvely.<br />

where<br />

o<br />

p<br />

Δ = 0 (TH) were obtained as free parameters <strong>of</strong><br />

o<br />

o<br />

Δ S , and the van't H<strong>of</strong>f enthalpy change H vH<br />

Discrimination between the linear van't H<strong>of</strong>f equation:<br />

ln K<br />

as<br />

o<br />

Δ H vH ,<br />

Snedecor's F-test 197 .<br />

o<br />

o vH<br />

Δ were calculated from<br />

ΔS<br />

ΔH<br />

= − , Eq. 3-50<br />

R RT<br />

o<br />

Δ S = const, and the non-linear model (eq. 4) was based on the statistical<br />

3.5.5.1. Coupling Between Binding and Conformational Transition <strong>of</strong> Ligand<br />

<strong>Thermodynamic</strong> parameters describing intramolecular base stacking <strong>of</strong> dinucleotide<br />

cap-analogues in the cationic form, i.e. entropy ( ΔS°<br />

1 ) and enthalpy ( 1 ΔH°<br />

) changes 198 ,<br />

were used for calculating the following quantities at four temperatures: stacking/unstacking


equilibrium constants ( 1 K ), contributions to enthalpy ( Δ H ), entropy ( Δ S ), and heat<br />

o<br />

psst<br />

capacity ( Δ C ) changes, which result from an induced shift in the self-stacking<br />

equilibrium, and intrinsic enthalpy ( Δ H ) and entropy ( Δ S ) changes <strong>of</strong> the 7-<br />

39<br />

o<br />

0<br />

methylGpppG − <strong>eIF4E</strong> association, according to the equations for the case <strong>of</strong> mandatory<br />

coupling <strong>of</strong> cap-<strong>eIF4E</strong> binding to cap stacking. 122 It is assumed that the entropy changes<br />

are approximately additive.<br />

1<br />

⎛ o<br />

⎜ 1ΔS<br />

1ΔH<br />

K = exp −<br />

⎜<br />

⎝ R RT<br />

o<br />

o<br />

⎞<br />

⎟<br />

⎠<br />

o<br />

sst<br />

o<br />

0<br />

o<br />

sst<br />

Eq. 3-51<br />

o 1ΔH<br />

⋅1<br />

K<br />

ΔHsst = −<br />

Eq. 3-52<br />

1+<br />

K<br />

1<br />

o<br />

o<br />

1ΔH<br />

⋅1<br />

K<br />

ΔSsst = −R<br />

ln(<br />

1+<br />

1K<br />

) −<br />

Eq. 3-53<br />

T<br />

( 1+<br />

K)<br />

1<br />

( ) 2<br />

o 2<br />

o ( 1ΔH<br />

) ⋅1<br />

K<br />

ΔCp = −<br />

Eq. 3-54<br />

sst 2<br />

RT 1+<br />

1K<br />

o<br />

0<br />

o<br />

cal ion<br />

o<br />

sst<br />

ΔH = ΔH<br />

− − ΔH<br />

Eq. 3-55<br />

o o o<br />

ΔS0 = ΔS<br />

− ΔSsst<br />

Eq. 3-56<br />

3.5.5.2. Enthalpy-Entropy Compensation in Congener Series – General Model 199<br />

For a system in which the number <strong>of</strong> states <strong>with</strong> an energy from U to U+δU is<br />

ω(U)δU, the mean energy (which is approximately the enthalpy <strong>of</strong> the system) is:<br />

∫ U ⋅ ϖ(<br />

U)<br />

⋅ e<br />

E = U =<br />

∞<br />

∫ ϖ(<br />

U)<br />

⋅ e<br />

−∞<br />

−U<br />

kT<br />

−U<br />

kT<br />

dU<br />

dU<br />

Eq. 3-57<br />

If some energy levels <strong>with</strong>in energy range from U' to U'+δU are significantly<br />

perturbed by ΔU > 3 kT, then the change in the energy due to the perturbance is:<br />

where<br />

ΔE ≈ (E − U')⋅P(U')dU, Eq. 3-58<br />

P(<br />

U'<br />

)<br />

−U'<br />

kT<br />

= ϖ(<br />

U'<br />

) ⋅ e is the probability <strong>of</strong> finding the system in a state <strong>with</strong> energy U'<br />

in the unperturbed system. Large and complex protein systems have many closely spaced<br />

energy levels, i. e. P(U')dU


ΔG ≈ kT⋅P(U')dU. Eq. 3-59<br />

The entropy change is then:<br />

TΔS ≈ (E − U' − kT)⋅P(U')dU. Eq. 3-60<br />

Enthalpy-entropy compensation can be thus described by a compensation temperature (Tc)<br />

that is related to the difference between the mean energy and the energy <strong>of</strong> the perturbed<br />

states:<br />

T c<br />

ΔH°<br />

ΔE°<br />

T<br />

= ≈ ≈<br />

Eq. 3-61<br />

ΔS°<br />

ΔS°<br />

RT<br />

1 −<br />

E°<br />

− U°<br />

'<br />

3.6. Isothermal Titration Calorimetry<br />

Microcalorimetry provides a direct route to thermodynamic characterization <strong>of</strong><br />

bimolecular equilibrium interactions. 200 When heat is generated or absorbed <strong>with</strong>in the<br />

sample cell a compensation appears in the feedback power to equilibrate the temperatures<br />

in the sample cell and in the reference cell. The time integral <strong>of</strong> the power deflection is a<br />

measure <strong>of</strong> the heat. The molar calorimetric enthalpy (ΔH°cal) and the association constant<br />

could be determined in one experiment by direct measurement <strong>of</strong> the heat <strong>of</strong> interaction<br />

when one component is titrated into the other. However, two conditions must be<br />

simultaneously satisfied so that determination <strong>of</strong> Kas is reliable and accurate: 201<br />

1. A product <strong>of</strong> the number <strong>of</strong> the interacting molecules ([Pact] expressed in M) and the<br />

association constant (Kas expressed in M -1 ) should be in the order <strong>of</strong> 100.<br />

2. Each injection should have at least 40 μJ <strong>of</strong> heat exchanged <strong>with</strong> the cell <strong>of</strong> the<br />

microcalorimeter.<br />

It was entirely impossible to meet these conditions for <strong>eIF4E</strong> and those cap analogues,<br />

which were described by the non-linear van't H<strong>of</strong>f plots. Numerical simulations <strong>of</strong> such<br />

experiments for m3 2,2,7 GTP <strong>with</strong> added Gaussian noise at the lowest possible level showed<br />

that fitting <strong>of</strong> the binding curve would be hopeless. Thus, a modified "single injection"<br />

method was proposed to obtain the most accurate values <strong>of</strong> the molar calorimetric<br />

enthalpies at several temperatures.<br />

40


3.6.1. Calorimetric Measurements for m 7 GpppG<br />

ITC experiments were run on OMEGA Ultrasensitive Titration Calorimeter<br />

(MicroCal, MA, U.S.A.), calibrated by 18-crown-6 titration <strong>with</strong> BaCl2. * The jacket <strong>of</strong> the<br />

microcalorimeter was filled <strong>with</strong> dry nitrogen to prevent condensation on the outer surface<br />

<strong>of</strong> the cells below room temperature. A refrigerated circulated bath (7 °C) was connected<br />

to the microcalorimeter to keep the surroundings <strong>of</strong> the cells cooler than the temperature <strong>of</strong><br />

a given experiment. The reference cell was filled <strong>with</strong> deionized water. Slow stirring at 240<br />

rpm was applied to avoid protein precipitation. The system was allowed to equilibrate until<br />

a stable baseline was observed (≥ 30 min), before an automated titration was initiated.<br />

Suitable buffers (50 mM Hepes/KOH, 100 mM KCl, 1 mM EDTA) were prepared to<br />

keep pH 7.20 ± 0.02 at 288.1, 293.1, 298.4, and 303.2 K. The protein sample buffer was<br />

exchanged by 4-fold centrifugation on 5 kDa Centricon filters (Millipore, MA, USA).<br />

After last centrifugation, the flow-through buffer was collected to dissolve 7-methylGpppG<br />

and to make control measurements <strong>of</strong> heat <strong>of</strong> the ligand dilution in the buffer. The<br />

concentration <strong>of</strong> the injected ligand was 1.00 ± 0.07 mM in each case. Samples were<br />

degassed, then filtered through 0.22 μm filter (Millipore, USA) directly before using. The<br />

main part from each protein solution was used for ITC and the remaining part for the<br />

control fluorescence titration. This ensured consistency for all the measurements. The<br />

active protein concentrations at four temperatures were 8.97 μM, 7.42 μM, 3.61 μM and<br />

5.35 μM, respectively.<br />

3.6.1.1. Modified "Single Injection" Experiment<br />

Low solubility <strong>of</strong> <strong>eIF4E</strong> (28-217) hampered also direct determination <strong>of</strong> Kas for 7-<br />

methylGpppG by the ITC measurements. The very first injections into the <strong>eIF4E</strong> solution<br />

results in ~ 35 μJ <strong>of</strong> the evolved heat, so at the verge <strong>of</strong> the instrument sensitivity, ~ 40<br />

μJ. 201 The subsequent heat signals decrease <strong>with</strong> the course <strong>of</strong> the titration due to the<br />

negative value <strong>of</strong> ΔH°, thus becoming indiscernible from the noise. The titrations were<br />

performed several times for different injection volumes (data not shown) but<br />

unsuccessfully. However, it was possible to determine the calorimetric enthalpies by a<br />

modified "single injection" method. The ligand solution was injected into the calorimetric<br />

cell (1386 μl volume) filled <strong>with</strong> <strong>eIF4E</strong> solution. Next, the 7-methylGpppG solution was<br />

* Calibration was done by Dr. Ma³gorzata Wszelaka-Rylik (Inst. Phys. Chem. PAS, Warszawa)<br />

41


injected into the buffer to measure the heat <strong>of</strong> dilution. Each experiment consisted <strong>of</strong> the<br />

main 40 μl injection, preceded by two 1 μl injections to calculate the correction for the<br />

initial outflow from the syringe, and followed by two 4 μl injections to check the protein<br />

saturation <strong>with</strong> the ligand. It was therefore possible to determine the total emitted heat in<br />

the most reliable way.<br />

3.6.2. Calorimetric Data Treatment<br />

After integration <strong>of</strong> all signals, the corresponding values from the protein titration<br />

<strong>with</strong> the ligand, and from the buffer titration <strong>with</strong> the ligand were subtracted from each<br />

other to yield the total calorimetric enthalpy<br />

o<br />

Δ H . As the heat which is exchanged at the<br />

42<br />

μl injections,<br />

which was apparently decreased by the common instrumental artifact (leakage from the<br />

syringe) during the baseline equilibration.<br />

Calorimetric<br />

<strong>of</strong> the enthalpies:<br />

o<br />

p<br />

o cal<br />

3.6.2.1. Buffer Ionization Heats<br />

o<br />

ion<br />

o o<br />

o o<br />

Δ Hion<br />

, Hcal ion = ΔHcal<br />

− ΔHion<br />

Δ − (see below).<br />

o Buffer ionization heats ( Δ Hion<br />

) <strong>of</strong> Hepes at pH =7.2 were calculated as:<br />

−pKa<br />

10<br />

−pKa<br />

−pH<br />

o<br />

H−diss<br />

ΔH =<br />

⋅ ΔH<br />

Eq. 3-62<br />

10 + 10<br />

o from the molar ionization heat for Hepes, H H−diss<br />

o temperature-dependence <strong>of</strong> H H−diss<br />

Δ = +20.95 kJ⋅mol -1 at 298.2 K and<br />

o<br />

Δ estimated as δ Δ H H−diss<br />

/δT = +0.0648 kJ⋅mol -1 K -1 , 202<br />

and pKa = 7.35 ± 0.05 for 7-methylGpppG. 177 The pKa changes were negligible over the<br />

temperature range used.<br />

3.7. NMR Spectroscopy<br />

Nuclear magnetic resonance (NMR) spectroscopy 203 is concerned <strong>with</strong> the magnetic<br />

energy <strong>of</strong> nuclei in a strong, static, magnetic field <strong>of</strong> strength Bo, (2 T to 21 T), and the<br />

transitions in the wave region <strong>of</strong> 2 - 900 MHz. Significance <strong>of</strong> the NMR spectroscopy in<br />

dynamic studies <strong>of</strong> molecules is related to interaction <strong>of</strong> a nucleus <strong>with</strong> surrounding


electrons (chemical shift) and other nuclei via the electrons (scalar coupling). The former<br />

leads to shifts <strong>of</strong> the resonance frequencies <strong>of</strong> the nuclei located in different chemical<br />

environment since the magnetic field is shielded by the electrons to B = Bo(1-σ), where σ<br />

is the shielding constant. Scalar coupling results in splitting <strong>of</strong> the resonance lines<br />

according to values <strong>of</strong> the coupling constants for the nuclear pairs.<br />

Formation <strong>of</strong> molecular complexes can be followed by changes <strong>of</strong> the chemical shifts<br />

<strong>of</strong> the nuclei directly engaged in the interactions. The total chemical shift can be<br />

represented as a sum <strong>of</strong> local effects (diamagnetic, d, and paramagnetic, p, contributions)<br />

and long-range effects, direct anisotropy (a) <strong>of</strong> the surrounding chemical bonds, ring<br />

currents (r), electric field (e) and solvent (s):<br />

δ = δd(local) + δp(local) + δa + δr + δe + δs<br />

43<br />

Eq. 3-63<br />

Substantial deshielding effect observed for protons upon formation a hydrogen bond<br />

(high value <strong>of</strong> δ) arises from an interplay <strong>of</strong> the increased electron density in the vicinity <strong>of</strong><br />

the proton (increase <strong>of</strong> δd(local)) and predominating effects <strong>of</strong> the additional restraint<br />

placed on field-induced electron circulations by the acceptor (increase <strong>of</strong> δp(local)) and its<br />

electric field effect (δe). Stacking <strong>of</strong> π-electron rings is usually accompanied by an upfield<br />

shift <strong>of</strong> the protons attached to the rings (δr), since the additional magnetic field from the<br />

circulating π-electrons reinforces Bo in the planes <strong>of</strong> the rings.<br />

To detect the NMR signal <strong>of</strong> the exchangeable protons in aqueous solution, e.g. H(8)<br />

<strong>of</strong> m 7 G, H2O instead <strong>of</strong> 2 H2O must be applied. The strong water signal in the 1 H NMR<br />

measurements needs to be suppressed.<br />

3.7.1. NMR Spectra Recording<br />

1 H NMR spectra <strong>of</strong> tryptophan N-acetylamid, a dodecapeptide DGIEPMWEDEKN,<br />

and cap analogues were recorded on a VarianUNITYplus 500 MHz spectrometer, in 1/15<br />

M phosphate buffer, pH 5.6 or 5.2, <strong>with</strong> sodium 3-trimethylsilyl-[2,2,3,3- 2 H4]propionate<br />

(TSP) as internal standard, and 10% 2 H2O for spin locking. The assignment <strong>of</strong> protons in<br />

the cap analogues and tryptophan were made on the basis <strong>of</strong> the splitting patterns <strong>of</strong> the<br />

proton resonances and their chemical shifts values (accuracy <strong>of</strong> ± 0.001 ppm). The<br />

unambiguous assignment for the H(4)/H(7) and H(5)/H(6) proton pairs in the tryptophan<br />

indole ring were done earlier from analysis <strong>of</strong> the cross-peaks in 2D ROESY spectrum by<br />

Dr. Ryszard Stolarski. 204 An effective method (~1000-fold suppression) <strong>of</strong> a binominal


1331 experiment was used for water suppression. 205 It tailors the excitation pr<strong>of</strong>iles so that<br />

there is zero net excitation <strong>of</strong> water. The hard pulse sequence was p1-τ-p2-τ-p3-τ-p4 <strong>with</strong><br />

the length ratio 1 : 3 : 3 : 1. The sum <strong>of</strong> the pulse lengths was 50 μs giving rotation <strong>of</strong> the<br />

proton magnetization by π/2 at B1 <strong>of</strong> 35 dB (attenuation <strong>of</strong> the maximal B1 strength). The<br />

<strong>of</strong>fset <strong>of</strong> the maximal excitation in respect to the carrier frequency put on the water signal,<br />

determined by the period τ, was 1500 Hz for maximal excitation in the aromatic region.<br />

3.7.2. NMR Data <strong>Analysis</strong><br />

3.7.2.1. Concentration Dependence<br />

Observed differences <strong>of</strong> 1 H chemical shifts (Δδ = δ(mixture) - δ(free)) <strong>of</strong> tryptophan<br />

N-acetylamid or <strong>of</strong> the dodecapeptide tryptophan (concentration ctrp) due to stacking upon<br />

titration <strong>with</strong> a cap analogue at 298 K depend on the total concentration <strong>of</strong> the cap<br />

analogue (ccap) according to equation:<br />

Δδ<br />

= Δδ<br />

u<br />

c<br />

⋅<br />

c<br />

u<br />

trp<br />

+ Δδ<br />

st<br />

⎛<br />

⎜<br />

c<br />

⋅ 1−<br />

⎜<br />

⎝<br />

c<br />

u<br />

trp<br />

⎞<br />

⎟ , Eq. 3-64<br />

⎟<br />

⎠<br />

where Δδu is the difference <strong>of</strong> the chemical shift <strong>of</strong> the unstacked tryptophan proton (fixed<br />

as 0), Δδst is the difference <strong>of</strong> the chemical shift <strong>of</strong> the stacked tryptophan proton (fitted as<br />

a free parameter), and equilibrium concentration <strong>of</strong> the unbound tryptophan proton (cu) is<br />

determined by a fitted association constant (K):<br />

c<br />

u<br />

2<br />

2<br />

K ⋅(<br />

ctrp<br />

− ccap<br />

) −1+<br />

K ⋅(<br />

ctrp<br />

− ccap<br />

) + 2⋅<br />

K(<br />

ctrp<br />

+ ccap)<br />

+ 1<br />

= Eq. 3-65<br />

2⋅<br />

K<br />

3.7.2.2. Temperature Dependence<br />

Temperature dependence <strong>of</strong> observed differences <strong>of</strong> 1 H chemical shifts <strong>of</strong> tryptophan<br />

N-acetylamid at 1 mM in the presence <strong>of</strong> 29.8 mM 7-methylGMP, and <strong>of</strong> the<br />

dodecapeptide tryptophan at 2.8 mM in the presence <strong>of</strong> 17.2 mM 7-methylGpppG is<br />

described by equation:<br />

Δδ<br />

= Δδ<br />

st<br />

⎛ ⎞<br />

⎜<br />

cu<br />

⋅ 1 − ⎟ , Eq. 3-66<br />

⎜ ⎟<br />

⎝<br />

ctrp<br />

⎠<br />

where cu is defined above, but the association constant is temperature dependent according<br />

to the van't H<strong>of</strong>f isobaric equation:<br />

44


⎛ ΔS°<br />

ΔH°<br />

vH ⎞<br />

K = exp⎜<br />

− ⎟ , Eq. 3-67<br />

⎝ R R ⋅T<br />

⎠<br />

R – the gas constant, T – absolute temperature. The standard molar entropy change (ΔS°)<br />

and enthalpy change (ΔH°vH) are fitted, constant parameters or treated as functions <strong>of</strong><br />

temperature. Then the fitted parameter is the heat capacity change ( Δ C ) that is involved<br />

through Eq. 1-11 and 1-12.<br />

3.8. Dynamic Light Scattering<br />

Light is scattered by interactions <strong>of</strong> electrons <strong>of</strong> molecules dissolved in a solution<br />

<strong>with</strong> the incident radiation. The scattering is quasi-elastic. The oscillating electric field<br />

causes a vibration <strong>of</strong> electrons turning them into oscillating dipoles which reemit radiation.<br />

The power spectrum is broadened in the frequency domain due to the Doppler effect on the<br />

Brownian motion <strong>of</strong> scattering macromolecules and hence due to their diffusion coefficient<br />

(D). D is related to the size <strong>of</strong> the molecules, for instance to the hydrodynamic radius (Rh)<br />

for approximately globular proteins. Fluctuations in the intensity <strong>of</strong> light scattered by a<br />

small volume <strong>of</strong> a solution in the microsecond time-scale are described by an<br />

autocorrelation function (g1(t)) which is registered by the instrument. For small<br />

monodispersed particles and homogeneous spheres the normalized autocorrelation function<br />

<strong>of</strong> scattered electric field is: 206<br />

g<br />

1<br />

( t)<br />

v<br />

−DQ<br />

2<br />

t<br />

= e , Eq. 3-68<br />

where Q v is the scattering vector resulting from the difference <strong>of</strong> the scattered and the<br />

incident wave vectors, t is the time interval for displacement <strong>of</strong> the scattering particle. D is<br />

related to the reciprocal <strong>of</strong> the characteristic decay time (Γ):<br />

2<br />

DQ v<br />

Γ =<br />

Eq. 3-69<br />

For a continuous polydispersed systems <strong>with</strong> the Γ distribution function (G(Γ)):<br />

∞<br />

−Γt g1<br />

( t)<br />

= a ∫ G(<br />

Γ)<br />

e dΓ<br />

, where a = const. Eq. 3-70<br />

0<br />

An inverse Laplace transform generates the distribution <strong>of</strong> decay rates (G(Γ)), from which<br />

the diffusion coefficients distribution and next the particle sizes distribution can be<br />

determined using the Stokes-Einstein equation:<br />

45<br />

o<br />

p


k BT<br />

D = Eq. 3-71<br />

6πηR<br />

h<br />

Molecular weight calculations are performed by means <strong>of</strong> a logarithmic function<br />

based upon experimental data on several well-characterized proteins, provided by the<br />

manufacturer <strong>of</strong> the DLS equipment.<br />

The DLS measurements were run on a DynaPro-801 Molecular Size Detector<br />

(<strong>Protein</strong> Solutions Inc., Charlottesville, VA) for <strong>eIF4E</strong> (residues 33-217) at a concentration<br />

<strong>of</strong> 1 mg/ml, in the absence and in the presence <strong>of</strong> 50-fold excess <strong>of</strong> m 7 GTP, p-Cl-bz 7 GTP,<br />

m3 227 GTP, m 7 GpppG, m 7 GppppG, or m 7 Gppppm 7 G, at 20ºC, in the standard buffer.<br />

3.9. Statistical <strong>Analysis</strong><br />

Goodness <strong>of</strong> fits was described by the R 2 value, which was not less than 0.999 and 0.99 for<br />

binding isotherms <strong>of</strong> cap analogues and peptides, respectively:<br />

∑ ( y exp − yfit<br />

)<br />

2<br />

R = 1−<br />

, Eq. 3-72<br />

2<br />

∑(<br />

y − y)<br />

exp<br />

2<br />

where yexp are measured values, yfit are fitted values, and y is the mean <strong>of</strong> the measured<br />

values. Other R 2 values are given in the text.<br />

Statistical analysis was done on the basis <strong>of</strong> the runs test and the P value 197 which<br />

denotes the probability that distribution <strong>of</strong> the fitting residuals <strong>of</strong> the y-values is random<br />

and does not depend on the x-values.<br />

Discrimination between two models <strong>of</strong> different numbers <strong>of</strong> degrees <strong>of</strong> freedom (ν1<br />

and ν2) was based <strong>of</strong> the statistical two-parameter Snedecor's F-test. 197 The test allows for<br />

a consistent and systematic assessment whether the results obtained from more<br />

complicated model <strong>with</strong> a greater number <strong>of</strong> fitted parameters are statistically important in<br />

comparison <strong>with</strong> the simpler model. The F-ratio quantifies the relationship between the<br />

relative decrease in sum-<strong>of</strong>-squares and the relative decrease in degrees <strong>of</strong> freedom:<br />

F =<br />

2<br />

2<br />

( ∑( y − y ) − ∑(<br />

y − y ) )<br />

exp<br />

fit _1<br />

exp<br />

fit _ 2<br />

( ν1<br />

− ν2<br />

) ν2<br />

46<br />

∑(<br />

y<br />

exp<br />

− y<br />

fit _ 2<br />

)<br />

2<br />

Eq. 3-73<br />

The P(ν1, ν2) value in this case is the probability that the improvement <strong>of</strong> the fit <strong>with</strong> the<br />

greater number <strong>of</strong> fitted parameters (i. e. <strong>with</strong> smaller ν) expressed by the F-ratio could be<br />

obtained randomly. The significance level was assumed as P(ν1, ν2) < 0.1 or < 0.05.


Errors <strong>of</strong> reported values (one standard deviation) were calculated according to the<br />

appropriate propagation rules 207 on the basis <strong>of</strong> the numerical uncertainty resulting from<br />

the fitting. Thanks to avoidance <strong>of</strong> linear transformations <strong>of</strong> the fluorescence titration data,<br />

their experimental uncertainties were constant <strong>with</strong>in the whole range <strong>of</strong> x-axis and almost<br />

always negligibly small (less than the size <strong>of</strong> graphical symbols). Hence, they did not<br />

influence much the final results (Kas). Experimental errors were taken into account on<br />

integration <strong>of</strong> the calorimetric data and on analyses <strong>of</strong> thermodynamic parameters.<br />

Regressions were performed by means <strong>of</strong> linear or non-linear, least-squares method,<br />

using PRISM 3.02 from GraphPad S<strong>of</strong>tware Inc., U.S.A. or ORIGIN 6.0 from Microcal<br />

S<strong>of</strong>tware Inc., U.S.A.<br />

3.10. Additional Information − Crystallography<br />

The crystallographic structure <strong>of</strong> murine <strong>eIF4E</strong> (residues 28-217) in complex <strong>with</strong> 7-<br />

methyl-GpppG to which the thesis <strong>of</strong>ten refers was resolved by Drs. Joseph Marcotrigiano<br />

and Stephen K. Burley (The Rockefeller University, N.Y., U.S.A.) similarly as reported<br />

previously 75 and was published in a joint paper. 1 Atomic coordinates have been submitted<br />

to the protein databank, accession code: 1L8B.<br />

47


4. Results and Discussion<br />

4.1. Methodological Aspects <strong>of</strong> Studies <strong>of</strong> Non-Enzymatic<br />

<strong>Protein</strong><br />

Biophysical analysis <strong>of</strong> intermolecular interactions aimed at rational design <strong>of</strong><br />

therapeutic agents requires reliable measurements <strong>of</strong> the protein-ligand association in<br />

solution as a necessary counterpart to high resolution structural studies. For <strong>eIF4E</strong>, such<br />

consistent information was lacking. The association constants have been changing<br />

significantly <strong>with</strong> each new publication due to several neglected sources <strong>of</strong> errors, both<br />

experimental and numerical. First, the lack <strong>of</strong> quantitative control <strong>of</strong> the protein activity.<br />

All association constants (Kas) reported in the previous studies were derived assuming<br />

100% activity <strong>of</strong> the protein. 5,156,161-164 As <strong>eIF4E</strong> is known to be a highly unstable protein,<br />

the prevailing assumption that entire population <strong>of</strong> <strong>eIF4E</strong> is active at every conditions was<br />

groundless. Second, the presence <strong>of</strong> practically unremovable contamination <strong>of</strong> affinity-<br />

purified <strong>eIF4E</strong> <strong>with</strong> the cap-analogue that was used for the elution from the cap-affinity<br />

column. Such <strong>eIF4E</strong> can be up to 60% m 7 GTP-bound, even after application <strong>of</strong> ionic<br />

exchange chromatography 208 or repeated buffer exchange on a 5 kDa NMWL filter. Third,<br />

dilution <strong>of</strong> a sample in the course <strong>of</strong> the titration, inner filter effect, and contribution <strong>of</strong> a<br />

free ligand to total fluorescence, were only partially eliminated in the previous<br />

publications. Dilution was considered in most <strong>of</strong> them, 5,156,161-163 while the inner filter<br />

effect and estimated emission <strong>of</strong> a free ligand were taken into account properly in the<br />

single paper. 5 The results <strong>of</strong> studies on <strong>eIF4E</strong> fused <strong>with</strong> GST, which molecular mass is<br />

comparable <strong>with</strong> that <strong>of</strong> <strong>eIF4E</strong>, were useful only for qualitative interpretation. 164<br />

All above-mentioned experimental problems are rigorously analysed herein.<br />

Purification <strong>of</strong> <strong>eIF4E</strong> via refolding from inclusion bodies guarantees that the protein is cap-<br />

free. Since <strong>eIF4E</strong> is a non-enzymatic protein, it is impossible to determine the specific<br />

activity <strong>of</strong> each sample from independent experiments. The active protein concentration<br />

([Pact]) is thus introduced as a free parameter in the numerical analysis. Addition <strong>of</strong> the<br />

next fitting parameter requires a novel, very precise measuring method <strong>of</strong> fluorescence<br />

quenching. <strong>Interaction</strong>s <strong>of</strong> the apo-protein <strong>with</strong> cap analogues are studied by the time-<br />

synchronized titration, yielding the goodness <strong>of</strong> fit R 2 = 0.999-0.9999, what assures the<br />

48


accuracy enough to extract the lacking information on the actual protein activity from each<br />

specific titration. Only this approach, combining a rigorous gathering <strong>of</strong> experimental data<br />

<strong>with</strong> simultaneous check-up on the protein activity, provides results <strong>with</strong>out systematic<br />

deviation from the theoretical model (P 0.05), and leads to concordance among Kas and<br />

other fitted parameters.<br />

4.1.1. Solution <strong>of</strong> Spectroscopic Difficulties<br />

Absorption<br />

0.012<br />

0.009<br />

0.006<br />

0.003<br />

0.000<br />

250 280 310 340 370<br />

λ [nm]<br />

Figure 4-1. Absorption (left) and fluorescence (right, λex = 280 nm) spectra <strong>of</strong> <strong>eIF4E</strong> (solid line,<br />

0.2μM) and m 7 GTP (dotted line, 4μM) at 20°C. Typical <strong>eIF4E</strong> fluorescence quenching spectra<br />

upon increasing concentration <strong>of</strong> m 7 GTP (0.7nM−4μM). Fluorescence spectra for higher<br />

concentrations <strong>of</strong> m 7 GTP exhibit an increasing emission <strong>of</strong> free m 7 GTP present in solution.<br />

All fluorescent species in solution are explicitly included in the numerical analysis.<br />

The complete overlapping <strong>of</strong> the absorption spectra for <strong>eIF4E</strong> and the cap analogues (Fig.<br />

4-1) makes it impossible to excite the protein selectively. The emission spectra are also<br />

overlapped; therefore the contribution from the increasing concentration <strong>of</strong> the free cap<br />

analogues to the total fluorescence in the course <strong>of</strong> the titration must not be neglected. This<br />

makes the spectroscopic analysis <strong>of</strong> intermolecular interactions complicated. Subtraction <strong>of</strong><br />

the free cap fluorescence is groundless, since the equilibrium concentration <strong>of</strong> the free<br />

ligand depends on the association constant to be determined. The spectra overlapping leads<br />

to erroneous attribution <strong>of</strong> the signal plateau to the saturation state, while the apparent<br />

plateau originates from two effects that cancel each other out: the quenching <strong>of</strong> the<br />

49<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

Fluorescence a. u.


intrinsic protein fluorescence and the increasing emission <strong>of</strong> the free cap analogues. Both<br />

effects are now taken into account. As a result, the observed maximal fluorescence<br />

quenching is different for various cap analogues (Fig. 4-2), and smaller than the calculated<br />

intrinsic <strong>eIF4E</strong> fluorescence quenching (Q, see Eq. 3-27, p. 34), which is the same for<br />

different cap analogues (~ 65% for the fresh protein). The association constants determined<br />

from the titration curves <strong>with</strong> the neglected free ligand emission would be erroneous,<br />

especially for weakly binding and strongly emitting ligands. The present method <strong>of</strong> data<br />

analysis eliminates these problems.<br />

Fluorescence (a. u.)<br />

Residuals<br />

100<br />

80<br />

60<br />

40<br />

20<br />

1<br />

-1<br />

10 -3 10 -2 10 -1 10 0 10 1 10 2<br />

cap concentration (μM) (μ<br />

Figure 4-2. Titration curves for m 7 GTP ( ), m 7 GpppG ( ), and m3 2,2,7 GTP ( ) at 20°C, and<br />

fitting residuals. An increasing fluorescence signal at higher cap concentrations originates from<br />

emission <strong>of</strong> the free-cap.<br />

50


4.1.2. Activity <strong>of</strong> Non-Enzymatic <strong>Protein</strong><br />

Quantitative control <strong>of</strong> the protein activity belongs to the fundamental state-<strong>of</strong>-the-art<br />

biochemical methods usually applied to enzymatic proteins, and was to date impossible to<br />

do <strong>with</strong> non-enzymatic proteins. Determination <strong>of</strong> the actual concentration <strong>of</strong> the active<br />

<strong>eIF4E</strong> factor upon each titration was the key, which made it possible to interpret the results<br />

<strong>of</strong> the binding studies in a quantitative manner.<br />

For previously frozen samples <strong>of</strong> apo-protein, the active fraction decreases down to<br />

even less than 10 % (Table 4-1, Fig. 4-3). The inactive fraction aggregates and precipitates,<br />

in spite <strong>of</strong> low protein concentration <strong>of</strong> the stored stock solution (< 0.5 mg/ml), the<br />

presence <strong>of</strong> 10 % (v/v) glycerol, 0.5 mM EDTA, and flash freezing <strong>of</strong> the aliquots (from<br />

10 μL to 50 μL). This is in contrast to the cap-saturated <strong>eIF4E</strong> that is stabilized by the<br />

ligand (see 4.2.3.1.6.1., p. 82). The experiments reported in this thesis were performed<br />

temporarily <strong>with</strong> frozen protein and then repeated <strong>with</strong> freshly prepared <strong>eIF4E</strong> (not frozen,<br />

stored at 4 °C), for which the quantity <strong>of</strong> active protein was satisfactory and varied mostly<br />

from 80 % to 113 % (± 12%) <strong>of</strong> the concentration estimated from the absorption spectra<br />

(Table 4-1). The statistics <strong>of</strong> the results proves that determination <strong>of</strong> the Kas values is<br />

reliable, and is unaffected by both the actual concentration <strong>of</strong> the active protein in the<br />

sample and the accompanying amount <strong>of</strong> the inactive protein. This indicates that the<br />

inactivation by freezing, e. g. by cold denaturation, 137 or by elevated temperatures during<br />

titrations do not partially diminish the affinity <strong>of</strong> the whole protein population, but removes<br />

a part <strong>of</strong> the population from the binding reaction, while the native protein molecules still<br />

retain their specific affinity.<br />

51


Table 4-1. Concentration <strong>of</strong> the active <strong>eIF4E</strong> protein ([Pact]) in the cuvette during fluorescence<br />

titration <strong>with</strong> m 7 GTP, equilibrium association constants (Kas), and percentage <strong>of</strong> the active protein<br />

in respect to the total concentration estimated from absorbance <strong>of</strong> the fresh or frozen stock solution<br />

([Ptotal]), at 20 °C, pH 7.2, 100 mM KCl.<br />

[Pact] (μM) Kas ⋅ 10 -6 (M -1 ) [Pact]/[Ptotal] (%)<br />

Fresh protein<br />

0.0283 ± 0.0023 115 ± 14 28<br />

0.0405 ± 0.0091 179 ± 103 110<br />

0.1026 ± 0.0053 106 ± 19 59<br />

0.1426 ± 0.0028 143 ± 18 93<br />

0.1551 ± 0.0037 72.3 ± 6.9 80<br />

0.1720 ± 0.0057 113 ± 24 105<br />

0.1798 ± 0.0014 103.2 ± 5.1 90<br />

0.1861 ± 0.0074 76 ± 15 113<br />

0.1893 ± 0.0056 108 ± 23 89<br />

0.2049 ± 0.0013 113.3 ± 6.0 103<br />

0.2258 ± 0.0067 158 ± 47 98<br />

0.2990 ± 0.0075 95 ± 23 60<br />

Frozen protein<br />

0.0005 ± 0.0053 89 ± 26 0.2<br />

0.0167 ± 0.0013 190 ± 19 4.0<br />

0.0239 ± 0.0012 83.3 ± 4.6 8.1<br />

0.0254 ± 0.0025 129 ± 18 20<br />

0.0265 ± 0.0037 70 ± 10 6.8<br />

0.0385 ± 0.0020 157 ± 23 21<br />

0.0466 ± 0.0035 111 ± 19 18<br />

0.0677 ± 0.0054 93 ± 24 53<br />

0.0720 ± 0.0022 75.9 ± 6.9 19<br />

0.0762 ± 0.0034 170 ± 44 42<br />

0.0867 ± 0.0029 118 ± 15 43<br />

0.0967 ± 0.0140 125 ± 71 76<br />

0.1896 ± 0.0180 73 ± 34 94<br />

52


Active protein (%)<br />

N<br />

100<br />

50<br />

10<br />

0<br />

0.0 0.1 0.2 0.3 0.4 0.5<br />

8<br />

6<br />

4<br />

2<br />

0<br />

Total protein concentration (μM)<br />

70 90 110 130 150 170 190<br />

K as ⋅ 10 -6 (M -1 )<br />

(a) (b)<br />

(c) (d)<br />

Figure 4-3. Statistics <strong>of</strong> titration results obtained for frozen ( ), and fresh protein () from one<br />

purification batch, used <strong>with</strong>in several subsequent days, at 20°C. (a) Active protein fraction<br />

([Pact]/[Ptotal]) vs total protein concentration during fluorescence titration. The frozen protein tends<br />

to precipitate the more the higher is the concentration <strong>of</strong> the solution, while [Pact]/[Ptotal] <strong>of</strong> the fresh<br />

samples seems to be clustered about 100% activity. (b) Values <strong>of</strong> association constant (Kas) for<br />

m 7 GTP binding to the active fraction <strong>of</strong> <strong>eIF4E</strong> are invariant irrespectively <strong>of</strong> the presence <strong>of</strong><br />

various amount <strong>of</strong> inactive protein in the frozen or fresh samples. (c) Distribution (N) <strong>of</strong> Kas values<br />

obtained from 26 repeats <strong>of</strong> titration experiments. (d) Gaussian distributions <strong>of</strong> the Gibbs free<br />

energy <strong>of</strong> binding <strong>of</strong> m 7 GTP to frozen ( , thin solid line) and fresh ( , broken line) <strong>eIF4E</strong>;<br />

summary ( , thick solid line, = -45.11 kJ/mol, SD = 0.87 kJ/mol).<br />

The influence <strong>of</strong> the concentration <strong>of</strong> active protein as fitting parameter on quality <strong>of</strong><br />

final results is illustrated in Fig. 4-4. Titration <strong>of</strong> a frozen sample <strong>with</strong> m 7 GTP in the<br />

nanomolar concentration range, i.e. where fluorescence <strong>of</strong> the ligand is negligible at the<br />

spectral bandwidths suitable for observation <strong>of</strong> the protein fluorescence changes, allows for<br />

analysis <strong>of</strong> the protein activity in the simple case <strong>with</strong> φlig-free = 0.<br />

53<br />

Kas (M -1 )<br />

N<br />

10 9<br />

10 8<br />

10 7<br />

16<br />

12<br />

8<br />

4<br />

0 20 40 60 80 100 120<br />

Active protein (%)<br />

0<br />

-47 -46 -45 -44 -43<br />

ΔG° (kJ/mol)


Fluorescence (a. u.)<br />

Residuals<br />

800<br />

600<br />

400<br />

200<br />

8<br />

0<br />

-8<br />

(a) (b)<br />

10 -3 10 -2 10 -1<br />

m 7 GTP (μM)<br />

10 -3 10 -2 10 -1<br />

Figure 4-4. Binding isotherms for interaction <strong>of</strong> m 7 GTP <strong>with</strong> frozen (a) and fresh (b) <strong>eIF4E</strong> <strong>with</strong><br />

protein concentration fixed as 100% (broken lines, as residuals) or treated as a free parameter <strong>of</strong><br />

the fitting (solid lines, and as residuals).<br />

It is clear that the fit <strong>with</strong> the fixed protein concentration determined from<br />

absorbance before freezing <strong>of</strong> the sample ([Pact] = 0.295 μM) does not satisfy the assumed<br />

model, <strong>with</strong> R 2 = 0.2957 and P = 0.00027. The apparent association constant attains an<br />

accidental value <strong>of</strong> 18 ± 130 ⋅ 10 6 M -1 , while the fit <strong>with</strong> the active protein concentration<br />

dealt as a free parameter shows an excellent goodness <strong>of</strong> fit (R 2 = 0.9996 and P = 0.27) and<br />

yields Kas = 107.1 ± 7.8 ⋅ 10 6 M -1 but ([Pact] = 0.0266 ± 0.0014 μM. This indicates that the<br />

active fraction is only ~ 8 %! Attempts for determination <strong>of</strong> Kas in the full range <strong>of</strong> the<br />

ligand concentration <strong>with</strong> the fixed [Pact] value give still worse, worthless results.<br />

Moreover, the parameter describing the actual concentration <strong>of</strong> the active protein is crucial<br />

not only for the evidently inactivated, frozen samples, but can also change the results<br />

significantly even when the fitted curve is apparently "nice" (which is <strong>of</strong>ten the case in<br />

many literature reports, especially in those using a linear scale for ligand concentrations).<br />

Fig. 4-4(b) presents a slight difference between two fits: <strong>with</strong> [Pact] fixed as 0.2 μM and<br />

<strong>with</strong> [Pact] fitted as 0.1798 ± 0.0014 μM. Although the graphical difference between solid<br />

(R 2 = 0.9999) and broken (R 2 = 0.9990) lines is almost indistinguishable at the first sight,<br />

and the protein activity is pretty high (90%), the fit assuming the fixed [Pact] value yields a<br />

function <strong>with</strong> a systematic deviation from the experimental data, which is revealed by the<br />

residuals and the P value <strong>of</strong> < 0.0001. The latter parameter indicates that it is only less than<br />

0.01% chance that this deviation is not caused by fixing <strong>of</strong> [Pact] at the level <strong>of</strong> 100%. The<br />

Fluorescence (a. u.)<br />

Residuals<br />

54<br />

100<br />

80<br />

60<br />

40<br />

10 -3 10 -2 10 -1 10 0 10 1<br />

2<br />

0<br />

-2<br />

m 7 GTP (μM)<br />

10 -3 10 -2 10 -1 10 0 10 1


Kas value obtained <strong>with</strong> free [Pact] is 103.2 ± 5.18 ⋅ 10 6 M -1 . Neglect <strong>of</strong> this small, 10%<br />

difference in the protein activity upon non-linear fitting <strong>with</strong> fixed [Pact] results in as many<br />

as 2-fold greater Kas <strong>of</strong> 188 ± 26 ⋅ 10 6 M -1 . Hence, the protein activity must be controlled<br />

upon each individual titration. This is especially important for strongly binding ligands,<br />

such as m 7 GTP, p-Cl-bz 7 GTP or m 7 Gppppm 7 G (Fig. 4-5). The temperatures at which the<br />

activity <strong>of</strong> <strong>eIF4E</strong> drops by 50% (34 – 41 °C) are in the physiological range, as estimated<br />

from the Boltzmann sigmoidal curves. These results could suggest that the unstable <strong>eIF4E</strong><br />

protein is usually complexed <strong>with</strong> another macromolecule (<strong>mRNA</strong> 5' cap or other proteins)<br />

in the living cell to avoid inactivation in the apo state, which is consistent <strong>with</strong> the<br />

observation that most <strong>of</strong> the <strong>eIF4E</strong> population is blocked by 4E-BPs, and biological<br />

potency <strong>of</strong> <strong>eIF4E</strong> is regulated by its accessibility. 42<br />

Active protein (%)<br />

Active protein (%)<br />

100<br />

50<br />

0<br />

100<br />

50<br />

0<br />

(a) 0 (b)<br />

0 5 10 15 20 25 30 35 40 45<br />

Temperature (°C)<br />

(c) (d)<br />

0 5 10 15 20 25 30 35 40 45<br />

Temperature (°C)<br />

Figure 4-5. Temperature dependence <strong>of</strong> the active fraction <strong>of</strong> fresh protein ( , , , ) obtained<br />

upon titration <strong>with</strong> ligands <strong>of</strong> different affinity for <strong>eIF4E</strong>. (a) m 7 GTP, Kas = 108.7 ⋅ 10 6 M -1 ; (b)<br />

m 7 Gppppm 7 G, Kas = 47.0 ⋅ 10 6 M -1 ; (c) p-Cl-bz 7 GTP, Kas = 44.6 ⋅ 10 6 M -1 , results for frozen protein<br />

at 20°C is shown for comparison (); (d) m 7 GpppC, Kas = 3.86 ⋅ 10 6 M -1 ; the Kas values at 20 °C.<br />

55<br />

Active protein (%)<br />

log(% <strong>of</strong> active protein)<br />

100<br />

50<br />

10<br />

2<br />

0<br />

0 5 10 15 20 25 30 35 40 45<br />

Temperature (°C)<br />

-10<br />

0 5 10 15 20 25 30 35 40 45<br />

Temperature (°C)


For the ligands that bind to the protein less strongly (Kas < 10 7 M -1 ), the<br />

concentration <strong>of</strong> active protein looses its numerical importance. Determination <strong>of</strong> [Pact]<br />

becomes inaccurate due to weakly pronounced curvature <strong>of</strong> the fitted function F([L]) in the<br />

ligand concentration range close to [Pact].<br />

The population <strong>of</strong> active protein is also ionic strength-dependent (Fig. 4-6). Relative<br />

stabilization is achieved at the ionic strength characteristic for the intracellular fluid (~200<br />

mM). Determination <strong>of</strong> the active fraction <strong>of</strong> protein as a function <strong>of</strong> [KCl] was necessary<br />

to obtain association and dissociation rate constants from the kinetic traces registered<br />

during stopped-flow experiments for interaction <strong>of</strong> <strong>eIF4E</strong> <strong>with</strong> m 7 GpppG. 4 Single-<br />

relaxation fits yielded overall kinetic constants and dual-relaxation fits provided evidence<br />

for the two-step character <strong>of</strong> the complex formation.<br />

Active protein (%)<br />

100<br />

50<br />

0<br />

0 100 200 300 400<br />

KCl (mM)<br />

Figure 4-6. Dependence <strong>of</strong> the average active fraction <strong>of</strong> fresh protein on KCl concentration after<br />

storage by ~6 hours at 25 °C. Results obtained from titrations <strong>of</strong> <strong>eIF4E</strong> <strong>with</strong> m 7 GTP to control the<br />

protein activity during kinetic stopped-flow experiments <strong>with</strong> use <strong>of</strong> m 7 GpppG. 4<br />

The knowledge about the actual concentration <strong>of</strong> active non-enzymatic protein is<br />

crucial for normalization <strong>of</strong> every quantity being determined experimentally, analogically<br />

as specific activity <strong>of</strong> enzymes is used. 209 In particular, long-lasting thermodynamic<br />

measurements performed by means <strong>of</strong> techniques other than emission spectroscopy require<br />

concurrent control fluorescence titrations so that the results are free from systematic errors<br />

caused by temperature-dependent protein inactivation.<br />

56


Fluorescence (a. u.)<br />

Figure 4-7. Determination <strong>of</strong> protein active fractions at 30 °C (303.2 K, , broken lines) and 15 °C<br />

(288.1 K, , solid lines). Titration curves determine both K as and [Pact] as free parameters <strong>of</strong> the<br />

fitting (Eq. 3-26, p. 34). [Pact] is graphically represented by the point where the curve for<br />

hypothetical infinite Kas would attain maximal quenching Qmax (Eq. 3-27, p. 34).<br />

Changes in the amount <strong>of</strong> active <strong>eIF4E</strong> incubated at 288 K and 303 K during a time<br />

corresponding to duration <strong>of</strong> isothermal calorimetric titration are shown in Fig. 4-7. Total<br />

<strong>eIF4E</strong> concentration was 0.2 μM at both temperatures. Beside obvious differences between<br />

the association constants, the control fluorescence experiments have shown that <strong>eIF4E</strong><br />

retained 89.7 % activity at 288 K but only 53.5 % at 303 K. Further determination <strong>of</strong> the<br />

standard molar binding enthalpies and heat capacity changes had to be based on this<br />

information.<br />

100<br />

80<br />

60<br />

40<br />

303K<br />

Qmax Q max 288K<br />

10 -3 10 -2 10-1 10 0 101 [Pact] 20<br />

303K [Pact] 288K<br />

m 7 GTP (μM)<br />

The methodology <strong>of</strong> determination <strong>of</strong> the active protein fraction was initially<br />

designed to resolve the problem <strong>of</strong> activity control <strong>of</strong> the non-enzymatic <strong>eIF4E</strong> translation<br />

factor. 4,6 However, it was further proposed to study stoichiometry <strong>of</strong> ligand binding to an<br />

enzymatic oligoprotein and proved successful in finding that a hexameric enzyme (PNP)<br />

binds only three substrate molecules per one hexamer. 210,211<br />

57


4.1.3. Incorrectness <strong>of</strong> Data Linearization in Case <strong>of</strong> Strong <strong>Interaction</strong>s<br />

Independently from skipping the fundamental problem <strong>of</strong> the protein activity, the<br />

previous reports were based mostly on linearized forms <strong>of</strong> the equilibrium equation which<br />

are not suitable for the analysis <strong>of</strong> strong interactions. The commonly used Eadie-H<strong>of</strong>stee<br />

linear transformation requires an assumption that protein concentration is negligible in<br />

comparison <strong>with</strong> ligand concentration over the entire course <strong>of</strong> the titration. 156,161-163 This is<br />

not satisfied for molecular systems <strong>of</strong> a higher affinity. When Kas ~ 10 μM -1 , saturation <strong>of</strong><br />

~ 40% occurs already for [L] [Pact], at typical [Pact] ~ 0.1 μM. This makes the<br />

transformed data non-linear, even after corrections for dilution, inner filter effect, and free<br />

ligand emission. Additionally, the experimental uncertainty is hidden in the x-axis (Fig. 4-<br />

8). Inverse representation <strong>of</strong> the data reveals the huge errors which are always neglected<br />

during the Eadie-H<strong>of</strong>stee analysis.<br />

ΔΔF (a. u.)<br />

400<br />

300<br />

200<br />

100<br />

0<br />

2500 5000 7500 10000<br />

ΔF/[L] (a. u./μM)<br />

Figure 4-8. (a) Eadie-H<strong>of</strong>stee transformation <strong>of</strong> experimental data from the same fluorescence<br />

titration as in Fig. 4-4(a), apparent Kas = 39 ± 12 ⋅ 10 6 M -1 and (b) inverse Eadie-H<strong>of</strong>stee<br />

transformation <strong>of</strong> these data, apparent Kas = 17.2 ± 5.0 ⋅ 10 6 M -1 .<br />

A modified linear Eadie-H<strong>of</strong>stee representation 5 requires the active protein<br />

concentration and the maximal quenching (ΔFmax) as known constants (Fig. 4-9). Even<br />

though they are guessed properly, the most important data points, which should determine<br />

the binding constant, are compressed by the transformation into a very narrow numerical<br />

range (the stronger binding the more narrow the range). This leads to loss <strong>of</strong> their<br />

numerical importance, and consequently, to significantly biased Kas values.<br />

ΔF/[L] (a. u./ μM)<br />

58<br />

15000<br />

10000<br />

5000<br />

0<br />

-5000<br />

-10000<br />

0 100 200 300<br />

ΔF (a. u.)


[m 7 GTP]/ΔF (μM/a. u.)<br />

0.0006<br />

0.0004<br />

0.0002<br />

0.0000<br />

0.00 0.01 0.02 0.03<br />

1/(ΔF max-ΔF) (1/a. u.)<br />

Figure 4-9. Modified linear Eadie-H<strong>of</strong>stee representation 5 <strong>of</strong> the data from Fig. 4-4(a) and Fig. 4-8.<br />

Although the required constant parameters ([Pact] and ΔFmax) <strong>of</strong> the linear data transformation are<br />

taken from the previous fit <strong>with</strong> the free active protein concentration, shown in Fig. 4-4(a), the<br />

apparent Kas value is significantly decreased, 72.3 ± 1.2 ⋅ 10 6 M -1 , due to contraction <strong>of</strong> the most<br />

important data at the low concentrations <strong>of</strong> the ligand to a very narrow range <strong>of</strong> the x variable.<br />

The non-linear analysis is free from the systematic errors caused by the linear<br />

transformations. The methodological improvements afford possibilities for a reliable<br />

comparison <strong>of</strong> the results for the protein from different purification batches, previously<br />

frozen or not, from experiments at different temperatures, ionic strengths, and pH. The Kas<br />

values up to 10 8 M -1 , accompanied by the systematic and self-consistent structure-affinity<br />

relationship, provide an exact, quantitative test for the proper fold <strong>of</strong> the renatured protein.<br />

This assures for the first time that the equilibrium association constants reflect the true,<br />

intrinsic affinity <strong>of</strong> <strong>eIF4E</strong> for <strong>mRNA</strong> 5' cap.<br />

59


4.2. Molecular Mechanism <strong>of</strong> Recognition <strong>of</strong> <strong>mRNA</strong> 5' Cap<br />

Structure by <strong>eIF4E</strong> Cap-Binding <strong>Protein</strong><br />

4.2.1. Affinity <strong>of</strong> Cap Analogues for <strong>eIF4E</strong><br />

4.2.1.1. Equilibrium Binding Constants<br />

Kas values determined for a wide class <strong>of</strong> cap analogues (Table 4-2) vary from 10 8 M -<br />

1 for specific 7-substituted cap analogues to 10 2 M -1 for 7-unsubstituted ligands (Scheme 2-<br />

1). The human and murine proteins have the same binding affinities for <strong>eIF4E</strong>. The murine<br />

protein was used for the further studies because <strong>of</strong> technical matters, i. e. more efficient<br />

expression and purification. However, thanks to almost 100% identity, the conclusions<br />

drawn from the data are common for both proteins.<br />

The substantial differences <strong>of</strong> the Kas values follow structural modifications <strong>of</strong> cap,<br />

m 7 , 2 , 2<br />

e.g. a 760-fold reduction <strong>of</strong> Kas for 3 GTP and up to a 5000-fold drop <strong>of</strong> Kas for GTP<br />

as compared to m 7 GTP. The presence <strong>of</strong> any N(7)-substituent enhances binding, but the<br />

methyl substituent appears to be optimal. Among several elements important for specific<br />

<strong>eIF4E</strong>-cap binding, the negative electrostatic charge <strong>of</strong> the phosphate chain, which depends<br />

both on the number <strong>of</strong> phosphate groups and the presence <strong>of</strong> a second nucleotide, is <strong>of</strong><br />

primary importance. The decrease <strong>of</strong> Kas <strong>with</strong> single step-wise reduction <strong>of</strong> the phosphate<br />

Fluorescence (a. u.)<br />

100<br />

80<br />

60<br />

40<br />

GMP<br />

GDP<br />

GTP<br />

m 7 GMP<br />

m 7 GDP<br />

m 7 GTP<br />

10 -3 10 -2 10 -1 10 0 10 1 10 2<br />

20<br />

Concentration <strong>of</strong> cap analogue (μM)<br />

Figure 4-10. Comparison <strong>of</strong> binding isotherms for <strong>eIF4E</strong> complexes <strong>with</strong> unmethylated and<br />

methylated at N(7) guanosine mono-, di-, and triphosphates, at 20°C, pH 7.2, 100 mM KCl.<br />

60


Table 4-2. Binding free energies (ΔG°) and equilibrium association constants (Kas) for complexes <strong>of</strong><br />

murine <strong>eIF4E</strong> (28-217) and human <strong>eIF4E</strong> (1-217) <strong>with</strong> various cap analogues, at 20°C, 50 mM<br />

Hepes/KOH pH 7.2, 100 mM KCl, 1 mM DTT, 0.5 mM EDTA.<br />

Cap-analogue Murine <strong>eIF4E</strong> (28-217) Human <strong>eIF4E</strong> (1-217)<br />

ΔG° (kJ/mol) Kas ⋅ 10 -6 (M -1 ) Kas ⋅ 10 -6 (M -1 )<br />

m 7 GTP -45.099 ± 0.088 108.7 ± 4.0 97 ± 30 1.17 ± 0.12 a<br />

m 7 GDP -41.02 ± 0.18 20.4 ± 1.5<br />

m 7 GMP -33.15 ± 0.20 0.806 ± 0.067<br />

m 7 G -20.54 ± 0.63 0.0046 ± 0.0012<br />

et 7 GTP -39.38 ± 0.20 10.41 ± 0.86<br />

bz 7 GTP -40.660 ± 0.059 17.59 ± 0.43<br />

p-Cl-bz 7 GTP -42.93 ± 0.21 44.6 ± 3.8<br />

m 7 , 2<br />

2 GTP<br />

-41.89 ± 0.20 29.2 ± 2.4<br />

m 7 , 2 , 2<br />

3 GTP -28.9 ± 1.4 0.143 ± 0.080<br />

m 7 Gpppm 2’O G -38.75 ± 0.16 8.04 ± 0.51<br />

m 7 GpppG -38.55 ± 0.15 7.39 ± 0.46 9.6 ± 0.8 0.69 ± 0.07 a<br />

m 7 GpppA -37.43 ± 0.27 4.68 ± 0.52<br />

m 7 GpppC -36.41 ± 0.40 3.08 ± 0.51 4.8 ± 0.4 0.48 ± 0.05 a<br />

m 7 Gpppm 7 (micro) b<br />

G -35.28 ± 0.36 1.93 ± 0.28<br />

m 7 Gpppm 7 (macro) b<br />

G 3.86 ± 0.56<br />

m 7 GppppG -44.96 ± 0.11 102.8 ± 4.4<br />

m 7 Gppppm 7 (micro) b<br />

G -41.37 ± 0.15 23.5 ± 1.4<br />

m 7 Gppppm 7 (macro) b<br />

G 47.0 ± 2.7<br />

GTP -24.30 ± 0.11 0.0214 ± 0.0010<br />

GDP -20.71 ± 0.25 0.0049 ± 0.0005<br />

GMP -12.93 ± 3.8 0.0002 ± 0.0003<br />

(micro) b<br />

GpppG -20.66 ± 0.21 0.0048 ± 0.0004<br />

(macro) b<br />

GpppG 0.0095 ± 0.0007<br />

a values for m 7 G-affinity purified full length human <strong>eIF4E</strong> 5<br />

b microscopic and macroscopic association constants, see text for details.<br />

chain in the series <strong>of</strong> 7-methylated mononucleotides is followed by the same changes <strong>of</strong><br />

Kas for the unmethylated compounds (~5- and ~25-fold for removal <strong>of</strong> the γ-phosphate and<br />

the β-phosphate, respectively). Removal <strong>of</strong> the α-phosphate leads to a drastic fall <strong>of</strong> Kas,<br />

which is 175-fold lower for m 7 G than that <strong>of</strong> m 7 GMP (Fig. 4-10).<br />

61


The dinucleotide triphosphate cap analogues also bind to <strong>eIF4E</strong> less strongly (~20-<br />

fold) than m 7 GTP. The identity <strong>of</strong> the second base seems to be <strong>of</strong> minor importance.<br />

Generally, either base <strong>of</strong> the dinucleotide cap analogues penetrate the binding slot <strong>of</strong><br />

<strong>eIF4E</strong>, but since the association constants <strong>of</strong> unmethylated molecules are far lower than<br />

those <strong>of</strong> the methylated ones, such dual binding is negligible for asymmetrical cap<br />

analogues. By contrast, for symmetrical molecules (m 7 Gpppm 7 G, m 7 Gppppm 7 G and<br />

GpppG) the observed Kas must be divided by 2 because <strong>of</strong> entropic effects. Only this<br />

normalized Kas represents a true microscopic association constant Kas (micro) , reflecting the<br />

intrinsic stabilization energy <strong>of</strong> the complex. The macroscopic Kas, however, corresponds<br />

to the effective biological potency <strong>of</strong> the cap analogues.<br />

In the pioneering work <strong>of</strong> Carberry et al., 156 the association constants <strong>of</strong> two<br />

fundamental cap analogues, m 7 GTP and m 7 GpppG, for <strong>eIF4E</strong> from human erythrocytes<br />

purified by cap-affinity chromatography were determined as 3.87 ⋅ 10 5 M -1 and 3.70 ⋅ 10 5<br />

M -1 , respectively (23°C, pH 7.6, <strong>with</strong>out KCl in the buffer). Similar Kas values in the range<br />

<strong>of</strong> 10 5 -10 6 M -1 were reported for recombinant human <strong>eIF4E</strong> purified by the same<br />

method. 5,161-163 In those studies also no essential difference in affinity between m 7 GTP and<br />

m 7 GpppG was observed. While the association constants reported here for the murine<br />

protein are at variance <strong>with</strong> these data, they are very close to those obtained for<br />

recombinant full length human <strong>eIF4E</strong>, which was also purified via refolding from inclusion<br />

bodies <strong>with</strong>out any prior contact <strong>with</strong> cap (Table 4-2). These Kas values are up to 500-fold<br />

higher than those reported previously.<br />

4.2.1.2. Binding Affinity vs Inhibitory Properties <strong>of</strong> Cap Analogues<br />

The affinities represented by the present Kas values (Table 4-2) reveal large<br />

differences among cap analogues possessing different functional groups and correspond<br />

well to the inhibitory properties <strong>of</strong> the cap analogues observed in a rabbit reticulocyte<br />

lysate (Fig. 4-11). In a kinetic model developed for the in vitro translational<br />

system, 157,212,213 the inhibition constant (KI) was determined as the overall dissociation<br />

constant <strong>of</strong> the cap analogue from the 48S initiation complex. Results <strong>of</strong> both approaches,<br />

biophysical and biochemical, point out the significance <strong>of</strong> the same structural features <strong>of</strong><br />

the cap responsible for specific recognition, e.g. the presence and type <strong>of</strong> N(7)-substituent<br />

<strong>of</strong> guanine, the negative charge density in the phosphate chain, and the N 2 amino group<br />

<strong>with</strong> at least one proton capable <strong>of</strong> forming a hydrogen bond. Interestingly, the observed<br />

62


values <strong>of</strong> KI are ~100 times greater than those <strong>of</strong> 1/Kas for all cap analogues. This<br />

apparently larger amount <strong>of</strong> cap analogue required to inhibit translation in vitro than that<br />

necessary to displace <strong>eIF4E</strong> from the <strong>mRNA</strong> cap most likely arises from the following<br />

reasons. As opposed to HeLa cells, 41 the <strong>eIF4E</strong> concentration in reticulocyte lysate is not<br />

limiting for translation, and the majority <strong>of</strong> <strong>eIF4E</strong> is not engaged in this process. 214 The<br />

assumption underlying the determination <strong>of</strong> KI values, that <strong>eIF4E</strong> concentration is<br />

negligible in comparison <strong>with</strong> that <strong>of</strong> a competitive inhibitor, 213 is not applicable here. The<br />

concentration <strong>of</strong> <strong>eIF4E</strong> is up to 50-fold higher than previously thought, 215 so much more<br />

cap analogue is required to inhibit translation due to the quadratic form <strong>of</strong> the equilibrium<br />

equation. Besides, protein synthesis results from a large number <strong>of</strong> catalytic reactions and<br />

depends not only on the <strong>eIF4E</strong> activity, e. g. in the in vitro experiments, the <strong>mRNA</strong><br />

binding activity <strong>of</strong> eIF4F complex is enhanced by the presence <strong>of</strong> eIF4G. 216<br />

KI [μμM]<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

10 -2 10 -1 10 0 10 1 10 2<br />

K as [μM -1 ]<br />

Figure 4-11. Correlation between inhibition constants <strong>of</strong> the cap analogues (KI), obtained from in<br />

vitro translation (30°C 157 ), and their equilibrium association constants (Kas, 20°C; macroscopic Kas<br />

were used for the symmetrical caps) by 4 orders <strong>of</strong> magnitude. Correlation coefficient r 2 = 0.795.<br />

63


4.2.2. Parsing the Free Energy <strong>of</strong> <strong>eIF4E</strong> Binding to <strong>mRNA</strong> 5' Cap<br />

A comparison <strong>of</strong> the binding free energies rather than that <strong>of</strong> the association<br />

constants reflects the biophysical basis <strong>of</strong> the ligand affinity for the protein. The data<br />

collected for twenty structurally different cap analogues permitted analysis <strong>of</strong> the influence<br />

<strong>of</strong> single structural modifications on the cap-<strong>eIF4E</strong> binding in terms <strong>of</strong> the standard Gibbs<br />

free energies ΔG° (Table 4-2), which were calculated either from the binding constants Kas,<br />

Δ G° = −RT<br />

ln K as , or from Kas (micro) ,<br />

64<br />

( micro)<br />

as<br />

Δ G° = −RT<br />

ln K , for the symmetrical cap<br />

analogues. Contribution <strong>of</strong> a given chemical alteration is represented by the corresponding<br />

change <strong>of</strong> the standard Gibbs energy <strong>of</strong> association (ΔΔG°), defined individually for each<br />

group <strong>of</strong> compounds in Table 4-3.<br />

It should be noted here that parsing is generally not applicable for every molecular<br />

system. 152,153 However, direct comparison <strong>of</strong> the ΔG° values surprisingly but distinctly<br />

showed that the total binding free energy resolved itself into individual contributions from<br />

interactions <strong>of</strong> the phosphate chain and into common contribution from all interactions <strong>of</strong><br />

the m 7 guanosine moiety. This limited additivity is a kind <strong>of</strong> empirical verification <strong>of</strong><br />

correctness <strong>of</strong> parsing employment in this particular case, and, on the other hand, points to<br />

cooperativity <strong>of</strong> interactions involving the m 7 guanosine moiety (stacking and hydrogen<br />

bonding) that may not be decomposed into their contributions.


Table 4-3. Changes in the standard Gibbs free energy (ΔΔG°) on <strong>eIF4E</strong> binding to the<br />

structurally modified cap analogues, at 20°C.<br />

Structural alteration ΔΔG° ΔΔ kJ/mol<br />

replacement <strong>of</strong> N(7)-methyl for larger substituents:<br />

ΔΔG° = ΔG°(x 7 GTP) − ΔG°(m 7 GTP)<br />

m → x : m → et m → bz m → p-Cl-bz<br />

x 7 GTP +5.73 ± 0.21 +4.44 ± 0.13 +2.18 ± 0.21<br />

methylation at N(7) for n phosphate groups:<br />

ΔΔG° = ΔG°(m 7 Gpn(G)) − ΔG°(Gpn(G))<br />

n: 1 2 3<br />

Gpn → m 7 Gpn −20.2 ± 3.8 −20.29 ± 0.29 −20.79 ± 0.13<br />

GpnG a → m 7 GpnG −17.91 ± 0.25<br />

addition or alteration <strong>of</strong> the second nucleoside:<br />

ΔΔG° = ΔG°(m 7 GpnY) − ΔG°(m 7 GpnX)<br />

X → Y: none → G G → m 7 G a<br />

G → A G → C G → m 2’O G<br />

m 7 Gp3X → m 7 Gp3Y +6.57 ± 0.17 +3.26 ± 0.38 +1.13 ± 0.29 +2.13 ± 0.42 −0.2 ± 0.2<br />

m 7 Gp4X → m 7 Gp4Y +3.60 ± 0.17<br />

Gp3X → Gp3Y a<br />

+3.64 ± 0.25<br />

successive addition <strong>of</strong> the 5'-phosphate groups:<br />

ΔΔG° = ΔG°(n+1) − ΔG°(n)<br />

n → n+1: 0 → 1 1 → 2 2 → 3 3 → 4<br />

m 7 Gpn −12.59 ± 0.67 −7.87 ± 0.25 −4.10 ± 0.21<br />

Gpn −7.8 ± 3.8 −3.60 ± 0.29<br />

m 7 GpnG −6.40 ± 0.17<br />

m 7 Gpnm 7 G a<br />

−6.11 ± 0.38<br />

successive addition <strong>of</strong> the N 2 -methyl groups:<br />

ΔΔG° = ΔG°(n+1) − ΔG°(n)<br />

n → n+1: 0 → 1 1 → 2<br />

2,<br />

2 7<br />

m n m GTP +0.77 ± 0.05 +3.10 ± 0.33<br />

a for symmetrical cap analogues ΔΔG° calculated from the microscopic association constant Kas (micro)<br />

65


4.2.2.1. Energetic Cost <strong>of</strong> 7-Substituent Alteration<br />

The presence <strong>of</strong> each type <strong>of</strong> substituent at the N(7) position <strong>of</strong> the guanine moiety<br />

produces a comparable effect on the <strong>eIF4E</strong> binding enhancement (Fig. 4-12), in<br />

comparison <strong>with</strong> the non-substituted cap analogue, by ΔΔG° <strong>of</strong> −15 to −21 kJ/mol (Tables<br />

4-2, 4-3). The main gain <strong>of</strong> the stacking energy is reached irrespective <strong>of</strong> the nature <strong>of</strong> the<br />

substituent. This indicates that the delocalized positive charge resulting from N(7)-<br />

substitution is <strong>of</strong> primary importance. The only direct interaction involving the 7-methyl<br />

group is a non-specific van der Waals contact <strong>with</strong> Trp166 in the complex <strong>with</strong> m 7 GDP<br />

(3.93 Å, Fig. 4-13a), which is lost in the complex <strong>with</strong> m 7 GpppG (4.17 Å, Fig. 4-13b).<br />

Replacement <strong>of</strong> the methyl group for larger substituent (ethyl, benzyl) can cause steric<br />

hindrance, and can interfere <strong>with</strong> creation <strong>of</strong> the water-mediated hydrogen bonds<br />

stabilizing the α-phosphate. Thus, it is energetically unfavourable by ΔΔG° <strong>of</strong> +2.2 to +5.7<br />

kJ/mol.<br />

Fluorescence (a. u.)<br />

100<br />

80<br />

60<br />

40<br />

20<br />

et 7 GTP<br />

p-Cl-bz 7 GTP<br />

m 7 GTP<br />

bz 7 GTP<br />

10 -3 10 -2 10 -1 10 0 10 1<br />

Concentration <strong>of</strong> cap analogue (μM)<br />

Figure 4-12. Binding isotherms for <strong>eIF4E</strong> complexes <strong>with</strong> N(7)-substituted guanosine<br />

triphosphates, at 20°C, pH 7.2, 100 mM KCl.<br />

66


Trp-166<br />

Trp-166<br />

Glu-103<br />

Trp-102<br />

Glu-103<br />

Trp-102<br />

Arg-112<br />

Arg-112<br />

c<br />

n2<br />

Figure 4-13. Interatomic contacts in (a) m 7 GDP-<strong>eIF4E</strong> complex, determined on the basis <strong>of</strong> the<br />

crystallographic structure (PDB ID code: 1EJ1), 75 and (b) m 7 GpppG-<strong>eIF4E</strong> complex. Hydrogen<br />

bonds, salt bridges and van der Waals interactions are indicated <strong>with</strong> dotted lines, <strong>with</strong> the lengths<br />

expressed in Å. Trapped water molecules supporting the hydrogen-bonded network are shown as<br />

magenta spheres. The fragments <strong>of</strong> <strong>eIF4E</strong> distant from the cap by 4 Å are shown as spheres.<br />

67<br />

Lys-162<br />

n2<br />

Lys-162<br />

Lys-206<br />

Trp-56<br />

c5'<br />

Trp-56<br />

c5'<br />

(a)<br />

Arg-157<br />

(b)<br />

Arg-157


4.2.2.2. General Separation <strong>of</strong> the Binding Free Energy<br />

The binding free energy <strong>of</strong> the 7-methylated analogues association <strong>with</strong> <strong>eIF4E</strong><br />

appears to be totally separated into two groups <strong>of</strong> interactions: stacking together <strong>with</strong><br />

hydrogen bonding and binding through the phosphates. Irrespective <strong>of</strong> the level <strong>of</strong> the cap<br />

analogue phosphorylation, methylation at N(7) is accompanied by almost constant binding<br />

free energy change, ΔΔG° about –20.5 kJ/mol (Table 4-3). The gain from N(7)-<br />

methylation slightly decreases for the shorter phosphate chains. This minute effect may<br />

result from the fact that acidic-basic properties <strong>of</strong> N(1)-H proton slightly depend on the<br />

number <strong>of</strong> the phosphate groups 177 (Scheme 3-1, p. 23). Therefore, the m 7 G moiety<br />

assumes the required cationic form (see 4.2.3.2., p. 83) more readily in analogues <strong>with</strong> a<br />

longer phosphate chain. However, both the pKa-shifts 177 and differences among ΔΔG°<br />

values related to methylation <strong>of</strong> mono-, di- and triphosphates (Table 4-3) are very small,<br />

indicating that the presence <strong>of</strong> the phosphates influences the charge <strong>of</strong> m 7 G to a negligible<br />

extent. The charge flow via the ribose σ-bonds does not occur.<br />

Surprisingly, the measured values <strong>of</strong> ΔG° satisfy the relationship: ΔG°(m 7 Gpn) =<br />

ΔG°(m 7 G) + ΔG°(G pn), for n = 1, 2 or 3. This means that the 7-methylguanosine binding<br />

is separate from binding <strong>of</strong> the phosphate chain. The possible entropic cost arising from a<br />

change in the degrees <strong>of</strong> freedom due to linkage <strong>of</strong> the two moieties is either negligible or<br />

can be partially cancelled out if the bound phosphates facilitate to some extent penetration<br />

<strong>of</strong> the cap-binding slot by m 7 G. The resultant entropic effect is, however, below the best<br />

experimental accuracy: δ(ΔG°(m 7 Gpn) − ΔG°(m 7 G) − ΔG°(Gpn)) = ± 0.67 kJ/mol. The<br />

total binding free energy <strong>of</strong> 7-methylguanosine, ΔG°(m 7 G) = −20.54 ± 0.63 kJ/mol (Table<br />

4-2), is the same as ΔΔG° resulting from methylation at N(7), which demonstrates that the<br />

ribose does not contribute to the complex stabilization. The 3D structures show only one<br />

non-specific van der Waals contact between the ribose and Trp56 (Fig. 4-13).<br />

The stabilization energy <strong>of</strong> the GMP-<strong>eIF4E</strong>, GDP-<strong>eIF4E</strong> and GTP-<strong>eIF4E</strong> complexes<br />

originates only from the interactions involving the phosphate groups, independently<br />

pointing to the minute role for the ribose in stabilization <strong>of</strong> the complex. It is directly seen<br />

from the comparison <strong>of</strong> the total binding energy, ΔG°(GMP) = −12.93 ± 3.8 kJ/mol (Table<br />

4-2), <strong>with</strong> the contribution <strong>of</strong> the single α-phosphate, ΔΔG° = −12.59 ± 0.67 kJ/mol (Table<br />

4-3). These findings reveal that although unmethylated guanosine phosphates are<br />

potentially capable to form three Watson-Crick like hydrogen bonds, such bonds are not<br />

68


created. It is only when the guanine moiety possesses a substituent at the N(7) position that<br />

the hydrogen bonds are formed. As shown above, the positive charge at m 7 G determines<br />

the cap affinity for <strong>eIF4E</strong> being a cause <strong>of</strong> the stacking enhancement.<br />

The most striking conclusion is that the cation-π stacking enhancement is a<br />

precondition for the hydrogen bond formation, and a double role should be attributed to the<br />

presence <strong>of</strong> the conserved tryptophans 56 and 102 in the <strong>eIF4E</strong> binding site: stabilization<br />

<strong>of</strong> the 7-methylguanine ring by stacking itself and enabling 7-methylguanine to form the<br />

hydrogen bonds <strong>with</strong> the carboxylate oxygen atoms <strong>of</strong> Glu-103 and the backbone amino<br />

group <strong>of</strong> Trp-102 (Fig. 4-13). Consequently, the system <strong>of</strong> m 7 G-<strong>eIF4E</strong> works according to<br />

the "all-or-nothing" rule.<br />

4.2.2.3. Relation <strong>of</strong> Stacking - Hydrogen Bonding Cooperativity to Other Cap-<br />

Binding Molecules<br />

Previous studies using small model peptides showed that the stacking interactions <strong>of</strong><br />

N(7)-methylated guanine <strong>with</strong> the aromatic rings <strong>of</strong> Trp, Tyr, Phe are stronger in<br />

comparison <strong>with</strong> the unmethylated base. However, the difference in the Kas values was<br />

lesser than 1.5-fold. 217 On the other hand, spectroscopic studies on the interaction between<br />

cap analogues and small model peptides, containing Trp and Glu, showed simultaneous<br />

formation <strong>of</strong> hydrogen bonds between 7-methylguanine and the Glu carboxyl side group,<br />

and stacking <strong>with</strong> the Trp indole ring. 218-220 These investigations alone, however, do not<br />

explain why compounds containing unmethylated guanine moiety, capable to form three<br />

hydrogen bonds involving O 6 , N(1) and N 2 , do not competitively inhibit translation 157 and<br />

are unable to bind <strong>eIF4E</strong> as well as an other cap-binding protein, the <strong>mRNA</strong> 5' cap-specific<br />

viral methyltransferase VP39. 221,222 Enhanced stacking <strong>of</strong> the methylated, positively<br />

charged bases <strong>with</strong> two parallel aromatic side chains <strong>of</strong> VP39 (Tyr22 and Phe180) plays a<br />

dominant role in the cap-recognition, 221 and hydrogen bonding is <strong>of</strong> secondary<br />

importance. 223 The single- and double-substitutions <strong>of</strong> Tyr22 and Phe180 by more strongly<br />

stacking tryptophans were shown to be associated <strong>with</strong> increased affinity for 7-<br />

methylguanosine by factors <strong>of</strong> 10 and 50, respectively. 224 The structural requirements for<br />

the specific cap recognition were qualitatively examined by means <strong>of</strong> the fluorometric<br />

titration applied to GST-fused <strong>eIF4E</strong> and His-tagged VP39, <strong>with</strong> amino acid substitutions<br />

in the cap-binding sites <strong>of</strong> the proteins. 164 The results suggest that, in order to permit the<br />

selective binding <strong>of</strong> m 7 GDP, <strong>eIF4E</strong> and VP39 require both the aromatic residues and a<br />

69


glutamic acid in a specific mutual orientation. Although VP39 and <strong>eIF4E</strong> do not share a<br />

common phylogenetical ancestor, they use a similar mode <strong>of</strong> cap binding.<br />

Application <strong>of</strong> the exact quantitative approach makes it possible now to elucidate this<br />

strict stacking/hydrogen bonding cooperativity. The observed entirely different affinity <strong>of</strong><br />

the 7-methylated nucleotides as compared <strong>with</strong> their unmethylated counterparts does not<br />

arise from different acidic/basic properties <strong>of</strong> the two classes <strong>of</strong> compounds, as was<br />

suggested previously, 156 but originates from their different stacking ability.<br />

4.2.2.4. Energetic Cost <strong>of</strong> Methylation <strong>of</strong> the N(2)-Amino Group<br />

Large ΔΔG° <strong>of</strong> +13 kJ/mol that accompanies substitution <strong>of</strong> the second amino-proton<br />

m 7 , 2 , 2<br />

for the methyl group at N 2 <strong>of</strong> guanine in 3 GTP points to disruption <strong>of</strong> possibly more<br />

than one hydrogen bond between <strong>eIF4E</strong> and the cap. If one subtracts the steric and<br />

hydrophobic costs (ΔΔG° <strong>of</strong> about +3 kJ/mol, as observed for the methyl substitution <strong>of</strong><br />

m 7 , 2<br />

2<br />

the first amino-proton in GTP , Table 4-3), the resulting energy (about +10 kJ/mol)<br />

suggests breaking <strong>of</strong> not only the Glu103–COO...HN 2 –m 7 GTP hydrogen bond but also the<br />

second hydrogen bond, Glu103–COO...HN(1)–m 7 GTP, most likely due to steric effects <strong>of</strong><br />

the bulk dimethylamino group, which can move the acceptor group <strong>of</strong> Glu103 away.<br />

4.2.2.5. Energetic Cost <strong>of</strong> the Second Nucleoside Addition and Modification<br />

All <strong>of</strong> the dinucleoside triphosphates bind to <strong>eIF4E</strong> less strongly than m 7 GTP, by<br />

ΔΔG° <strong>of</strong> +6.6 to +10 kJ/mol, and even <strong>with</strong> lesser strength than m 7 GDP (Tables 4-2, 4-3).<br />

The decreased affinity suggests that the second nucleoside does not bind tightly to <strong>eIF4E</strong>.<br />

Independently from the affinity studies, the same conclusion arises from the cocrystal<br />

structure <strong>of</strong> m 7 GpppG bound to <strong>eIF4E</strong>. Figures 4-13b and 4-14 show the only part <strong>of</strong><br />

m 7 GpppG, which can be detected from the crystallographic data. There is no defined<br />

electron density for the second guanine nucleoside. 1<br />

The presence <strong>of</strong> the unbound fragment <strong>of</strong> the cap analogue can destabilize the<br />

remaining intermolecular contacts. Addition <strong>of</strong> the second guanosine to GTP gives<br />

disadvantageous entropic effect by ΔΔG° <strong>of</strong> about +3.6 kJ/mol, while the second<br />

guanosine attached to m 7 GTP results in ΔΔG° about +6.6 kJ/mol (Table 4-3). This larger<br />

value, together <strong>with</strong> the former finding that GTP is bound only through the phosphates (see<br />

4.2.2.2., p.68), can indicate that the interactions <strong>of</strong> the 7-methylguanine moiety in<br />

70


Figure 4-14. Stereo view <strong>of</strong> the geometry <strong>of</strong> the <strong>eIF4E</strong> cap-binding site <strong>with</strong> the only part <strong>of</strong><br />

m 7 GpppG, which was detected by crystallography. 1 Molecular electrostatic potential <strong>of</strong> the <strong>eIF4E</strong><br />

surface at pH 6.0 in the absence <strong>of</strong> the ligand was calculated by <strong>Protein</strong> Explorer 225 (red and blue<br />

represent negatively and positively charged regions, respectively). The ligand is represented as<br />

sticks (carbon – green, other atoms – CPK colours). Water molecules are shown as magenta balls.<br />

m 7 GpppG are also disturbed by the unbound second nucleoside "hanging" on the<br />

phosphate chain, by ΔΔG° <strong>of</strong> about +3 kJ/mol. The crystallographic structure <strong>of</strong> the<br />

m 7 GpppG-<strong>eIF4E</strong> complex shows clearly, that the hydrogen bonds <strong>with</strong> Glu103 (donor-<br />

acceptor distances > 3Å) and the van der Waals interaction <strong>of</strong> the 7-methyl group <strong>with</strong><br />

Trp166 (distance > 4Å) are destabilized (Fig. 4-13b).<br />

In the other crystal complex <strong>of</strong> m 7 GpppA <strong>with</strong> full length human <strong>eIF4E</strong>, 48 the second<br />

adenosine is visible in the electron density map. It is stabilized by a direct hydrogen bond<br />

between adenosine N 6 -amino group and the backbone carbonyl <strong>of</strong> Thr205. Guanosine<br />

possesses a hydrogen acceptor at C(6), which precludes formation <strong>of</strong> this bond, and the<br />

protein loop containing Thr205 is disordered. The binding free energy (ΔG°) <strong>of</strong> m 7 GpppA<br />

is slightly less than that <strong>of</strong> m 7 GpppG (Table 4-2), which was recently confirmed by<br />

subsequent studies. 74 Hence, the energetic gain from the additional stabilization contact <strong>of</strong><br />

the adenosine does not compensate for the entropic cost related to the ordering <strong>of</strong> the C-<br />

terminal loop.<br />

71


Methylation <strong>of</strong> the second guanine at N(7) decreases the affinity <strong>of</strong> tri- and<br />

tetraphosphate dinucleotide caps to the same extent, ΔΔG° about +3.4 kJ/mol, consistent<br />

<strong>with</strong> attenuation <strong>of</strong> the negative charge <strong>of</strong> the phosphate chain (see below) and, at least<br />

partially, <strong>with</strong> the entropic cost <strong>of</strong> ordering the larger molecule. Exchange <strong>of</strong> the second<br />

base for adenine or cytosine is energetically less favourable. A minute advantageous effect<br />

in comparison <strong>with</strong> m 7 GpppG appears after methylation <strong>of</strong> the second ribose at O(2'),<br />

which is however in the range <strong>of</strong> the experimental error.<br />

4.2.2.6. Contributions <strong>of</strong> Phosphate Groups and Trapped Water Molecules<br />

Data in Table 4-3 show a clear accordance <strong>of</strong> the ΔΔG° values related to the<br />

phosphate interactions between two series <strong>of</strong> compounds: 7-methylated and nonmethylated<br />

guanosine phosphates. Addition <strong>of</strong> the phosphate groups one after another is accompanied<br />

by formation <strong>of</strong> direct or water-mediated hydrogen bonds and salt bridges, yielding the<br />

same free energies in the two groups <strong>of</strong> compounds. This finding shows again that there is<br />

no influence <strong>of</strong> the positive charge resulting from N(7)-methylation on the charge<br />

distribution at the phosphate chain. ΔΔG° for each step agrees <strong>with</strong> the number <strong>of</strong><br />

intermolecular contacts revealed by the crystallographic structures <strong>of</strong> the <strong>eIF4E</strong>-cap<br />

complexes (Fig. 4-13). The α-phosphate forms a direct salt bridge <strong>with</strong> Nχ2 <strong>of</strong> Arg157 as<br />

well as three indirect hydrogen bonds <strong>with</strong> Nε1 <strong>of</strong> Trp102, Nε1 <strong>of</strong> Trp166 and Nχ2 <strong>of</strong><br />

Arg112 via five water molecules trapped in the cavity <strong>of</strong> the binding centre. All water-<br />

mediated hydrogen bonds satisfy a requirement <strong>of</strong> tetrahedral geometry <strong>of</strong> the water<br />

molecules and the Pα-oxygen anion. These five bridging water molecules support the same<br />

hydrogen-bonded network both in the <strong>eIF4E</strong>-m 7 GDP complex 75 and the <strong>eIF4E</strong>-m 7 GpppG<br />

complex, and are present also in the crystallographically independent second complexes in<br />

the asymmetric units. It should be noted here that such specific network <strong>of</strong> water molecules<br />

in the protein interior is sometimes very rigid, and the water molecules buried in the<br />

internal cavities increase the protein stability. 226 Moreover, the internal water scaffold<br />

which is directly involved in ligand binding contributes to the protein specificity attained<br />

by the evolutionary specialization. 227<br />

The total free energy <strong>of</strong> the α-phosphate stabilization is ΔΔG° = −12.59 ± 0.67<br />

kJ/mol (Table 4-3). The β-phosphate group makes a salt bridge <strong>with</strong> Nζ <strong>of</strong> Lys162 and a<br />

hydrogen bond <strong>with</strong> Nε <strong>of</strong> Arg157, <strong>with</strong> the total ΔΔG° = −7.87 ± 0.25 kJ/mol, i.e. about<br />

72


−4 kJ/mol per one direct bond. This indicates average ΔΔG° <strong>of</strong> about –3 kJ/mol per one<br />

water-mediated hydrogen bond in case <strong>of</strong> the α-phosphate. The latter value is slightly less<br />

than typical magnitudes <strong>of</strong> hydrogen bonds and salt bridges in a solvent accessible region<br />

in proteins (–5.4 ± 2.5 kJ/mol) 141-143 due to the entropic cost <strong>of</strong> ordering the water<br />

molecules. The γ-phosphate <strong>of</strong> m 7 GpppG is stabilized by an additional, bifurcated salt<br />

bridge <strong>with</strong> Nζ <strong>of</strong> Lys206 <strong>of</strong> a similar energy about –3.9 kJ/mol (Fig. 4-13b).<br />

Extension <strong>of</strong> the phosphate chain by the δ-moiety in the dinucleotide cap analogues<br />

gives an energetic gain <strong>of</strong> about −6.3 kJ/mol, what cancels out the energetic cost <strong>of</strong> the<br />

second guanosine addition to m 7 GTP (Table 4-3). Most likely the δ-phosphate does not<br />

interact directly <strong>with</strong> <strong>eIF4E</strong> but the four-membered chain secures the necessary<br />

conformational freedom for the second nucleoside which is not bound, enabling the<br />

remaining m 7 Gppp-part <strong>of</strong> the cap to bind tightly. Thus, the presence <strong>of</strong> the δ-phosphate in<br />

m 7 GppppG leads to formation <strong>of</strong> the same intermolecular contacts as for m 7 GTP (ΔG° =<br />

−45.099 ± 0.088 kJ/mol) and restores the affinity <strong>of</strong> the dinucleotide cap analogue to the<br />

same level (ΔG° = −44.96 ± 0.11 kJ/mol). However, it cannot be excluded that the δ-<br />

phosphate is additionally stabilized, and the entropic cost due to loss <strong>of</strong> some<br />

conformational degrees <strong>of</strong> freedom <strong>of</strong> the second nucleoside partially cancels the<br />

stabilization effect, thus leading to accidental agreement.<br />

4.2.3. <strong>Interaction</strong>s in Context <strong>of</strong> Environment<br />

Interpretation <strong>of</strong> structural and <strong>of</strong> thermodynamic data obtained under a single set <strong>of</strong><br />

solution conditions is complicated by the fact that ligand binding by proteins is an<br />

exchange reaction. Functional groups <strong>of</strong> both protein and the ligand are capable <strong>of</strong><br />

interacting not only <strong>with</strong> each other, but also <strong>with</strong> water and solutes. Hydrogen bonds <strong>with</strong><br />

water and Coulombic interactions <strong>of</strong> phosphates <strong>with</strong> cations may be comparable to the<br />

protein-ligand interactions that replace them in the specific complex. In order to evaluate<br />

the role ions and water upon the <strong>eIF4E</strong>-cap complex formation the dependence <strong>of</strong> the Kas<br />

values <strong>of</strong> several cap analogues on salt concentration and pH have been investigated.<br />

73


4.2.3.1. <strong>Interaction</strong> <strong>of</strong> <strong>eIF4E</strong>-Cap in the Presence <strong>of</strong> Electrolyte<br />

The process <strong>of</strong> <strong>eIF4E</strong>-cap binding has been examined regarding the salt effects in<br />

two complementary ways: according to Davies-Stockes-Robinson electrostatic screening<br />

theory 191,192 and according to stoichiometric Wyman linkage analysis. 193<br />

4.2.3.1.1. Davies-Stockes-Robinson Electrostatic Screening Approach<br />

4.2.3.1.1.1. Electrostatic Effects<br />

Fluorescence (a. u.)<br />

100<br />

80<br />

60<br />

40<br />

10 -3 10 -2 10 -1 10 0 10 1 10 2<br />

m 7 GppppG (μM)<br />

Figure 4-15. Exemplary <strong>eIF4E</strong> binding isotherms for m 7 GppppG at different KCl concentrations<br />

added to Hepes 50 mM, pH 7.2, at 20 °C (, 20 mM; , 100 mM; , 201 mM; , 500 mM).<br />

At elevated ionic strength (I) the interaction is strongly attenuated (Fig. 4-15, Table<br />

4-4). Screening <strong>of</strong> the electrostatic attraction between <strong>eIF4E</strong> and the cap analogues <strong>of</strong><br />

increasing negative charges, i.e. m 7 GDP, m 7 GpppG, m 7 GTP and m 7 GppppG, is<br />

qualitatively shown in Fig. 4-16, on the basis <strong>of</strong> the simple Brønstes-Bjerrum-Debye-<br />

Hückel limited law. ∗ This approximation, although too far-reaching, is useful to show a<br />

certain experimental regularity. The negative slope <strong>of</strong> the regression line is approximately<br />

proportional to the charge at the phosphate chain, assuming that the terminal phosphates in<br />

the mononucleotide cap analogues are not fully ionized at pH 7.2 (pKa phosph ~6.1-6.5,<br />

depending on the number <strong>of</strong> phosphate groups 178 ).<br />

∗ A linear dependence <strong>of</strong> logKas on the square root <strong>of</strong> ionic strength (I 1/2 ); assumptions: very low salt<br />

concentration (< 0.01 M), dissociation rate constant the same for each I value, point charges (aj = 0), no<br />

hydration (ΔN = 0). None <strong>of</strong> the assumptions is fulfilled in the case <strong>of</strong> macromolecular interactions at the salt<br />

concentration close to physiological conditions, hence no quantitative conclusions can be drawn.<br />

74


log Kas - log Kas(0)<br />

0<br />

-1<br />

-2<br />

-3<br />

-4<br />

m 7 m<br />

GppppG<br />

7 m<br />

Gppp<br />

7 Gpp<br />

m 7 GpppG<br />

0.0 0.2 0.4 0.6 0.8<br />

I 1/2 (M 1/2 )<br />

Figure 4-16. The ionic strength-dependent decrease <strong>of</strong> the relative affinity for <strong>eIF4E</strong> <strong>of</strong> four<br />

selected cap-analogues <strong>with</strong> increasing negative charges. The approximate slopes are the steepest<br />

the more negative are the charges <strong>of</strong> the phosphate chains. Kas(0) is the value <strong>of</strong> the association<br />

constant at I = 0, extrapolated on the basis <strong>of</strong> the linear regression.<br />

Table 4-4. Equilibrium association constants for binding <strong>of</strong> <strong>eIF4E</strong> to the cap analogues at 20 °C,<br />

pH 7.2, in Hepes/KOH 50 mM (pKa = 7.5), and at the concentration <strong>of</strong> added KCl as listed below.<br />

Ionic strength was calculated taking into account contribution from partially ionized buffer. The<br />

slopes and the correlation coefficients (r 2 ) have been obtained by linear regression based on the<br />

Brønstes-Bjerrum-Debye-Hückel limited law.<br />

[KCl] m 7 GDP m 7 GpppG m 7 GTP m 7 GppppG<br />

(mM) Kas ⋅ 10 -6 (M -1 )<br />

5.3 1900 ± 3200<br />

20 40.7 ± 5.6 34.4 ± 3.8 1000 ± 460 1112 ± 530<br />

50 17.9 ± 1.0 192 ± 23<br />

54 20.2 ± 2.4 263 ± 32<br />

104 20.4 ± 1.5 7.39 ± 0.46 121.0 ± 3.0 102.8 ± 4.4<br />

150 3.92 ± 0.18 73.0 ± 4.0<br />

201 4.94 ± 0.71 2.99 ± 0.32 21.5 ± 1.9 27.6 ± 3.1<br />

250 1.96 ± 0.13 36.5 ± 1.5<br />

300 1.21 ± 0.18<br />

350 0.88 ± 0.10 3.98 ± 0.79<br />

500 1.83 ± 0.31 0.47 ± 0.12 3.85 ± 0.33 1.83 ± 0.33<br />

slope -2.58 ± 0.12 -3.619 ± 0.044 -4.47 ± 0.14 -5.08 ± 0.29<br />

r 2<br />

0.9553 0.9895 0.9592 0.9901<br />

75


Monovalent cations were reported to significantly increase in vitro translation in<br />

rabbit reticulocyte lysate and K + appeared to have the most suitable radius to support<br />

translation. 228 As the efficiency <strong>of</strong> <strong>eIF4E</strong>-<strong>mRNA</strong> 5'-cap binding decreases at higher K +<br />

concentration, the positive effect caused by K + should rather be attributed to the regulation<br />

<strong>of</strong> other translation components.<br />

Figure 4-17. Determination <strong>of</strong> hydration number ΔN from KCl-dependence <strong>of</strong> the <strong>eIF4E</strong>m<br />

7 GpppG complex formation, at 20°C. <strong>Analysis</strong> according to electrostatic screening theory (solid<br />

line) and stoichiometric cation release <strong>with</strong> the cation number c = 1 (broken line).<br />

An extended analysis, including finite dimensions <strong>of</strong> the interacting molecular<br />

spheres and the hydration effects revealed by the osmotic stress, has been performed for<br />

m 7 GpppG ∗ (Fig. 4-17). The optimal result has been obtained <strong>with</strong>in Davies approximation<br />

for the effective ionic diameter a j = 8 Å, which describes the sum <strong>of</strong> the radii <strong>of</strong> the<br />

interacting species, i.e. three-membered phosphate chain <strong>of</strong> m 7 GpppG and the group <strong>of</strong><br />

four positively charged amino acids localized at the entry <strong>of</strong> the <strong>eIF4E</strong> cap-binding slot<br />

(Arg112, Arg157, Lys162 and Lys206, Fig. 4-13). The product <strong>of</strong> mutually attracting<br />

charges is z = −11.<br />

54 ± 0.<br />

95e.<br />

Previous studies showed that the resultant protein charge<br />

z1 2<br />

log Kas (m 7 GpppG)<br />

8<br />

7<br />

6<br />

5<br />

10 -2 10 -1 10 0<br />

KCl (M)<br />

is only +0.58 e, while a dipole moment is substantial (579 D). 4 Hence, the cap recognizes<br />

directly the binding site <strong>of</strong> <strong>eIF4E</strong> and not the protein as a whole (Fig. 4-18), and the charge<br />

∗ The extended, non-linear analysis required very accurate experimental data for Kas. Those were only<br />

available for m 7 GpppG. The binding constants <strong>of</strong> the remaining cap analogues were determined <strong>with</strong> too<br />

much scattering due to the greater extent <strong>of</strong> wear <strong>of</strong> the xenon impulse lamp <strong>of</strong> the spectr<strong>of</strong>luorometer and<br />

thus insufficient signal-to- noise ratio (< 100:1).<br />

76


localized on the guanine moiety does not by itself play any important role in the first step<br />

<strong>of</strong> the encounter. The negatively charged phosphate chain serves as a molecular anchor,<br />

enabling the cap to form further contacts <strong>with</strong>in the binding site. This conclusion is in a<br />

very good agreement <strong>with</strong> our independent finding that the free energies corresponding to<br />

the phosphate chain interactions are separate from the energy <strong>of</strong> stacking together <strong>with</strong><br />

hydrogen bonds (Table 4-3).<br />

77<br />

Figure 4-18. Molecular electrostatic potential<br />

<strong>of</strong> the <strong>eIF4E</strong> surface at pH 6.0 75 calculated by<br />

<strong>Protein</strong> Explorer 225 in the absence <strong>of</strong> the cap<br />

analogue. The ligand is represented as green<br />

sticks.<br />

4.2.3.1.1.2. Two-Step Mechanism <strong>of</strong> the Cap-<strong>eIF4E</strong> Association<br />

Taken together, the above results provide a molecular interpretation <strong>of</strong> the two-step<br />

mechanism observed for the cap-<strong>eIF4E</strong> association by means <strong>of</strong> stopped-flow<br />

spectroscopy 4 . The first step can be now attributed to the anchoring the phosphate on the<br />

external basic amino acids, and the second step to the interactions deeply inside the m 7 G-<br />

binding slot, i.e. cooperative cation-π stacking and hydrogen bonding <strong>with</strong> Trp56, Trp102<br />

and Glu103. As all these residues are located at the flexible loops, 75 this second step <strong>of</strong> the<br />

complex formation is expected to involve a conformation change.<br />

The two-step model could be verified by mutant analysis, a mirror reflected<br />

methodological approach to the current cap analogue studies. The mutations eliminating all<br />

phosphate interactions should yield the equilibrium Kas for m 7 GTP close to that for m 7 G<br />

(Table 1) and considerably decrease the kinetic encounter rate constant k+1. The complete<br />

cap-slot mutations should retain only the phosphate binding affinity, <strong>with</strong> Kas as for GTP<br />

and the k+1 unchanged. However, some mutations <strong>of</strong> crucial residues can change not only<br />

the nature <strong>of</strong> amino acids, but also the internal structure <strong>of</strong> the protein (local or even global


fold), thus obscuring the investigated effect. By contrast, the studies based on the<br />

chemically modified cap analogues, where each chemical alteration can be individually<br />

controlled, are free from such pitfalls.<br />

4.2.3.1.1.3. Relation to the "Clamping Cycle"<br />

The analysis <strong>of</strong> electrostatic screening confirms a biological role for a putative<br />

phosphate bridge connecting Ser-209 <strong>with</strong> Lys-159. It is thought to "clamp" the previously<br />

bound <strong>mRNA</strong> chain 229 when <strong>eIF4E</strong> is phosphorylated by the eIF4G-associated Mnk1<br />

kinase. 230,231 As the phosphate chain <strong>of</strong> the cap is electrostatically steered toward the<br />

binding site, the mechanism <strong>of</strong> a "clamping cycle", where the phosphorylation happens<br />

after the cap-<strong>eIF4E</strong> binding, 229 is most likely. Otherwise, if the phosphorylation at Ser-209<br />

occurred prior to the cap binding, it would attenuate the electrostatic attraction between<br />

<strong>eIF4E</strong> and the cap. Therefore, the affinity <strong>of</strong> the cap to the previously phosphorylated<br />

protein would be significantly decreased. Recent data showing the effect <strong>of</strong> <strong>eIF4E</strong><br />

phosphorylation on the affinity <strong>of</strong> cap for <strong>eIF4E</strong> 73,74 confirm this conclusion.<br />

4.2.3.1.1.4. Osmotic Stress<br />

Molecular crowding <strong>of</strong> the solution causes a shortage <strong>of</strong> the free water molecules.<br />

The Kas at elevated salt concentration decreases more than it could be explained by<br />

electrostatic screening only (Fig. 4-17), indicating that the binding <strong>of</strong> the cap to <strong>eIF4E</strong><br />

requires additional hydration <strong>of</strong> the molecular surfaces. It should be noted that, although<br />

the activity <strong>of</strong> water at the maximal salt-concentration in the present study was not very far<br />

from unity (aw ~0.97 at 0.5 M KCl), the hydration effects appear quite distinctly, and the<br />

considerable number <strong>of</strong> water molecules that are taken up to the cap-<strong>eIF4E</strong> complex has<br />

been detected: ΔN = 68 ± 16.<br />

The association constant <strong>of</strong> m 7 GpppG extrapolated to the optimal hydration<br />

conditions, i.e. <strong>with</strong>out osmotic stress and electrostatic screening, is<br />

K<br />

as<br />

( 0)<br />

9<br />

−1<br />

= 1.<br />

12 ± 0.<br />

18⋅10<br />

M . It is 150-fold greater than Kas measured in 100 mM KCl,<br />

and is an evidence <strong>of</strong> the strong electrostatic contribution to <strong>eIF4E</strong>-cap binding. On the<br />

other hand, the association constant extrapolated to the pseudostandard state <strong>of</strong> 1M KCl is<br />

K<br />

as<br />

( 1)<br />

4<br />

−1<br />

= 4.<br />

8 ± 1.<br />

9 ⋅10<br />

M . In spite <strong>of</strong> strong electrostatic screening, the intrinsic<br />

thermodynamic preference <strong>of</strong> <strong>eIF4E</strong> for the methylated cap is still efficient. This<br />

78


completely screened interaction <strong>of</strong> <strong>eIF4E</strong> <strong>with</strong> m 7 GpppG at 1 M KCl has exactly the same<br />

association constant as m 7 G, the analogue insensitive toward electrostatic interactions,<br />

K<br />

as<br />

7<br />

4<br />

−1<br />

( m G)<br />

= 4.<br />

6 ± 1.<br />

2 ⋅10<br />

M (Table 4-2). Hence, at ionic strength <strong>of</strong> 1 M all the<br />

interactions <strong>with</strong> the phosphates are lost.<br />

4.2.3.1.2. Wyman Linkage <strong>Analysis</strong><br />

In terms <strong>of</strong> Wyman linkage analysis, stoichiometrical ionic interactions are taken<br />

into account in the same manner as excluded volume effects. The most probable process<br />

accompanying the <strong>eIF4E</strong>-cap binding in addition to the change <strong>of</strong> preferential hydration is<br />

potassium cation release 118 from the phosphate chain <strong>of</strong> the cap analogue:<br />

cap( c⋅K + ) + <strong>eIF4E</strong> + ΔN⋅H2O cap<strong>eIF4E</strong>( ΔN⋅H2O) + c⋅K +<br />

4.2.3.1.2.1. Electrolyte Effect<br />

The number <strong>of</strong> K + released from the ligand, c = 0.826 ± 0.092 (Fig. 4-17), indicates<br />

that possibly one K + is displaced from the phosphate chain <strong>of</strong> m 7 GpppG upon binding to<br />

<strong>eIF4E</strong>. Cation release (polyelectrolyte effect) was extensively studied for the entropy-<br />

driven B-DNA interactions <strong>with</strong> oligocations. 118 Large contribution <strong>of</strong> the counterions<br />

displacement into bulk solvent was also reported for anthracycline antibiotic binding to<br />

DNA. 120 Because the cap analogues lacked substantial axial charge distribution, such as<br />

that observed in the helical B-DNA polyanion, the cation dilution effect is much less<br />

pronounced. The contribution <strong>of</strong> the electrolyte effect to the stability <strong>of</strong> the m 7 GpppG-<br />

<strong>eIF4E</strong> complex at 100 mM KCl is ΔG°el = −6.40 ± 0.63 kJ/mol, in comparison <strong>with</strong> the<br />

total free energy <strong>of</strong> m 7 GpppG binding upon these conditions, ΔG° = −38.55 ± 0.15 kJ/mol<br />

(Table 4-2), so the entropic contribution to the association driving force, resulting from the<br />

cation release, is rather moderate (ΔS°el = + 21.8 ± 2.1 J/mol⋅K). The association constant<br />

calculated from the stoichiometrical model and extrapolated to 1M KCl is the same as that<br />

derived from the approach <strong>of</strong> the continuous screening, Kas(1) = 4.45 ± 0.34 ⋅ 10 4 M -1 .<br />

Relatively high Kas ~10 4 <strong>of</strong> m 7 GpppG at 1M KCl and the moderate influence <strong>of</strong> the<br />

electrolyte effect on the binding distinguish the electrostatically attracting cap-<strong>eIF4E</strong><br />

system from the counterion release-driven DNA-oligocation system, for which the<br />

corresponding Kas value was <strong>with</strong>in an order <strong>of</strong> magnitude <strong>with</strong> unity. 118<br />

79


4.2.3.1.2.2. Osmotic Stress<br />

The number <strong>of</strong> water molecules that have to be taken up from the bulk is ΔN = 95 ±<br />

18. Keeping the constant value <strong>of</strong> c = 1 does not cause significant worsening <strong>of</strong> the fit,<br />

according to the Snedecor's F-test (P = 0.11), and gives ΔN = 63.5 ± 9.3. Both <strong>of</strong> these<br />

values agree <strong>with</strong>in the numerical accuracy <strong>with</strong> the hydration effect (ΔN) derived from the<br />

complementary model <strong>of</strong> electrostatic screening, so the two theoretical treatments indicate<br />

the same extent <strong>of</strong> water participation in the cap binding event.<br />

4.2.3.1.3. Comparison <strong>of</strong> the Two Approaches<br />

In respect <strong>of</strong> the direct interaction <strong>with</strong> the electrolyte, the Snedecor's F-test<br />

determines that the screening theory is better (R 2 = 0.9967) than the cation release<br />

approach (R 2 = 0.9906 for c = 1), on the significance level <strong>of</strong> P = 0.0155. Hence, the<br />

description in terms <strong>of</strong> continuous screening by the ionic atmosphere <strong>of</strong> the excess salt is<br />

more relevant. The possible stoichiometric interaction <strong>with</strong> the counterion contributes to<br />

the screening, and additionally provides the advantageous entropic effect to the binding.<br />

4.2.3.1.4. Reasons for KCl-Dependence <strong>of</strong> Internal Rearrangement Rate Constants<br />

Significance <strong>of</strong> the electrostatic forces in kinetics <strong>of</strong> the <strong>eIF4E</strong>-cap complex<br />

formation was also shown by the fluorescence stopped-flow measurements. 4 The encounter<br />

rate constant (k+1) measured experimentally and confirmed by Brownian dynamics<br />

simulations was markedly diminished by the ionic strength (from 4.58 ± 0.18 ⋅10 8 M -1 ⋅s -1 at<br />

50 mM KCl to 0.68 ± 0.04 ⋅10 8 M -1 ⋅s -1 at 350 mM KCl), and the dissociation rate constant<br />

(k-1) was slightly increased (from 30 ± 8 s -1 to 45 ± 7 s -1 , respectively). However, estimates<br />

<strong>of</strong> the second-step rate constants for the internal rearrangement <strong>of</strong> the initial complex (k+2<br />

and k−2) revealed surprisingly analogous ionic strength dependence (from 124 to 73 s -1 , and<br />

from 1 to 12 s -1 , respectively). This intriguing effect could not be explained when<br />

neglecting the excluded volume effects. Now, in light <strong>of</strong> the osmotic stress analysis, the<br />

decrease <strong>of</strong> the second-step rate constant k+2 and the increase <strong>of</strong> k−2 is readily explicable in<br />

terms <strong>of</strong> the conformational change, upon which the significant number <strong>of</strong> water molecules<br />

have to hydrate the protein surface, which becomes more difficult at the elevated osmotic<br />

stress.<br />

80


4.2.3.1.5. Discussion <strong>of</strong> Hydration Effects<br />

For several years increasing attention was paid both to the activity <strong>of</strong> water and to the<br />

ions that can participate in the intermolecular interactions. 117,126,232,233 Additional hydration<br />

can accompany the ligand binding by the proteins that show simultaneous conformational<br />

changes. For instance, the hydration <strong>of</strong> haemoglobin by 60–65 water molecules occurs<br />

<strong>with</strong> the transition from the deoxy T to fully oxygenated R state. 119 Only one chlorine<br />

anion is released upon this transition. Recently, hydration changes that accompany the<br />

binding <strong>of</strong> several intercalators to DNA were exhaustively studied 121 . It was found that<br />

from 0 to 30 water molecules are taken up during complex formation, depending on the<br />

intercalator type. On the other hand, the entropy-driven association <strong>of</strong> lac repressor to lac<br />

operator DNA is accompanied by the release <strong>of</strong> ~200 water molecules. 114 In the cap-<strong>eIF4E</strong><br />

system, the additional hydration similar to that accompanying the haemoglobin<br />

oxygenation is observed. The total number <strong>of</strong> water molecules taken up from the bulk<br />

solvent (~ 65) is much greater than that trapped inside the cap-binding centre, found by<br />

crystallography (9 and 16, in the complex <strong>with</strong> m 7 GDP 75 and m 7 GpppG, 1 respectively, Fig.<br />

4-19), and apparently comparable <strong>with</strong> the amount <strong>of</strong> total water bound per one protein<br />

molecule. 75,91 It should be noted, however, that the water found in the X-ray diffraction<br />

structures is usually an integral part <strong>of</strong> the macromolecular structure and is located in its<br />

internal cavities. The water taken up from the bulk solvent during the complex formation<br />

refers not only to the trapped water but also to the changes <strong>of</strong> the first-layer hydration shell<br />

<strong>of</strong> the molecular surfaces. 234 Such peripheral water is highly unstructured, due mainly to<br />

fluctuations <strong>of</strong> the water molecules imposed by the protein hydrophobic groups.<br />

Figure 4-19. 16 water molecules (magenta spheres, van der Waals radii) were found by<br />

crystallography in the <strong>eIF4E</strong>-m 7 GpppG complex deeply inside the cap-binding pocket. Molecular<br />

solvent accessible surfaces (beige for <strong>eIF4E</strong> and bluish for m 7 GpppG, probe radius <strong>of</strong> 1.4 Å) show<br />

that these water molecules are tightly closed in the protein cavity by the ligand which serves as a<br />

molecular stopper, and thus cannot be regarded as a part <strong>of</strong> the dynamic first-layer hydration shell.<br />

81


4.2.3.1.6. Conformational Change <strong>of</strong> <strong>eIF4E</strong> upon Cap Binding<br />

Such extensive hydration <strong>of</strong> <strong>eIF4E</strong> would respond to exposure <strong>of</strong> additional ~450 Å 2<br />

<strong>of</strong> the molecular surface area to contact <strong>with</strong> water 232 . The exposure <strong>of</strong> the surface requires<br />

a substantial conformational change <strong>of</strong> the protein. This hypothesis is supported by several<br />

independent facts, i.e. by the rearrangement <strong>of</strong> the encounter complex, 4 by the pr<strong>of</strong>ound<br />

(~65%) quenching <strong>of</strong> the intrinsic fluorescence <strong>of</strong> <strong>eIF4E</strong> upon cap binding, by significantly<br />

different fluorescence quenching patterns when comparing interactions <strong>of</strong> apo-<strong>eIF4E</strong> and<br />

cap-saturated <strong>eIF4E</strong> <strong>with</strong> the 4E-BP1 and eIF4G peptides (see 4.3., p. 86), and by<br />

deaggregation <strong>of</strong> <strong>eIF4E</strong> molecules (residues 33-217) forced by the cap analogues binding,<br />

observed by means <strong>of</strong> Dynamic Light Scattering (Fig. 4-20). Conformational changes <strong>of</strong> a<br />

homologous plant eIF(iso)4F complex upon binding <strong>of</strong> m 7 GpppG were also observed by<br />

means <strong>of</strong> the fluorescence stopped-flow measurements. 235<br />

4.2.3.1.6.1. Deaggregation <strong>of</strong> <strong>eIF4E</strong> Induced by Cap Binding<br />

The apo-form <strong>of</strong> <strong>eIF4E</strong> <strong>with</strong> an N-terminal tail shorter by 5 amino acids than that<br />

used for the fluorescence studies formed huge aggregates and the solution was totally<br />

polydispersed, as checked by DLS. Only rough estimates <strong>of</strong> the molecular weight (MW)<br />

and the hydrodynamic radius (Rh) <strong>of</strong> the <strong>eIF4E</strong> aggregates could be registered. After<br />

addition <strong>of</strong> one <strong>of</strong> cap analogues (m 7 GTP, p-Cl-bz 7 GTP, m3 227 GTP, m 7 GpppG,<br />

m 7 GppppG, or m 7 Gppppm 7 G) to the <strong>eIF4E</strong> samples progressive deaggregation <strong>with</strong> the<br />

time <strong>of</strong> incubation <strong>with</strong> the cap analogue has been observed. Exemplary results are shown<br />

in Fig. 4-20. The estimates <strong>of</strong> the molecular weight and the hydrodynamic radius <strong>of</strong> the<br />

aggregates dropped after several hours down to the well defined, reliably measured values.<br />

The efficiency <strong>of</strong> the process correlated <strong>with</strong> the affinity <strong>of</strong> the cap analogue for <strong>eIF4E</strong>.<br />

After 24 hours <strong>of</strong> incubation <strong>with</strong> m 7 GTP the protein molecules was sufficiently<br />

monodispersed, characterized by MW = 23.3 kDa and Rh = 2.36 nm, which corresponded<br />

to the monomeric state <strong>of</strong> the protein in the complex <strong>with</strong> m 7 GTP (MWcx ≈ 22.2 kDa,<br />

dimensions: 41 Å × 36 Å × 45 Å, as determined by crystallography 75 ), while <strong>eIF4E</strong> in the<br />

presence <strong>of</strong> m3 227 GTP was polydispersed, and the apo-protein remained still totally<br />

aggregated. Hence, deaggregation <strong>of</strong> <strong>eIF4E</strong> has occurred under the influence <strong>of</strong> interaction<br />

<strong>with</strong> cap.<br />

82


Autocorrelation<br />

coefficient<br />

1.6<br />

1.4<br />

1.2<br />

1.0<br />

10 -6<br />

10 -5<br />

10 -4<br />

Delay time (μs)<br />

(a) 10 (b)<br />

5<br />

Figure 4-20. Deaggregation <strong>of</strong> <strong>eIF4E</strong> (33-217) at 1 mg/ml induced by the cap analogue binding.<br />

(a) The autocorrelation function becomes well defined <strong>with</strong> the time <strong>of</strong> incubation <strong>with</strong> the 50-fold<br />

excess <strong>of</strong> m 7 GTP ( , apo-protein; , after 1-hour incubation; , after 24-hour incubation). (b) The<br />

estimates <strong>of</strong> the molecular weight (MW, ) and <strong>of</strong> the hydrodynamic radius (Rh, ) <strong>of</strong> non-specific<br />

protein aggregates decrease systematically in the time <strong>of</strong> incubation <strong>with</strong> m 7 GTP by several orders<br />

<strong>of</strong> magnitude to the final values that are characteristic for the single <strong>eIF4E</strong>-cap complex.<br />

4.2.3.2. Ionic Equilibria in <strong>eIF4E</strong>-Cap Complex<br />

7-methylguanosine is a mixture <strong>of</strong> positively charged and zwitterionic forms (due to<br />

dissociation <strong>of</strong> N(1)-H, pKa ~7.24-7.54), the proportions being slightly dependent on the<br />

secondary substituents and the length <strong>of</strong> the phosphate chain. 177 The drop <strong>of</strong> the pKa value<br />

by approximately 2 units, which accompanies the 7-substitution <strong>of</strong> guanosine was<br />

suspected to have a biological significance. The zwitterionic form <strong>of</strong> 7-methylated guanine<br />

was initially postulated to be responsible for interaction <strong>with</strong> <strong>eIF4E</strong>. 156 In contrast,<br />

structural data indicated that the cationic form <strong>of</strong> m 7 GDP binds to <strong>eIF4E</strong>, because <strong>of</strong> the<br />

spatial distances suitable for formation <strong>of</strong> a hydrogen bond between solvent accessible<br />

Glu103 and N(1)-H <strong>of</strong> 7-methylguanosine in the crystal structures (PDB ID codes: 1EJ1,<br />

1AP8) 75,76 . This hydrogen bond is also unambiguously present in the ternary complex <strong>of</strong><br />

<strong>eIF4E</strong> bound to m 7 GDP and eIF4GII peptide, although the crystals were obtained at pH 8.5<br />

(PDB ID code: 1EJH) 91 . However, the N(1)-H proton was not directly observed during the<br />

NMR experiment, 76 so the issue was still not entirely resolved.<br />

10 -3<br />

10 -2<br />

The affinity <strong>of</strong> <strong>eIF4E</strong> for m 7 GTP as a function <strong>of</strong> pH (Table 4-5, Fig. 4-21) shows<br />

the optimal binding at pH 7 . 24 ± 0.<br />

13,<br />

close to the physiological pH <strong>of</strong> 7.4, and lower than<br />

the pH 7.6 reported for the full length <strong>eIF4E</strong> from human erythrocytes, purified by m 7 G-<br />

affinity chromatography. 156 The observed pH-dependence <strong>of</strong> Kas is rather flat, providing a<br />

wide pH-range <strong>of</strong> efficient cap-<strong>eIF4E</strong> binding. From the Wyman model, 193 comprising one<br />

site to be protonated and one site to be deprotonated, two effective acidic dissociation<br />

83<br />

Rh (nm), MW (kDa)<br />

10 6<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

MW<br />

R h<br />

0 6 12 18 24<br />

Time <strong>of</strong> incubation <strong>with</strong> m 7 GTP (h)<br />

23.3<br />

2.36


constants have been determined: = 7.<br />

99 ± 0.<br />

14 for N(1)-H <strong>of</strong> the ligand base moiety<br />

pK L<br />

in the presence <strong>of</strong> the protein, and = 6.<br />

49 ± 0.<br />

13 for the hydroxyl group <strong>of</strong> Glu103 in<br />

pK P<br />

the binding site <strong>of</strong> <strong>eIF4E</strong>, in the presence <strong>of</strong> the ligand (the errors <strong>of</strong> pKs are numerical).<br />

Taking into account experimental uncertainties as well as simplifications <strong>of</strong> the model, the<br />

real accuracy <strong>of</strong> pKs is close to 1 pH unit. Making the number <strong>of</strong> protonated and<br />

deprotonated sites as free parameters <strong>of</strong> the fitted function leads to negligible improvement<br />

<strong>of</strong> the results, on the basis <strong>of</strong> the statistical Snedecor's F-test. The elevated effective pKL in<br />

relation to the pKL value in the absence <strong>of</strong> the interacting counterpart shows that m 7 G<br />

exists in a cationic form in the <strong>eIF4E</strong>-bound state and donates N(1)-H proton for the<br />

hydrogen bond <strong>with</strong> the deprotonated carboxyl group <strong>of</strong> Glu103. Such an increase <strong>of</strong> pKL<br />

caused by the chemical microenvironment <strong>of</strong> the negatively charged Glu103 was<br />

postulated on the basis <strong>of</strong> the crystallographic structure. 75<br />

Table 4-5. Association constants for m 7 GTP<br />

binding to <strong>eIF4E</strong> as a function <strong>of</strong> pH, at<br />

20 °C, 100 mM KCl.<br />

pH Kas ⋅ 10 -6 (M -1 )<br />

6.32 53.7 ± 4.4<br />

6.56 74.4 ± 7.7<br />

6.78 96.9 ± 8.4<br />

7.01 97.9 ± 8.4<br />

7.20 108.7 ± 9.2<br />

7.42 102.3 ± 6.5<br />

7.61 82.1 ± 11.5<br />

7.84 74.2 ± 4.3<br />

8.02 74.7 ± 6.7<br />

log Kas (m 7 GTP)<br />

84<br />

Figure 4-21. Binding affinity <strong>of</strong> m 7 GTP to<br />

<strong>eIF4E</strong> as a function <strong>of</strong> pH, at 20°C.<br />

6.0 6.5 7.0 7.5 8.0 8.5<br />

The free energy <strong>of</strong> binding between the species in their appropriate ionic states (pH-<br />

independent ΔG°) is: ΔG° pH-ind = −45.765 ± 0.017 kJ/mol, and the energetic costs <strong>of</strong><br />

keeping the cationic state <strong>of</strong> m 7 GTP and the anionic state <strong>of</strong> Glu103 in the <strong>eIF4E</strong> binding<br />

site are very small: ΔG°L = +0.368 ± 0.11 kJ/mol and ΔG°P = +0.435 ± 0.12 kJ/mol,<br />

respectively. Even after the increase <strong>of</strong> the pKP <strong>of</strong> Glu103 to ~ 6 units inside the protein,<br />

pKL and pKP are separated enough to allow tight m 7 GTP-<strong>eIF4E</strong> binding at pH 7.2,<br />

characterized by ΔG° = −45.099 ± 0.088 kJ/mol (Table 4-2) very close to ΔG° pH-ind .<br />

8.5<br />

8.0<br />

7.5<br />

7.0<br />

pH


In summary, the decrease <strong>of</strong> pKa <strong>of</strong> guanosine as a result <strong>of</strong> methylation at N(7) is a<br />

side-effect, which must be cancelled by the microenvironment <strong>of</strong> the protein binding site.<br />

Instead, the biological importance <strong>of</strong> the 7-methylation consists in strong enhancement <strong>of</strong><br />

stacking between the cationic m 7 G and the aromatic side chains (cation - π stacking),<br />

which stabilizes the base moiety enough to allow formation <strong>of</strong> the three hydrogen bonds in<br />

the binding site (see 4.2.2., p. 64).<br />

4.2.4. Conclusions<br />

The results reveal individual role <strong>of</strong> various structural elements <strong>of</strong> the cap in the two-<br />

step process <strong>of</strong> <strong>eIF4E</strong>-cap recognition. The 5'-phosphate chain is the primary anchor to<br />

<strong>eIF4E</strong>. The <strong>eIF4E</strong>-cap binding is accomplished by the specific contacts <strong>of</strong> three phosphates<br />

and m 7 G moiety but not <strong>of</strong> the second guanosine. The biological importance <strong>of</strong><br />

methylation <strong>of</strong> the first guanosine at N(7) has been elucidated by showing that the<br />

thermodynamic origin <strong>of</strong> the strong discrimination between 7-methylated and<br />

unmethylated counterparts (5000-fold difference in Ka at 20°C) consists in absolute<br />

cooperativity <strong>of</strong> cation-π stacking and hydrogen bonding during the second step <strong>of</strong> the<br />

complex formation. The enhanced stacking between the cationic m 7 G and the aromatic Trp<br />

side chains is necessary to stabilize effectively the base moiety, which is a precondition for<br />

formation <strong>of</strong> three hydrogen bonds inside the cap-binding slot. <strong>eIF4E</strong> selects precisely<br />

between the 7-monomethylguanine and 2,2,7-trimethylguanine analogues (760-fold<br />

difference in Kas). Possibilities <strong>of</strong> cation release and water exchange that could accompany<br />

the binding have been also checked. The estimated water uptake <strong>of</strong> ~65 water molecules to<br />

the cap-binding slot and to the hydration shell <strong>of</strong> the complex is relevant to conformational<br />

rearrangement <strong>of</strong> entire <strong>eIF4E</strong> upon the second step <strong>of</strong> cap binding.<br />

85


4.3. Molecular <strong>Interaction</strong>s <strong>with</strong>in Ternary Complexes<br />

Involving <strong>eIF4E</strong>, a Cap Analogue, and an <strong>eIF4E</strong>-Binding Motif<br />

from eIF4G or 4E-BP1 <strong>Protein</strong>s<br />

The fluorescence affinity measurements have been extended to include ternary<br />

complexes which consisted <strong>of</strong> <strong>eIF4E</strong>, a cap-analogue, and a synthetic peptide<br />

encompassing the <strong>eIF4E</strong> recognition motif from the mammalian proteins eIF4GI, eIF4GII<br />

and 4E-BP1. The 4E-BP1 peptides were either unphosphorylated or monophosphorylated<br />

at Ser65 and diphosphorylated at Ser65/Thr70. The eIF4G and 4E-BP1 proteins bind to the<br />

convex dorsal surface <strong>of</strong> <strong>eIF4E</strong>, on the opposite side <strong>of</strong> the protein in respect to the cap-<br />

binding slot (Fig. 4-22). Application <strong>of</strong> the precise synchronized titration method allowed<br />

to test the putative cooperation between the two distant binding sites <strong>of</strong> <strong>eIF4E</strong>, 166,167 and<br />

the influence <strong>of</strong> phosphorylation at Ser65 and Thr70 on regulation <strong>of</strong> 4E-BP1 binding to<br />

<strong>eIF4E</strong>.<br />

(a) (b)<br />

Figure 4-22. The dorsal surface <strong>of</strong> <strong>eIF4E</strong> coloured according to the molecular electrostatic<br />

potential 225 in the absence <strong>of</strong> ligands (red – negative, blue – positive). The attached peptide is<br />

shown in the ribbon representation and corresponds to the <strong>eIF4E</strong> recognition sequence <strong>of</strong> (a) 4E-<br />

BP1 (yellow, at pH 7.5) or (b) eIF4GII (orange, at pH 8.5). 91 The peptide tyrosine is shown as balls<br />

and sticks. Trp73 <strong>of</strong> <strong>eIF4E</strong> is shown in the van der Waals spacefill representation and coloured<br />

green. Orientation <strong>of</strong> the two aromatic residues is almost ideally perpendicular. A fragment <strong>of</strong><br />

bound m 7 GDP (oxygens <strong>of</strong> the β phosphate group, red spheres) is visible in the bottom right part <strong>of</strong><br />

each figure.<br />

86


4.3.1. <strong>eIF4E</strong> Fluorescence Quenching Pattern upon Peptide Binding<br />

Association <strong>of</strong> the eIF4GI, eIF4GII or 4E-BP1 peptides <strong>with</strong> <strong>eIF4E</strong> leads to<br />

quenching <strong>of</strong> the <strong>eIF4E</strong> intrinsic fluorescence, mostly at longer wavelengths (355 – 360<br />

nm) than quenching upon association <strong>with</strong> cap (330 – 340 nm). This is caused primarily by<br />

interaction <strong>of</strong> the peptides <strong>with</strong> the water accessible Trp73 <strong>of</strong> <strong>eIF4E</strong>. This amino acid is<br />

thought to be the most important residue <strong>of</strong> the phylogenetically invariant dorsal surface<br />

region, since it interacts <strong>with</strong> three peptide amino acid side chains: Leu(5), Leu(6) and<br />

Gln(9) <strong>of</strong> eIF4GII, or Leu(5), Met(6) and Arg(9) <strong>of</strong> 4E-BP1 91 (Fig. 4-23, numbering in<br />

relation to the invariant Tyr). Mutation <strong>of</strong> Trp73 to Ala was found to prevent productive<br />

interactions <strong>with</strong> 4E-BP1 and eIF4GI in a biological assay. 230<br />

(a)<br />

(b)<br />

Figure 4-23. interatomic contacts between Trp73 <strong>of</strong> <strong>eIF4E</strong> (green sticks) and the peptides in the<br />

crystal structures. 91 The peptide amino acid side chains that interact <strong>with</strong> Trp73 are shown as thick<br />

sticks in the CPK colours. (a) Leu(5), Leu(6) and Gln(9) <strong>of</strong> the eIF4GII peptide; (b) Leu(5), Met(6)<br />

and Arg(9) <strong>of</strong> 4E-BP1. Remaining parts <strong>of</strong> the peptides are represented by the backbone. The<br />

hydrogen bond and the van der Waals bonds <strong>with</strong> their lengths are drawn as dotted lines.<br />

87


To analyse the <strong>eIF4E</strong> binding affinity <strong>of</strong> the peptides and check whether cap-binding<br />

and 4E-BP/eIF4G-binding centres are cooperative, cross-titration experiments <strong>with</strong> each<br />

peptide have been performed (Fig. 4-24). Comparison <strong>of</strong> the emission band changes for<br />

saturation <strong>of</strong> <strong>eIF4E</strong> <strong>with</strong> a cap analogue or <strong>with</strong> an oligopeptide, and <strong>with</strong> both <strong>of</strong> them is<br />

shown in Fig. 4-25. Titration curves for formation <strong>of</strong> <strong>eIF4E</strong> binary and ternary complexes<br />

<strong>with</strong> m 7 GTP and the oligopeptides are presented in Fig. 4-26, 27, 28. Irrespective <strong>of</strong> the<br />

succession <strong>of</strong> the cap and the peptide added, the total final quenching after saturation <strong>of</strong><br />

both <strong>eIF4E</strong> binding sites is always identical. However, the partial quenching (Q) <strong>of</strong> the<br />

total initial fluorescence signal, which corresponds to saturation <strong>of</strong> one <strong>of</strong> the <strong>eIF4E</strong><br />

binding sites, <strong>with</strong> and <strong>with</strong>out the prior saturation <strong>of</strong> the other, is different (Fig. 4-25,<br />

Table 4-6). The fluorescence quenching upon titration <strong>with</strong> m 7 GTP <strong>of</strong> the peptide-saturated<br />

<strong>eIF4E</strong> is < 60 %, while the apo-protein reveals ~ 65 % <strong>of</strong> quenching. The quenching<br />

resulting from the peptides binding is much smaller, ~ 12 % for m 7 GTP-saturated <strong>eIF4E</strong><br />

and ~ 19 % for the apo-protein. The differences between Q(apo) and Q(sat) are ~ 3 % for<br />

the unphosphorylated 4E-BP1 peptide, ~ 9 % for the phosphorylated 4E-BP1 peptides and<br />

~ 7 % for eIF4G peptides. This succession- and structure-dependent behaviour <strong>of</strong> the<br />

quenching patterns supports previous conclusions pointing to the conformational change <strong>of</strong><br />

<strong>eIF4E</strong> upon cap binding, and a possible conformational change upon the peptide binding.<br />

Fluorescence (a. u.)<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

0 10 20 30 40 50 60 70<br />

Time (min)<br />

<strong>eIF4E</strong> + 4GI + m 7 GTP<br />

<strong>eIF4E</strong> + m 7 GTP + 4GI<br />

Figure 4-24. A typical time course <strong>of</strong> the fluorescence intensity for the cross-titration experiment<br />

<strong>of</strong> the ternary complex formation, at 20°C.<br />

Data points ( ): 0-5 min, initial baseline; 6-38 min, titration <strong>with</strong> eIF4GI peptide; 39 min,<br />

intermediate baseline; 40-71 min, titration <strong>with</strong> m 7 GTP; 72-76 min, final baseline.<br />

Data points ( ): 0-5 min, initial baseline; 6-37 min, titration <strong>with</strong> m 7 GTP; 38 min, intermediate<br />

baseline; 39-71 min, titration <strong>with</strong> eIF4GI peptide; 72-76 min, final baseline.<br />

88


Fluorescence (a. u.)<br />

ΔΔ F (%)<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

300 320 340 360 380 400<br />

20<br />

0<br />

-20<br />

-40<br />

-60<br />

Wavelength (nm)<br />

Figure 4-25. (a) Fluorescence emission spectra excited at 290 nm <strong>of</strong> apo-<strong>eIF4E</strong> (——), <strong>eIF4E</strong> in<br />

the binary complex <strong>with</strong> the 4E-BP1 P-Ser65 peptide (-----), <strong>eIF4E</strong> in the binary complex <strong>with</strong><br />

m 7 GTP (− − −), <strong>eIF4E</strong> in the ternary complex <strong>with</strong> m 7 GTP and the peptide (− - −), and <strong>of</strong> the pure<br />

peptide (). Application <strong>of</strong> the emission wavelength <strong>of</strong> 355 nm allows for elimination <strong>of</strong> the<br />

direct fluorescence signal originating from peptide tyrosine during titration. (b) Differential spectra<br />

corresponding to saturation <strong>of</strong> the <strong>eIF4E</strong>-m 7 GTP complex <strong>with</strong> the 4E-BP1 P-Ser65 peptide (− - −),<br />

saturation <strong>of</strong> apo-<strong>eIF4E</strong> <strong>with</strong> the peptide (-----), saturation <strong>of</strong> the <strong>eIF4E</strong>-peptide complex <strong>with</strong><br />

m 7 GTP (− − −), saturation <strong>of</strong> apo-<strong>eIF4E</strong> <strong>with</strong> m 7 GTP(− - - −), and saturation <strong>of</strong> apo-<strong>eIF4E</strong> <strong>with</strong><br />

both m 7 GTP and the peptide (——).<br />

89<br />

(a)<br />

300 320 340 360 380 400<br />

Wavelength (nm)<br />

(b)


Table 4-6. Equilibrium association constants (Kas) and partial quenching (Q) <strong>of</strong> total initial<br />

fluorescence upon formation <strong>of</strong> the binary complexes <strong>of</strong> the apo-form <strong>of</strong> <strong>eIF4E</strong> <strong>with</strong> the eIF4GI,<br />

eIF4GII and 4E-BP1 peptides, and <strong>of</strong> the ternary complexes (m 7 GTP-<strong>eIF4E</strong>-peptide) <strong>of</strong> previously<br />

m 7 GTP- or peptide-saturated <strong>eIF4E</strong>, at 20°C. For reference the association constant <strong>of</strong> m 7 GTP to<br />

apo-<strong>eIF4E</strong> Kas = 108.7 ± 4.0 ⋅ 10 6 M -1 (Table 4-2), quenching Q ~ 65 %.<br />

Peptides: eIF4GI eIF4GII<br />

Kas ⋅ 10 -6 (M -1 ) for the peptides<br />

<strong>eIF4E</strong>(m 7 GTP-saturated) 21.43 ± 2.15 5.35 ± 0.20<br />

<strong>eIF4E</strong>(apo) 13.65 ± 0.86 6.51 ± 0.19<br />

Kas ⋅ 10 -6 (M -1 ) for m 7 GTP<br />

<strong>eIF4E</strong>(peptide-saturated) 120.5 ± 7.2 137.9 ± 15.8<br />

Fluorescence quenching Q (% <strong>of</strong> total initial fluorescence)<br />

<strong>eIF4E</strong>(m 7 GTP-saturated) 12.1 ± 0.9 13.6 ± 2.1<br />

<strong>eIF4E</strong>(apo) 18.3 ± 1.2 20.6 ± 2.5<br />

Peptides: 4E-BP1 4E-BP1 4E-BP1<br />

(P-Ser65) (P-Ser65/Thr70)<br />

Kas ⋅ 10 -6 (M -1 ) for the peptides<br />

<strong>eIF4E</strong>(m 7 GTP-saturated) 9.47 ± 0.39 4.77 ± 0.40 5.72 ± 0.35<br />

<strong>eIF4E</strong>(apo) 10.59 ± 0.84 5.91 ± 0.44 5.74 ± 0.34<br />

Kas ⋅ 10 -6 (M -1 ) for m 7 GTP<br />

<strong>eIF4E</strong>(peptide-saturated) 101.6 ± 7.3 107.4 ± 5.2 101.0 ± 8.6<br />

Fluorescence quenching Q (% <strong>of</strong> total initial fluorescence)<br />

<strong>eIF4E</strong>(m 7 GTP-saturated) 15.1 ± 3.6 10.2 ± 5.5 11.3 ± 4.6<br />

<strong>eIF4E</strong>(apo) 18.2 ± 9.5 18.8 ± 9.0 20.2 ± 1.4<br />

4.3.2. Cooperativity in Ternary Peptide-<strong>eIF4E</strong>-Cap Complexes<br />

In general, the image <strong>of</strong> interactions <strong>with</strong>in ternary complexes is unsymmetrical,<br />

since binding <strong>of</strong> one ligand (e. g. the peptide) can modify the protein affinity for the other<br />

(the cap analogue), but this is not necessarily accompanied by the evident influence in the<br />

reverse direction. Hence, there is no simple thermodynamic cycle that would describe<br />

binding <strong>of</strong> the two ligands to <strong>eIF4E</strong>.<br />

90


4.3.2.1. Cooperativity <strong>of</strong> Cap and eIF4G Peptides Binding<br />

The fluorometrically determined values <strong>of</strong> the association constants have been used<br />

to address the cooperativity problem by a direct quantitative comparison. All the <strong>eIF4E</strong>-<br />

peptide Kas are in the order <strong>of</strong> 10 7 M -1 (Table 4-6). The Kas values for binding <strong>of</strong> cap-<br />

saturated <strong>eIF4E</strong> to eIF4GI and eIF4GII peptides (21.43 μM -1 and 5.35 μM -1 , respectively)<br />

are close to those derived by Isothermal Titration Calorimetry 91 (ITC) (37 μM -1 and 6.7<br />

μM -1 , respectively). The peptide which binds the tightest to <strong>eIF4E</strong> is eIF4GI and its<br />

interaction is few but unambiguously influenced by the apo- or cap-saturated state <strong>of</strong><br />

<strong>eIF4E</strong>, i.e. an 1.6-fold increase <strong>of</strong> Kas has been observed for the eIF4GI peptide binding to<br />

cap-saturated <strong>eIF4E</strong>. A similar effect, albeit to a lesser extent, appears also for the eIF4GII<br />

peptide. Reversely, a slight enhancement <strong>of</strong> the m 7 GTP-affinity (only up to 1.3-fold,<br />

quantitatively beyond one standard deviation but <strong>with</strong>in two standard deviations) has been<br />

noted after previous saturation <strong>of</strong> <strong>eIF4E</strong> <strong>with</strong> eIF4GI or eIF4GII peptides.<br />

Fluorescence (a. u.)<br />

100<br />

80<br />

60<br />

40<br />

20<br />

Figure 4-26. Titration <strong>with</strong> m 7 GTP <strong>of</strong> apo-<strong>eIF4E</strong> () and <strong>of</strong> the <strong>eIF4E</strong>-eIF4GI peptide complex<br />

(); 20 °C, excitation wavelength 290 nm, emission wavelength 355 nm.<br />

The case for the cooperativity is based on several experimental results:<br />

1. The eIF4GI protein was found to strongly enhance the crosslinking effectivity <strong>of</strong> <strong>eIF4E</strong><br />

to the <strong>mRNA</strong> 5'-cap structure. 216<br />

10 -3 10 -2 10 -1 10 0 10 1<br />

m 7 GTP (μM)<br />

2. Mutation <strong>of</strong> Trp166 to Ala in the cap-binding slot resulted in the protein which was<br />

unable to bind either eIF4G or 4E-BPs at its dorsal surface. Hence, it can be concluded<br />

91<br />

<strong>eIF4E</strong>+eIF4GI<br />

<strong>eIF4E</strong>


Fluorescence (a. u.)<br />

Fluorescence (a. u.)<br />

100<br />

90<br />

80<br />

10 -3 10 -2 10 -1 10 0 10 1<br />

eIF4GI peptide ( μM)<br />

100<br />

90<br />

80<br />

(a)<br />

(b)<br />

Figure 4-27. (a) Titration <strong>with</strong> the eIF4GI peptide <strong>of</strong> <strong>eIF4E</strong> ( ) and <strong>of</strong> the <strong>eIF4E</strong>-m 7 GTP complex<br />

( ). (b) Titration <strong>with</strong> the eIF4GII peptide <strong>of</strong> <strong>eIF4E</strong> ( ) and <strong>of</strong> the <strong>eIF4E</strong>-m 7 GTP complex ( ). 20<br />

°C, excitation wavelength 290 nm, emission wavelength 355 nm.<br />

that this cap-binding residue plays a role in maintaining the internal structure <strong>of</strong> <strong>eIF4E</strong><br />

necessary for recognition <strong>of</strong> the eIF4G proteins and <strong>of</strong> 4E-BPs. 236<br />

3. Among nine crystallographically independent structures <strong>of</strong> the m 7 GDP-<strong>eIF4E</strong> binary 75<br />

and ternary 91 complexes, and the m 7 GpppG complex, 1 there are two cases where the C-<br />

terminal loop (206-213) located in the vicinity <strong>of</strong> the cap-binding site has fully ordered<br />

backbone, and the N-terminus (28-35) close to the eIF4G/4E-BP binding site is<br />

concurrently ordered. In the remaining cases, both the loop and the N-tail are<br />

92<br />

<strong>eIF4E</strong>+m<br />

<strong>eIF4E</strong><br />

7 GTP<br />

<strong>eIF4E</strong>+m<br />

<strong>eIF4E</strong><br />

7 GTP<br />

10 -3 10 -2 10 -1 10 0 10 1<br />

eIF4GII peptide ( μM)


unordered. These could indirectly suggest a possibility <strong>of</strong> cooperation between the two<br />

distant binding sites.<br />

On the other hand:<br />

1. Comparison <strong>of</strong> the crystallographic structures <strong>of</strong> binary and ternary complexes revealed<br />

no important structural differences <strong>with</strong>in the <strong>eIF4E</strong>-cap binding slot in the absence or<br />

presence <strong>of</strong> either the eIF4GII or 4E-BP1 peptide. 75,91<br />

2. Preliminary fluorescence study suggested no significant changes <strong>of</strong> the association<br />

constant for binding <strong>of</strong> <strong>eIF4E</strong> <strong>with</strong> the cap analogue in the presence <strong>of</strong> a 17-aminoacid<br />

mammalian eIF4GII peptide. 237<br />

Solution <strong>of</strong> the apparent discrepancies came from two sets <strong>of</strong> experiments.<br />

1. Recently, it was shown by the gel-shift experiments and m 7 GTP-Sepharose binding<br />

that a 17-mer fragment <strong>of</strong> yeast eIF4GI exerted a negligible influence on the <strong>eIF4E</strong>-cap<br />

interaction, while complete eIF4GI and its larger fragments produced distinct<br />

enhancement <strong>of</strong> the <strong>eIF4E</strong>-cap binding. 168<br />

2. Finally, the quantitative results reported herein confirm that the putative positive<br />

cooperativity is possible, and that the minimal <strong>eIF4E</strong>-binding motifs <strong>of</strong> the eIF4G<br />

proteins can be alone insufficient to induce such spatial changes which could be<br />

evident in the crystal structures <strong>of</strong> the cap-binding site. 91<br />

4.3.2.2. Cooperativity <strong>of</strong> Cap and 4E-BP1 Peptides Binding<br />

In case <strong>of</strong> the 4E-BP1 peptides, the affinity for the apo-protein is the same as that for<br />

the cap-saturated <strong>eIF4E</strong> (Table 4-6, p. 90). The Kas values for the unphosphorylated 17-mer<br />

peptide 4E-BP1 (9.47 ⋅ 10 6 M -1 and 10.59 ⋅ 10 6 M -1 ) agree well <strong>with</strong> Kas calculated from<br />

ITC 91 (20 ⋅ 10 6 M -1 ). ITC for the full length 4E-BP1 yielded Kas 7-fold higher (67 ⋅ 10 6 M -<br />

1 ). In terms <strong>of</strong> the binding free energy, however, this difference between ΔG° for the full<br />

length protein (−43.93 kJ/mol) and the peptide (−39.16 kJ/mol) is small, what indicates<br />

that majority <strong>of</strong> the intermolecular contacts required for stabilization <strong>of</strong> the 4E-BP1-<strong>eIF4E</strong><br />

complex are accomplished in the model system <strong>with</strong> participation <strong>of</strong> the short peptide.<br />

Binding <strong>of</strong> full length 4E-BP1 and 4E-BP2 to <strong>eIF4E</strong> was reported to enhance the<br />

cap-<strong>eIF4E</strong> association, as suggested from m 7 GTP-affinity chromatography and Surface<br />

Plasmon Resonance 166 (SPR). The reversed cooperativity, i.e. an increase <strong>of</strong> the 4E-BP2<br />

affinity for the cap-saturated <strong>eIF4E</strong> in comparison <strong>with</strong> apo-<strong>eIF4E</strong> was also found using<br />

SPR. 238 The differences in the fluorescence quenching patterns upon formation <strong>of</strong> the<br />

93


ternary complexes (Fig. 4-25, 26, 27, 28, Table 4-6), and arguments discussed above for<br />

the eIF4G peptides points to a putative cooperation between the two binding sites.<br />

However, the binary 4E-BP1 peptide-<strong>eIF4E</strong> and cap-<strong>eIF4E</strong> interactions do not affect the<br />

association constant values for the further formation <strong>of</strong> the ternary peptide-<strong>eIF4E</strong>-cap<br />

complexes, <strong>with</strong>in experimental errors. No measurable affinity changes for the 17-mer and<br />

25-mer fragments <strong>of</strong> 4E-BP1 could reflect the inability <strong>of</strong> the peptides to completely<br />

mimic the influence <strong>of</strong> full length 4E-BP1 on the conformation <strong>of</strong> the <strong>eIF4E</strong> protein. It is<br />

worth noting here that free 4E-BP1 is unfolded in solution unless gets bound to <strong>eIF4E</strong>. 239<br />

4.3.3. Regulation <strong>of</strong> 4E-BP1 Activity by Phosphorylation<br />

Phosphorylation <strong>of</strong> specific serine and threonine residues <strong>of</strong> 4E-binding proteins<br />

modulates their affinity for <strong>eIF4E</strong>. 229 A two-step mechanism <strong>of</strong> 4E-BP1 phosphorylation<br />

was proposed, involving phosphorylation on Thr37 and Thr46 as a priming event for<br />

subsequent phosphorylation <strong>of</strong> Ser65 and Thr70. 92 Recently, it was shown by<br />

phosphopeptide mapping that both <strong>of</strong> these phosphorylation events were insufficient to<br />

disrupt the 4E-BP1 binding <strong>with</strong> <strong>eIF4E</strong>, 3 and the possible sources <strong>of</strong> disagreement <strong>with</strong> the<br />

previous publications were thoroughly discussed. 93-95<br />

The fluorescence quenching experiments for cap-free and cap-saturated <strong>eIF4E</strong><br />

confirm these results (Table 4-6). The monophosphorylation <strong>of</strong> the 4E-BP1 peptide at<br />

Ser65 causes only ~ 2-fold decrease in the affinity for <strong>eIF4E</strong> (ΔΔG° less than +1.7 kJ/mol),<br />

and diphosphorylation at both Ser65/Thr70 does not reduce the affinity further. Mono- and<br />

diphosphorylated 4E-BP1 peptides still retain the high ability to interact <strong>with</strong> <strong>eIF4E</strong>, <strong>with</strong><br />

Kas about 5 ⋅ 10 6 M -1 and ΔG° about −38 kJ/mol. Thus, phosphorylation <strong>of</strong> both Ser65 and<br />

Thr70 is insufficient to abolish the <strong>eIF4E</strong> binding.<br />

However, several authors reported substantial reduction <strong>of</strong> the <strong>eIF4E</strong> association<br />

<strong>with</strong> 4E-BP1 phosphorylated at Ser65 in vitro. 93-95 A nearly two orders <strong>of</strong> magnitude<br />

decrease <strong>of</strong> Kas was observed by means <strong>of</strong> SPR due to phosphorylation at Ser65 <strong>of</strong> the full<br />

length 4E-BP1 mutant in which alanine had been substituted at four out <strong>of</strong> the five<br />

Thr/Ser-P sites, leaving Ser65 susceptible to phosphorylation. 94 A value <strong>of</strong> Kas for the<br />

unphosphorylated mutant was determined as 278 ⋅ 10 6 M -1 . ITC yielded a lower Kas value<br />

<strong>of</strong> 67 ⋅ 10 6 M -1 for the full length 4E-BP1, 91 but at this level <strong>of</strong> the equilibrium constants<br />

94


Fluorescence (a. u.)<br />

Fluorescence (a. u.)<br />

Fluorescence (a. u.)<br />

100<br />

90<br />

80<br />

100<br />

(a)<br />

10 -3 10 -2 10 -1 10 0 10 1<br />

90<br />

80<br />

100<br />

(b)<br />

4E-BP1 peptide ( μM)<br />

Figure 4-28. () Titration <strong>of</strong> <strong>eIF4E</strong> <strong>with</strong> (a) the 4E-BP1 peptide, (b) the 4E-BP1 P-Ser65 peptide,<br />

(c) the 4E-BP1 P-Ser65/Thr70 peptide. () Titration <strong>of</strong> the <strong>eIF4E</strong>-m 7 GTP complex <strong>with</strong> the<br />

peptides, as indicated above. 20 °C, excitation wavelength 290 nm, emission wavelength 355 nm.<br />

95<br />

<strong>eIF4E</strong>+m<br />

<strong>eIF4E</strong><br />

7 GTP<br />

<strong>eIF4E</strong>+m<br />

<strong>eIF4E</strong><br />

7 GTP<br />

10 -3 10 -2 10 -1 10 0 10 1<br />

90<br />

80<br />

4E-BP1 P-Ser65 peptide ( μM)<br />

(c)<br />

<strong>eIF4E</strong>+m 7 GTP<br />

<strong>eIF4E</strong><br />

10 -3 10 -2 10 -1 10 0 10 1<br />

4E-BP1 P-Ser65/Thr70 peptide ( μM)


(corresponding to ΔG° about –46 kJ/mol) the 4-fold difference between the Kas values<br />

determined by two different methods is not very significant (ΔΔG° about 3.7 kJ/mol). Even<br />

taking into account the 75-fold drop <strong>of</strong> the association constant to 3.7 ⋅ 10 6 M -1 upon<br />

phosphorylation at Ser65 according to SPR results, 94 the corresponding energy difference<br />

ΔΔG° would be about +10.5 kJ/mol, which is not much in comparison <strong>with</strong> the remaining<br />

binding free energy <strong>of</strong> the monophosphorylated protein (ΔG° about –36.8 kJ/mol).<br />

Discrepancies could partially arise from some experimental problems <strong>with</strong><br />

unphosphorylated 4E-BP1 protein, which lacks folded structure in solution, 239 so the<br />

cysteine residues are not protected against the oxidation. Many times the presence <strong>of</strong> the<br />

4E-BP1 peptide dimers in the stock solution (see 3.2., p. 24) resulted in apparently much<br />

higher values <strong>of</strong> the association constants for the 4E-BP1 peptide than reported in Table 4-<br />

6 (up to 2 orders <strong>of</strong> magnitude, data not shown).<br />

The peptides seem not to be the perfect mimetic models for all properties <strong>of</strong> the full<br />

length protein. However, the previous studies by phosphopeptide mapping demonstrated<br />

that full-length 4E-BP1 monophosphorylated on Ser65 could bind to <strong>eIF4E</strong> efficiently,<br />

while no binding could be detected for 4E-BP1 deleted in the <strong>eIF4E</strong>-binding site. 3 Hence,<br />

the peptides are good models for investigations how single phosphorylation modulates the<br />

intermolecular interactions.<br />

4.3.4. Concluding Remarks<br />

Experiments based on the fluorescence titration studies reconcile the biochemical,<br />

sometimes contradictory, observations that were done hitherto. The results suggest that<br />

cooperation between cap- and eIF4G/4E-BP-binding sites <strong>of</strong> <strong>eIF4E</strong> is possible but hardly<br />

detectable for the short peptides. The <strong>mRNA</strong> cap can be caught more tightly by <strong>eIF4E</strong> in<br />

complex <strong>with</strong> the full length eIF4Gs, and eIF4GI can bind more efficiently to <strong>eIF4E</strong> in<br />

complex <strong>with</strong> RNA.<br />

4E-BP1 monophosphorylated at Ser-65 still possesses the high, submicromolar<br />

affinity for <strong>eIF4E</strong>. Diphosphorylation at Ser-65 and Thr-70 <strong>of</strong> 4E-BP1 has been also<br />

proved insufficient to abrogate 4E-BP1 binding <strong>with</strong> both apo-<strong>eIF4E</strong> and cap-saturated<br />

<strong>eIF4E</strong>. Thus, phosphorylation <strong>of</strong> additional residues, most likely the priming sites Thr37<br />

and Thr46, 92 is required to release 4E-BP1 from <strong>eIF4E</strong>.<br />

96


4.4. <strong>Thermodynamic</strong>s <strong>of</strong> <strong>mRNA</strong> 5' Cap Binding by <strong>eIF4E</strong><br />

Hydrophobic interactions are widely believed to make important contributions to the<br />

stability <strong>of</strong> protein-ligand complexes. A lot <strong>of</strong> studies were devoted to evaluation <strong>of</strong> their<br />

structural and thermodynamic basis for the purpose <strong>of</strong> drug design. Considering that salt<br />

bridges, cation-π stacking, and hydrogen bonds play a dominant role in <strong>mRNA</strong> 5' cap −<br />

<strong>eIF4E</strong> binding 75 , this molecular system is significantly distinct from the hydrophobic ones<br />

which are usually a focus <strong>of</strong> the thermodynamic studies. The cap − <strong>eIF4E</strong> binding is also<br />

accompanied by other processes, i.e. the conformational change <strong>of</strong> the protein and the<br />

solvent effects, like additional water uptake to the complex and partial protonation (see<br />

4.2.3., p. 73).<br />

In an effort to elucidate the structural features <strong>of</strong> <strong>mRNA</strong> 5' cap analogues responsible<br />

for the observed <strong>eIF4E</strong> binding energetics, thermodynamic investigations have been<br />

undertaken. The aim <strong>of</strong> the present chapter is to analyse the specific binding in terms <strong>of</strong><br />

quantitative thermodynamic parameters determined independently by the van't H<strong>of</strong>f<br />

method and isothermal titration calorimetry (ITC). Such quantitative data are <strong>of</strong> primary<br />

importance for rational design <strong>of</strong> new cap analogues <strong>of</strong> potential therapeutic activity since<br />

the high <strong>eIF4E</strong> cellular level is relevant to malignancy and apoptosis 13 .<br />

In this chapter the absolute temperature scale will be used, T (K) = t (°C) + 273.16.<br />

4.4.1. Van't H<strong>of</strong>f <strong>Analysis</strong> <strong>of</strong> Binding <strong>of</strong> <strong>eIF4E</strong> to Cap Analogues<br />

Fluorescence titration experiments were performed to obtain values <strong>of</strong> equilibrium<br />

association constants for nine structurally altered cap analogues, m 7 GMP, m 7 GDP, m 7 GTP,<br />

m3 2,2,7 GTP, bz 7 GTP, p-Cl-bz 7 GTP, m 7 GpppG, m 7 GpppC, and m 7 Gppppm 7 G in the<br />

temperature range <strong>of</strong> 280 – 314 K. Exemplary binding isotherms for <strong>eIF4E</strong> association<br />

<strong>with</strong> m 7 GpppG at six selected temperatures are presented in Fig. 4-29. The equilibrium<br />

association constants (Kas) and the corresponding standard molar free energies <strong>of</strong><br />

formation <strong>of</strong> the complexes (ΔG°) are gathered in Table 4-7.<br />

97


Table 4-7. Association constants (Kas ± SD) and standard molar free energies (ΔG° ± SD) for<br />

binding <strong>of</strong> selected mononucleotide and dinucleotide cap analogues to <strong>eIF4E</strong> at different<br />

temperatures (T), at pH 7.2.<br />

m 7 GMP<br />

T (K)<br />

279.0<br />

280.2<br />

283.9<br />

288.2<br />

293.2<br />

297.9<br />

301.2<br />

304.2<br />

306.3<br />

309.7<br />

313.2<br />

m 7 GDP<br />

T (K)<br />

280.2<br />

283.2<br />

286.2<br />

286.7<br />

287.7<br />

288.7<br />

289.2<br />

293.2<br />

297.8<br />

304.2<br />

309.7<br />

313.2<br />

⋅10 −6 (Μ −1 )<br />

Κ as ⋅10<br />

Y SD<br />

2.734 0.605<br />

2.492 0.201<br />

1.260 0.390<br />

1.430 0.090<br />

0.766 0.090<br />

0.870 0.160<br />

0.780 0.110<br />

0.475 0.157<br />

0.580 0.181<br />

0.473 0.250<br />

0.838 0.328<br />

⋅10 −6 (Μ −1 )<br />

Κ as ⋅10<br />

Y SD<br />

74.33 10.57<br />

74.58 6.75<br />

46.64 4.08<br />

38.25 3.36<br />

43.18 6.05<br />

38.57 4.16<br />

37.50 4.46<br />

20.40 1.54<br />

17.48 3.15<br />

10.48 1.32<br />

5.04 0.74<br />

6.00 1.70<br />

98<br />

ΔG° (kJ/mol)<br />

Y SD<br />

-34.38 0.51<br />

-34.31 0.19<br />

-33.15 0.73<br />

-33.96 0.15<br />

-33.02 0.29<br />

-33.87 0.46<br />

-33.97 0.35<br />

-33.06 0.84<br />

-33.79 0.79<br />

-33.64 1.36<br />

-35.51 1.02<br />

ΔG° (kJ/mol)<br />

Y SD<br />

-42.22 0.33<br />

-42.68 0.21<br />

-42.01 0.21<br />

-41.61 0.21<br />

-42.05 0.34<br />

-41.92 0.26<br />

-41.93 0.29<br />

-41.02 0.18<br />

-41.29 0.45<br />

-40.88 0.32<br />

-39.73 0.38<br />

-40.64 0.74


m 7 GTP<br />

T (K)<br />

281.7<br />

283.6<br />

286.0<br />

288.2<br />

290.6<br />

293.2<br />

295.1<br />

296.5<br />

297.9<br />

299.3<br />

301.2<br />

303.4<br />

305.2<br />

309.0<br />

311.4<br />

313.5<br />

m3 2,2,7 GTP<br />

T (K)<br />

281.3<br />

284.9<br />

286.2<br />

290.2<br />

293.0<br />

296.1<br />

298.6<br />

301.6<br />

304.4<br />

309.7<br />

313.2<br />

Kas ⋅ 10 -6 (M -1 )<br />

Y SD<br />

389.4 45.6<br />

333.5 157.4<br />

352.3 97.1<br />

161.4 35.7<br />

139.7 61.8<br />

116.4 13.0<br />

112.3 22.4<br />

105.3 16.8<br />

54.1 12.1<br />

50.7 10.0<br />

67.7 13.6<br />

44.6 6.8<br />

63.0 8.4<br />

28.5 7.2<br />

17.1 2.5<br />

16.6 9.3<br />

Kas ⋅ 10 -6 (M -1 )<br />

Y SD<br />

0.291 0.155<br />

0.175 0.098<br />

0.155 0.084<br />

0.129 0.110<br />

0.143 0.080<br />

0.112 0.145<br />

0.112 0.100<br />

0.138 0.113<br />

0.126 0.040<br />

0.209 0.087<br />

0.290 0.408<br />

99<br />

ΔG° (kJ/mol)<br />

Y SD<br />

-46.32 0.27<br />

-46.27 1.11<br />

-46.79 0.66<br />

-45.28 0.53<br />

-45.31 1.07<br />

-45.47 0.25<br />

-45.47 0.49<br />

-45.53 0.39<br />

-44.10 0.55<br />

-44.14 0.49<br />

-45.15 0.50<br />

-44.43 0.38<br />

-45.56 0.34<br />

-44.10 0.65<br />

-43.11 0.37<br />

-43.32 1.47<br />

ΔG° (kJ/mol)<br />

Y SD<br />

-29.42 1.24<br />

-28.60 1.32<br />

-28.44 1.29<br />

-28.38 2.07<br />

-28.92 1.36<br />

-28.62 3.18<br />

-28.85 2.23<br />

-29.68 2.04<br />

-29.72 0.80<br />

-31.54 1.07<br />

-32.75 3.67


z 7 GTP<br />

T (K)<br />

281.6<br />

289.1<br />

291.2<br />

293.2<br />

299.2<br />

304.2<br />

309.7<br />

313.2<br />

p-Cl-bz 7 GTP<br />

T (K)<br />

281.4<br />

283.7<br />

285.9<br />

287.4<br />

288.8<br />

290.8<br />

292.0<br />

293.2<br />

297.0<br />

301.6<br />

304.7<br />

308.5<br />

309.1<br />

312.0<br />

313.8<br />

Kas ⋅ 10 -6 (M -1 )<br />

Y SD<br />

58.60 6.80<br />

32.94 3.20<br />

24.20 1.90<br />

17.50 0.50<br />

12.88 0.40<br />

9.88 0.95<br />

7.46 1.03<br />

4.26 0.96<br />

Kas ⋅ 10 -6 (M -1 )<br />

Y SD<br />

91.2 13.2<br />

84.8 12.9<br />

80.0 12.4<br />

113.3 39.6<br />

57.3 4.6<br />

51.5 4.7<br />

85.0 9.8<br />

47.9 7.8<br />

76.6 17.4<br />

29.9 3.8<br />

32.4 4.1<br />

29.7 5.4<br />

16.4 3.5<br />

21.3 4.2<br />

19.7 4.0<br />

100<br />

ΔG° (kJ/mol)<br />

Y SD<br />

-41.87 0.27<br />

-41.60 0.23<br />

-41.16 0.19<br />

-40.65 0.07<br />

-40.72 0.08<br />

-40.74 0.24<br />

-40.74 0.36<br />

-39.75 0.59<br />

ΔG° (kJ/mol)<br />

Y SD<br />

-42.88 0.34<br />

-43.06 0.36<br />

-43.25 0.37<br />

-44.31 0.83<br />

-42.90 0.19<br />

-42.93 0.22<br />

-44.32 0.28<br />

-43.10 0.40<br />

-44.82 0.56<br />

-43.16 0.32<br />

-43.81 0.32<br />

-44.13 0.46<br />

-42.69 0.55<br />

-43.77 0.51<br />

-43.82 0.53


m 7 GpppG<br />

T (K)<br />

m 7 GpppC<br />

280.6<br />

283.9<br />

288.2<br />

293.2<br />

293.2<br />

297.6<br />

301.4<br />

304.9<br />

309.8<br />

312.6<br />

T (K)<br />

280.0<br />

284.3<br />

287.4<br />

290.2<br />

292.8<br />

296.6<br />

299.4<br />

304.0<br />

308.6<br />

313.1<br />

m 7 Gppppm 7 G<br />

T (K)<br />

281.4<br />

283.8<br />

287.8<br />

290.4<br />

292.7<br />

293.2<br />

296.8<br />

302.2<br />

304.9<br />

308.7<br />

312.6<br />

⋅10 −6 (Μ −1 )<br />

Κ as ⋅10<br />

Y SD<br />

24.94 1.05<br />

23.70 4.35<br />

12.94 1.20<br />

7.39 0.46<br />

6.50 0.50<br />

5.13 0.81<br />

4.01 0.61<br />

2.29 0.52<br />

2.41 0.27<br />

2.56 0.22<br />

⋅10 −6 (Μ −1 )<br />

Κ as ⋅10<br />

Y SD<br />

12.25 1.22<br />

8.11 0.82<br />

7.09 0.86<br />

5.16 0.64<br />

2.76 0.48<br />

3.39 0.52<br />

2.44 0.50<br />

2.42 0.43<br />

1.27 0.56<br />

2.74 1.80<br />

micro<br />

Kas ⋅10 -6 (M -1 )<br />

Y SD<br />

125.85 46.68<br />

75.00 20.50<br />

48.70 5.45<br />

35.70 5.20<br />

23.95 1.95<br />

23.48 2.40<br />

20.00 2.10<br />

10.72 1.10<br />

6.62 1.57<br />

5.88 1.07<br />

4.91 0.27<br />

101<br />

ΔG° (kJ/mol)<br />

Y SD<br />

-39.73 0.10<br />

-40.08 0.43<br />

-39.23 0.22<br />

-38.55 0.15<br />

-38.24 0.19<br />

-38.22 0.39<br />

-38.10 0.38<br />

-37.12 0.58<br />

-37.84 0.29<br />

-38.35 0.22<br />

ΔG° (kJ/mol)<br />

Y SD<br />

-37.99 0.23<br />

-37.60 0.24<br />

-37.69 0.29<br />

-37.29 0.30<br />

-36.10 0.42<br />

-37.07 0.38<br />

-36.61 0.51<br />

-37.15 0.45<br />

-36.06 1.13<br />

-38.58 1.71<br />

ΔG° (kJ/mol)<br />

Y SD<br />

-43.63 0.87<br />

-42.78 0.64<br />

-42.35 0.27<br />

-41.98 0.35<br />

-41.34 0.20<br />

-41.37 0.25<br />

-41.48 0.26<br />

-40.67 0.26<br />

-39.81 0.60<br />

-40.00 0.47<br />

-40.04 0.14


Fluorescence (a. u.)<br />

100<br />

80<br />

60<br />

40<br />

10 -3 10 -2 10 -1 10 0 10 1<br />

m 7 GpppG (μM)<br />

Figure 4-29. Quenching <strong>of</strong> <strong>eIF4E</strong> intrinsic fluorescence upon titration <strong>with</strong> 7-methylGpppG.<br />

Increasing fluorescence intensity at higher concentration <strong>of</strong> 7-methylGpppG originates from the<br />

free ligand in solution. Binding isotherms for different temperatures: 312.6 K, 304.9 K, 297.6 K,<br />

293.2 K, 288.2 K, 280.6 K. Titrations were performed in 50 mM Hepes/KOH (pH 7.2), 100 mM<br />

KCl, 1 mM DTT, 0.5 mM EDTA.<br />

The results <strong>of</strong> the fluorescence titrations have been analysed by the van't H<strong>of</strong>f<br />

method. For the symmetrical cap analogue, m 7 Gppppm 7 G, the microscopic association<br />

constant was taken into the analysis. Standard molar enthalpy changes (ΔH°) and standard<br />

molar entropy changes (ΔS°) at 293 K determined from the van’t H<strong>of</strong>f dependencies are<br />

collected in Table 4-8. The association <strong>of</strong> <strong>eIF4E</strong> to <strong>mRNA</strong> 5' cap at 273 K has been found<br />

to be generally enthalpy-driven for the whole series, and entropy-opposed or –driven,<br />

depending on the cap analogue (Fig. 4-30). Strong, specific interactions may be attributed<br />

to fairly high enthalpy <strong>of</strong> association, from −50 to −81 kJ⋅mol -1 , while for the less specific<br />

ligands ΔH° is −17 to −36 kJ⋅mol -1 . One exception among the most tightly binding<br />

analogues is p-Cl-bz 7 GTP that has a small binding enthalpy <strong>of</strong> only –38.5 kJ/mol but a<br />

large positive binding entropy <strong>of</strong> +16.7 J/mol⋅K. The binding entropy <strong>of</strong> the remaining cap<br />

analogues ranges from +40 to −136 J⋅mol -1 ⋅K -1 .<br />

The enthalpy change plays a more pronounced role for binding <strong>of</strong> the analogues <strong>with</strong><br />

longer phosphate chains, which is entropy-opposed. An enthalpy decrease that results from<br />

comparison <strong>of</strong> the binding enthalpies for m 7 GMP and m 7 GDP equals ΔΔH°m7GMPm7GDP =<br />

−25.9 ± 8.4 kJ/mol, and elongation <strong>of</strong> the phosphate chain to three members gives the<br />

further enthalpy decrease <strong>of</strong> ΔΔH°m7GDPm7GTP = −12.4 ± 4.6 kJ/mol. These enthalpy<br />

differences, when divided by the number <strong>of</strong> hydrogen bonds or salt bridges formed by the<br />

102<br />

312.6 K<br />

304.9 K<br />

297.6 K<br />

293.2 K<br />

288.2 K<br />

280.6 K


β and γ phosphate groups (2 and 1, respectively), yield almost equal values <strong>of</strong> the unitary<br />

enthalpy change per one bond, ΔΔH°B ≈ −12.8 ± 9.6 kJ/mol, which is a typical value. 240<br />

Although in general the binding entropy cannot be theoretically decomposed into<br />

individual atom-atom interactions, in this case the entropic penalties, that equal<br />

ΔΔS°m7GMPm7GDP = −61 ± 11 and ΔΔS°m7GDPm7GTP = −30 ± 16 J/mol⋅K, appear to yield<br />

an average <strong>of</strong> ΔΔS°B ≈ −30 ± 19 J/mol⋅K per bond. This observation can serve as an<br />

indicator that binding <strong>of</strong> the subsequent phosphate group is approximately independent<br />

from binding <strong>of</strong> the previous one, i. e. there is no visible effect associated <strong>with</strong> linkage <strong>of</strong><br />

the phosphates, arising from a change in the number <strong>of</strong> degrees <strong>of</strong> freedom.<br />

The binding entropy is more conducive to association <strong>of</strong> the analogues possessing<br />

more or larger substituents at the guanine moiety. For the larger and larger substituents in<br />

the series <strong>of</strong> triphosphates: 7-methyl-, 7-benzyl, 7-para-chloro-benzylGTP, the ΔS° value<br />

increases subsequently by +46 ± 18 and +69 ± 20 J/mol⋅K. This is caused most likely by<br />

expulsion <strong>of</strong> several water molecules from the depth <strong>of</strong> the cap-binding slot into the bulk<br />

solvent, proportionally to the N(7)-substituent volume. The large and electronegative<br />

chloride atom at the N 7 -benzyl ring can attenuate the van der Waals interaction <strong>with</strong> Trp-<br />

166 and destroy the specific water network inside the <strong>eIF4E</strong> cap-binding centre due to<br />

steric and electrostatic effects (Fig. 4-13, p. 67). Consequently, the binding enthalpy for p-<br />

Cl-bz 7 GTP is significantly less negative than that for bz 7 GTP by ~ 18 kJ/mol. The higher<br />

affinity <strong>of</strong> p-Cl- bz 7 GTP for <strong>eIF4E</strong> arises from the more favourable entropy change.<br />

ΔΔH°° (kJ/mol)<br />

50<br />

0<br />

-50<br />

-100<br />

-150<br />

-45 -40 -35 -30<br />

ΔG° (kJ/mol)<br />

Figure 4-30. Enthalpic (ΔH°) and entropic (−TΔS°) contributions to Gibbs free energy (ΔG°) <strong>of</strong><br />

<strong>eIF4E</strong> binding to structurally different cap analogues at 293 K.<br />

103<br />

-TΔ S°° (kJ/mol)<br />

50<br />

0<br />

-50<br />

-100<br />

-150<br />

-45 -40 -35 -30<br />

ΔG° (kJ/mol)


m 7 , 2 , 2<br />

3<br />

In the case <strong>of</strong> GTP , the relative contribution <strong>of</strong> the hydrophobic interactions to<br />

ΔG° increases due to fewer hydrogen bonds stabilizing the complex (see 4.2.2., p. 64) and<br />

much weaker cation-π stacking 241 which itself would provide a large enthalpic component<br />

(discussed below, 4.4.9.3., p. 124). Hence, the thermodynamic driving force for GTP<br />

becomes more entropic.<br />

104<br />

m 7 , 2 , 2<br />

3<br />

Table 4-8. Standard molar enthalpy changes (ΔH°) and entropy changes (ΔS°) at 293 K obtained<br />

from the van't H<strong>of</strong>f isobaric equation for binding <strong>of</strong> <strong>eIF4E</strong> to selected <strong>mRNA</strong> 5' cap analogues.<br />

Standard molar Gibbs free energy changes (ΔG°) calculated from the association constants at 392<br />

K are shown for reference, pH 7.2.<br />

Cap analogue ΔH° ΔS° ΔG°<br />

(kJ·mol -1 ) (J·mol -1 ·K -1 ) (kJ·mol -1 )<br />

m 7 GMP a<br />

m 7 GDP b<br />

m 7 GTP b<br />

bz 7 GTP b<br />

p-Cl-bz 7 GTP b<br />

m3 2,2,7 GTP a<br />

m 7 Gpppp(m 7 G)<br />

m 7 GpppG a<br />

m 7 GpppC a<br />

a, c<br />

-36.0 ± 7.9 -9.3 ± 3.4 -33.15 ± 0.20<br />

-61.9 ± 2.9 -69.8 ± 10.5 -41.031 ± 0.179<br />

-74.3 ± 3.6 -98.7 ± 12.1 -45.109 ± 0.090<br />

-56.5 ± 3.8 -52.7 ± 13.0 -40.669 ± 0.060<br />

-38.5 ± 4.6 +16.7 ± 15.5 -42.94 ± 0.21<br />

-16.6 ± 2.5 +40.3 ± 12.7 -28.94 ± 1.36<br />

-81 ± 54 -136 ± 88 -41.377 ± 0.146<br />

-65 ± 31 -91 ± 58 -38.555 ± 0.152<br />

-50 ± 28 -45 ± 29 -36.42 ± 0.40<br />

a temperature-dependent ΔH° and ΔS° (non-linear van’t H<strong>of</strong>f plot)<br />

b constant ΔH° and ΔS° (deviation <strong>of</strong> van’t H<strong>of</strong>f plot from linearity not detected, P(ν1, ν2) > 0.6)<br />

c the microscopic association constant (Kas micro ) has been taken into account<br />

Careful examination <strong>of</strong> the van’t H<strong>of</strong>f plots revealed that the linear character is lost<br />

for the analogues <strong>of</strong> moderate affinity for <strong>eIF4E</strong> (Fig. 4-31). The thermodynamic<br />

parameters that are related to the non-linearity, i. e. standard heat capacity change under<br />

constant pressure (ΔC°p), and the critical temperatures where ΔH° = 0 and ΔS° = 0 (TH and<br />

TS, respectively) are gathered in Table 4-9. The large<br />

o<br />

Δ C p values resulting from the non-<br />

linearity are surprisingly positive, from +1.66 up to +5.12 kJ⋅mol -1 ⋅K -1 for the m3 2,2,7 GTP -<br />

<strong>eIF4E</strong> association, and TH is higher than TS. The results derived from the non-linear fitting<br />

are statistically unambiguously superior to those derived from the linear one on the<br />

significance level (P) resulting from the Snedecor's F-test, from 0.049 to even less than<br />

0.0001. Involvement <strong>of</strong> the next fitting parameter that was related to a possible linear<br />

dependence <strong>of</strong> ΔC°p on temperature did not render further improvement <strong>of</strong> the results.


ln Kas<br />

ln Kas<br />

22<br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

22<br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

m 7 GTP<br />

m 7 GDP<br />

m 7 GMP<br />

m 7 m<br />

GpppC<br />

7 m<br />

GpppG<br />

7 Gppppm 7 G<br />

3.2 3.3 3.4 3.5 3.6<br />

T -1 ⋅ 10 3 (K -1 )<br />

(a)<br />

(c)<br />

105<br />

bz 7 p-Cl-bz<br />

GTP<br />

7 GTP<br />

2,2,7<br />

m3 GTP<br />

3.2 3.3 3.4 3.5 3.6<br />

T -1 ⋅ 10 3 (K -1 )<br />

(b)<br />

Figure 4-31. Dependence <strong>of</strong> van’t H<strong>of</strong>f<br />

plots for <strong>eIF4E</strong> - cap binding on structural<br />

alterations <strong>of</strong> the cap analogue:<br />

(a) elongation <strong>of</strong> the phosphate chain,<br />

(b) replacement <strong>of</strong> the 7-methyl group for<br />

larger substituents and dimethylation <strong>of</strong> 2amino<br />

group,<br />

(c) change <strong>of</strong> the second base and number <strong>of</strong><br />

the phosphate groups.


Table 4-9. Critical temperatures TH (where ΔH° = 0), TS (ΔS° = 0), and standard molar heat<br />

capacity changes under constant pressure (ΔCp°) obtained from fitting <strong>of</strong> the non-linear van't H<strong>of</strong>f<br />

equation to the equilibrium association constants for selected <strong>mRNA</strong> 5' cap analogues, at pH 7.2.<br />

The relative decrease in sum-<strong>of</strong>-squares (F) for the number <strong>of</strong> degrees <strong>of</strong> freedom (ν2) in<br />

comparison <strong>with</strong> that for the linear model (ν1), and the probability <strong>of</strong> random improvement <strong>of</strong><br />

goodness <strong>of</strong> fit P(ν1, ν2). Calculated Gibbs free energies at TS, ΔG°TS = ΔH°TS, and at TH, ΔG°TH =<br />

−TH⋅ΔS°TH.<br />

Cap analogue TH TS ΔCp° F P(ν1, ν2)<br />

(K) (K) (kJ·mol -1 ·K -1 )<br />

m 7 Gpppp(m 7 G) a 342.0 ± 16.0 318.0 ± 7.8 +1.66 ± 0.57 8.54 0.019<br />

m 7 GpppG 327.1 ± 15.2 307.4 ± 6.0 +1.92 ± 0.93 6.36 0.040<br />

m 7 GpppC 310.0 ± 6.2 297.6 ± 2.1 +2.96 ± 1.25 5.65 0.049<br />

m 7 GMP 306.9 ± 4.8 294.20 ± 1.68 +2.62 ± 0.97 7.34 0.027<br />

m3 2,2,7 GTP 296.41 ± 0.43 290.86 ± 0.69 +5.12 ± 0.48 115 TH, which is most <strong>of</strong>ten<br />

attributed to burial <strong>of</strong> solvent accessible hydrophobic molecular surface upon complex<br />

formation. 131<br />

To confirm validity <strong>of</strong> the exceptional results derived from the van't H<strong>of</strong>f method for<br />

interaction <strong>of</strong> <strong>eIF4E</strong> <strong>with</strong> the cap analogues, direct calorimetric measurements have been<br />

o carried out. The standard enthalpy changes ( Δ Hcal<br />

) have been determined for one selected<br />

cap analogue, m 7 GpppG, at four different temperatures by means <strong>of</strong> Isothermal Titration<br />

Calorimetry (ITC) (Fig. 4-32). In order to control the activity <strong>of</strong> <strong>eIF4E</strong> during the ITC<br />

106


experiment, concurrent fluorometric titrations have been run <strong>with</strong> use <strong>of</strong> the same protein<br />

samples, incubated at a given temperature for the same time. The synchronised<br />

measurements <strong>of</strong> the protein intrinsic fluorescence quenching allowed to determine the<br />

actual concentration <strong>of</strong> the active protein ([Pact]) at each temperature. [Pact] at 288.1, 293.1,<br />

298.4, and 303.2 K was 8.97, 7.42, 3.61, and 5.35 μM, respectively. The enthalpy<br />

estimates for the assumed 100 % protein activity (10 μM), and those corrected for the<br />

active protein concentration determined from fluorometric titrations, are shown in<br />

comparison <strong>with</strong> the van't H<strong>of</strong>f enthalpy change (Fig. 4-33). The corrected calorimetric<br />

enthalpies yield a slope <strong>of</strong><br />

the result <strong>of</strong> the van't H<strong>of</strong>f analysis (<br />

o<br />

Δ C p = +1.966 ± 0.061 kJ⋅mol -1 K -1 (Table 4-10), which confirms<br />

o<br />

Δ C p = +1.92 ± 0.93 kJ⋅mol -1 K -1 ).<br />

Table 4-10. The van't H<strong>of</strong>f and calorimetric thermodynamic parameters for binding <strong>of</strong> m 7 GpppG to<br />

<strong>eIF4E</strong> in Hepes buffer at pH 7.2<br />

ΔH<br />

ΔH<br />

o a<br />

cal<br />

o b<br />

ion<br />

o<br />

cal ion<br />

d<br />

ΔH<br />

−<br />

o<br />

ΔH<br />

vH<br />

o<br />

ΔS e<br />

Temperature (K)<br />

288.1 293.1 298.4 303.2 ΔCp°(kJ⋅mol -1 K -1 )<br />

Enthalpy change (kJ⋅mol -1 )<br />

−65.5 ± 3.2 −54.4 ± 4.0 −47.3 ± 8.4 −35.5 ± 1.7 +1.966 ± 0.061 f<br />

+8.53 +8.66 +8.81 +8.94 +0.0273 ± 0.0003 f<br />

c −74.0 ± 3.2 −63.1 ± 4.0 −56.1 ± 8.4 −44.4 ± 1.7 +1.941 ± 0.059 f<br />

−75 ± 36 −65 ± 31 −56 ± 26 −46 ± 22 +1.92 ± 0.93 g<br />

Entropy change (J⋅mol -1 ⋅K -1 )<br />

−124 ± 71 −91 ± 58 −57 ± 47 −26 ± 40<br />

a<br />

directly measured total enthalpy changes<br />

b o<br />

buffer ionization heats at pH 7.2; estimated δ Δ Hion<br />

= ± 0.09<br />

c o<br />

o<br />

net calorimetric enthalpy changes, Δ Hcal−ion = Δ Hcal<br />

− ΔH<br />

d van't H<strong>of</strong>f enthalpy changes<br />

e entropy changes from the van't H<strong>of</strong>f method<br />

f calculated from linear temperature-dependence <strong>of</strong> corresponding<br />

g calculated from non-linear van't H<strong>of</strong>f dependence<br />

107<br />

o<br />

ion<br />

o<br />

ΔH<br />

x


Power ( μ J⋅ s -1 )<br />

0<br />

-2<br />

-4<br />

-6<br />

-8<br />

-10<br />

-12<br />

-14<br />

0<br />

-2<br />

-4<br />

-6<br />

-8<br />

-10<br />

-12<br />

-14<br />

Time (s)<br />

Time (s)<br />

Figure 4-32. Modified "single injection" ITC curves at 288.1 K (left panels) and 303.2 K (right<br />

panels). Each experiment consisted <strong>of</strong> a main 40 μl injection, preceded by two 1 μl injections<br />

enabling to calculate a correction for the instrumental artifacts, and followed by two 4 μl injections<br />

to check the protein saturation <strong>with</strong> the ligand. Measured heat <strong>of</strong> mixing <strong>of</strong> 7-methylGpppG <strong>with</strong><br />

<strong>eIF4E</strong> solution (—,—), and <strong>of</strong> 7-methylGpppG dilution in buffer (—,—) (upper panels). Calculated<br />

reaction heat <strong>of</strong> binding <strong>of</strong> 7-methylGpppG to <strong>eIF4E</strong> at pH 7.2 in Hepes buffer (—) (bottom<br />

panels).<br />

4.4.3. Protonation Equilibrium <strong>of</strong> m 7 GpppG<br />

o A systematic positive shift <strong>of</strong> the calorimetric enthalpies ( Δ Hcal<br />

) in comparison <strong>with</strong><br />

o their van't H<strong>of</strong>f counterparts ( Δ H vH ) appears from the data (Fig. 4-33, Table 4-10).<br />

Although the numerical uncertainty <strong>of</strong><br />

the difference between<br />

288.1 K<br />

500 1000 1500<br />

o<br />

Δ H vH is about ± 30 kJ⋅mol -1 , the average value <strong>of</strong><br />

o o<br />

Δ Hcal<br />

and Δ H vH <strong>of</strong> about +9.7 kJ⋅mol -1 seems to be well specified.<br />

Differences between the calorimetric and the van't H<strong>of</strong>f enthalpy estimates were the<br />

subject <strong>of</strong> empirical and theoretical analyses for various association processes 244-250 . The<br />

observed discrepancies were ascribed to contributions <strong>of</strong> usually unknown molecular<br />

108<br />

303.2 K<br />

500 1000 1500


Δ H° cal (kJ ⋅mol -1 )<br />

Figure 4-33. Total calorimetric enthalpies corrected for ligand dilution, assuming 100 % protein<br />

activity ( ,------), and those corrected also for active protein concentration, ( ,——). Temperature<br />

dependence <strong>of</strong> the van't H<strong>of</strong>f enthalpy is shown for comparison ( ).<br />

transitions or coupled processes, other than the net complex formation, to<br />

erroneous apparent values <strong>of</strong><br />

0<br />

-20<br />

-40<br />

-60<br />

-80<br />

283 288 293 298 303 308<br />

o<br />

Δ H vH and<br />

Temperature (K)<br />

109<br />

o<br />

p<br />

o<br />

cal<br />

Δ H , and/or<br />

Δ C , arising from the experimental noise. In the<br />

case <strong>of</strong> cap − <strong>eIF4E</strong> binding the latter obscuring effect is eliminated, as testified by the<br />

accordance <strong>of</strong> the van't H<strong>of</strong>f and calorimetric<br />

o<br />

p<br />

Δ C values. This approximately constant<br />

difference can be analysed in terms <strong>of</strong> the protonation equilibria 251 . 7-methylGpppG exists<br />

as a mixture <strong>of</strong> cationic (58%) and zwitterionic (42%) forms at pH 7.2 (pKa = 7.35 ± 0.05<br />

for the N(1)-H proton <strong>of</strong> 7-methylguanine moiety 177 , Scheme 3-1, p. 23). Several<br />

contradictory conclusions were reported regarding the ionic state <strong>of</strong> the <strong>mRNA</strong> 5' cap that<br />

binds to <strong>eIF4E</strong> most tightly. Initially, the zwitterionic form <strong>of</strong> 7-methylguanine was<br />

postulated to interact <strong>with</strong> <strong>eIF4E</strong> 155,156 . In contrast, the crystallographic and NMR<br />

structures revealed the spatial distances suitable for formation <strong>of</strong> a hydrogen bond between<br />

Glu103 and N(1)-H <strong>of</strong> 7-methylguanine 1,48,75,76 , pointing to the cationic form <strong>of</strong> 7-<br />

methylguanine. This hydrogen bond is also present in the ternary 7-methylGDP − <strong>eIF4E</strong> −<br />

eIF4GII peptide complex at pH 8.5 91 . However, the N(1)-H proton has not been directly<br />

observed by NMR 76 . The binding studies as a function <strong>of</strong> pH showed the upward shift <strong>of</strong><br />

pKa <strong>of</strong> N(1)-H inside the <strong>eIF4E</strong> binding site, suggesting that the tightest binding is<br />

accomplished through the cationic form <strong>of</strong> 7-methylguanosine (see 4.2.3.2., p. 83). The<br />

<strong>eIF4E</strong> − cap association that is accompanied by the partial protonation <strong>of</strong> the ligand must<br />

be equilibrated by additional deprotonation <strong>of</strong> the buffer to keep the cation-zwitterion<br />

equilibrium <strong>of</strong> the free ligand at constant pH. As shown in Table 4-10, the calculated


o<br />

ion<br />

contribution <strong>of</strong> the Hepes ionization heat ( Δ H ) to the total reaction heat is in a very good<br />

agreement <strong>with</strong> the difference between<br />

o<br />

cal<br />

Δ H and<br />

110<br />

o<br />

Δ H vH . This demonstrates the linkage<br />

between 7-methylGpppG − <strong>eIF4E</strong> binding and concomitant protonation <strong>of</strong> the residue<br />

which has pKa = 7.35 in the unbound state. Taking into account the net calorimetric<br />

enthalpy changes related to cap binding by <strong>eIF4E</strong>,<br />

o<br />

cal ion<br />

Δ H − = Δ Hcal<br />

− Δ Hion<br />

,<br />

o<br />

o<br />

the calorimetric heat capacity change corrected for buffer ionization is<br />

0.059 kJ⋅mol -1 K -1 .<br />

o<br />

p<br />

Δ C = +1.941 ±<br />

The coupling <strong>of</strong> the binding process <strong>with</strong> the acidic-basic equilibrium <strong>of</strong> the ligand is<br />

nonmandatory 122 , since the zwitterionic form <strong>of</strong> 7-methylGpppG is also able to bind to the<br />

protein via weaker interactions <strong>of</strong> 7-methylG moiety and almost unchanged interactions <strong>of</strong><br />

the phosphate chain. It was shown that the protein-induced shift in the two-state transition<br />

<strong>of</strong> the ligand could contribute to the observed heat capacity change 122 . This contribution<br />

may be either positive or negative, depending on the unknown values <strong>of</strong> the intrinsic<br />

enthalpy changes and equilibrium constants that describe binding <strong>of</strong> the cationic and the<br />

zwitterionic forms <strong>of</strong> 7-methylGpppG to <strong>eIF4E</strong> independently.<br />

4.4.4. Conformational Equilibrium <strong>of</strong> Cap Analogues<br />

Association <strong>of</strong> <strong>eIF4E</strong> <strong>with</strong> the <strong>mRNA</strong> 5' terminus is also directly coupled <strong>with</strong><br />

intramolecular self-stacking <strong>of</strong> the cap. The question arises whether and to how extent this<br />

coupled process influences binding to <strong>eIF4E</strong>. Self-stacking provides for additional<br />

o o enthalpic ( Δ Hsst<br />

) and entropic ( Ssst<br />

o psst<br />

C<br />

Δ ) contributions, and an apparent molar heat capacity<br />

change ( Δ ), resulting from an induced shift in the conformational equilibrium <strong>of</strong> the<br />

ligand upon binding to the protein. 122 As only the unstacked form <strong>of</strong> dinucleotide cap<br />

analogues can penetrate the <strong>eIF4E</strong> cap-binding slot, the coupling is mandatory.<br />

The known thermodynamic parameters describing self-stacking <strong>of</strong> the cationic form<br />

(pH 5.2) and zwitterionic form (pH 9.0) <strong>of</strong> some dinucleotide cap analogues at 293 K 198<br />

gave a possibility to assess the influence <strong>of</strong> this coupled process on the cap − <strong>eIF4E</strong><br />

complex formation. On the basis <strong>of</strong> these parameters, required values <strong>of</strong> equilibrium<br />

constants (1K) and standard molar enthalpy changes (1ΔH°) for self-stacking at pH 7.2 have<br />

been calculated, and next the contributions from self-stacking to the observed<br />

thermodynamic parameters have been determined (Table 4-11). Both<br />

o o<br />

Δ Hsst<br />

and Ssst<br />

Δ for


each dinucleotide analogue are positive and relatively large. They reduce the negative<br />

values <strong>of</strong> the intrinsic thermodynamic parameters ( Δ H , Δ S , Table 4-12) to a substantial<br />

extent, leading to the apparently less negative values <strong>of</strong> the observed ΔH°cal-ion, ΔH°vH and<br />

ΔS°, as reported above (Table 4-8). In case <strong>of</strong> m 7 GpppG, the Δ H and Δ S values change<br />

by 12 – 17 %, and even by 17 – 46 %, respectively, depending on temperature.<br />

Table 4-11. Equilibrium constants (1K) and standard molar enthalpy changes (1ΔH°) for<br />

intramolecular self-stacking <strong>of</strong> dinucleotide cap analogues; and apparent contributions to standard<br />

molar enthalpy (ΔH°sst), entropy (ΔS°sst), and heat capacity (ΔCp°sst) changes <strong>of</strong> <strong>eIF4E</strong> association<br />

<strong>with</strong> cap analogues, resulting from the self-stacking equilibrium shift that is induced by the<br />

mandatory coupling <strong>of</strong> unstacking to the binding, at 293 K.<br />

Cap analogue 1K 1ΔH° ΔH°sst ΔS°sst ΔCp°sst<br />

(kJ·mol -1 ) (kJ·mol -1 ) (J·mol -1 ·K -1 ) (J·mol -1 ·K -1 )<br />

pH 5.2 a<br />

m 7 GpppG 1.40 -15.2 ± 0.5<br />

m 7 GpppC<br />

pH 9.0<br />

0.18 -12.0 ± 0.8<br />

a<br />

m 7 GpppG 1.58 -18.6 ± 0.7<br />

m 7 Gppp(m 7 G)<br />

pH 7.2<br />

0.48 -21.1 ± 0.9<br />

m 7 GpppG b<br />

111<br />

o<br />

0<br />

1.47 -16.6 ± 0.4 +9.89 ± 1.17 +30.5 ± 2.6 -93.2 ± 10.4<br />

m 7 Gppp(m 7 G) c 0.32 -15.8 ± 0.7 +3.85 ± 0.59 +10.81 ± 1.38 -64.5 ± 8.3<br />

m 7 GpppC d<br />

0.18 -12.0 ± 0.8 +1.83 ± 0.44 +4.87 ± 0.82 -26.1 ± 5.8<br />

a data from 198<br />

b values calculated from pKa = 7.35 ± 0.05 177 , 1K and 1ΔH° at pH 5.2 and 9.0<br />

c values estimated from the assumptions: pKa = 7.2 for both ends, no stacking for the double<br />

cationic form, stacking for the zwitterionic-cationic form the same as for the double zwitterionic<br />

form (pH 9.0)<br />

d values estimated from the assumptions: 1K and 1ΔH° the same as for the cationic form (pH 5.2)<br />

Table 4-12. Calculated standard molar enthalpy (ΔH°0), entropy (ΔS°0), and Gibbs free energy<br />

changes (ΔG°0) at 293 K, pH 7.2, for intrinsic binding <strong>of</strong> the unstacked dinucleotide cap analogues<br />

to <strong>eIF4E</strong>, at 293 K, pH 7.2.<br />

Cap analogue ΔH°0 ΔS°0 ΔG°0<br />

(kJ·mol -1 ) (J·mol -1 ·K -1 ) (kJ·mol -1 )<br />

m 7 Gpppp(m 7 G) b<br />

-85 ± 54 -147 ± 88 -42 ± 60<br />

m 7 GpppG -75 ± 31 -122 ± 58 -39 ± 35<br />

m 7 GpppC -52 ± 28 -50 ± 29 -37 ± 29<br />

b microscopic association constant has been taken into account<br />

o<br />

0<br />

o<br />

0<br />

o<br />

0


Table 4-13. Calculated critical temperatures TH 0 and TS 0 (where ΔH°0 = 0 and ΔS°0 = 0,<br />

respectively) and standard molar heat capacity changes under constant pressure (ΔCp°0) for the<br />

intrinsic binding <strong>of</strong> <strong>eIF4E</strong> to the unstacked dinucleotide cap analogues, at pH 7.2.<br />

Cap analogue TH 0 TS 0 ΔCp°0<br />

(K) (K) (kJ·mol -1 ·K -1 )<br />

m 7 Gpppp(m 7 G) a<br />

342 ± 35 319.2 ± 18.5 +1.73 ± 0.57<br />

m 7 GpppG 330 ± 23 311.5 ± 12.5 +2.01 ± 0.93<br />

m 7 GpppC 310.6 ± 11.8 298.1 ± 3.6 +2.99 ± 1.25<br />

a microscopic association constant has been taken into account<br />

While enthalpy and entropy changes <strong>of</strong> binding <strong>of</strong> the dinucleotide cap analogues to<br />

<strong>eIF4E</strong> are pr<strong>of</strong>oundly affected by the coupling between the binding and the unstacking, the<br />

intrinsic binding free energy for the unstacked caps ( Δ G , Table 4-12) are only negligibly<br />

enhanced in comparison <strong>with</strong> the resultant ΔG° values (constant <strong>with</strong>in SD). The negative<br />

o<br />

Δ C psst contributions to the resultant heat capacity changes are very small and almost<br />

temperature-independent. Assuming constant<br />

112<br />

o<br />

0<br />

o<br />

Δ C psst values, they encompass a range from<br />

–26 to −93 J⋅mol -1 K -1 for the three analogues (Table 4-11), and shift negligibly the heat<br />

capacity changes <strong>of</strong> the overall association <strong>with</strong> <strong>eIF4E</strong> from the intrinsic values for the<br />

o<br />

p0<br />

unstacked caps ( Δ C , Table 4-13) to the resultant values reported in Table 4-9. The<br />

critical temperatures TS0 and TH0 are also not significantly changed. Although the<br />

indirectly determined contributions from self-stacking as well as the intrinsic<br />

thermodynamic parameters for <strong>eIF4E</strong> binding to the unstacked cap are charged by<br />

significant uncertainties, they can serve as general indicators whether and how the complex<br />

formation is influenced by the coupled equilibrium <strong>of</strong> cap self-stacking.<br />

The determined small<br />

o<br />

Δ C psst values are very close to a value found as a contribution<br />

due to the coupling <strong>of</strong> adenine base unstacking to binding between single-stranded<br />

dA(pA)34 and the E. coli SSB protein, −62.8 ± 2.5 J⋅mol -1 K -1 per one stack 252 . This<br />

suggests that unstacking <strong>of</strong> the dinucleotide <strong>mRNA</strong> 5' cap analogue, i.e. between 7-<br />

methylguanosine and guanosine moieties linked via the symmetric 5'-5' triphosphate<br />

bridge, can be energetically considered in a similar way as oligodeoxyadenylates<br />

unstacking that accompanies specific binding <strong>of</strong> proteins to single-stranded DNA.


4.4.5. Could the Positive Values <strong>of</strong> ΔCp° Result from <strong>Protein</strong><br />

Deaggregation ?<br />

Two <strong>eIF4E</strong> (28-217) molecules are sticked together through the 4E-BP and eIF4G<br />

binding hydrophobic dorsal surface in the crystallographic asymmetric unit. 75,91 On the<br />

other hand, binding <strong>of</strong> cap forces the <strong>eIF4E</strong> (33-217) deaggregation (see 4.2.3.1.6.1., p.<br />

82). Therefore, some additional non-polar surface area would be exposed to water due to<br />

the binding-induced deaggregation. One could thus suspect that the positive<br />

113<br />

o<br />

p<br />

Δ C is related<br />

only to this side effect. Hence, the question arises whether the possible <strong>eIF4E</strong> (28-217)<br />

aggregation – deaggregation equilibrium coupled to the cap binding could be responsible<br />

for the observed positive heat capacity changes.<br />

The <strong>eIF4E</strong> (28-217) protein was filtered through 100 kD pore size immediately<br />

before experiments, and the probability <strong>of</strong> appearance <strong>of</strong> dimers in a dilute <strong>eIF4E</strong> solution<br />

was minimal. DLS measurements <strong>of</strong> this apo-<strong>eIF4E</strong> (28-217) showed that it was<br />

monodispersed. 253 Moreover, the main argument against this hypothesis is that then the<br />

putative deaggregation-related heat capacity change should be maximal for the strongest<br />

binding analogue (m 7 GTP), for which the deaggregation was the most effective. The<br />

situation is entirely opposite, since m 7 GTP and the other most tightly binding compounds<br />

yield<br />

o<br />

p<br />

Δ C = 0. Hence, it's impossible that the phenomenon which was independently<br />

detected by means <strong>of</strong> both emission spectroscopy and calorimetry at the <strong>eIF4E</strong><br />

concentration <strong>of</strong> ~ 0.1 μM and ~ 10 μM, respectively, could be a trivial one, resulting from<br />

a series <strong>of</strong> systematic experimental mistakes and wrong interpretation.<br />

4.4.6. Enthalpy-Entropy Compensation for Individual Cap Analogues<br />

Having verified the surprisingly positive and unusually high value <strong>of</strong> the heat<br />

capacity change for m 7 GpppG, one can ask about its consequences regarding the<br />

mechanism <strong>of</strong> cap – <strong>eIF4E</strong> interaction. As a result <strong>of</strong> the non-zero values <strong>of</strong><br />

o<br />

p<br />

Δ C , which<br />

are large in comparison <strong>with</strong> ΔS°, temperature-dependent enthalpy-entropy compensation<br />

occurs, i.e. the values <strong>of</strong> both the linear<br />

o<br />

Δ H vH term and the logarithmic TΔS° term in the<br />

expression for ΔG° increase <strong>with</strong> temperature, <strong>with</strong> nearly identical slopes (Fig. 4-<br />

34). 118,196 The large positive<br />

o<br />

p<br />

ΔC values for the <strong>eIF4E</strong> − cap association cause that the<br />

binding is enthalpy-driven at temperatures below TS (Table 4-9), while the entropy is


unfavourable, confirming the importance <strong>of</strong> electrostatic stabilization <strong>of</strong> the complex.<br />

Then, the interaction changes its thermodynamic character to both enthalpy- and entropy-<br />

driven between TS and TH, and finally, above TH the association becomes entropy-driven<br />

and enthalpy-opposed.<br />

Although the ΔG° function attains its maximum at TS, the stabilization <strong>of</strong> the<br />

complex does not decrease rapidly but is still efficient. For the two analogues which are <strong>of</strong><br />

frequent occurrence as natural <strong>mRNA</strong> 5' termini, m 7 GpppG and m 7 GpppC, the maximal<br />

values are about −38 and −37 kJ⋅mol -1 , respectively (Table 4-9). The fact <strong>of</strong> occurrence <strong>of</strong><br />

the non-zero heat capacity changes for the two natural dinucleotide cap analogues has three<br />

salient consequences:<br />

1. the interaction <strong>with</strong> <strong>eIF4E</strong> is both enthalpy and entropy favourable at biological<br />

temperatures (~ 310 K, 37 °C);<br />

2. the affinities <strong>of</strong> the two analogues for <strong>eIF4E</strong> are almost equal at ~ 310 K;<br />

3. the free energies <strong>of</strong> the complexes stabilization are relatively temperature-invariant<br />

over this range.<br />

On the contrary, the results obtained for the mononucleotide triphosphates show that<br />

thermodynamic driving forces make a distinction between the monomethylguanosine cap<br />

(m 7 GTP) and the trimethylguanosine cap ( m GTP<br />

7 , 2 , 2<br />

3<br />

114<br />

) at the biological temperature range.<br />

The association constant <strong>of</strong> the former is ~ 110-fold greater than that <strong>of</strong> the latter at 310 K<br />

(37 °C). The temperature-dependent thermodynamic parameters for m GTP<br />

7 , 2 , 2<br />

3 are ΔH°310K<br />

= +69.6 ± 6.9 kJ/mol and ΔS°310 K = +326 ± 33 J/mol⋅K. It is worth noting that due to the<br />

positive heat capacity change the m GTP<br />

7 , 2 , 2<br />

3 binding enthalpy attains the same absolute<br />

value as that for m 7 GTP but <strong>with</strong> the opposite sign, in spite <strong>of</strong> the fact that both analogues<br />

have the charged triphosphate chains. Hence, elevation <strong>of</strong> temperature at this range causes<br />

an increase <strong>of</strong> the <strong>eIF4E</strong> affinity for m GTP<br />

7 , 2 , 2<br />

3 and an affinity decrease for m 7 GTP. This<br />

phenomenon diminishes the predominance <strong>of</strong> m 7 GTP affinity in relation to m GTP<br />

7 , 2 , 2<br />

value <strong>of</strong> ~ 57 already at 313 K (40 °C).<br />

3 to a


Contributions to ΔΔG°° (kJ⋅⋅mol -1 )<br />

Contributions to ΔΔG o (kJ⋅mol -1 )<br />

20<br />

-30<br />

-40<br />

-50<br />

-60<br />

-70<br />

-80<br />

0<br />

-20<br />

-40<br />

-60<br />

-80<br />

-100<br />

TΔS°<br />

m 7 GpppG<br />

ΔH° cal-ion<br />

m 7 GTP<br />

ΔG o<br />

ΔH o<br />

T S<br />

TΔS o<br />

ΔG°<br />

280 290 300 310 320 330<br />

Temperature (K)<br />

ΔH° vH<br />

280 290 300 310 320 330<br />

Temperature (K)<br />

T H<br />

Figure 4-34. Temperature-dependent enthalpy-entropy compensation that accompanies the binding<br />

<strong>of</strong> m 7 GpppG, m 7 GpppC, and m3 2,2,7 GTP to <strong>eIF4E</strong>. Theoretical non-linear fit (—) to the binding free<br />

energies ΔG° (, , ), the entropic (− − −), and the van't H<strong>of</strong>f enthalpic (----) contributions to<br />

ΔG° as well as the calorimetric enthalpies corrected for ligand dilution, protein activity and buffer<br />

ionization heat ( ) <strong>with</strong> the linear regression (—) are plotted as a function <strong>of</strong> temperature. ΔS° = 0<br />

at TS and ΔH° = 0 at TH. Linear contributions to ΔG° <strong>of</strong> m 7 GTP () are shown for comparison.<br />

115<br />

Contributions to ΔG o (kJ⋅mol -1 )<br />

Contributions to ΔG o (kJ⋅mol -1 )<br />

60<br />

30<br />

0<br />

-30<br />

-60<br />

-90<br />

150<br />

100<br />

50<br />

0<br />

-50<br />

-100<br />

m 7 GpppC<br />

TΔS o<br />

ΔG o<br />

ΔH o<br />

280 290 300 310 320 330<br />

Temperature (K)<br />

m 3 2,2,7 GTP<br />

TΔS o<br />

ΔG o<br />

ΔH o<br />

280 290 300 310 320 330<br />

Temperature (K)


4.4.7. Biological Implications <strong>of</strong> Non-Linear van't H<strong>of</strong>f <strong>Thermodynamic</strong>s<br />

<strong>Thermodynamic</strong> features <strong>of</strong> the process <strong>of</strong> recognition <strong>of</strong> the 5' cap structure by<br />

<strong>eIF4E</strong> can provide an <strong>eIF4E</strong>-dependent switch for several in vivo metabolic pathways, as<br />

follows:<br />

1. The temperature-invariability <strong>of</strong> ΔG° for the natural dinucleotide cap analogues is <strong>of</strong><br />

special biological importance, because the regulation <strong>of</strong> <strong>eIF4E</strong> and its inhibitory<br />

binding protein, 4E-BP1, is affected by heat shock which causes an increase <strong>of</strong> the<br />

association between <strong>eIF4E</strong> and 4E- BP1 81 . Increased binding <strong>of</strong> 4E-BP1 to <strong>eIF4E</strong><br />

interferes <strong>with</strong> the formation <strong>of</strong> the active eIF4F translation initiation complex 31<br />

during heat shock 41,254-256 . As <strong>mRNA</strong>s <strong>of</strong> heat shock proteins are relatively cap-<br />

independent 257,258 , the enhanced inhibition <strong>of</strong> <strong>eIF4E</strong> by 4E-BP1 together <strong>with</strong> the<br />

temperature-invariant cap-binding affinity <strong>of</strong> <strong>eIF4E</strong> could provide a mechanism for<br />

the selective up-regulation <strong>of</strong> the synthesis <strong>of</strong> heat shock proteins.<br />

2. The distinct thermodynamic parameters <strong>of</strong> m 7 GTP and m GTP<br />

7 , 2 , 2<br />

116<br />

3 can shift the<br />

biochemical equilibria related to splicing in eukaryotic cells during heat shock. Two<br />

types <strong>of</strong> capped RNAs exist in eukaryotes, <strong>mRNA</strong>s and small nuclear RNAs<br />

(snRNAs). Their 5' termini are methylated in the nucleus to yield m 7 GpppN<br />

structure. 259 After nuclear export to cytosol, the cap <strong>of</strong> snRNAs is further methylated<br />

the at the amino group <strong>of</strong> the 7-methylguanosine moiety, forming m GpppN<br />

7 , 2 , 2<br />

The trimethylguanosine cap is a fragment <strong>of</strong> a nuclear localization signal (NLS),<br />

which is responsible for import <strong>of</strong> the snRN-protein spliceosomal complexes<br />

(snRNPs) into the nucleus, 23,24 where they take part in pre-<strong>mRNA</strong> splicing. 25 The<br />

trimethylguanosine cap-dependent transport is enhanced by a m GpppN<br />

7 , 2 , 2 -specific<br />

nuclear import receptor, a protein called snurportin1, which binds the trimethyl-<br />

guanosine cap structure approximately three orders <strong>of</strong> magnitude more strongly than<br />

the monomethylguanosine cap. 260 Since the preference <strong>of</strong> <strong>eIF4E</strong> for m 7 GTP vs.<br />

GTP<br />

, 2 , 2<br />

3 is significantly diminuted at higher temperatures, snRNAs could be also<br />

m 7<br />

recruited to the translational machinery, thus contributing to inhibition <strong>of</strong> translation<br />

initiation. Moreover, the binding <strong>of</strong> <strong>eIF4E</strong> to the trimethylguanosine cap <strong>of</strong> snRNPs<br />

prevents recognition <strong>of</strong> this structure by snurportin1, hence the nuclear import and<br />

consequently splicing could be disturbed. This competition could also lead to<br />

significant inhibition <strong>of</strong> protein biosynthesis.<br />

3<br />

3<br />

. 22


3. In some primitive eukaryotes 29,261,262 and in some chordate species 28 the process <strong>of</strong><br />

trans-splicing causes that RNAs contain the trimethylguanosine cap structure. Even<br />

~ 70% <strong>of</strong> the <strong>mRNA</strong> population has the original cap replaced by 3 GpppN in the<br />

117<br />

m 7 , 2 , 2<br />

C. elegans nematode. Five <strong>eIF4E</strong>-like translation initiation factors were found in this<br />

organism, and three <strong>of</strong> them could bind both to m 7 GTP and m GTP<br />

7 , 2 , 2<br />

3<br />

. 55,263,264 The<br />

cellular role for the individual <strong>eIF4E</strong> is<strong>of</strong>orms is still not known, although some <strong>of</strong><br />

them are essential for viability <strong>of</strong> the worm embryos, as shown by RNA<br />

interference. 264<br />

4.4.8. Isothermal Enthalpy-Entropy Compensation in Congener Series<br />

The enthalpic and entropic contributions to the Gibbs free energies <strong>of</strong> binding <strong>of</strong><br />

<strong>eIF4E</strong> to the series <strong>of</strong> the cap analogues at 293.16 K (Table 4-8) appear to satisfy a linear<br />

relationship <strong>with</strong> the high correlation coefficient, R 2 = 0.975 (Fig. 4-35). The slope yields a<br />

compensation temperature, Tc = 399 ± 24 K. The result is exactly the same, irrespective <strong>of</strong><br />

the data set which is chosen for the dinucleotide cap analogues, i. e. either the couples <strong>of</strong><br />

o o the intrinsic ( Δ H 0 , S0<br />

Δ ) or observed (ΔH°, ΔS°) parameters. This means that although the<br />

resultant enthalpy and the entropy <strong>of</strong> interaction are significantly influenced by self-<br />

stacking, the general thermodynamic property <strong>of</strong> recognition <strong>of</strong> <strong>mRNA</strong> 5' cap by <strong>eIF4E</strong> is<br />

insensitive to it.<br />

Enthalpy-entropy compensation is a widely reported phenomenon, which is <strong>of</strong>ten<br />

called an extra-thermodynamic feature <strong>of</strong> complex molecular systems that fluctuate or<br />

interact <strong>with</strong> the aqueous medium. However, several authors made detailed and critical<br />

reviews <strong>of</strong> compensation in congener series, demonstrating that it is in majority a trivial<br />

observation or an artifact that follows immediately from applied experimental<br />

conditions. 199,265-267 There are three categories <strong>of</strong> isothermal enthalpy-entropy<br />

compensation that can be detected in a series <strong>of</strong> homologous compounds:


TΔΔS° (kJ mol -1 )<br />

40<br />

20<br />

Figure 4-35. Isothermal enthalpy-entropy compensation for the <strong>mRNA</strong> 5’ cap congener series at<br />

293 K. The enthalpy gain is always greater than the entropy loss that accompanies the association<br />

<strong>of</strong> more and more tightly binding cap analogues, irrespective <strong>of</strong> whether the binding is described by<br />

a zero () or non-zero () heat capacity change. The linear fitting was performed <strong>with</strong> weighting by<br />

errors <strong>of</strong> both TΔS° and ΔH° (Table 4-8). The errors are not shown for clarity.<br />

1. In case <strong>of</strong> the solvation thermodynamics <strong>of</strong> a solute series, or the binding<br />

thermodynamics <strong>of</strong> a ligand series, when the congeners differ in the number or size<br />

<strong>of</strong> similar substituents, then the linear compensation is a trivial manifestation <strong>of</strong> the<br />

presence <strong>of</strong> a single source <strong>of</strong> additivity. The series <strong>of</strong> the studied nine cap analogues<br />

contains several sources <strong>of</strong> additivity: phosphate groups at the ribose ring, methyl<br />

substituents at the 2-amino group <strong>of</strong> the 7-methylguanine moiety, different aromatic<br />

substituents for the 7-methyl group, and different second nucleosides. Hence, this<br />

case can be ruled out.<br />

2. If the range <strong>of</strong> binding free energy (ΔG°) that can be observed during experiments is<br />

small in relation to the binding enthalpy (ΔH°), due to some intrinsic features <strong>of</strong> an<br />

investigated system or due to experimental conditions, then the linear correlation is<br />

simply a consequence <strong>of</strong> the equation:<br />

ΔH° − TΔS° = ΔG° ≈ constant .<br />

0<br />

-20<br />

-40<br />

-60<br />

m 7 GTP<br />

bz 7 GTP<br />

p-Cl-bz 7 GTP<br />

m 7 GDP<br />

m 7 GpppG<br />

m 7 Gppppm 7 G<br />

-100 -80 -60 -40 -20 0<br />

The ΔG° values <strong>of</strong> the <strong>eIF4E</strong>-cap association are always greater than 50 % <strong>of</strong> the<br />

corresponding ΔH° values (Fig. 4-36), which allows to exclude this case, too.<br />

118<br />

2,2,7<br />

m3 GTP<br />

m 7 GMP<br />

m 7 GpppC<br />

ΔH° (kJ⋅mol -1 )


ΔG° /ΔH°<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

119<br />

Figure 4-36. Proportion <strong>of</strong> ΔG° to<br />

ΔH° for association <strong>of</strong> cap analogues<br />

to <strong>eIF4E</strong> at 293 K<br />

3. True extra-thermodynamic compensation that reflects some additional information<br />

about a system, e. g. about the distribution <strong>of</strong> energy levels available to the system or<br />

about interactions between components, can occur when the two above mentioned<br />

cases were eliminated. The only problem that remains to be resolved consists in an<br />

answer to the question whether enthalpy-entropy compensation does not result<br />

fortuitously from high correlation between errors <strong>of</strong> ΔH° and ΔS° obtained by the<br />

van't H<strong>of</strong>f method.<br />

-45 -40 -35 -30 -25<br />

ΔG° (kJ⋅mol -1 )<br />

A statistical test has been used to determine the significance <strong>of</strong> the observed ΔH° vs.<br />

ΔS° plot for the series <strong>of</strong> the cap analogues. 266,267 The most exact criterion is to check<br />

whether the experimental temperature (Texp) lies outside the 95 % confidence interval for<br />

the compensation temperature: Texp < Tc − 2⋅σ or Tc + 2⋅σ < Texp. The 95 % interval is 351<br />

– 447 K, while the harmonic mean value <strong>of</strong> the experimental temperatures applied for the<br />

van't H<strong>of</strong>f analysis is ~ 297 K, as calculated from Table 4-7. Hence, the difference between<br />

Tc and Texp is even more than 4⋅σ, which unambiguously points to strong, non-trivial,<br />

isothermal enthalpy-entropy compensation.<br />

4.4.8.1. Molecular Interpretation <strong>of</strong> Isothermal Enthalpy-Entropy Compensation<br />

It is <strong>of</strong>ten suggested that enthalpy-entropy compensation is an intrinsic property <strong>of</strong><br />

complex systems that have many s<strong>of</strong>t modes <strong>of</strong> fluctuations, e. g. aqueous solutions and<br />

soluble proteins. A statistical mechanical analysis <strong>of</strong> an entirely general model <strong>of</strong> a<br />

significantly perturbed complex system revealed that the compensation temperature<br />

depended on where the energy <strong>of</strong> the perturbed state (U') lay <strong>with</strong> respect to the mean


energy <strong>of</strong> the unperturbed system (E). 199 However, any clear interpretation for the Tc value<br />

was not proposed.<br />

It seems that Tc can be directly related to the difference in stability <strong>of</strong> the system in<br />

the unperturbed and perturbed states. In other words, the difference between the<br />

compensation temperature and the harmonic mean experimental temperature can be a<br />

measure <strong>of</strong> the fluctuations <strong>of</strong> the unperturbed system, which are modulated by the<br />

perturbation. According as Tc is greater or lower than Texp, the fluctuations are silenced or<br />

enhanced, so the perturbation acts either stabilizing or destabilising, respectively. Thus, the<br />

Tc value <strong>of</strong> 399 ± 24 K shows that the energy (U') <strong>of</strong> the state <strong>of</strong> <strong>eIF4E</strong>, which binds to the<br />

cap structure lies much below the mean energy <strong>of</strong> the apo-protein (E):<br />

E − U' = 9.66 ± 1.7 kJ/mol.<br />

This rather large value, far exceeding RT (~ 2.5 kJ/mol), and almost comparable <strong>with</strong> the<br />

stabilization energy <strong>of</strong> globular proteins <strong>of</strong> similar size and hydrophobicity (~ 50 kJ/mol), 11<br />

suggests that the apo-<strong>eIF4E</strong> is a highly fluctuating, unstable protein, and only the specific<br />

cap binding provides enough stiffening <strong>of</strong> the global protein structure to make it stable on<br />

an usual level. This conclusion is in very good accordance <strong>with</strong> the observations gathered<br />

on other fields:<br />

1. biochemical: the great instability index calculated from the <strong>eIF4E</strong> amino acid<br />

sequence; 44<br />

2. biophysical: extensive conformational changes <strong>of</strong> <strong>eIF4E</strong> upon cap binding; the<br />

significant decrease <strong>of</strong> the amount <strong>of</strong> the active protein even at medium temperature;<br />

aggregation;<br />

3. biological: much lower concentration <strong>of</strong> <strong>eIF4E</strong> than that <strong>of</strong> the other translation<br />

initiation factors; the presence <strong>of</strong> <strong>eIF4E</strong> bound to 4E-BP or to eIF4G <strong>with</strong>in the<br />

eIF4F complex in the living cell; regulation <strong>of</strong> <strong>eIF4E</strong> biological activity by its<br />

cellular accessibility. 42<br />

In summary, the thermodynamics <strong>of</strong> the <strong>eIF4E</strong> association <strong>with</strong> the cap structure<br />

elucidates basic properties <strong>of</strong> this unusual protein, which are key to its biological activity.<br />

120


4.4.9. Inquiries about Origins <strong>of</strong> the Unusual Positive Heat Capacity<br />

Change<br />

Usually observed large, negative values <strong>of</strong> standard heat capacity change are<br />

characteristic for specific hydrophobic binding between proteins and nucleic acids as well<br />

as for protein folding. 131,134,194,196,242,268-271 The main negative contribution to<br />

121<br />

o<br />

p<br />

Δ C comes<br />

from the hydrophobic effect, i.e. the removal <strong>of</strong> non-polar and aromatic molecular surface<br />

from water upon complex formation. 131 On the contrary, the reduction <strong>of</strong> the polar surface<br />

area gives positive contribution to<br />

o<br />

Δ Cp<br />

, although to less extent, 2.3-fold 134 or 1.7-<br />

fold 130,268 , depending on the proposed model. Other contributions to negative<br />

o<br />

ΔCp<br />

comprise changes in s<strong>of</strong>t internal vibrational modes and temperature-dependent<br />

conformational changes which lead to the overall decrease in mobility <strong>of</strong> residues and to<br />

stabilization <strong>of</strong> relative arrangement <strong>of</strong> domains, and/or protein aggregation 131 . For<br />

instance, a large<br />

o<br />

p<br />

Δ C value <strong>of</strong> −1.58 ± 0.06 kJ/mol⋅K for water mediated antigen –<br />

antibody association is completely unrelated to the hydrophobic effect. Instead, this results<br />

only from an interdomain induced fit <strong>of</strong> the heavy and light polypeptide chains, which was<br />

proved by resolving <strong>of</strong> the crystal structures <strong>of</strong> the antibody in the free and antigen-bound<br />

forms. 243<br />

A process <strong>of</strong> protein-ligand interaction characterized itself by<br />

o<br />

p<br />

Δ C = 0 can also give<br />

rise to a non-zero, positive or negative heat capacity change, due to thermodynamic<br />

coupling <strong>with</strong> other possible equilibrium transitions 122,272 : proton uptake or dissociation,<br />

binding <strong>of</strong> the second ligand, a conformational change <strong>of</strong> the protein and/or the ligand.<br />

While the positive<br />

o<br />

p<br />

Δ C relevant to the protein unfolding is a common observation,<br />

examples reported for intermolecular interactions are extremely rare and contain few data,<br />

e.g. the formation <strong>of</strong> the phosph<strong>of</strong>ructokinase tetramer 123 , the interaction <strong>of</strong> the brain<br />

natriuretic peptide <strong>with</strong> heparin 273 , cobalt hexamine and spermidine binding to DNA, 274<br />

anion-exchange adsorption <strong>of</strong> cytochrome b5 and its surface-charge mutants. 275 On the<br />

other hand, it was shown for the interaction <strong>of</strong> c-Myb DNA-binding domain (R2R3) <strong>with</strong><br />

its target DNA, that the heat capacity change can be strongly temperature- and ionic<br />

strength-dependent, leading even to the sign inversion <strong>of</strong><br />

o<br />

p<br />

Δ C <strong>with</strong>in some ranges <strong>of</strong> those<br />

parameters 276 . In each case, Coulombic interactions were involved into the binding<br />

processes.


4.4.9.1. Changes <strong>of</strong> Solvent Accessible Surfaces Upon Binding<br />

The kinetic studies <strong>of</strong> the <strong>eIF4E</strong> − cap interaction by means <strong>of</strong> stopped-flow<br />

fluorescence spectroscopy as well as Brownian molecular dynamics simulations 4 revealed<br />

two-step character <strong>of</strong> the complex formation. The first step is the diffusionally and<br />

electrostatically controlled encounter <strong>of</strong> the protein and the ligand and the second step is<br />

the internal rearrangement <strong>of</strong> the encounter complex. Detailed analysis <strong>of</strong> the salt effect on<br />

the equilibrium association constants revealed that an uptake <strong>of</strong> roughly 65 water<br />

molecules to the first-layer hydration shell is necessary for the complex formation (see<br />

4.2.3.1., p. 74). This large hydration effect is relevant to significant conformational change<br />

<strong>of</strong> the protein upon the second step <strong>of</strong> the cap binding (see 4.2.3.1.6., p. 82), and leads to<br />

an increase <strong>of</strong> the protein solvent accessible surface area. The cooperativity <strong>of</strong> cap-binding<br />

site <strong>with</strong> the eIF4G/4E-BP-binding site (see 4.3.2., p. 90) suggests that the hydrophobic<br />

dorsal surface recognized by the eIF4G and 4E-BP1 peptides becomes more accessible<br />

after the cap has been bound. To estimate the upper limit <strong>of</strong> the positive contribution to<br />

ΔCp° from hydration <strong>of</strong> the protein hydrophobic surface, one should subtract the number <strong>of</strong><br />

water molecules trapped inside the cap-binding centre (at least 16, as detected by<br />

crystallography 1 ) from the total number <strong>of</strong> waters taken up to the complex (ΔN ≈ 65), since<br />

water in internal cavities is usually stiffly bound. 226,227 Since the surface hydrated by one<br />

water molecule is ~ 9.5 Å 2 , 119,232 then, even if one assumes that all remaining 49 water<br />

molecules hydrate a purely aliphatic surface <strong>of</strong> ~ 465 Å 2 (which is unlikely), and if one<br />

takes the maximal increment <strong>of</strong> ΔCp° per 1 Å <strong>of</strong> aliphatic surface (+2.14 J/mol⋅K 130 ), the<br />

o<br />

p<br />

calculated contribution to Δ C would equal only about +1 kJ/mol⋅K.<br />

Some additional contribution to the positive<br />

122<br />

o<br />

p<br />

Δ C arises from the burial <strong>of</strong> the<br />

uncharged and charged polar groups in the protein binding centre. Heat capacity changes<br />

arising from dehydration <strong>of</strong> nucleic acid components were calculated recently on the basis<br />

<strong>of</strong> polar and apolar accessible surface areas as +52.63 J/mol⋅K for guanine, −36.07 J/mol⋅K<br />

for ribose, and +46.02 J/mol⋅K for a single phosphate group. 277 Neglecting neighbour<br />

effects between base, sugar, and phosphate groups in the entire cap analogue, and the<br />

unknown resultant influence <strong>of</strong> the guanosine methylation which provides both the<br />

additional aliphatic group and the positive charge to the ring, the estimated range <strong>of</strong><br />

contribution from burial <strong>of</strong> 7-methylguanosine triphosphate would be about +0.2 kJ/mol⋅K.


4.4.9.2. Electrostatic Contributions to Heat Capacity Changes<br />

Careful inspection <strong>of</strong> the crystallographic 1,48,75 and NMR 76 structural data shows that<br />

stabilization <strong>of</strong> the <strong>mRNA</strong> 5' cap – <strong>eIF4E</strong> complexes is accomplished to great extent by<br />

charge-related interactions and partially complemented by the van der Waals contacts (see<br />

4.2., p. 60). Many charged groups are removed from water upon the association. The<br />

negatively charged 5'-5' phosphate chain <strong>of</strong> 7-methylGpppG is a primary anchor to the<br />

positively charged Arg112, Lys162, Arg157 and Lys159 side-chains <strong>of</strong> <strong>eIF4E</strong>. This<br />

charge-to-charge anchoring makes it possible to form further specific polar contacts in the<br />

narrow binding slot: cation-π sandwich stacking <strong>of</strong> the 7-methylguanine moiety <strong>with</strong><br />

Trp56 and Trp102, and three hydrogen bonds <strong>with</strong> Glu103 and Trp102.<br />

Hydrophobic area-based models <strong>of</strong> the heat capacity changes take into account only<br />

short-rang effects, while there are direct electrostatic contributions to heat capacity, which<br />

do not scale <strong>with</strong> the non-polar and polar surface areas. As shown by the finite-difference<br />

Poisson-Boltzmann method, three components <strong>of</strong> the total electrostatic contribution to<br />

o<br />

p<br />

Δ C are related to: (1) the rearrangement <strong>of</strong> water dipoles upon binding, (2) the<br />

redistribution <strong>of</strong> mobile ions in the solvent upon binding, and (3) the coupling between the<br />

dipolar and ionic terms. 278 The overall contribution <strong>of</strong> electrostatic interactions to<br />

123<br />

o<br />

p<br />

Δ C due<br />

to dehydration <strong>of</strong> charged residues upon protein – ligand binding is dominated by the<br />

positive term arising from the dielectric behaviour <strong>of</strong> water, which is opposed by the<br />

contribution <strong>of</strong> mobile solvent ions. Therefore, in addition to burial <strong>of</strong> polar or polarizable<br />

surface, changes in the water structure that accompany dehydration <strong>of</strong> ionized groups can<br />

also partially account for the observed positive<br />

o<br />

p<br />

Δ C . However, the possible electrostatic<br />

contribution from the phosphate groups and the basic amino acid side-chains is still at least<br />

one order <strong>of</strong> magnitude too small to explain the huge positive heat capacity changes for the<br />

cap analogues.<br />

The cation - π stacking itself has also a great electrostatic component, resulting from<br />

the attraction between the cation and the quadrupole charge distribution <strong>of</strong> the aromatic<br />

ring, which dominates the polarizability and dispersive forces. 129,148,149,279,280 In light <strong>of</strong> the<br />

above mentioned possible explanation <strong>of</strong> the positive<br />

seemed to be very interesting to check whether the unusual values <strong>of</strong><br />

o<br />

p<br />

Δ C by Coulombic contributions, it<br />

stacking <strong>of</strong> tryptophan <strong>with</strong> the non-typical, large, 7-methylguanosine cation.<br />

o<br />

p<br />

Δ C could rise from


4.4.9.3. Tryptophan Stacking <strong>with</strong> Cationic 7-Methylguanosine Moiety<br />

In order to extract information concerning the thermodynamics <strong>of</strong> the cationic 7-<br />

methylguanosine moiety interactions, NMR spectroscopy has been applied to examine two<br />

model systems: 7-methylGMP and tryptophan N-acetylamid, and a dinucleotide cap<br />

analogue, 7-methylGpppG, interacting <strong>with</strong> a dodecapeptide <strong>of</strong> the sequence related to a<br />

part <strong>of</strong> the <strong>eIF4E</strong> cap-binding site around Trp102 (DGIEPMWEDEKN). Stacking between<br />

7-methylG and tryptophan shifts the 1 H signals upfield, yielding microscopic equilibrium<br />

association constants (K), listed for the tryptophan protons at 298 K in Table 4-14,<br />

according to the titration curves shown in Fig. 4-37.<br />

Similar microscopic association constants for the individual tryptophan protons,<br />

related to stacking <strong>with</strong> m 7 GMP suggest that configuration <strong>of</strong> the two heteroaromatic rings<br />

is almost parallel. The dodecapeptide tryptophan interacts only <strong>with</strong> the methylated base <strong>of</strong><br />

m 7 GpppG, and the complex assumes both parallel and perpendicular orientations <strong>of</strong> the 7-<br />

methylguanosine moiety in relation to the indol ring in a dynamic equilibrium. 7,281 The<br />

presence <strong>of</strong> multiple forms <strong>of</strong> the complexes causes that the chemical environment <strong>of</strong> the<br />

tryptophan protons is affected in different manner, which yields the microscopic K for<br />

Table 4-14. Microscopic equilibrium association constants for tryptophan protons, related to<br />

stacking <strong>of</strong> tryptophan N-acetylamid <strong>with</strong> m 7 GMP at pH 5.6, and to binding <strong>of</strong> tryptophancontaining<br />

dodecapeptide <strong>with</strong> m 7 GpppG at pH 5.2; temperature 298 K.<br />

Tryptophan protons: H(2) H(4) H(5) H(6) H(7)<br />

Model system K (M -1 )<br />

Trp(N-aa)-m 7 GMP 15.9 ± 3.8 6.5 ± 1.6 6.7 ± 2.0 5.9 ± 1.8 7.8 ± 1.3<br />

Trp(pept)-m 7 GpppG 242 ± 67 61.8 ± 7.5 51.4 ± 7.2 39.4 ± 9.5 n. d.<br />

a not determined because <strong>of</strong> too weak changes <strong>of</strong> H(7) chemical shift <strong>with</strong> m 7 GpppG concentration<br />

Δδ HTrp (ppm)<br />

0.00<br />

-0.02<br />

-0.04<br />

-0.06<br />

-0.08<br />

H(2)<br />

H(6)<br />

H(5)<br />

H(4)<br />

H(7)<br />

0 5 10 15 20 25 30<br />

m 7 GMP (mM)<br />

Figure 4-37. Differences <strong>of</strong> 1 H chemical shifts <strong>of</strong> tryptophan N-acetylamid at 1 mM due to<br />

stacking upon titration <strong>with</strong> m 7 GMP at pH 5.6 (left); and <strong>of</strong> the dodecapeptide tryptophan at 2.8<br />

mM due to stacking upon titration <strong>with</strong> m 7 GpppG at pH 5.2 (right) at 298 K.<br />

Δδ HTrp (ppm)<br />

124<br />

0.01<br />

0.00<br />

-0.01<br />

-0.02<br />

-0.03<br />

-0.04<br />

0 3 6 9 12 15<br />

m 7 GpppG (mM)


Δδ HTrp (ppm)<br />

Figure 4-38. Temperature dependence <strong>of</strong> differences <strong>of</strong> 1 H chemical shifts <strong>of</strong> tryptophan Nacetylamid<br />

at 1 mM in the presence <strong>of</strong> 29.8 mM m 7 GMP at pH 5.6 (left); and <strong>of</strong> the dodecapeptide<br />

tryptophan at 2.8 mM in the presence <strong>of</strong> 17.2 mM m 7 GpppG at pH 5.2 (right).<br />

H(2) about 5-fold greater than K for the remaining protons, except H(7). The H(7) proton<br />

<strong>of</strong> the dodecapeptide tryptophan exercises a slight downfield shift.<br />

Although stacking is thought <strong>of</strong> as a hydrophobic interaction, binding <strong>of</strong> tryptophan<br />

<strong>with</strong> both cationic 7-methylGMP and 7-methylGpppG is enthalpy-driven and entropy-<br />

opposed, <strong>with</strong>out significant heat capacity changes in the temperature range <strong>of</strong> 278–320 K,<br />

as determined on the basis <strong>of</strong> the temperature-dependent differences <strong>of</strong> 1 H chemical shifts<br />

<strong>of</strong> the tryptophan protons (Fig. 4-38, Table 4-15). Satisfactory goodness <strong>of</strong> fit in terms <strong>of</strong><br />

the sum-<strong>of</strong>-squares <strong>of</strong> deviations (R 2 ) has been obtained when it is assumed that the<br />

stacking is described by the constant values <strong>of</strong> the van't H<strong>of</strong>f enthalpy change (ΔH°vH) and<br />

the entropy change (ΔS°). However, the P-value that refers to the probability <strong>of</strong> random<br />

distribution <strong>of</strong> fitting residuals along the fitted curve is slightly improved if a non-zero<br />

o<br />

Δ Cp<br />

value is allowed for the m 7 GMP-Trp interaction (Fig. 4-39). The heat capacity change<br />

estimated in this way is positive but small, from +0.12 to +0.18 kJ/mol⋅K for different<br />

tryptophan protons, charged by the relative numerical error <strong>of</strong> ~ 100 %. It seems unlikely<br />

that the 7-methylguanosine sandwich stacking between Trp102 and Trp56 <strong>with</strong>in the cap-<br />

binding site <strong>of</strong> <strong>eIF4E</strong> could alone provide a positive contribution to<br />

o<br />

p<br />

explain the observed Δ C values <strong>of</strong> the order <strong>of</strong> +5 kJ/mol⋅K.<br />

R 2 , P value<br />

-0.02<br />

-0.06<br />

-0.10<br />

-0.14<br />

1.0<br />

0.5<br />

0.0<br />

280 290 300 310 320<br />

Temperature (K)<br />

-1000 -500 0 500 1000<br />

ΔC p° (J/mol⋅K)<br />

H(2)<br />

H(6)<br />

H(5)<br />

H(4)<br />

H(7)<br />

125<br />

Δδ HTrp (ppm)<br />

-0.02<br />

-0.04<br />

-0.06<br />

-0.08<br />

280 290 300 310<br />

Temperature (K)<br />

o<br />

p<br />

H(2)<br />

H(7)<br />

H(6)<br />

H(5)<br />

H(4)<br />

Δ C great enough to<br />

Figure 4-39. Goodness <strong>of</strong> fit (R2, ) and<br />

probability <strong>of</strong> random distribution <strong>of</strong><br />

fitting residuals (P value, ) for curves<br />

fitted to temperature dependence <strong>of</strong><br />

ΔδH(7) <strong>of</strong> tryptophan N-acetylamid<br />

stacking <strong>with</strong> m 7 GMP at pH 5.6 <strong>with</strong> fixed<br />

ΔCp° values indicated in figure.


Table 4-15. <strong>Thermodynamic</strong> parameters for stacking <strong>of</strong> <strong>mRNA</strong> 5' cap analogues <strong>with</strong> tryptophan<br />

N-acetylamid and the tryptophan-containing dodecapeptide. The van't H<strong>of</strong>f enthalpy change<br />

(ΔH°vH) and the entropy change (ΔS°) are approximately constant <strong>with</strong> temperature. For reference<br />

the parameters <strong>of</strong> <strong>eIF4E</strong>-m 7 GpppG interaction at 293 K: ΔH°vH = −65 ± 31 kJ/mol, ΔS° = −91 ± 58<br />

J/mol⋅K (Table 4-8)<br />

Tryptophan protons: H(2) H(4) H(5) H(6) H(7)<br />

Model system<br />

Trp(N-aa)-m<br />

ΔH°vH (kJ/mol)<br />

7 GMP<br />

Trp(pept)-m<br />

-26.6 ± 1.8 -25.87 ± 0.78 -26.3 ± 1.4 -25.7 ± 1.1 -26.44 ± 0.79<br />

7 GpppG n. d. a<br />

-29.6 ± 3.2 -33.8 ± 4.1 -31.8 ± 4.6 -35.1 ± 4.1<br />

ΔS° (J/mol⋅K)<br />

Trp(N-aa)-m 7 GMP -65.8 ± 4.6 -64.0 ± 1.9 -65.1 ± 3.4 -62.4 ± 2.7 -64.1 ± 2.0<br />

Trp(pept)-m 7 GpppG n. d. a<br />

-71.3 ± 8.6 -81 ± 12 -75 ± 13 -89 ± 12<br />

a<br />

not determined due to H(2) and H(6) signal overlapping at higher temperatures<br />

The ΔH°vH values for stacking <strong>of</strong> the 7-methylG moiety <strong>with</strong> tryptophan are ~2-fold<br />

greater than those reported for adenine and uracil base stacking (−12.6 ÷ −14.2 kJ/mol per<br />

stack), 252,282,283 and for intramolecular self-stacking <strong>of</strong> the dinucleotide cap analogues 198<br />

(1ΔH°, Table 4-11). The enthalpy and entropy changes <strong>of</strong> the m 7 G-Trp stacking seem to<br />

provide a significant contribution to the overall thermodynamic parameters <strong>of</strong> cap binding<br />

to <strong>eIF4E</strong> upon translation initiation. The 7-methylG moiety represents a unique example <strong>of</strong><br />

a cation which is concurrently a heteroaromatic ring. The large, negative values <strong>of</strong> both<br />

ΔH°vH and ΔS° for stacking <strong>of</strong> 7-methylG <strong>with</strong> the protein tryptophan or <strong>with</strong> the second<br />

base <strong>with</strong>in the dinucleotide cap should be attributed to the dominating Coulombic<br />

character <strong>of</strong> the cation - π interactions. 129,148,149,279,280<br />

Figure 4-40. Sandwich stacking <strong>of</strong> 7-methylG moiety in between two absolutely conserved<br />

tryptophans in the <strong>mRNA</strong> 5’ cap-binding centres <strong>of</strong> murine 75 (left) and yeast 76 (right) translation<br />

initiation factor <strong>eIF4E</strong>. The structure <strong>of</strong> the human complex 48 is identical to the murine one.<br />

126


The results obtained for the model systems point to a diversity <strong>of</strong> permissible spatial<br />

structures <strong>of</strong> the stacked complexes. Similar diversity is observed among the sandwich<br />

configurations in the <strong>eIF4E</strong> cap-binding site. The structures are becoming more parallel in<br />

the course <strong>of</strong> evolution, from primitive eukaryotes like yeast 76 to higher ones, like mouse 75<br />

or human 48 (Fig. 4-40).<br />

4.4.9.4. Coupling <strong>of</strong> Other Intermolecular Equilibria to <strong>eIF4E</strong> – Cap Binding<br />

The possible origin <strong>of</strong> the positive heat capacity changes can be also connected to<br />

coupled equilibrium transitions. If coupled equilibria are occurring and/or significant<br />

concentrations <strong>of</strong> intermediate complexes are present at equilibrium, then measured<br />

association constants may depend on the total concentration <strong>of</strong> protein and/or ligand. 118,284<br />

The question arises why the Kas values for <strong>eIF4E</strong> – cap interactions do not depend on the<br />

concentration range (see 4.1.2., p. 51).<br />

The processes that are coupled to the <strong>eIF4E</strong> – cap binding take place <strong>with</strong><br />

participation <strong>of</strong> environment components. The partial protonation is immediately<br />

equilibrated by buffer ionization (Hepes at 50 mM), and the molar concentration <strong>of</strong> water<br />

(~ 50 M) and potassium cations (0.1 M) is much greater than those <strong>of</strong> the protein and the<br />

ligand (nM to μM range). Self-stacking <strong>of</strong> the dinucleotide cap analogues is weak and the<br />

conformational change is fast.<br />

The intermediate complex life-time can be calculated from the kinetic data obtained<br />

by the stopped-flow method. 4 The encounter rate constant for the association <strong>of</strong> <strong>eIF4E</strong> <strong>with</strong><br />

m 7 GpppG was k+1 = 1.59 ± 0.07 ⋅ 10 8 M -1 ⋅s -1 (at 150 mM KCl), the rate for dissociation <strong>of</strong><br />

the encounter complex was k−1 = 50 ± 8 s -1 , the rate for internal rearrangement to the final<br />

conformation was k+2 = 99 s -1 , and the rate for the reverse process was k−2 = 6 s -1 . Hence,<br />

at typical protein concentration range that was applied for the Kas determination ([Pact]<br />

from 0.01 μM to 1 μM) and at the saturating concentration <strong>of</strong> the ligand, the equilibrium <strong>of</strong><br />

the second step is much more shifted toward the final product than the equilibrium <strong>of</strong> the<br />

first step that leads to the intermediate state:<br />

k + 2 k + 1 ⋅[<br />

Pact<br />

]<br />

>><br />

k k<br />

−2<br />

−1<br />

Additionally, the cap analogue concentration is far from the saturation state during the<br />

whole course <strong>of</strong> the titration, so the inequality is still stronger. Therefore, the only<br />

significant rate-limiting step during the overall association is the diffusionally and<br />

127


electrostatically driven first encounter <strong>of</strong> <strong>eIF4E</strong> and cap, and the observed Kas does not<br />

depend on the protein concentration.<br />

Having removed the doubts regarding the lack <strong>of</strong> influence <strong>of</strong> the coupled equilibria<br />

on the values <strong>of</strong> the association constants one can consider putative contributions to<br />

from the thermodynamic coupling.<br />

128<br />

o<br />

ΔCp<br />

1. Self stacking <strong>of</strong> dinucleotide cap analogues is mandatory coupled to <strong>eIF4E</strong><br />

association <strong>with</strong> cap, and yields a negligible negative contribution to the observed<br />

o<br />

p<br />

Δ C (see 4.4.4., p. 110).<br />

2. Partial protonation <strong>of</strong> the N(1) position <strong>of</strong> the cap analogue, the potassium cation<br />

release from the phosphate chain, and uptake <strong>of</strong> ~ 65 water that accompanies<br />

conformational changes <strong>of</strong> the protein upon cap binding (see 4.2.3.1., p. 74)<br />

represent the case <strong>of</strong> non-mandatory coupling, and hence can provide positive<br />

o<br />

p<br />

contributions to the overall Δ C .<br />

In summary, the following effects can together yield the positive<br />

several kJ/mol⋅K: 131<br />

• the hydration <strong>of</strong> the hydrophobic dorsal surface (up to ~ +1 kJ/mol⋅K),<br />

o<br />

p<br />

Δ C at the level <strong>of</strong><br />

• the dehydration <strong>of</strong> the ligand and protein ionized groups in the binding site (up to ~ 0.4<br />

kJ/mol⋅K for cation-π stacking, and up to ~ 0.5 kJ/mol⋅K for dehydration <strong>of</strong> the other<br />

charged and uncharged polar surfaces),<br />

• and the thermodynamic coupling <strong>with</strong> accompanying equilibrium transitions<br />

A kind <strong>of</strong> difficulty in interpretation is related to the lack <strong>of</strong> the apo-<strong>eIF4E</strong> structure,<br />

which is necessary for the complete analysis.<br />

4.4.10. Linear Correlation between ΔCp° and ΔG°<br />

The values <strong>of</strong><br />

o<br />

Δ C p appear to correlate linearly <strong>with</strong> the intrinsic free energy <strong>of</strong><br />

the binding (ΔG°0): the stronger the binding the less positive<br />

o<br />

Δ C p (Fig. 4-41). The<br />

apparent correlation can be fortuitous but it can also consist in any causality, as the<br />

correlation coefficient (r 2 ) equals 0.91. The cap analogues <strong>of</strong> the highest binding affinity,<br />

e.g. 7-methylGTP, do not reveal any observable curvature <strong>of</strong> the van't H<strong>of</strong>f plot.<br />

Systematic ΔC°p dependence on the cap affinity for <strong>eIF4E</strong> suggests that the common<br />

positive contribution can be compensated to different extent by the negative contribution


elated to the binding specificity. Most probably, this can result from tightening <strong>of</strong> the<br />

protein global fold upon binding, that yields the negative contribution to<br />

129<br />

o<br />

p<br />

Δ C , similarly as<br />

in the case <strong>of</strong> the specific antigen-antibody association. 131,243 This hypothesis is supported<br />

by the presence <strong>of</strong> the non-trivial enthalpy-entropy compensation discussed above. It was<br />

shown that the preferential ligand binding to the lower microstates <strong>of</strong> a protein led to a<br />

shift in the distribution <strong>of</strong> the microstates, and, consequently, reduced width <strong>of</strong> the<br />

distribution implied a decrease in heat capacity upon ligand binding. 122 Thus, for the most<br />

specific cap analogues the negative stabilization effect related to the specificity <strong>of</strong> the<br />

binding can make the<br />

experimental data. 248<br />

ΔCp°° (kJ/mol⋅K)<br />

6<br />

3<br />

0<br />

-3<br />

-6<br />

-9<br />

o<br />

p<br />

Δ C value too small to be discerned <strong>with</strong>in the noise <strong>of</strong> the<br />

-70 -60 -50 -40 -30 -20 -10 0 10<br />

ΔG° (kJ/mol)<br />

<strong>eIF4E</strong> + <strong>mRNA</strong> 5' cap analogues<br />

DNA + intercalators<br />

L-isoleucine tRNA ligase + substrates<br />

serine proteases + Na +<br />

protein + DNA (specific)<br />

protein + DNA (nonspecific)<br />

Figure 4-41. Correlation <strong>of</strong> heat capacity changes (ΔC°p) <strong>with</strong> standard molar binding free energies<br />

(ΔG°) for different intermolecular interactions.<br />

Linear ΔG° - ΔCp° correlation is rarely found, since it requires systematic studies on<br />

a consistent ligand series, while the suitable data obtained by time- and protein-consuming<br />

thermodynamic measurements are rather scattered. Several available examples are<br />

collected in Fig. 4-41. A general tendency is common but only two enthalpy driven<br />

processes involving charged ligands, i. e. specific sodium cation binding by serine<br />

proteases 285 and the <strong>eIF4E</strong> – cap binding yield remarkable slopes (0.93 ± 0.14 K -1 , and<br />

0.275 ± 0.044 K -1 , respectively). Binding <strong>of</strong> the L-isoleucine tRNA ligase to L-isoleucine<br />

and L-valine is not accompanied by any conformational changes <strong>of</strong> the large (120 kD),<br />

multidomain protein, which is known from the crystal structures <strong>of</strong> the native enzyme and


its complexes. 286 The heat capacity changes for binding <strong>of</strong> various substrates to the enzyme<br />

are approximately constant, though the range <strong>of</strong> the corresponding free energy changes<br />

encompasses almost 20 kJ/mol. 287 Drug - DNA intercalation is related partially to the<br />

binding-induced changes in non-polar and polar solvent-accessible surface areas, and<br />

consequently the binding free energy is proportional to the hydrophobic effect (slight<br />

dependence). 120,121,288 Specific interactions <strong>of</strong> DNA <strong>with</strong> repressors and RNA polymerase,<br />

driven by hydrophobicity <strong>of</strong> the specific sites, are characterized by the wide-spread protein<br />

conformational changes and large negative heat capacity changes, while non-specific<br />

interactions, driven only by the counterion release, are accomplished <strong>with</strong><br />

117,196,289-293<br />

130<br />

o<br />

Δ C p = 0. 112-<br />

However, no systematic literature data for the ΔG°-dependent positive heat capacity<br />

changes are available. Searching for origins <strong>of</strong> the positive<br />

o<br />

Δ C p and for the causal<br />

explanation <strong>of</strong> the apparent linear relationship in terms <strong>of</strong> statistical physics will be the<br />

subject <strong>of</strong> further investigations.<br />

4.4.11. Discussion <strong>with</strong> Other Authors<br />

The results described herein are contradictory to the first conclusions drawn from<br />

studies <strong>of</strong> 7-methylGTP and 7-methylGpppG binding to <strong>eIF4E</strong> from human<br />

erythrocytes 156 . Although the human 49 and murine 50 proteins are almost identical (98%<br />

homology for the full length proteins, and 100% for truncated (28-217) <strong>eIF4E</strong>), those<br />

results suggested that the <strong>eIF4E</strong> − cap binding was entropy-driven and enthalpy-opposed,<br />

<strong>with</strong> the constant thermodynamic parameters in the 278-308 K temperature range (ΔS° =<br />

+219 ± 11 J⋅mol -1 K -1 , ΔH° = +33.9 ± 1.7 kJ⋅mol -1 for 7-methylGpppG 156 ). However, the<br />

human protein was purified by means <strong>of</strong> the cap-affinity chromatography, what could<br />

result in up to 60% <strong>of</strong> the unremovable cap analogue bound to <strong>eIF4E</strong>. 1,208 This purification<br />

method yielded the apparent association constants up to 300-fold lower 5 then those<br />

measured for <strong>eIF4E</strong> purified <strong>with</strong>out contact <strong>with</strong> cap 1 (see Table 4-2, p. 61). Moreover,<br />

neither the decreasing temperature-dependence <strong>of</strong> the human <strong>eIF4E</strong> activity nor<br />

fluorescence <strong>of</strong> the cap analogues were taken into account. The authors also reported the<br />

fluorescence intensity <strong>of</strong> <strong>eIF4E</strong> and the analogues as invariant over the temperature range<br />

studied, which is impossible. Taken together, these main reasons could lead to<br />

misinterpretation <strong>of</strong> the experimental data.


The same reservation but even to much greater extent concerns the work <strong>of</strong> Shen et<br />

al. 294 Apparently higher association constants for dinucleotides than that for 7-methylGTP<br />

results simply from the inner filter effect, as no corrections were applied in spite <strong>of</strong> the<br />

huge concentration <strong>of</strong> <strong>eIF4E</strong> (10 μM) and the cap (22 μM). <strong>Thermodynamic</strong> parameters<br />

derived in such manner are fortuitous and do not allow a reasonable analysis.<br />

Recently, the human <strong>eIF4E</strong> − 7-methylGpppG binding affinity was fluorometrically<br />

studied in the context <strong>of</strong> cell growth suppression, and the association constant <strong>of</strong> 0.83 ±<br />

0.14 μM -1 was obtained <strong>with</strong> use <strong>of</strong> some corrections (at 296.2 K, pH 7.5, 300 mM<br />

KCl). 295 The corresponding Kas value reported herein for truncated murine <strong>eIF4E</strong> (28-217)<br />

is 5.13 ± 0.81 μM -1 (at 297.6 K, pH 7.2, 100 mM KCl, Table 4-7, p. 101). These results are<br />

in an excellent agreement, since elevation <strong>of</strong> KCl concentration from 100 to 300 mM<br />

causes a significant, ~ 6-fold decrease <strong>of</strong> the association constant due to screening <strong>of</strong> the<br />

electrostatic attraction between the basic amino acids in the cap-binding site <strong>of</strong> <strong>eIF4E</strong> and<br />

the phosphate chain <strong>of</strong> 7-methylGpppG (see 4.2.3.1., p. 74).<br />

4.4.12. Conclusions<br />

The positive heat capacity change at constant pressure appears to be a characteristic<br />

feature <strong>of</strong> <strong>eIF4E</strong> binding to <strong>mRNA</strong> 5' cap. The chemical cap analogues <strong>of</strong> the highest<br />

specificity exhibit the heat capacity change proportionally shifted toward less positive or<br />

undetectable values. Data obtained from two independent methods, fluorescence<br />

quenching and isothermal titration calorimetry provided excellently accordant<br />

thermodynamic parameters. The microcalorimetric results additionally yielded a<br />

quantitative confirmation <strong>of</strong> partial protonation <strong>of</strong> the ligand, which is necessary to form<br />

one <strong>of</strong> the crucial hydrogen bonds in the protein binding centre.<br />

Isothermal enthalpy-entropy compensation among the cap analogues <strong>of</strong> different<br />

specificity for <strong>eIF4E</strong> at 293 K points to several thermodynamic features: the dominating<br />

enthalpic character for the whole congener series, a great instability <strong>of</strong> the apo-form <strong>of</strong><br />

<strong>eIF4E</strong>, and to conformational rearrangement <strong>of</strong> the whole protein upon the complex<br />

formation, leading to the final, stable, cap-bound state.<br />

The exceptionally large positive<br />

131<br />

o Δ Cp<br />

<strong>of</strong> the binding can be attributed to<br />

simultaneous interplay <strong>of</strong> various intermolecular processes: the extensive additional<br />

hydration <strong>of</strong> the <strong>eIF4E</strong> hydrophobic dorsal surface, dehydration <strong>of</strong> the ionized groups,


including the 7-methylguanosine cation, and the burial <strong>of</strong> polar groups <strong>of</strong> the interacting<br />

molecules, as well as to the thermodynamically coupled differential ligand protonation and<br />

the potassium cation release from the ligand. The negative contribution cancelling the<br />

positive one for the most tightly binding analogues can arise from the global protein<br />

conformational change that stiffen the complex structure. The induced shift in the self-<br />

stacking equilibrium <strong>of</strong> the dinucleotide cap analogues gives a negligible negative<br />

o<br />

p<br />

contribution to the overall Δ C .<br />

Because <strong>of</strong> the non-zero heat capacity changes, the interactions <strong>of</strong> <strong>eIF4E</strong> <strong>with</strong> natural<br />

dinucleotide cap analogues are both enthalpy- and entropy-driven over the range <strong>of</strong><br />

biological temperatures, and stability <strong>of</strong> the <strong>eIF4E</strong> − cap complex is almost temperature-<br />

independent.<br />

The general description <strong>of</strong> the cap – <strong>eIF4E</strong> association is intricate due to strict<br />

synergy <strong>of</strong> charge-related and hydrophobic interactions <strong>with</strong>in the binding site and on the<br />

solvent accessible molecular surface <strong>of</strong> the complex. These properties make this molecular<br />

system unique in the thermodynamic sense.<br />

132


5. Summary<br />

The thesis was aimed at revealing the molecular mechanism <strong>of</strong> recognition <strong>of</strong> <strong>mRNA</strong><br />

5' cap structure by <strong>eIF4E</strong>, previously limited to a static view from the crystal structures.<br />

Chemically modified cap-analogues proved to be a valuable tool to probe contribution <strong>of</strong><br />

various types <strong>of</strong> molecular interaction to the <strong>eIF4E</strong>-cap complex stability via emission<br />

spectroscopy measurements. A fast and accurate method <strong>of</strong> synchronized fluorescence<br />

titration has been designed to obtain reliable equilibrium association constant values,<br />

which are not only relative to one another, but for the first time have an absolute meaning,<br />

and thus can be interpreted in terms <strong>of</strong> the free energy <strong>of</strong> binding. The mechanism <strong>of</strong> the<br />

complex formation at physiological pH can be now presented as an interplay <strong>of</strong> several<br />

intermolecular processes (Scheme 5-1). Recognition <strong>of</strong> the 5' cap structure by <strong>eIF4E</strong><br />

binding site begins <strong>with</strong> electrostatic attraction <strong>of</strong> the cap phosphate chain, from which one<br />

potassium cation is being removed. In the initial encounter complex ([<strong>eIF4E</strong>•cap]*) the<br />

cap is first anchored by the phosphate chain and next, cooperative cation - π sandwich<br />

stacking together <strong>with</strong> hydrogen bonding <strong>of</strong> 7-methylguanine occurs. Differential<br />

protonation at the N(1) position <strong>of</strong> 7-methylguanine is necessary to enhance stacking,<br />

which in turn conditions formation <strong>of</strong> three hydrogen bonds. The second step <strong>of</strong><br />

association is accompanied by preferential hydration involving ~ 65 water molecules.<br />

Some <strong>of</strong> them are trapped inside the cap-binding slot, while several tens supply the first-<br />

layer hydration shell <strong>of</strong> the complex. This extensive water reorganization implies a<br />

significant conformational change <strong>of</strong> the whole protein. The latter conclusion comes<br />

independently <strong>of</strong> several other sets <strong>of</strong> experimental results: isothermal enthalpy-entropy<br />

compensation, deaggregation <strong>of</strong> <strong>eIF4E</strong> forced by association <strong>with</strong> cap, cooperativity <strong>of</strong><br />

cap-binding and eIF4G/4E-BP1-binding sites.<br />

- 1⋅ K + + 0.5 ⋅ H +<br />

<strong>eIF4E</strong> + cap [<strong>eIF4E</strong> cap]* <strong>eIF4E</strong> cap<br />

133<br />

+ 65 ⋅ H2O<br />

Scheme 5-1. Proposed model <strong>of</strong> two-step association <strong>of</strong> <strong>eIF4E</strong> <strong>with</strong> the <strong>mRNA</strong> 5' cap


The appropriate equilibrium equation for <strong>eIF4E</strong>-cap binding has the following form:<br />

K<br />

t<br />

[ <strong>eIF4E</strong><br />

• cap][<br />

K ]<br />

=<br />

65 +<br />

[ <strong>eIF4E</strong>]<br />

[ cap]<br />

[ H O]<br />

[ H ]<br />

0<br />

0<br />

2<br />

+<br />

0.<br />

5<br />

(<strong>with</strong>out terms for buffer; index "0" denotes equilibrium concentrations <strong>of</strong> free species).<br />

Comparison <strong>of</strong> the <strong>eIF4E</strong>-binding affinity <strong>of</strong> natural vs structurally modified cap<br />

analogues together <strong>with</strong> the structures <strong>of</strong> the <strong>eIF4E</strong>-cap complexes makes a foundation to<br />

rational design <strong>of</strong> new cap analogues <strong>with</strong> better inhibitory properties than those<br />

synthesized hitherto. Quantitative analysis <strong>of</strong> the binding free energy shows that the cap<br />

triphosphate chain contributes in one half to it. Since the negative charge at the cap<br />

phosphate chain is so crucial for the efficient binding to <strong>eIF4E</strong>, it should be indispensably<br />

included in the designed agents. Hence, thinking about inhibition <strong>of</strong> the <strong>eIF4E</strong> cellular<br />

activity by 5' cap analogues must involve an active transport <strong>of</strong> the charged species through<br />

the cellular bilayer phospholipid membrane. One hopeful analogue is the p-Ch-bz7GTP,<br />

which strongly binds to <strong>eIF4E</strong> <strong>with</strong> the affinity <strong>of</strong> 10 7 – 10 8 M -1 . The equilibrium binding<br />

constant is relatively weakly temperature-dependent, since the association is both enthalpy-<br />

and entropy-driven. This artificial cap analogue should not be easily metabolized or<br />

trapped by other proteins, and thus it can be expected that its half lifetime could be long<br />

enough to render significant translation inhibition. This hope is supported by the<br />

observation that the overall inhibition constant (KI) <strong>of</strong> p-Ch-bz7GTP determined from<br />

biological in vitro assay in a rabbit reticulocyte lizate 68 is shifted toward lower values in<br />

comparison <strong>with</strong> that resulting from the average correlation for all cap analogues (Fig. 4-<br />

11, p. 63).<br />

The thermodynamic studies revealed unique, large and positive heat capacity<br />

changes that accompany the <strong>eIF4E</strong> – cap interactions. The binding <strong>of</strong> natural dinucleotide<br />

cap analogues are both enthalpy- and entropy-driven over the range <strong>of</strong> biological<br />

temperatures, and stability <strong>of</strong> the <strong>eIF4E</strong> − cap complex is almost temperature-independent.<br />

These findings are suitable for further development <strong>of</strong> quantitative interpretation <strong>of</strong><br />

intermolecular recognition specificity, and are also a kind <strong>of</strong> challenge for theoretical<br />

interpretation.<br />

The author hopes that the presented thesis contribute to more pr<strong>of</strong>ound insights into<br />

how <strong>eIF4E</strong> interacts <strong>with</strong> other components <strong>of</strong> the cytoplasmic machinery responsible for<br />

cap-dependent translation initiation.<br />

134


6. References<br />

1. Niedzwiecka,A., J.Marcotrigiano, J.Stepinski, M.Jankowska-Anyszka, A.Wyslouch-Cieszynska, M.Dadlez,<br />

A.C.Gingras, P.Mak, E.Darzynkiewicz, N.Sonenberg, S.K.Burley, and R.Stolarski. (2002). Biophysical<br />

studies <strong>of</strong> <strong>eIF4E</strong> cap-binding protein: recognition <strong>of</strong> <strong>mRNA</strong> 5' cap structure and synthetic fragments <strong>of</strong> eIF4G<br />

and 4E-BP1 proteins. J. Mol. Biol., 319, 615-635.<br />

2. Niedzwiecka,A., J.Stepinski, E.Darzynkiewicz, N.Sonenberg, and R.Stolarski. (2002). Positive heat capacity<br />

change upon specific binding <strong>of</strong> translation initiation factor <strong>eIF4E</strong> to <strong>mRNA</strong> 5' cap. Biochemistry, 41, 12140-<br />

12148.<br />

3. Gingras,A.C., B.Raught, S.P.Gygi, A.Niedzwiecka, M.Miron, S.K.Burley, R.D.Polakiewicz, A.Wyslouch-<br />

Cieszynska, R.Aebersold, and N.Sonenberg. (2001). Hierarchical phosphorylation <strong>of</strong> the translation inhibitor<br />

4E-BP1. Genes Dev., 15, 2852-2864.<br />

4. Blachut-Okrasinska,E., E.Bojarska, A.Niedzwiecka, L.Chlebicka, E.Darzynkiewicz, R.Stolarski, J.Stepinski, and<br />

J.M.Antosiewicz. (2000). Stopped-flow and Brownian dynamics studies <strong>of</strong> electrostatic effects in the kinetics<br />

<strong>of</strong> binding <strong>of</strong> 7-methyl-GpppG to the protein <strong>eIF4E</strong>. Eur. Biophys. J., 29, 487-498.<br />

5. Wieczorek,Z., A.Niedzwiecka-Kornas, L.Chlebicka, M.Jankowska, K.Kiraga, J.Stepinski, M.Dadlez, R.Drabent,<br />

E.Darzynkiewicz, and R.Stolarski. (1999). Fluorescence studies on association <strong>of</strong> human translation initiation<br />

factor <strong>eIF4E</strong> <strong>with</strong> <strong>mRNA</strong> cap-analogues. Z. Naturforsch. [C. ], 54, 278-284.<br />

6. Niedzwiecka-Kornas,A., L.Chlebicka, J.Stepinski, M.Jankowska-Anyszka, Z.Wieczorek, E.Darzynkiewicz,<br />

R.E.Rhoads, and R.Stolarski. (1999). Spectroscopic studies on association <strong>of</strong> <strong>mRNA</strong> cap-analogues <strong>with</strong><br />

human translation factor <strong>eIF4E</strong>. From modelling <strong>of</strong> interactions to inhibitory properties. Collect. Symp. Ser.,<br />

2, 214-218.<br />

7. Niedzwiecka-Kornas,A., R.Przedmojski, L.Balaspiri, Z.Wieczorek, J.Stepinski, M.Jankowska, H.Lonnberg,<br />

E.Darzynkiewicz, and R.Stolarski. (1999). Studies on association <strong>of</strong> <strong>mRNA</strong> cap-analogues <strong>with</strong> a synthetic<br />

dodecapeptide DGIEPMWEDEKN. Nucleosides & Nucleotides, 18, 1105-1106.<br />

8. Crick,F. (1970). Central dogma <strong>of</strong> molecular biology. Nature, 227, 561-563.<br />

9. Raven, P. H. and Johnson, G. B. (1996). Biology. WCB Publishers.<br />

10. Proudfoot,N.J., A.Furger, and M.J.Dye. (2002). Integrating <strong>mRNA</strong> processing <strong>with</strong> transcription. Cell, 108, 501-<br />

512.<br />

11. Creighton, T. E. (1993). <strong>Protein</strong>s: structures and molecular properties. W. H. Freeman and Co., New York.<br />

12. Sonenberg,N., M.A.Morgan, W.C.Merrick, and A.J.Shatkin. (1978). A polypeptide in eukaryotic initiation factors<br />

that crosslinks specifically to the 5'-terminal cap in <strong>mRNA</strong>. Proc. Natl. Acad. Sci. U. S. A, 75, 4843-4847.<br />

13. Gingras,A.C., B.Raught, and N.Sonenberg. (1999). eIF4 initiation factors: effectors <strong>of</strong> <strong>mRNA</strong> recruitment to<br />

ribosomes and regulators <strong>of</strong> translation. Annu. Rev. Biochem., 68, 913-963.<br />

14. Muthukrishnan,S., G.W.Both, Y.Furuichi, and A.J.Shatkin. (1975). 5'-Terminal 7-methylguanosine in eukaryotic<br />

<strong>mRNA</strong> is required for translation. Nature, 255, 33-37.<br />

15. Hershey, J. W. and Merrick, W. C. (2000). The pathway and mechanism <strong>of</strong> initiation <strong>of</strong> protein synthesis.<br />

Sonenberg, N, Hershey, J. W., and Mathews, M. B. [39], 33-88, CSHL Press, Cold Spring Harbor, NY.<br />

Translational control <strong>of</strong> gene expression.<br />

16. Konarska,M.M., R.A.Padgett, and P.A.Sharp. (1984). Recognition <strong>of</strong> cap structure in splicing in vitro <strong>of</strong> <strong>mRNA</strong><br />

precursors. Cell, 38, 731-736.<br />

17. Izaurralde,E., J.Lewis, C.McGuigan, M.Jankowska, E.Darzynkiewicz, and I.W.Mattaj. (1994). A nuclear cap<br />

binding protein complex involved in pre-<strong>mRNA</strong> splicing. Cell, 78, 657-668.<br />

135


18. Izaurralde,E., J.Lewis, C.Gamberi, A.Jarmolowski, C.McGuigan, and I.W.Mattaj. (1995). A cap-binding protein<br />

complex mediating U snRNA export. Nature, 376, 709-712.<br />

19. Izaurralde,E., J.Stepinski, E.Darzynkiewicz, and I.W.Mattaj. (1992). A cap binding protein that may mediate<br />

nuclear export <strong>of</strong> RNA polymerase II-transcribed RNAs. J. Cell Biol., 118, 1287-1295.<br />

20. Lewis,J.D. and E.Izaurralde. (1997). The role <strong>of</strong> the cap structure in RNA processing and nuclear export. Eur. J.<br />

Biochem., 247, 461-469.<br />

21. Sonenberg,N. (1988). Cap-binding proteins <strong>of</strong> eukaryotic messenger RNA: functions in initiation and control <strong>of</strong><br />

translation. Prog. Nucleic Acid Res. Mol. Biol., 35, 173-207.<br />

22. Mattaj,I.W. (1986). Cap trimethylation <strong>of</strong> U snRNA is cytoplasmic and dependent on U snRNP protein binding.<br />

Cell, 46, 905-911.<br />

23. Hamm,J., E.Darzynkiewicz, S.M.Tahara, and I.W.Mattaj. (1990). The trimethylguanosine cap structure <strong>of</strong> U1<br />

snRNA is a component <strong>of</strong> a bipartite nuclear targeting signal. Cell, 62, 569-577.<br />

24. Gorlich,D. and I.W.Mattaj. (1996). Nucleocytoplasmic transport. Science, 271, 1513-1518.<br />

25. Sharp,P.A. (1994). Split genes and RNA splicing. Cell, 77, 805-815.<br />

26. Eschenfeldt,W.H., B.G.Cohen, and R.E.Rhoads. (1983). Structure <strong>of</strong> the 5' terminus <strong>of</strong> hen oviduct lysozyme<br />

messenger ribonucleic acid. J. Biol. Chem., 258, 13076-13081.<br />

27. Maroney,P.A., J.A.Denker, E.Darzynkiewicz, R.Laneve, and T.W.Nilsen. (1995). Most <strong>mRNA</strong>s in the nematode<br />

Ascaris lumbricoides are trans-spliced: a role for spliced leader addition in translational efficiency. RNA., 1,<br />

714-723.<br />

28. Vandenberghe,A.E., T.H.Meedel, and K.E.Hastings. (2001). <strong>mRNA</strong> 5'-leader trans-splicing in the chordates.<br />

Genes Dev., 15, 294-303.<br />

29. Van Doren,K. and D.Hirsh. (1990). <strong>mRNA</strong>s that mature through trans-splicing in Caenorhabditis elegans have a<br />

trimethylguanosine cap at their 5' termini. Mol. Cell Biol., 10, 1769-1772.<br />

30. Sachs,A.B., P.Sarnow, and M.W.Hentze. (1997). Starting at the beginning, middle, and end: translation initiation<br />

in eukaryotes. Cell, 89, 831-838.<br />

31. Sonenberg,N. (1996). <strong>mRNA</strong> 5' cap-binding protein <strong>eIF4E</strong> and control <strong>of</strong> cell growth. In Translational Control<br />

(J.W.B.Hershey, M.B.Mathews, and N.Sonenberg, editors), pp.245-69, Cold Spring Harbor Laboratory, Cold<br />

Spring Harbor, New York.<br />

32. Pause,A., N.Methot, Y.Svitkin, W.C.Merrick, and N.Sonenberg. (1994). Dominant negative mutants <strong>of</strong><br />

mammalian translation initiation factor eIF-4A define a critical role for eIF-4F in cap-dependent and capindependent<br />

initiation <strong>of</strong> translation. EMBO J., 13, 1205-1215.<br />

33. Hentze,M.W. (1997). eIF4G: a multipurpose ribosome adapter? Science, 275, 500-501.<br />

34. Imataka,H., A.Gradi, and N.Sonenberg. (1998). A newly identified N-terminal amino acid sequence <strong>of</strong> human<br />

eIF4G binds poly(A)-binding protein and functions in poly(A)-dependent translation. EMBO J., 17, 7480-<br />

7489.<br />

35. Gradi,A., H.Imataka, Y.V.Svitkin, E.Rom, B.Raught, S.Morino, and N.Sonenberg. (1998). A novel functional<br />

human eukaryotic translation initiation factor 4G. Mol. Cell Biol., 18, 334-342.<br />

36. Wei,C.C., M.L.Balasta, J.Ren, and D.J.Goss. (1998). Wheat germ poly(A) binding protein enhances the binding<br />

affinity <strong>of</strong> eukaryotic initiation factor 4F and (iso)4F for cap analogues. Biochemistry, 37, 1910-1916.<br />

37. Lejbkowicz,F., C.Goyer, A.Darveau, S.Neron, R.Lemieux, and N.Sonenberg. (1992). A fraction <strong>of</strong> the <strong>mRNA</strong> 5'<br />

cap-binding protein, eukaryotic initiation factor 4E, localizes to the nucleus. Proc. Natl. Acad. Sci. U. S. A, 89,<br />

9612-9616.<br />

136


38. Strudwick,S. and K.L.Borden. (2002). The emerging roles <strong>of</strong> translation factor <strong>eIF4E</strong> in the nucleus.<br />

Differentiation, 70, 10-22.<br />

39. Dostie,J., F.Lejbkowicz, and N.Sonenberg. (2000). Nuclear eukaryotic initiation factor 4E (<strong>eIF4E</strong>) colocalizes<br />

<strong>with</strong> splicing factors in speckles. J. Cell Biol., 148, 239-247.<br />

40. Dostie,J., M.Ferraiuolo, A.Pause, S.A.Adam, and N.Sonenberg. (2000). A novel shuttling protein, 4E-T, mediates<br />

the nuclear import <strong>of</strong> the <strong>mRNA</strong> 5' cap-binding protein, <strong>eIF4E</strong>. EMBO J., 19, 3142-3156.<br />

41. Duncan,R., S.C.Milburn, and J.W.Hershey. (1987). Regulated phosphorylation and low abundance <strong>of</strong> HeLa cell<br />

initiation factor eIF-4F suggest a role in translational control. Heat shock effects on eIF-4F. J. Biol. Chem.,<br />

262, 380-388.<br />

42. Raught,B. and A.C.Gingras. (1999). <strong>eIF4E</strong> activity is regulated at multiple levels. Int. J. Biochem. Cell Biol., 31,<br />

43-57.<br />

43. Jones,R.M., J.Branda, K.A.Johnston, M.Polymenis, M.Gadd, A.Rustgi, L.Callanan, and E.V.Schmidt. (1996). An<br />

essential E box in the promoter <strong>of</strong> the gene encoding the <strong>mRNA</strong> cap- binding protein (eukaryotic initiation<br />

factor 4E) is a target for activation by c-myc. Mol. Cell Biol., 16, 4754-4764.<br />

44. Guruprasad,K., B.V.Reddy, and M.W.Pandit. (1990). Correlation between stability <strong>of</strong> a protein and its dipeptide<br />

composition: a novel approach for predicting in vivo stability <strong>of</strong> a protein from its primary sequence. <strong>Protein</strong><br />

Eng, 4, 155-161.<br />

45. Altmann,M., N.Schmitz, C.Berset, and H.Trachsel. (1997). A novel inhibitor <strong>of</strong> cap-dependent translation<br />

initiation in yeast: p20 competes <strong>with</strong> eIF4G for binding to <strong>eIF4E</strong>. EMBO J., 16, 1114-1121.<br />

46. Kabsch,W. and C.Sander. (1983). Dictionary <strong>of</strong> protein secondary structure: pattern recognition <strong>of</strong> hydrogenbonded<br />

and geometrical features. Biopolymers, 22, 2577-2637.<br />

47. Rost,B. and C.Sander. (1994). Conservation and prediction <strong>of</strong> solvent accessibility in protein families. <strong>Protein</strong>s,<br />

20, 216-226.<br />

48. Tomoo,K., X.Shen, K.Okabe, Y.Nozoe, S.Fukuhara, S.Morino, T.Ishida, T.Taniguchi, H.Hasegawa,<br />

A.Terashima, M.Sasaki, Y.Katsuya, K.Kitamura, H.Miyoshi, M.Ishikawa, and K.Miura. (2002). Crystal<br />

structures <strong>of</strong> 7-methylguanosine 5'-triphosphate (m(7)GTP)- and P(1)-7-methylguanosine-P(3)-adenosine-<br />

5',5'-triphosphate (m(7)GpppA)- bound human full-length eukaryotic initiation factor 4E: biological<br />

importance <strong>of</strong> the C-terminal flexible region. Biochem. J., 362, 539-544.<br />

49. Rychlik,W., L.L.Domier, P.R.Gardner, G.M.Hellmann, and R.E.Rhoads. (1987). Amino acid sequence <strong>of</strong> the<br />

<strong>mRNA</strong> cap-binding protein from human tissues. Proc. Natl. Acad. Sci. U. S. A, 84, 945-949.<br />

50. Altmann,M., P.P.Muller, J.Pelletier, N.Sonenberg, and H.Trachsel. (1989). A mammalian translation initiation<br />

factor can substitute for its yeast homologue in vivo. J. Biol. Chem., 264, 12145-12147.<br />

51. Rychlik,W. and R.E.Rhoads. (1992). Nucleotide sequence <strong>of</strong> rabbit eIF-4E cDNA. Nucleic Acids Res., 20, 6415.<br />

52. Wakiyama,M., M.Saigoh, K.Shiokawa, and K.Miura. (1995). <strong>mRNA</strong> encoding the translation initiation factor<br />

eIF-4E is expressed early in Xenopus embryogenesis. FEBS Lett., 360, 191-193.<br />

53. Dyer,J.R., A.M.Pepio, S.K.Yanow, and W.S.Sossin. (1998). Phosphorylation <strong>of</strong> <strong>eIF4E</strong> at a conserved serine in<br />

Aplysia. J. Biol. Chem., 273, 29469-29474.<br />

54. Hernandez,G. and J.M.Sierra. (1995). Translation initiation factor eIF-4E from Drosophila: cDNA sequence and<br />

expression <strong>of</strong> the gene. Biochim. Biophys. Acta, 1261, 427-431.<br />

55. Jankowska-Anyszka,M., B.J.Lamphear, E.J.Aamodt, T.Harrington, E.Darzynkiewicz, R.Stolarski, and<br />

R.E.Rhoads. (1998). Multiple is<strong>of</strong>orms <strong>of</strong> eukaryotic protein synthesis initiation factor 4E in Caenorhabditis<br />

elegans can distinguish between mono- and trimethylated <strong>mRNA</strong> cap structures. J. Biol. Chem., 273, 10538-<br />

10542.<br />

56. Manjunath,S., A.J.Williams, and J.Bailey-Serres. (1999). Oxygen deprivation stimulates Ca2+-mediated<br />

phosphorylation <strong>of</strong> <strong>mRNA</strong> cap- binding protein <strong>eIF4E</strong> in maize roots. Plant J., 19, 21-30.<br />

137


57. Aliyeva,E., A.M.Metz, and K.S.Browning. (1996). Sequences <strong>of</strong> two expressed sequence tags (EST) from rice<br />

encoding different cap-binding proteins. Gene, 180, 221-223.<br />

58. Metz,A.M., R.T.Timmer, and K.S.Browning. (1992). Isolation and sequence <strong>of</strong> a cDNA encoding the cap binding<br />

protein <strong>of</strong> wheat eukaryotic protein synthesis initiation factor 4F. Nucleic Acids Res., 20, 4096.<br />

59. Rodriguez,C.M., M.A.Freire, C.Camilleri, and C.Robaglia. (1998). The Arabidopsis thaliana cDNAs coding for<br />

<strong>eIF4E</strong> and eIF(iso)4E are not functionally equivalent for yeast complementation and are differentially<br />

expressed during plant development. Plant J., 13, 465-473.<br />

60. Ono N. and Sudoh M. (2003). Candida glabrata <strong>eIF4E</strong> gene. EMBL/GenBank/DDBJ databases,<br />

61. Altmann,M., C.Handschin, and H.Trachsel. (1987). <strong>mRNA</strong> cap-binding protein: cloning <strong>of</strong> the gene encoding<br />

protein synthesis initiation factor eIF-4E from Saccharomyces cerevisiae. Mol. Cell Biol., 7, 998-1003.<br />

62. Ptushkina,M., I.Fierro-Monti, J.van den Heuvel, S.Vasilescu, R.Birkenhager, K.Mita, and J.E.McCarthy. (1996).<br />

Schizosaccharomyces pombe has a novel eukaryotic initiation factor 4F complex containing a cap-binding<br />

protein <strong>with</strong> the human <strong>eIF4E</strong> C- terminal motif KSGST. J. Biol. Chem., 271, 32818-32824.<br />

63. Thompson,J.D., D.G.Higgins, and T.J.Gibson. (1994). CLUSTAL W: improving the sensitivity <strong>of</strong> progressive<br />

multiple sequence alignment through sequence weighting, position-specific gap penalties and weight matrix<br />

choice. Nucleic Acids Res., 22, 4673-4680.<br />

64. Risler,J.L., M.O.Delorme, H.Delacroix, and A.Henaut. (1988). Amino acid substitutions in structurally related<br />

proteins. A pattern recognition approach. Determination <strong>of</strong> a new and efficient scoring matrix. J. Mol. Biol.,<br />

204, 1019-1029.<br />

65. Joshi,B., A.L.Cai, B.D.Keiper, W.B.Minich, R.Mendez, C.M.Beach, J.Stepinski, R.Stolarski, E.Darzynkiewicz,<br />

and R.E.Rhoads. (1995). Phosphorylation <strong>of</strong> eukaryotic protein synthesis initiation factor 4E at Ser-209. J.<br />

Biol. Chem., 270, 14597-14603.<br />

66. Miyamoto,S., S.R.Kimball, and B.Safer. (2000). Signal transduction pathways that contribute to increased protein<br />

synthesis during T-cell activation. Biochim. Biophys. Acta, 1494, 28-42.<br />

67. Pause,A., G.J.Belsham, A.C.Gingras, O.Donze, T.A.Lin, J.C.Lawrence, Jr., and N.Sonenberg. (1994). Insulindependent<br />

stimulation <strong>of</strong> protein synthesis by phosphorylation <strong>of</strong> a regulator <strong>of</strong> 5'-cap function. Nature, 371,<br />

762-767.<br />

68. Whalen,S.G., A.C.Gingras, L.Amankwa, S.Mader, P.E.Branton, R.Aebersold, and N.Sonenberg. (1996).<br />

Phosphorylation <strong>of</strong> eIF-4E on serine 209 by protein kinase C is inhibited by the translational repressors, 4Ebinding<br />

proteins. J. Biol. Chem., 271, 11831-11837.<br />

69. Duncan,R.F., H.Peterson, C.H.Hagedorn, and A.Sevanian. (2003). Oxidative stress increases eukaryotic initiation<br />

factor 4E phosphorylation in vascular cells. Biochem. J., 369, 213-225.<br />

70. Minich,W.B., M.L.Balasta, D.J.Goss, and R.E.Rhoads. (1994). Chromatographic resolution <strong>of</strong> in vivo<br />

phosphorylated and nonphosphorylated eukaryotic translation initiation factor eIF-4E: increased cap affinity<br />

<strong>of</strong> the phosphorylated form. Proc. Natl. Acad. Sci. U. S. A, 91, 7668-7672.<br />

71. Morley,S.J. and S.Naegele. (2002). Phosphorylation <strong>of</strong> eukaryotic initiation factor (eIF) 4E is not required for de<br />

novo protein synthesis following recovery from hypertonic stress in human kidney cells. J. Biol. Chem., 277,<br />

32855-32859.<br />

72. Scheper,G.C. and C.G.Proud. (2002). Does phosphorylation <strong>of</strong> the cap-binding protein <strong>eIF4E</strong> play a role in<br />

translation initiation? Eur. J. Biochem., 269, 5350-5359.<br />

73. Scheper,G.C., B.van Kollenburg, J.Hu, Y.Luo, D.J.Goss, and C.G.Proud. (2002). Phosphorylation <strong>of</strong> Eukaryotic<br />

Initiation Factor 4E Markedly Reduces Its Affinity for Capped <strong>mRNA</strong>. J. Biol. Chem., 277, 3303-3309.<br />

74. Zuberek,J., A.Wyslouch-Cieszynska, A.Niedzwiecka, M.Dadlez, J.Stepinski, W.Augustyniak, A.C.Gingras,<br />

Z.Zhang, S.K.Burley, N.Sonenberg, R.Stolarski, and E.Darzynkiewicz. (2003). Phosphorylation <strong>of</strong> <strong>eIF4E</strong><br />

attenuates its interaction <strong>with</strong> <strong>mRNA</strong> 5' cap analogs by electrostatic repulsion: Intein-mediated protein<br />

ligation strategy to obtain phosphorylated protein. RNA, 9, 52-61.<br />

138


75. Marcotrigiano,J., A.C.Gingras, N.Sonenberg, and S.K.Burley. (1997). Cocrystal structure <strong>of</strong> the messenger RNA<br />

5' cap-binding protein (<strong>eIF4E</strong>) bound to 7-methyl-GDP. Cell, 89, 951-961.<br />

76. Matsuo,H., H.Li, A.M.McGuire, C.M.Fletcher, A.C.Gingras, N.Sonenberg, and G.Wagner. (1997). Structure <strong>of</strong><br />

translation factor <strong>eIF4E</strong> bound to m7GDP and interaction <strong>with</strong> 4E-binding protein. Nat. Struct. Biol., 4, 717-<br />

724.<br />

77. Mathews, M. B., Sonenberg, N, and Hershey, J. W. (2000). Origins and principles <strong>of</strong> translational control.<br />

Sonenberg, N, Hershey, J. W., and Mathews, M. B. [39], 1-31, CSHL Press, Cold Spring Harbor, NY.<br />

Translational control <strong>of</strong> gene expression.<br />

78. Tang,S.J., G.Reis, H.Kang, A.C.Gingras, N.Sonenberg, and E.M.Schuman. (2002). A rapamycin-sensitive<br />

signaling pathway contributes to long-term synaptic plasticity in the hippocampus. Proc. Natl. Acad. Sci. U. S.<br />

A, 99, 467-472.<br />

79. Martin,K.C., A.Casadio, H.Zhu, E.Yaping, J.C.Rose, M.Chen, C.H.Bailey, and E.R.Kandel. (1997). Synapsespecific,<br />

long-term facilitation <strong>of</strong> aplysia sensory to motor synapses: a function for local protein synthesis in<br />

memory storage. Cell, 91, 927-938.<br />

80. Casadio,A., K.C.Martin, M.Giustetto, H.Zhu, M.Chen, D.Bartsch, C.H.Bailey, and E.R.Kandel. (1999). A<br />

transient, neuron-wide form <strong>of</strong> CREB-mediated long-term facilitation can be stabilized at specific synapses by<br />

local protein synthesis. Cell, 99 , 221-237.<br />

81. Vries,R.G., A.Flynn, J.C.Patel, X.Wang, R.M.Denton, and C.G.Proud. (1997). Heat shock increases the<br />

association <strong>of</strong> binding protein-1 <strong>with</strong> initiation factor 4E. J. Biol. Chem., 272, 32779-32784.<br />

82. Gingras,A.C., Y.Svitkin, G.J.Belsham, A.Pause, and N.Sonenberg. (1996). Activation <strong>of</strong> the translational<br />

suppressor 4E-BP1 following infection <strong>with</strong> encephalomyocarditis virus and poliovirus. Proc. Natl. Acad. Sci.<br />

U. S. A, 93, 5578-5583.<br />

83. Ohlmann,T., D.Prevot, D.Decimo, F.Roux, J.Garin, S.J.Morley, and J.L.Darlix. (2002). In vitro cleavage <strong>of</strong><br />

eIF4GI but not eIF4GII by HIV-1 protease and its effects on translation in the rabbit reticulocyte lysate<br />

system. J. Mol. Biol., 318, 9-20.<br />

84. De Benedetti,A. and A.L.Harris. (1999). <strong>eIF4E</strong> expression in tumors: its possible role in progression <strong>of</strong><br />

malignancies. Int. J. Biochem. Cell Biol., 31, 59-72.<br />

85. Sorrells,D.L., G.E.Ghali, C.Meschonat, R.J.DeFatta, D.Black, L.Liu, A.De Benedetti, C.A.Nathan, and B.D.Li.<br />

(1999). Competitive PCR to detect <strong>eIF4E</strong> gene amplification in head and neck cancer. Head Neck, 21, 60-65.<br />

86. DeFatta,R.J., E.A.Turbat-Herrera, B.D.Li, W.Anderson, and A.De Benedetti. (1999). Elevated expression <strong>of</strong><br />

<strong>eIF4E</strong> in confined early breast cancer lesions: possible role <strong>of</strong> hypoxia. Int. J. Cancer, 80, 516-522.<br />

87. Hershey, J. W. and Miyamoto, S. (2000). Translational control and cancer. Sonenberg, N, Hershey, J. W., and<br />

Mathews, M. B. [39], 637-654, CSHL Press, Cold Spring Harbor, NY. Translational control <strong>of</strong> gene<br />

expression.<br />

88. Haghighat,A., S.Mader, A.Pause, and N.Sonenberg. (1995). Repression <strong>of</strong> cap-dependent translation by 4Ebinding<br />

protein 1: competition <strong>with</strong> p220 for binding to eukaryotic initiation factor-4E. EMBO J., 14, 5701-<br />

5709.<br />

89. Mader,S., H.Lee, A.Pause, and N.Sonenberg. (1995). The translation initiation factor eIF-4E binds to a common<br />

motif shared by the translation factor eIF-4 gamma and the translational repressors 4E-binding proteins. Mol.<br />

Cell Biol., 15, 4990-4997.<br />

90. Poulin,F., A.C.Gingras, H.Olsen, S.Chevalier, and N.Sonenberg. (1998). 4E-BP3, a new member <strong>of</strong> the<br />

eukaryotic initiation factor 4E-binding protein family. J. Biol. Chem., 273, 14002-14007.<br />

91. Marcotrigiano,J., A.C.Gingras, N.Sonenberg, and S.K.Burley. (1999). Cap-dependent translation initiation in<br />

eukaryotes is regulated by a molecular mimic <strong>of</strong> eIF4G. Mol. Cell, 3, 707-716.<br />

139


92. Gingras,A.C., S.P.Gygi, B.Raught, R.D.Polakiewicz, R.T.Abraham, M.F.Hoekstra, R.Aebersold, and<br />

N.Sonenberg. (1999). Regulation <strong>of</strong> 4E-BP1 phosphorylation: a novel two-step mechanism. Genes Dev., 13,<br />

1422-1437.<br />

93. Mothe-Satney,I., G.J.Brunn, L.P.McMahon, C.T.Capaldo, R.T.Abraham, and J.C.Lawrence, Jr. (2000).<br />

Mammalian target <strong>of</strong> rapamycin-dependent phosphorylation <strong>of</strong> PHAS-I in four (S/T)P sites detected by<br />

phospho-specific antibodies. J. Biol. Chem., 275 , 33836-33843.<br />

94. Karim,M.M., J.M.Hughes, J.Warwicker, G.C.Scheper, C.G.Proud, and J.E.McCarthy. (2001). A quantitative<br />

molecular model for modulation <strong>of</strong> mammalian translation by the <strong>eIF4E</strong>-binding protein 1. J. Biol. Chem.,<br />

95. Mothe-Satney,I., D.Yang, P.Fadden, T.A.Haystead, and J.C.Lawrence, Jr. (2000). Multiple mechanisms control<br />

phosphorylation <strong>of</strong> PHAS-I in five (S/T)P sites that govern translational repression. Mol. Cell Biol., 20, 3558-<br />

3567.<br />

96. Gingras,A.C., B.Raught, and N.Sonenberg. (2001). Regulation <strong>of</strong> translation initiation by FRAP/mTOR. Genes<br />

Dev., 15, 807-826.<br />

97. Chung,J., R.E.Bachelder, E.A.Lipscomb, L.M.Shaw, and A.M.Mercurio. (2002). Integrin (alpha 6 beta 4)<br />

regulation <strong>of</strong> eIF-4E activity and VEGF translation: a survival mechanism for carcinoma cells. J. Cell Biol.,<br />

158, 165-174.<br />

98. Nathan,C.A., L.Liu, B.D.Li, F.W.Abreo, I.Nandy, and A.De Benedetti. (1997). Detection <strong>of</strong> the proto-oncogene<br />

<strong>eIF4E</strong> in surgical margins may predict recurrence in head and neck cancer. Oncogene, 15, 579-584.<br />

99. Cohen,N., M.Sharma, A.Kentsis, J.M.Perez, S.Strudwick, and K.L.Borden. (2001). PML RING suppresses<br />

oncogenic transformation by reducing the affinity <strong>of</strong> <strong>eIF4E</strong> for <strong>mRNA</strong>. EMBO J., 20, 4547-4559.<br />

100. Nathan,C.A., P.Carter, L.Liu, B.D.Li, F.Abreo, A.Tudor, S.G.Zimmer, and A.De Benedetti. (1997). Elevated<br />

expression <strong>of</strong> <strong>eIF4E</strong> and FGF-2 is<strong>of</strong>orms during vascularization <strong>of</strong> breast carcinomas. Oncogene, 15, 1087-<br />

1094.<br />

101. Gradi,A., Y.V.Svitkin, H.Imataka, and N.Sonenberg. (1998). Proteolysis <strong>of</strong> human eukaryotic translation<br />

initiation factor eIF4GII, but not eIF4GI, coincides <strong>with</strong> the shut<strong>of</strong>f <strong>of</strong> host protein synthesis after poliovirus<br />

infection. Proc. Natl. Acad. Sci. U. S. A, 95, 11089-11094.<br />

102. Svitkin,Y.V., A.Gradi, H.Imataka, S.Morino, and N.Sonenberg. (1999). Eukaryotic initiation factor 4GII<br />

(eIF4GII), but not eIF4GI, cleavage correlates <strong>with</strong> inhibition <strong>of</strong> host cell protein synthesis after human<br />

rhinovirus infection. J. Virol., 73, 3467-3472.<br />

103. Morley,S.J., L.McKendrick, and M.Bushell. (1998). Cleavage <strong>of</strong> translation initiation factor 4G (eIF4G) during<br />

anti-Fas IgM-induced apoptosis does not require signalling through the p38 mitogen-activated protein (MAP)<br />

kinase. FEBS Lett., 438, 41-48.<br />

104. Bushell,M., L.McKendrick, R.U.Janicke, M.J.Clemens, and S.J.Morley. (1999). Caspase-3 is necessary and<br />

sufficient for cleavage <strong>of</strong> protein synthesis eukaryotic initiation factor 4G during apoptosis. FEBS Lett., 451,<br />

332-336.<br />

105. Thomas,G. and M.N.Hall. (1997). TOR signalling and control <strong>of</strong> cell growth. Curr. Opin. Cell Biol., 9, 782-787.<br />

106. Huang,S. and P.J.Houghton. (2002). Inhibitors <strong>of</strong> mammalian target <strong>of</strong> rapamycin as novel antitumor agents: from<br />

bench to clinic. Curr. Opin. Investig. Drugs, 3, 295-304.<br />

107. Schmelzle,T. and M.N.Hall. (2000). TOR, a central controller <strong>of</strong> cell growth. Cell, 103, 253-262.<br />

108. Huang,S. and P.J.Houghton. (2001). Resistance to rapamycin: a novel anticancer drug. Cancer Metastasis Rev.,<br />

20, 69-78.<br />

109. Huang,S. and P.J.Houghton. (2001). Mechanisms <strong>of</strong> resistance to rapamycins. Drug Resist. Updat., 4, 378-391.<br />

110. DeFatta,R.J., Y.Li, and A.De Benedetti. (2002). Selective killing <strong>of</strong> cancer cells based on translational control <strong>of</strong> a<br />

suicide gene. Cancer Gene Ther., 9, 573-578.<br />

140


111. Patikoglou,G. and S.K.Burley. (1997). Eukaryotic transcription factor-DNA complexes. Annu. Rev. Biophys.<br />

Biomol. Struct., 26, 289-325.<br />

112. DeHaseth,P.L., T.M.Lohman, and M.T.Record, Jr. (1977). Nonspecific interaction <strong>of</strong> lac repressor <strong>with</strong> DNA: an<br />

association reaction driven by counterion release. Biochemistry, 16, 4783-4790.<br />

113. Record,M.T., Jr., P.L.DeHaseth, and T.M.Lohman. (1977). Interpretation <strong>of</strong> monovalent and divalent cation<br />

effects on the lac repressor-operator interaction. Biochemistry, 16, 4791-4796.<br />

114. Record,M.T., Jr., C.F.Anderson, and T.M.Lohman. (1978). <strong>Thermodynamic</strong> analysis <strong>of</strong> ion effects on the binding<br />

and conformational equilibria <strong>of</strong> proteins and nucleic acids: the roles <strong>of</strong> ion association or release, screening,<br />

and ion effects on water activity. Q. Rev. Biophys., 11, 103-178.<br />

115. Lohman,T.M., P.L.DeHaseth, and M.T.Record, Jr. (1978). <strong>Analysis</strong> <strong>of</strong> ion concentration effects <strong>of</strong> the kinetics <strong>of</strong><br />

protein- nucleic acid interactions. Application to lac repressor-operator interactions. Biophys. Chem., 8, 281-<br />

294.<br />

116. Roe,J.H. and M.T.Record, Jr. (1985). Regulation <strong>of</strong> the kinetics <strong>of</strong> the interaction <strong>of</strong> Escherichia coli RNA<br />

polymerase <strong>with</strong> the lambda PR promoter by salt concentration. Biochemistry, 24, 4721-4726.<br />

117. Ha,J.H., M.W.Capp, M.D.Hohenwalter, M.Baskerville, and M.T.Record, Jr. (1992). <strong>Thermodynamic</strong><br />

stoichiometries <strong>of</strong> participation <strong>of</strong> water, cations and anions in specific and non-specific binding <strong>of</strong> lac<br />

repressor to DNA. Possible thermodynamic origins <strong>of</strong> the "glutamate effect" on protein-DNA interactions. J.<br />

Mol. Biol., 228, 252-264.<br />

118. Record,M.T., Jr., J.H.Ha, and M.A.Fisher. (1991). <strong>Analysis</strong> <strong>of</strong> equilibrium and kinetic measurements to determine<br />

thermodynamic origins <strong>of</strong> stability and specificity and mechanism <strong>of</strong> formation <strong>of</strong> site-specific complexes<br />

between proteins and helical DNA. Methods Enzymol., 208, 291-343.<br />

119. Parsegian,V.A., R.P.Rand, and D.C.Rau. (1995). Macromolecules and water: probing <strong>with</strong> osmotic stress.<br />

Methods Enzymol., 259, 43-94.<br />

120. Chaires,J.B., S.Satyanarayana, D.Suh, I.Fokt, T.Przewloka, and W.Priebe. (1996). Parsing the free energy <strong>of</strong><br />

anthracycline antibiotic binding to DNA. Biochemistry, 35, 2047-2053.<br />

121. Qu,X. and J.B.Chaires. (2001). Hydration Changes for DNA Intercalation Reactions. J. Am. Chem. Soc., 123, 1-7.<br />

122. Eftink,M.R., A.C.Anusiem, and R.L.Biltonen. (1983). Enthalpy-entropy compensation and heat capacity changes<br />

for protein- ligand interactions: general thermodynamic models and data for the binding <strong>of</strong> nucleotides to<br />

ribonuclease A. Biochemistry, 22, 3884-3896.<br />

123. Luther,M.A., G.Z.Cai, and J.C.Lee. (1986). <strong>Thermodynamic</strong>s <strong>of</strong> dimer and tetramer formations in rabbit muscle<br />

phosph<strong>of</strong>ructokinase. Biochemistry, 25, 7931-7937.<br />

124. Baker,B.M. and K.P.Murphy. (1996). Evaluation <strong>of</strong> linked protonation effects in protein binding reactions using<br />

isothermal titration calorimetry. Biophys. J., 71, 2049-2055.<br />

125. Guminski, K. (1974). Termodynamika. PWN, Warszawa.<br />

126. Parsegian,V.A., R.P.Rand, and D.C.Rau. (2000). Osmotic stress, crowding, preferential hydration, and binding: A<br />

comparison <strong>of</strong> perspectives. Proc. Natl. Acad. Sci. U. S. A, 97, 3987-3992.<br />

127. McNaught, A. D. and Wilkinson, A. (1997). Compendium <strong>of</strong> Chemical Terminology. Blackwell Science.<br />

128. Bloomfield,V.A. (2001). <strong>Thermodynamic</strong>s and Statistical <strong>Thermodynamic</strong>s. In Biophysics Textbook online<br />

(E.Freire, editor), Biophysical Society & IUPAB,<br />

129. Dougherty,D.A. (1996). Cation-pi interactions in chemistry and biology: a new view <strong>of</strong> benzene, Phe, Tyr, and<br />

Trp. Science, 271, 163-168.<br />

130. Makhatadze,G.I. and P.L.Privalov. (1995). Energetics <strong>of</strong> protein structure. Adv. <strong>Protein</strong> Chem., 47, 307-425.<br />

141


131. Sturtevant,J.M. (1977). Heat capacity and entropy changes in processes involving proteins. Proc. Natl. Acad. Sci.<br />

U. S. A, 74, 2236-2240.<br />

132. Murphy,K.P. and S.J.Gill. (1991). Solid model compounds and the thermodynamics <strong>of</strong> protein unfolding. J. Mol.<br />

Biol., 222, 699-709.<br />

133. Murphy,K.P. and E.Freire. (1992). <strong>Thermodynamic</strong>s <strong>of</strong> structural stability and cooperative folding behavior in<br />

proteins. Adv. <strong>Protein</strong> Chem., 43, 313-361.<br />

134. Spolar,R.S., J.R.Livingstone, and M.T.Record, Jr. (1992). Use <strong>of</strong> liquid hydrocarbon and amide transfer data to<br />

estimate contributions to thermodynamic functions <strong>of</strong> protein folding from the removal <strong>of</strong> nonpolar and polar<br />

surface from water. Biochemistry, 31, 3947-3955.<br />

135. (1994). The encyclopedia <strong>of</strong> molecular biology. Kendrew, J. and Lawrence, E. Blackwell Science, Oxford.<br />

136. Stryer, L. (1995). Biochemistry. W. H. Freeman & Co., New York.<br />

137. Fersht, A. R. (1999). Structure and mechanism in protein science. W. H. Freeman & Co., New York.<br />

138. Horovitz,A., L.Serrano, B.Avron, M.Bycr<strong>of</strong>t, and A.R.Fersht. (1990). Strength and co-operativity <strong>of</strong> contributions<br />

<strong>of</strong> surface salt bridges to protein stability. J. Mol. Biol., 216, 1031-1044.<br />

139. Kuntz,I.D., K.Chen, K.A.Sharp, and P.A.Kollman. (1999). The maximal affinity <strong>of</strong> ligands. Proc. Natl. Acad. Sci.<br />

U. S. A, 96, 9997-10002.<br />

140. Serrano,L., J.T.Kellis, Jr., P.Cann, A.Matouschek, and A.R.Fersht. (1992). The folding <strong>of</strong> an enzyme. II.<br />

Substructure <strong>of</strong> barnase and the contribution <strong>of</strong> different interactions to protein stability. J. Mol. Biol., 224,<br />

783-804.<br />

141. Shirley,B.A., P.Stanssens, U.Hahn, and C.N.Pace. (1992). Contribution <strong>of</strong> hydrogen bonding to the<br />

conformational stability <strong>of</strong> ribonuclease T1. Biochemistry, 31, 725-732.<br />

142. Hu,C.Q., J.M.Sturtevant, J.A.Thomson, R.E.Erickson, and C.N.Pace. (1992). <strong>Thermodynamic</strong>s <strong>of</strong> ribonuclease<br />

T1 denaturation. Biochemistry, 31, 4876-4882.<br />

143. Pace,C.N., B.A.Shirley, M.McNutt, and K.Gajiwala. (1996). Forces contributing to the conformational stability <strong>of</strong><br />

proteins. FASEB J., 10, 75-83.<br />

144. Sneddon,S.F., D.J.Tobias, and C.L.Brooks, III. (1989). <strong>Thermodynamic</strong>s <strong>of</strong> amide hydrogen bond formation in<br />

polar and apolar solvents. J. Mol. Biol., 209, 817-820.<br />

145. Pace,C.N. (1992). Contribution <strong>of</strong> the hydrophobic effect to globular protein stability. J. Mol. Biol., 226, 29-35.<br />

146. McGaughey,G.B., M.Gagne, and A.K.Rappe. (1998). pi-Stacking interactions. Alive and well in proteins. J. Biol.<br />

Chem., 273, 15458-15463.<br />

147. Serrano,L., M.Bycr<strong>of</strong>t, and A.R.Fersht. (1991). Aromatic-aromatic interactions and protein stability. Investigation<br />

by double-mutant cycles. J. Mol. Biol., 218, 465-475.<br />

148. Mecozzi,S., A.P.West, Jr., and D.A.Dougherty. (1996). Cation-pi interactions in aromatics <strong>of</strong> biological and<br />

medicinal interest: electrostatic potential surfaces as a useful qualitative guide. Proc. Natl. Acad. Sci. U. S. A,<br />

93, 10566-10571.<br />

149. Gallivan,J.P. and D.A.Dougherty. (1999). Cation-pi interactions in structural biology. Proc. Natl. Acad. Sci. U. S.<br />

A, 96, 9459-9464.<br />

150. Jayaram,B., K.J.McConnell, S.B.Dixit, and D.L.Beveridge. (1999). Free energy analysis <strong>of</strong> protein-DNA binding:<br />

the EcoRI endonuclease-DNA complex. J. Comput. Phys., 151, 333-357.<br />

151. Haq,I., J.E.Ladbury, B.Z.Chowdhry, T.C.Jenkins, and J.B.Chaires. (1997). Specific binding <strong>of</strong> hoechst 33258 to<br />

the d(CGCAAATTTGCG)2 duplex: calorimetric and spectroscopic studies. J. Mol. Biol., 271, 244-257.<br />

142


152. Boresch,S., G.Archontis, and M.Karplus. (1994). Free energy simulations: the meaning <strong>of</strong> the individual<br />

contributions from a component analysis. <strong>Protein</strong>s, 20, 25-33.<br />

153. Mark,A.E. and W.F.van Gunsteren. (1994). Decomposition <strong>of</strong> the free energy <strong>of</strong> a system in terms <strong>of</strong> specific<br />

interactions. Implications for theoretical and experimental studies. J. Mol. Biol., 240, 167-176.<br />

154. Brady,G.P., A.Szabo, and K.A.Sharp. (1996). On the decomposition <strong>of</strong> free energies. J. Mol. Biol., 263, 123-125.<br />

155. Rhoads,R.E., G.M.Hellmann, P.Remy, and J.P.Ebel. (1983). Translational recognition <strong>of</strong> messenger ribonucleic<br />

acid caps as a function <strong>of</strong> pH. Biochemistry, 22, 6084-6088.<br />

156. Carberry,S.E., R.E.Rhoads, and D.J.Goss. (1989). A spectroscopic study <strong>of</strong> the binding <strong>of</strong> m7GTP and m7GpppG<br />

to human protein synthesis initiation factor 4E. Biochemistry, 28, 8078-8083.<br />

157. Cai,A., M.Jankowska-Anyszka, A.Centers, L.Chlebicka, J.Stepinski, R.Stolarski, E.Darzynkiewicz, and<br />

R.E.Rhoads. (1999). Quantitative assessment <strong>of</strong> <strong>mRNA</strong> cap analogues as inhibitors <strong>of</strong> in vitro translation.<br />

Biochemistry, 38, 8538-8547.<br />

158. McCubbin,W.D., I.Edery, M.Altmann, N.Sonenberg, and C.M.Kay. (1988). Circular dichroism and fluorescence<br />

studies on protein synthesis initiation factor eIF-4E and two mutant forms from the yeast Saccharomyces<br />

cerevisiae. J. Biol. Chem., 263, 17663-17671.<br />

159. Carberry,S.E., E.Darzynkiewicz, J.Stepinski, S.M.Tahara, R.E.Rhoads, and D.J.Goss. (1990). A spectroscopic<br />

study <strong>of</strong> the binding <strong>of</strong> N-7-substituted cap analogues to human protein synthesis initiation factor 4E.<br />

Biochemistry, 29, 3337-3341.<br />

160. Goss,D.J., S.E.Carberry, T.E.Dever, W.C.Merrick, and R.E.Rhoads. (1990). Fluorescence study <strong>of</strong> the binding <strong>of</strong><br />

m7GpppG and rabbit globin <strong>mRNA</strong> to protein synthesis initiation factors 4A, 4E, and 4F. Biochemistry, 29,<br />

5008-5012.<br />

161. Ueda,H., H.Maruyama, M.Doi, M.Inoue, T.Ishida, H.Morioka, T.Tanaka, S.Nishikawa, and S.Uesugi. (1991).<br />

Expression <strong>of</strong> a synthetic gene for human cap binding protein (human IF- 4E) in Escherichia coli and<br />

fluorescence studies on interaction <strong>with</strong> <strong>mRNA</strong> cap structure analogues. J. Biochem. (Tokyo), 109, 882-889.<br />

162. Morino,S., M.Yasui, M.Doi, M.Inoue, T.Ishida, H.Ueda, and S.Uesugi. (1994). Direct expression <strong>of</strong> a synthetic<br />

gene in Escherichia coli: purification and physicochemical properties <strong>of</strong> human initiation factor 4E. J.<br />

Biochem. (Tokyo), 116, 687-693.<br />

163. Hagedorn,C.H., T.Spivak-Kroizman, D.E.Friedland, D.J.Goss, and Y.Xie. (1997). Expression <strong>of</strong> functional eIF-<br />

4Ehuman: purification, detailed characterization, and its use in isolating eIF-4E binding proteins. <strong>Protein</strong><br />

Expr. Purif., 9, 53-60.<br />

164. Hsu,P.C., M.R.Hodel, J.W.Thomas, L.J.Taylor, C.H.Hagedorn, and A.E.Hodel. (2000). Structural requirements<br />

for the specific recognition <strong>of</strong> an m7G <strong>mRNA</strong> cap. Biochemistry, 39, 13730-13736.<br />

165. Burley, S. K. (1998). Personal communication.<br />

166. Ptushkina,M., T.von der Haar, M.M.Karim, J.M.Hughes, and J.E.McCarthy. (1999). Repressor binding to a dorsal<br />

regulatory site traps human <strong>eIF4E</strong> in a high cap-affinity state. EMBO J., 18, 4068-4075.<br />

167. Ptushkina,M., T.von der Haar, S.Vasilescu, R.Frank, R.Birkenhager, and J.E.McCarthy. (1998). Cooperative<br />

modulation by eIF4G <strong>of</strong> <strong>eIF4E</strong>-binding to the <strong>mRNA</strong> 5' cap in yeast involves a site partially shared by p20.<br />

EMBO J., 17, 4798-4808.<br />

168. von der Haar,T., P.D.Ball, and J.E.McCarthy. (2000). Stabilization <strong>of</strong> eukaryotic initiation factor 4E binding to<br />

the <strong>mRNA</strong> 5'- Cap by domains <strong>of</strong> eIF4G. J. Biol. Chem., 275, 30551-30555.<br />

169. Darzynkiewicz,E., I.Ekiel, S.M.Tahara, L.S.Seliger, and A.J.Shatkin. (1985). Chemical synthesis and<br />

characterization <strong>of</strong> 7-methylguanosine cap analogues. Biochemistry, 24, 1701-1707.<br />

170. Jankowska,M., J.Stepinski, R.Stolarski, A.Temeriusz, and E.Darzynkiewicz. (1993). Synthesis and properties <strong>of</strong><br />

new NH 2 and N7 substituted GMP and GTP 5'-<strong>mRNA</strong> cap analogues. Collect. Czech. Chem. Commun., 58,<br />

138-141.<br />

143


171. Darzynkiewicz,E., J.Stepinski, I.Ekiel, Y.Jin, D.Haber, T.Sijuwade, and S.M.Tahara. (1988). Beta-globin <strong>mRNA</strong>s<br />

capped <strong>with</strong> m7G, m2.7(2)G or m2.2.7(3)G differ in intrinsic translation efficiency. Nucleic Acids Res., 16,<br />

8953-8962.<br />

172. Stepinski,J., M.Bretner, M.Jankowska, K.Felczak, R.Stolarski, Z.Wieczorek, A.L.Cai, R.E.Rhoads, A.Temeriusz,<br />

D.Haber, and E.Darzynkiewicz. (1995). Synthesis and properties <strong>of</strong> P 1 ,P 2 -, P 1 ,P 3 - and P 1 ,P 4 - dinucleotide di-,<br />

tri- and tetraphosphate <strong>mRNA</strong> 5'-cap analogues. Nucleosides & Nucleotides, 14, 717-721.<br />

173. Jankowska,M., J.Stepinski, R.Stolarski, Z.Wieczorek, A.Temeriusz, D.Haber, and E.Darzynkiewicz. (1996). H-1<br />

NMR and fluorescence studies <strong>of</strong> new <strong>mRNA</strong> 5'-cap analogues. Collect. Czech. Chem. Commun., 61, 197-<br />

202.<br />

174. Darzynkiewicz,E., J.Stepinski, S.M.Tahara, R.Stolarski, I.Ekiel, D.Haber, K.Neuvonen, P.Lehikoinen, I.Labadi,<br />

and H.Lonnberg. (1990). Synthesis, conformation and hydrolytic stability <strong>of</strong> P 1 ,P 3 -dinucleoside triphosphates<br />

related to <strong>mRNA</strong> 5'-cap, and comparative kinetic studies on their nucleoside and nucleoside monophosphate<br />

analogues. Nucleosides & Nucleotides, 9, 599-618.<br />

175. Ruszczynska, K. and Stolarski, R. (2000). Personal Communication.<br />

176. Schnolzer,M., P.Alewood, A.Jones, D.Alewood, and S.B.Kent. (1992). In situ neutralization in Boc-chemistry<br />

solid phase peptide synthesis. Rapid, high yield assembly <strong>of</strong> difficult sequences. Int. J. Pept. <strong>Protein</strong> Res., 40,<br />

180-193.<br />

177. Wieczorek,Z., J.Stepinski, M.Jankowska, and H.Lonnberg. (1995). Fluorescence and absorption spectroscopic<br />

properties <strong>of</strong> RNA 5'-cap analogues derived from 7-methyl-, N2,7-dimethyl- and N2,N2,7-trimethylguanosines.<br />

J. Photochem. Photobiol. B, 28, 57-63.<br />

178. Dawson, R. M., Elliott, D. C., Elliott, W. H., and Jones, K. M. (1969). Data for biochemical research. 2, 174-175,<br />

Clarendon Press, Oxford.<br />

179. Edery,I., M.Altmann, and N.Sonenberg. (1988). High-level synthesis in Escherichia coli <strong>of</strong> functional cap-binding<br />

eukaryotic initiation factor eIF-4E and affinity purification using a simplified cap-analog resin. Gene, 74, 517-<br />

525.<br />

180. Stern,B.D., M.Wilson, and R.Jagus. (1993). Use <strong>of</strong> nonreducing SDS-PAGE for monitoring renaturation <strong>of</strong><br />

recombinant protein synthesis initiation factor, eIF-4 alpha. <strong>Protein</strong> Expr. Purif., 4, 320-327.<br />

181. Webb,N.R., R.V.Chari, G.DePillis, J.W.Kozarich, and R.E.Rhoads. (1984). Purification <strong>of</strong> the messenger RNA<br />

cap-binding protein using a new affinity medium. Biochemistry, 23, 177-181.<br />

182. Lakowicz, J. R. (1999). Principles <strong>of</strong> fluorescence spectroscopy. Kluwer Academic/Plenum Publishers, New<br />

York.<br />

183. Parker, C. A. (1968). Photoluminescence <strong>of</strong> Solutions. Elsevier Publishing Company, Amsterdam.<br />

184. Callis,P.R. (1997). 1La and 1Lb transitions <strong>of</strong> tryptophan: applications <strong>of</strong> theory and experimental observations to<br />

fluorescence <strong>of</strong> proteins. Methods Enzymol., 278, 113-150.<br />

185. Antosiewicz, J. (2003). Personal communication.<br />

186. Wieczorek,Z., J.Stepinski, E.Darzynkiewicz, and H.Lonnberg. (1993). Association <strong>of</strong> nucleosides and their 5'monophosphates<br />

<strong>with</strong> a tryptophan containing tripeptide, Trp-Leu-Glu: the source <strong>of</strong> an overestimation by<br />

fluorescence spectroscopy. Biophys. Chem., 47, 233-240.<br />

187. Eadie,G.S. (1942). J. Biol. Chem., 146, 85-93.<br />

188. H<strong>of</strong>stee,B.H.J. (1959). Nature, Lond., 184, 1296.<br />

189. (2001). Database <strong>of</strong> osmotic pressures.<br />

190. Marcus, Y. (1997). Ion Properties. Marcel Dekker, Inc., New York.<br />

191. Robinson, R. A. and Stokes, R. H. (1970). Electrolyte Solutions. Butterworths, London.<br />

144


192. Davies, C. W. (1962). Ion Association. Butterworths, London.<br />

193. Wyman,J.Jr. (1964). Linked functions and reciprocal effects in hemoglobin: a second look. Adv. <strong>Protein</strong> Chem.,<br />

223-286.<br />

194. Baldwin,R.L. (1986). Temperature dependence <strong>of</strong> the hydrophobic interaction in protein folding. Proc. Natl.<br />

Acad. Sci. U. S. A, 83, 8069-8072.<br />

195. Becktel,W.J. and J.A.Schellman. (1987). <strong>Protein</strong> stability curves. Biopolymers, 26, 1859-1877.<br />

196. Ha,J.H., R.S.Spolar, and M.T.Record, Jr. (1989). Role <strong>of</strong> the hydrophobic effect in stability <strong>of</strong> site-specific<br />

protein- DNA complexes. J. Mol. Biol., 209, 801-816.<br />

197. Beyer, W. H. (1987). CRC Standard mathematical tables. 28th, 536, CRC Press, Boca Raton, FL.<br />

198. Wieczorek,Z., K.Zdanowski, L.Chlebicka, J.Stepinski, M.Jankowska, B.Kierdaszuk, A.Temeriusz,<br />

E.Darzynkiewicz, and R.Stolarski. (1997). Fluorescence and NMR studies <strong>of</strong> intramolecular stacking <strong>of</strong><br />

<strong>mRNA</strong> cap- analogues. Biochim. Biophys. Acta, 1354 , 145-152.<br />

199. Sharp,K. (2001). Entropy-enthalpy compensation: fact or artifact? <strong>Protein</strong> Sci., 10, 661-667.<br />

200. Ladbury,J.E. and B.Z.Chowdhry. (1996). Sensing the heat: the application <strong>of</strong> isothermal titration calorimetry to<br />

thermodynamic studies <strong>of</strong> biomolecular interactions. Chem. Biol., 3, 791-801.<br />

201. (1991). MicroCal OMEGA Manual. MicroCal, Inc., Northampton, MA, U.S.A.<br />

202. Christensen, J. J., Hansen, L. D., and Izatt, R. M. (1976). Handbook <strong>of</strong> proton ionization heats and related<br />

thermodynamic quantities. John Wiley & Sons, New York.<br />

203. Harris, R. K. (1983). Nuclear Magnetic Resonance Spectroscopy. A Physicochemical View. Pitman Books Ltd.,<br />

London.<br />

204. Brown,L.R. and B.T.Farmer. (1989). Rotating-frame nuclear Overhauser effect. Methods Enzymol., 176, 199-<br />

216.<br />

205. Hore,P.J. (1983). Solvent suppression in Fourier transform nuclear magnetic resonance. Journal <strong>of</strong> Magnetic<br />

Resonance, 55, 283-300.<br />

206. Brunner,H. and K.Dransfeld. (1983). Light scattering by macromolecules. In Biophysics (W.Hoppe, W.Lohmann,<br />

H.Markl, and H.Ziegler, editors), pp.93-100, Springer-Verlag, New York.<br />

207. Taylor, J. R. (1982). An introduction to error analysis. University Science Books, Mill Valley, CA.<br />

208. Shibata,S., S.Morino, K.Tomoo, Y.In, and T.Ishida. (1998). Effect <strong>of</strong> <strong>mRNA</strong> cap structure on eIF-4E<br />

phosphorylation and cap binding analyses using Ser209-mutated eIF-4Es. Biochem. Biophys. Res. Commun.,<br />

247, 213-216.<br />

209. (1989). Elementy enzymologii . Witwicki, J. and Ardelt, W. 2PWN, Warszawa.<br />

210. Koellner,G., A.Bzowska, B.Wielgus-Kutrowska, M.Luic, T.Steiner, W.Saenger, and J.Stepinski. (2002). Open<br />

and closed conformation <strong>of</strong> the E. coli purine nucleoside phosphorylase active center and implications for the<br />

catalytic mechanism. J. Mol. Biol., 315, 351-371.<br />

211. Wielgus-Kutrowska,B., A.Bzowska, J.Tebbe, G.Koellner, and D.Shugar. (2002). Purine nucleoside phosphorylase<br />

from Cellulomonas sp.: physicochemical properties and binding <strong>of</strong> substrates determined by ligand-dependent<br />

enhancement <strong>of</strong> enzyme intrinsic fluorescence, and by protective effects <strong>of</strong> ligands on thermal inactivation <strong>of</strong><br />

the enzyme. Biochim. Biophys. Acta, 1597, 320-334.<br />

212. Lodish,H.F. (1974). Model for the regulation <strong>of</strong> <strong>mRNA</strong> translation applied to haemoglobin synthesis. Nature,<br />

251, 385-388.<br />

145


213. Chu,L.Y. and R.E.Rhoads. (1980). Inhibition <strong>of</strong> cell-free messenger ribonucleic acid translation by 7methylguanosine<br />

5'-triphosphate: effect <strong>of</strong> messenger ribonucleic acid concentration. Biochemistry, 19, 184-<br />

191.<br />

214. Rau,M., T.Ohlmann, S.J.Morley, and V.M.Pain. (1996). A reevaluation <strong>of</strong> the cap-binding protein, <strong>eIF4E</strong>, as a<br />

rate-limiting factor for initiation <strong>of</strong> translation in reticulocyte lysate. J. Biol. Chem., 271, 8983-8990.<br />

215. Hiremath,L.S., N.R.Webb, and R.E.Rhoads. (1985). Immunological detection <strong>of</strong> the messenger RNA cap-binding<br />

protein. J. Biol. Chem., 260, 7843-7849.<br />

216. Haghighat,A. and N.Sonenberg. (1997). eIF4G dramatically enhances the binding <strong>of</strong> <strong>eIF4E</strong> to the <strong>mRNA</strong> 5'-cap<br />

structure. J. Biol. Chem., 272, 21677-21680.<br />

217. Ishida,T., K.Ohnishi, M.Doi, and M.Inoue. (1989). Proton nuclear magnetic resonance study on the aromatic<br />

amino acid- guanine nucleotide system: effect <strong>of</strong> base methylation on the stacking interaction <strong>with</strong> tyrosine<br />

and phenylalanine. Chem. Pharm. Bull. (Tokyo), 37, 1-4.<br />

218. Ueda,H., M.Doi, M.Inoue, T.Ishida, T.Tanaka, and S.Uesugi. (1988). A possible recognition mode <strong>of</strong> <strong>mRNA</strong> cap<br />

terminal structure by peptide: cooperative stacking and hydrogen-bond pairing interactions between<br />

m7GpppA and Trp-Leu-Glu. Biochem. Biophys. Res. Commun., 154, 199-204.<br />

219. Ueda,H., H.Iyo, M.Doi, M.Inoue, and T.Ishida. (1991). Cooperative stacking and hydrogen bond pairing<br />

interactions <strong>of</strong> fragment peptide in cap binding protein <strong>with</strong> <strong>mRNA</strong> cap structure. Biochim. Biophys. Acta,<br />

1075, 181-186.<br />

220. Ueda,H., H.Iyo, M.Doi, M.Inoue, T.Ishida, H.Morioka, T.Tanaka, S.Nishikawa, and S.Uesugi. (1991).<br />

Combination <strong>of</strong> Trp and Glu residues for recognition <strong>of</strong> <strong>mRNA</strong> cap structure. <strong>Analysis</strong> <strong>of</strong> m7G base<br />

recognition site <strong>of</strong> human cap binding protein (IF-4E) by site-directed mutagenesis. FEBS Lett., 280, 207-210.<br />

221. Hu,G., P.D.Gershon, A.E.Hodel, and F.A.Quiocho. (1999). <strong>mRNA</strong> cap recognition: dominant role <strong>of</strong> enhanced<br />

stacking interactions between methylated bases and protein aromatic side chains. Proc. Natl. Acad. Sci. U. S.<br />

A, 96, 7149-7154.<br />

222. Hodel,A.E., P.D.Gershon, X.Shi, and F.A.Quiocho. (1996). The 1.85 A structure <strong>of</strong> vaccinia protein VP39: a<br />

bifunctional enzyme that participates in the modification <strong>of</strong> both <strong>mRNA</strong> ends. Cell, 85, 247-256.<br />

223. Quiocho,F.A., G.Hu, and P.D.Gershon. (2000). Structural basis <strong>of</strong> <strong>mRNA</strong> cap recognition by proteins. Curr.<br />

Opin. Struct. Biol., 10, 78-86.<br />

224. Hu,G., A.Oguro, C.Li, P.D.Gershon, and F.A.Quiocho. (2002). The "cap-binding slot" <strong>of</strong> an <strong>mRNA</strong> cap-binding<br />

protein: quantitative effects <strong>of</strong> aromatic side chain choice in the double-stacking sandwich <strong>with</strong> cap.<br />

Biochemistry, 41, 7677-7687.<br />

225. Martz, E. (2002). <strong>Protein</strong> Explorer. [1.982 Beta] //www.proteinexplorer.org.<br />

226. Takano,K., J.Funahashi, Y.Yamagata, S.Fujii, and K.Yutani. (1997). Contribution <strong>of</strong> water molecules in the<br />

interior <strong>of</strong> a protein to the conformational stability. J. Mol. Biol., 274, 132-142.<br />

227. Lucke,C., S.Huang, M.Rademacher, and H.Ruterjans. (2002). New insights into intracellular lipid binding<br />

proteins: The role <strong>of</strong> buried water. <strong>Protein</strong> Sci., 11, 2382-2392.<br />

228. Hempel,R., J.Schmidt-Brauns, M.Gebinoga, F.Wirsching, and A.Schwienhorst. (2001). Cation radius effects on<br />

cell-free translation in rabbit reticulocyte lysate. Biochem. Biophys. Res. Commun., 283, 267-272.<br />

229. Raught,B., A.C.Gingras, and N.Sonenberg. (2000). Regulation <strong>of</strong> ribosomal recruitment in eukaryotes. In<br />

Translational Control <strong>of</strong> Gene Expression pp.245-93, Cold Spring Harbor Laboratory Press, New York.<br />

230. Pyronnet,S., H.Imataka, A.C.Gingras, R.Fukunaga, T.Hunter, and N.Sonenberg. (1999). Human eukaryotic<br />

translation initiation factor 4G (eIF4G) recruits mnk1 to phosphorylate <strong>eIF4E</strong>. EMBO J., 18, 270-279.<br />

231. Waskiewicz,A.J., J.C.Johnson, B.Penn, M.Mahalingam, S.R.Kimball, and J.A.Cooper. (1999). Phosphorylation <strong>of</strong><br />

the cap-binding protein eukaryotic translation initiation factor 4E by protein kinase Mnk1 in vivo. Mol. Cell<br />

Biol., 19, 1871-1880.<br />

146


232. Colombo,M.F., D.C.Rau, and V.A.Parsegian. (1992). <strong>Protein</strong> solvation in allosteric regulation: a water effect on<br />

hemoglobin. Science, 256, 655-659.<br />

233. Colombo,M.F., D.C.Rau, and V.A.Parsegian. (1994). Reevaluation <strong>of</strong> chloride's regulation <strong>of</strong> hemoglobin oxygen<br />

uptake: the neglected contribution <strong>of</strong> protein hydration in allosterism. Proc. Natl. Acad. Sci. U. S. A, 91,<br />

10517-10520.<br />

234. Saenger,W. (1987). Structure and dynamics <strong>of</strong> water surrounding biomolecules. Annu. Rev. Biophys. Biophys.<br />

Chem., 16, 93-114.<br />

235. Sha,M., Y.Wang, T.Xiang, A.van Heerden, K.S.Browning, and D.J.Goss. (1995). <strong>Interaction</strong> <strong>of</strong> wheat germ<br />

protein synthesis initiation factor eIF- (iso)4F and its subunits p28 and p86 <strong>with</strong> m7GTP and <strong>mRNA</strong><br />

analogues. J. Biol. Chem., 270, 29904-29909.<br />

236. Gingras, A. C. (2002). Personal Communication.<br />

237. Niedzwiecka-Kornas, A., Chlebicka, L., Jankowska-Anyszka, M., Stepinski, J., Dadlez, M., Gingras, A. C.,<br />

Marcotrigiano, J., Sonenberg, N., Burley, S. K., Stolarski, R., and Darzynkiewicz, E. (1999). RNA Meeting<br />

Abstract '99.<br />

238. Miyoshi,H., T.Youtani, H.Ide, H.Hori, K.Okamoto, M.Ishikawa, M.Wakiyama, T.Nishino, T.Ishida, and K.Miura.<br />

(1999). Binding analysis <strong>of</strong> Xenopus laevis translation initiation factor 4E (<strong>eIF4E</strong>) in initiation complex<br />

formation. J. Biochem. (Tokyo), 126, 897-904.<br />

239. Fletcher,C.M., A.M.McGuire, A.C.Gingras, H.Li, H.Matsuo, N.Sonenberg, and G.Wagner. (1998). 4E binding<br />

proteins inhibit the translation factor <strong>eIF4E</strong> <strong>with</strong>out folded structure. Biochemistry, 37, 9-15.<br />

240. Pimentel,G.C. and A.L.McClellan. (1971). Hydrogen bonding. Annu. Rev. Phys. Chem., 22, 347-385.<br />

241. Stolarski,R., A.Sitek, J.Stepinski, M.Jankowska, P.Oksman, A.Temeriusz, E.Darzynkiewicz, H.Lonnberg, and<br />

D.Shugar. (1996). 1H-NMR studies on association <strong>of</strong> <strong>mRNA</strong> cap-analogues <strong>with</strong> tryptophan- containing<br />

peptides. Biochim. Biophys. Acta, 1293, 97-105.<br />

242. Spolar,R.S. and M.T.Record, Jr. (1994). Coupling <strong>of</strong> local folding to site-specific binding <strong>of</strong> proteins to DNA.<br />

Science, 263, 777-784.<br />

243. Bhat,T.N., G.A.Bentley, G.Boulot, M.I.Greene, D.Tello, W.Dall'Acqua, H.Souchon, F.P.Schwarz, R.A.Mariuzza,<br />

and R.J.Poljak. (1994). Bound water molecules and conformational stabilization help mediate an antigenantibody<br />

association. Proc. Natl. Acad. Sci. U. S. A, 91, 1089-1093.<br />

244. Naghibi,H., A.Tamura, and J.M.Sturtevant. (1995). Significant discrepancies between van't H<strong>of</strong>f and calorimetric<br />

enthalpies. Proc. Natl. Acad. Sci. U. S. A, 92, 5597-5599.<br />

245. Horn,J.R., J.F.Brandts, and K.P.Murphy. (2002). van't H<strong>of</strong>f and calorimetric enthalpies II: effects <strong>of</strong> linked<br />

equilibria. Biochemistry, 41, 7501-7507.<br />

246. Liu,Y. and J.M.Sturtevant. (1995). Significant discrepancies between van't H<strong>of</strong>f and calorimetric enthalpies. II.<br />

<strong>Protein</strong> Sci., 4, 2559-2561.<br />

247. Liu,Y. and J.M.Sturtevant. (1997). Significant discrepancies between van't H<strong>of</strong>f and calorimetric enthalpies. III.<br />

Biophys. Chem., 64, 121-126.<br />

248. Chaires,J.B. (1997). Possible origin <strong>of</strong> differences between van't H<strong>of</strong>f and calorimetric enthalpy estimates.<br />

Biophys. Chem., 64, 15-23.<br />

249. Rouzina,I. and V.A.Bloomfield. (1999). Heat capacity effects on the melting <strong>of</strong> DNA. 1. General aspects.<br />

Biophys. J., 77, 3242-3251.<br />

250. Horn,J.R., D.Russell, E.A.Lewis, and K.P.Murphy. (2001). Van't H<strong>of</strong>f and calorimetric enthalpies from<br />

isothermal titration calorimetry: are there significant discrepancies? Biochemistry, 40, 1774-1778.<br />

251. Kavanoor,M. and M.R.Eftink. (1997). Characterization <strong>of</strong> the role <strong>of</strong> side-chain interactions in the binding <strong>of</strong><br />

ligands to apo trp repressor: pH dependence studies. Biophys. Chem., 66, 43-55.<br />

147


252. Kozlov,A.G. and T.M.Lohman. (1999). Adenine base unstacking dominates the observed enthalpy and heat<br />

capacity changes for the Escherichia coli SSB tetramer binding to single-stranded oligoadenylates.<br />

Biochemistry, 38, 7388-7397.<br />

253. Marcotrigiano, J. (1999). Personal communication.<br />

254. Panniers,R., E.B.Stewart, W.C.Merrick, and E.C.Henshaw. (1985). Mechanism <strong>of</strong> inhibition <strong>of</strong> polypeptide chain<br />

initiation in heat-shocked Ehrlich cells involves reduction <strong>of</strong> eukaryotic initiation factor 4F activity. J. Biol.<br />

Chem., 260, 9648-9653.<br />

255. Lamphear,B.J. and R.Panniers. (1991). Heat shock impairs the interaction <strong>of</strong> cap-binding protein complex <strong>with</strong> 5'<br />

<strong>mRNA</strong> cap. J. Biol. Chem., 266, 2789-2794.<br />

256. Morley,S.J. and V.M.Pain. (1995). Translational regulation during activation <strong>of</strong> porcine peripheral blood<br />

lymphocytes: association and phosphorylation <strong>of</strong> the alpha and gamma subunits <strong>of</strong> the initiation factor<br />

complex eIF-4F. Biochem. J., 312 ( Pt 2), 627-635.<br />

257. Sierra,J.M. and J.M.Zapata. (1994). Translational regulation <strong>of</strong> the heat shock response. Mol. Biol. Rep., 19, 211-<br />

220.<br />

258. Rhoads,R.E. and B.J.Lamphear. (1995). Cap-independent translation <strong>of</strong> heat shock messenger RNAs. Curr. Top.<br />

Microbiol. Immunol., 203, 131-153.<br />

259. Varani,G. (1997). A cap for all occasions. Structure., 5, 855-858.<br />

260. Huber,J., U.Cronshagen, M.Kadokura, C.Marshallsay, T.Wada, M.Sekine, and R.Luhrmann. (1998). Snurportin1,<br />

an m3G-cap-specific nuclear import receptor <strong>with</strong> a novel domain structure. EMBO J., 17, 4114-4126.<br />

261. Zorio,D.A., N.N.Cheng, T.Blumenthal, and J.Spieth. (1994). Operons as a common form <strong>of</strong> chromosomal<br />

organization in C. elegans. Nature, 372, 270-272.<br />

262. Blumenthal,T. and K.Steward. (1997). In C. Elegans II (D.L.Riddle, T.Blumenthal, B.J.Meyer, and J.R.Priess,<br />

editors), pp.117-45, Cold Spring Harbor Laboratory, Cold Spring Harbor, NY.<br />

263. Stachelska,A., Z.Wieczorek, K.Ruszczynska, R.Stolarski, M.Pietrzak, B.J.Lamphear, R.E.Rhoads,<br />

E.Darzynkiewicz, and M.Jankowska-Anyszka. (2002). <strong>Interaction</strong> <strong>of</strong> three Caenorhabditis elegans is<strong>of</strong>orms <strong>of</strong><br />

translation initiation factor <strong>eIF4E</strong> <strong>with</strong> mono- and trimethylated <strong>mRNA</strong> 5' cap analogues. Acta Biochim. Pol.,<br />

49, 671-682.<br />

264. Keiper,B.D., B.J.Lamphear, A.M.Deshpande, M.Jankowska-Anyszka, E.J.Aamodt, T.Blumenthal, and<br />

R.E.Rhoads. (2000). Functional characterization <strong>of</strong> five <strong>eIF4E</strong> is<strong>of</strong>orms in Caenorhabditis elegans. J. Biol.<br />

Chem., 275, 10590-10596.<br />

265. Lumry,R. and S.Rajender. (1970). Enthalpy-entropy compensation phenomena in water solutions <strong>of</strong> proteins and<br />

small molecules: a ubiquitous property <strong>of</strong> water. Biopolymers, 9, 1125-1227.<br />

266. Krug,R.R., W.G.Hunter, and R.A.Grieger. (1976). Enthalpy-entropy compensation. I. Some fundamental<br />

statistical problems associated <strong>with</strong> the analysis <strong>of</strong> van't H<strong>of</strong>f and Arrhenius data. J. Phys. Chem., 80, 2335-<br />

2341.<br />

267. Krug,R.R., W.G.Hunter, and R.A.Grieger. (1976). Statistical interpretation <strong>of</strong> enthalpy-entropy compensation.<br />

Nature, 261, 566-567.<br />

268. Murphy,K.P., V.Bhakuni, D.Xie, and E.Freire. (1992). Molecular basis <strong>of</strong> co-operativity in protein folding. III.<br />

Structural identification <strong>of</strong> cooperative folding units and folding intermediates. J. Mol. Biol., 227, 293-306.<br />

269. Spolar,R.S., J.H.Ha, and M.T.Record, Jr. (1989). Hydrophobic effect in protein folding and other noncovalent<br />

processes involving proteins. Proc. Natl. Acad. Sci. U. S. A, 86, 8382-8385.<br />

270. Murphy,K.P., P.L.Privalov, and S.J.Gill. (1990). Common features <strong>of</strong> protein unfolding and dissolution <strong>of</strong><br />

hydrophobic compounds. Science, 247, 559-561.<br />

271. Gomez,J., V.J.Hilser, D.Xie, and E.Freire. (1995). The heat capacity <strong>of</strong> proteins. <strong>Protein</strong>s, 22, 404-412.<br />

148


272. Bruzzese,F.J. and P.R.Connelly. (1997). Allosteric properties <strong>of</strong> inosine monophosphate dehydrogenase revealed<br />

through the thermodynamics <strong>of</strong> binding <strong>of</strong> inosine 5'-monophosphate and mycophenolic acid. Temperature<br />

dependent heat capacity <strong>of</strong> binding as a signature <strong>of</strong> ligand-coupled conformational equilibria. Biochemistry,<br />

36, 10428-10438.<br />

273. Hileman,R.E., R.N.Jennings, and R.J.Linhardt. (1998). <strong>Thermodynamic</strong> analysis <strong>of</strong> the heparin interaction <strong>with</strong> a<br />

basic cyclic peptide using isothermal titration calorimetry. Biochemistry, 37, 15231-15237.<br />

274. Matulis,D., I.Rouzina, and V.A.Bloomfield. (2000). <strong>Thermodynamic</strong>s <strong>of</strong> DNA binding and condensation:<br />

isothermal titration calorimetry and electrostatic mechanism. J. Mol. Biol., 296, 1053-1063.<br />

275. Gill,D.S., D.J.Roush, K.A.Shick, and R.C.Willson. (1995). Microcalorimetric characterization <strong>of</strong> the anionexchange<br />

adsorption <strong>of</strong> recombinant cytochrome b5 and its surface-charge mutants. J. Chromatogr. A, 715,<br />

81-93.<br />

276. Oda,M., K.Furukawa, K.Ogata, A.Sarai, and H.Nakamura. (1998). <strong>Thermodynamic</strong>s <strong>of</strong> specific and non-specific<br />

DNA binding by the c-Myb DNA-binding domain. J. Mol. Biol., 276, 571-590.<br />

277. Madan,B. and K.A.Sharp. (2001). Hydration heat capacity <strong>of</strong> nucleic acid constituents determined from the<br />

random network model. Biophys. J., 81, 1881-1887.<br />

278. Gallagher,K. and K.Sharp. (1998). Electrostatic contributions to heat capacity changes <strong>of</strong> DNA-ligand binding.<br />

Biophys. J., 75, 769-776.<br />

279. Ma,J.C. and D.A.Dougherty. (1997). The cation-pi interaction. Chemical Reviews, 97, 1303-1324.<br />

280. Mecozzi,S., A.P.West, and D.A.Dougherty. (1996). Cation-pi interactions in simple aromatics: Electrostatics<br />

provide a predictive tool. Journal <strong>of</strong> the American Chemical Society, 118, 2307-2308.<br />

281. Przedmojski, R. (1998). Masters thesis.<br />

282. Filimonov,V.V. and P.L.Privalov. (1978). <strong>Thermodynamic</strong>s <strong>of</strong> base interaction in (A)n and (A.U)n. J. Mol. Biol.,<br />

122, 465-470.<br />

283. Breslauer,K.J. and J.M.Sturtevant. (1977). A calorimetric investigation <strong>of</strong> single stranded base stacking in the<br />

ribo-oligonucleotide A7. Biophys. Chem., 7, 205-209.<br />

284. Berg,O.G., R.B.Winter, and P.H.von Hippel. (1981). Diffusion-driven mechanisms <strong>of</strong> protein translocation on<br />

nucleic acids. 1. Models and theory. Biochemistry, 20, 6929-6948.<br />

285. Griffon,N. and E.Di Stasio. (2001). <strong>Thermodynamic</strong>s <strong>of</strong> Na+ binding to coagulation serine proteases. Biophys.<br />

Chem., 90, 89-96.<br />

286. Nureki,O., D.G.Vassylyev, M.Tateno, A.Shimada, T.Nakama, S.Fukai, M.Konno, T.L.Hendrickson, P.Schimmel,<br />

and S.Yokoyama. (1998). Enzyme structure <strong>with</strong> two catalytic sites for double-sieve selection <strong>of</strong> substrate.<br />

Science, 280, 578-582.<br />

287. Hinz,H.J., K.Weber, J.Flossdorf, and M.R.Kula. (1976). <strong>Thermodynamic</strong> studies on the specificity <strong>of</strong> Lisoleucine-tRNA<br />

ligase <strong>of</strong> Escherichia coli MRE 600. Calorimetric investigations on binding <strong>of</strong> amino acids<br />

and isoleucinol to the enzyme. Eur. J. Biochem., 71, 437-442.<br />

288. Ren,J., T.C.Jenkins, and J.B.Chaires. (2000). Energetics <strong>of</strong> DNA intercalation reactions. Biochemistry, 39, 8439-<br />

8447.<br />

289. DeHaseth,P.L., T.M.Lohman, R.R.Burgess, and M.T.Record, Jr. (1978). Nonspecific interactions <strong>of</strong> Escherichia<br />

coli RNA polymerase <strong>with</strong> native and denatured DNA: differences in the binding behavior <strong>of</strong> core and<br />

holoenzyme. Biochemistry, 17, 1612-1622.<br />

290. Strauss,H.S., R.R.Burgess, and M.T.Record, Jr. (1980). Binding <strong>of</strong> Escherichia coli ribonucleic acid polymerase<br />

holoenzyme to a bacteriophage T7 promoter-containing fragment: evaluation <strong>of</strong> promoter binding constants as<br />

a function <strong>of</strong> solution conditions. Biochemistry, 19, 3504-3515.<br />

149


291. Takeda,Y., P.D.Ross, and C.P.Mudd. (1992). <strong>Thermodynamic</strong>s <strong>of</strong> Cro protein-DNA interactions. Proc. Natl.<br />

Acad. Sci. U. S. A, 89, 8180-8184.<br />

292. Jin,L., J.Yang, and J.Carey. (1993). <strong>Thermodynamic</strong>s <strong>of</strong> ligand binding to trp repressor. Biochemistry, 32, 7302-<br />

7309.<br />

293. Ladbury,J.E., J.G.Wright, J.M.Sturtevant, and P.B.Sigler. (1994). A thermodynamic study <strong>of</strong> the trp repressoroperator<br />

interaction. J. Mol. Biol., 238, 669-681.<br />

294. Shen,X., K.Tomoo, S.Uchiyama, Y.Kobayashi, and T.Ishida. (2001). Structural and thermodynamic behavior <strong>of</strong><br />

eukaryotic initiation factor 4E in supramolecular formation <strong>with</strong> 4E-binding protein 1 and <strong>mRNA</strong> cap<br />

analogue, studied by spectroscopic methods. Chem. Pharm. Bull. (Tokyo), 49, 1299-1303.<br />

295. Kentsis,A., E.C.Dwyer, J.M.Perez, M.Sharma, A.Chen, Z.Q.Pan, and K.L.Borden. (2001). The RING Domains <strong>of</strong><br />

the Promyelocytic Leukemia <strong>Protein</strong> PML and the Arenaviral <strong>Protein</strong> Z Repress Translation by Directly<br />

Inhibiting Translation Initiation Factor <strong>eIF4E</strong>. J. Mol. Biol., 312, 609-623.<br />

150


151

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!