Skip to main content
Log in

Changes in the level of poly(Phe) synthesis in Escherichia coli ribosomes containing mutants of L4 ribosomal protein from Thermus thermophilus can be explained by structural changes in the peptidyltransferase center: a molecular dynamics simulation analysis

  • Article
  • Published:
European Biophysics Journal Aims and scope Submit manuscript

Abstract

Data from polyphenylalanine [poly(Phe)] synthesis determination in the presence and in the absence of erythromycin have been used in conjunction with Molecular Dynamics Simulation analysis, in order to localize the functional sites affected by mutations of Thermus thermophilus ribosomal protein L4 incorporated in Escherichia coli ribosomes. We observed that alterations in ribosome capability to synthesize poly(Phe) in the absence of erythromycin were mainly correlated to shifts of A2062 and C2612 of 23S rRNA, while in the presence of erythromycin they were correlated to shifts of A2060 and U2584 of 23S rRNA. Our results suggest a means of understanding the role of the extended loop of L4 ribosomal protein in ribosomal peptidyltransferase center.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

Abbreviations

Poly(Phe):

Polyphenylalanine

PTase:

Peptidyltransferase

rRNA:

Ribosomal RNA

TthL4:

Ribosomal protein L4 from Thermus thermophilus

EcL22:

Ribosomal protein L22 from Escherichia coli

MDS:

Molecular dynamics simulation

wtTthL4:

Wild type L4 from Thermus thermophilus

TthL4-Glu55:

TthL4 with Gly55 replaced by Glu

TthL4-Ser55:

TthL4 with Gly55 replaced by Ser

TthL4-Asp56:

TthL4 with Glu56 replaced by Asp

TthL4-Gln56:

TthL4 with Glu56 replaced by Gln

TthL4- Glu55Gly56:

TthL4 with Gly55Glu56 inverted

RMSD:

Root mean square deviation

References

  • Agmon I, Amit M, Auerbach T, Bashan A, Baram D, Bartels H, Berisio R, Greenberg J, Harms J, Hansen HAS, Kessler M, Pyetan E, Schluenzen F, Sittner A, Yonath A, Zarivach R (2004) Ribosomal crystallography: aflexible nucleotide anchoring tRNA translocation, facilitates peptide-bond formation, chirality discrimination and antibiotics synergism. FEBS Lett 567:20–26

    Article  Google Scholar 

  • Ban N, Nissen P, Hansen J, Moore PB, Steitz TA (2000) The complete atomic structure of the large ribosomal subunit at 2.4 Å resolution. Science 289:905–920

    Article  ADS  Google Scholar 

  • Brodersen DE, Clemons Jr WM, Carter AP, Wimberly BT, Ramakrishnan V (2002) Crystal structure of the 30S ribosomal subunit from Thermus thermophilus: structure of the proteins and their interactions with 16S RNA. J Mol Biol 316:725–768

    Article  Google Scholar 

  • Douthwaite S, Aagaard C (1993) Erythromycin binding is reduced in ribosomes with conformational alterations in the 23S rRNA peptidyl transferase loop. J Mol Biol 232:725–731

    Article  Google Scholar 

  • Gabashvili IS, Gregory ST, Valle M, Grassucci R, Worbs M, Wahl MC, Dahlberg AE, Frank J (2001) The polypeptide tunnel system in the ribosome and its gating in erythromycin resistance mutants of L4 and L22. Mol Cell 8:181–188

    Article  Google Scholar 

  • Guex N, Peitsch MC (1997) SWISS-MODEL and the Swiss-PdbViewer: an environment for comparative protein modelling. Electrophoresis 18:2714–2723

    Article  Google Scholar 

  • Hansen JL, Ippolito JA, Ban N, Nissen P, Moore PB, Steitz TA (2002) The structure of four macrolide antibiotics bound to the large ribosomal subunit. Mol Cell 10:117–128

    Article  Google Scholar 

  • Hansen JL, Moore PB, Steitz TA (2003) Structures of five antibiotics bound at the peptidyl transferase center of the large ribosomal subunit. J Mol Biol 330:1061–1075

    Article  Google Scholar 

  • Harms J, Schluenzen F, Zarivach R, Bashan A, Gat S, Agmon I, Bartels H, Franceschi F, Yonath A (2001) High resolution structure of the large ribosomal subunit from a mesophilic eubacterium. Cell 107:679–688

    Article  Google Scholar 

  • Humphrey W, Dalke A, Schulten K: VMD (1996) Visual molecular dynamics. J Mol Graph 14:33–38

    Article  Google Scholar 

  • Jenni S, Ban N (2003) The chemistry of protein synthesis and voyage through the ribosomal tunnel. Curr Opin Struct Biol 13:212–219

    Article  Google Scholar 

  • Kalé L, Skeel R, Bhandarkar M, Brunner R, Gursoy A, Krawetz N, Phillips J, Shinozaki A, Varadarajan K, Schulten K (1999) NAMD2: greater scalability for parallel molecular dynamics. J Comput Phys 151:283–312

    Article  MATH  ADS  Google Scholar 

  • Leontiadou F, Xaplanteri MA, Papadopoulos G, Gerassimou C, Kalpaxis DL, Choli-Papadopoulou T (2003) On the structural and functional importance of the highly conserved Glu56 of Thermus thermophilus L4 ribosomal protein. J Mol Biol 332:73–84

    Article  Google Scholar 

  • Lovmar M, Tenson T, Ehrenberg M (2004) Kinetics of macrolide action: the josamycin and erythromycin cases. J Biol Chem 279:53506–53515

    Article  Google Scholar 

  • Moazed D, Noller HF (1987) Chloramphenicol, erythromycin, carbomycin and vernamycin B protect overlapping sites in peptidyl transferase region of 23S ribosomal RNA. Biochimie 69:879–884

    Article  Google Scholar 

  • Nakatogawa H, Ito K (2002) The ribosomal exit tunnel functions as discriminating gate. Cell 108:629–636

    Article  Google Scholar 

  • Nissen P, Hansen J, Ban N, Moore PB, Steitz TA (2000) The structural basis of ribosome activity in peptide bond synthesis. Science 289:920–930

    Article  ADS  Google Scholar 

  • Pfister P, Corti N, Hobbie S, Bruell C, Zarivach R, Yonath A, Böttger EC (2005) 23S rRNA base pair 2057–2611 determines ketolide susceptibility and fitness cost of the macrolide resistance mutation 2058A→G. Proc Natl Acad Sci USA 102:5180–5185

    Article  ADS  Google Scholar 

  • Rodriguez-Fonseca C, Phan H, Long KS, Porse BT, Kirillov SV, Amils R, Garrett RA (2000) Puromycin-rRNA interaction sites at the peptidyl transferase center. RNA 6:744–754

    Article  Google Scholar 

  • Schluenzen F, Tocilj A, Zarivach R, Harms R, Harms J, Gluehmann M, Janell D, Bashan A, Bartels H, Agmon I, Franceschi F, Yonath A (2000) Structure of functionally activated small ribosomal subunit at 3.3 angstroms resolution. Cell 102:615–623

    Article  Google Scholar 

  • Schluenzen F, Zarivach R, Harms J, Bashan A, Tocilj A, Albrecht A, Yonath A, Franceschi F (2001) Structural basis for the interaction of antibiotics with the peptidyl transferase center in eubacteria. Nature 413:814–821

    Article  ADS  Google Scholar 

  • Steiner G, Kuechler E, Barta A (1988) Photo-affinity labeling at the peptidyl transferase center reveals two different positions for the A- and P-sites in domain V of 23S rRNA. EMBO J 7:3949–3955

    Google Scholar 

  • Tenson T, Ehrenberg M (2002) Regulatory nascent peptides in the ribosomal tunnel. Cell 108:591–594

    Article  Google Scholar 

  • Tenson T, Lovmar M, Ehrenberg M (2003) The mechanism of action of macrolides, lincosamides and streptogramin B reveals the nascent peptide exit path in the ribosome. J Mol Biol 330:1005–1014

    Article  Google Scholar 

  • Tsagkalia A, Leontiadou F, Xaplanteri MA, Papadopoulos G, Kalpaxis DL, Choli-Papadopoulou T (2005) Ribosomes containing mutants of L4 ribosomal protein from Thermus thermophilus display multiple defects in ribosomal functions and sensitivity against erythromycin. RNA 11:1633–1639

    Article  Google Scholar 

  • Tu D, Blaha G, Moore PB, Steitz TA (2005) Structures of MLSBK antibiotics bond to mutated large ribosomal subunits provide a structural explanation for resistance. Cell 121:257–270

    Article  Google Scholar 

  • Vester B, Douthwaite S (2001) Macrolide resistance conferred by base substitutions in 23S rRNA. Antimicrob Agents Chemother 45:1–12

    Article  Google Scholar 

  • Wimberly BT, Brodersen DE, Clemons WM Jr, Morgan-Warren RJ, Carter AP, Vonrhein C, Hartsch T, Ramakrishnan V (2000) Structure of the 30S ribosomal subunit. Nature 407:327–339

    Article  ADS  Google Scholar 

  • Yusupov MM, Yusupova GZ, Baucom A, Lieberman K, Earnest TN, Cate JHD, Noller HF (2001) Crystal structure of the ribosome at 5.5 Å resolution. Science 292:883–896

    Article  ADS  Google Scholar 

  • Zengel JM, Jerauld A, Walker A, Wahl MC, Lindahl L (2003) The extended loops of ribosomal proteins L4 and L22 are not required for ribosomal assembly or L4-mediated autogenous control. RNA 9:1188–1197

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to G. Papadopoulos.

Appendix

Appendix

In this section, a justification is given for the application of a linear regression analysis to identify distant groups affected by the mutations. First, we start with a simplification of the complex process of the protein elongation step, by assuming that the binding of the aminoacyl-tRNA to the A-site pocket is the rate limiting step. According to the transition state theory,

$$ {\text{Kinetic scheme 1: L }} + {\text{ S }} \rightleftarrows {\text{ }}({\text{LS}})^{{ \ne {\text{ }}}} \to {\text{ LS}} $$

where L is the ligand (aminoacyl-tRNA), S is the acceptor molecule (mRNA-programmed ribosome, containing peptidyl-tRNA at the P-site), and LS is the encounter complex prior to peptide bond formation.

The rate constant k of the overall reaction is given by

$$ k = \frac{{k_{{\text{B}}} T}} {h}{\text{e}}^{{ - \Delta G^{{ \ne {\text{ }}}} {\text{/ }}RT}} , $$
(3)

where \( \Delta G^{{ \ne}} = G^{ \ne }_{{{\text{LS}}}} {\text{ }} - {\text{ }}(G_{{\text{L}}} {\text{ }} + {\text{ }}G_{{\text{S}}} ). \)

In the latter relationship, ΔG is the activation energy, G LS is the free energy of the activated complex, and G L, G S are the free energies of the ligand and the acceptor molecule, respectively. The symbols T, k B and h have their usual meaning. It might be expected that the activation energy depends on the conformation of the binding pocket and that of the ligand (aminoacyl-tRNA). Also, it might be expected that some conformations of the binding pocket favor binding (smaller ΔG ) while others prevent it (higher ΔG ). It is reasonable to assume that the major contribution to the dependence of ΔG on conformational changes comes from G LS . This is because aminoacyl-tRNA is not disturbed by the mutations and because the changes of G S are small compared with changes of the interaction energy between n groups at the (LS) interface. Lets denote by δ i the distance of an individual residue i of S at the interface with L from its counterpart at L. The interaction potential between these individual groups is assumed to be of the Lennard-Jones type. For small changes in δ i , leaving the system on the same side of the potential well, we can roughly adopt a linear dependence of the interaction potential on δ i . Other factors contributing to the activation energy do not significantly depend on δ i and therefore,

$$ \frac{{\partial \Delta G^{ \ne}}} {{\partial \delta _{i}}} \sim {\left\langle {\frac{{\partial U}} {{\partial \delta _{i} }}} \right\rangle} \sim\gamma _{i}. $$
(4)

In Eq. 4, U is the total interaction potential of all trans-interacting groups at the interface, and γ i are constants. The symbols “∼” and “ \( {\left\langle {} \right\rangle} \)” denote proportionality and ensemble averaging correspondingly.

After mutating an amino acid at the extended loop of L4, the disturbed system adopts a new equilibrium for residue positions (near the original), leading to a change dδ i in δ i . In turns, this leads to a change in the rate constant,

$$ {\text{d}}k = - k\frac{{\partial \Delta G^{ \ne } }} {{RT\partial \delta _{1} }}{\text{d}}\delta _{1} - k\frac{{\partial \Delta G^{ \ne } }} {{RT\partial \delta _{2} }}{\text{d}}\delta _{2} - \cdots - k\frac{{\partial \Delta G^{ \ne } }} {{RT\partial \delta _{n} }}{\text{d}}\delta _{n} . $$
(5)

Equation 5 combines with 4 to give the relationship

$$ {\text{d}}k = {\sum\limits_{i = 1}^n {\beta _{i} {\text{d}}\delta _{i} } }, $$
(6)

where \( \beta _{i} \approx -\frac{{k_{{{\text{wt}}}} \gamma _{i} }} {{RT}}, \) and k wt is the rate constant for the wild type.

We can use similar considerations to generalize the above analysis to include all N interacting groups in the ribosome and express the overall change of the experimentally measured level of poly(Phe) synthesis in 60 min, R, as a linear function of dδ i . We use as a measure of dδ i what we call here shifts, s i , given by the relationship

$$ s_{i} = {\sqrt {(x^{{\text{m}}}_{i} - x^{{{\text{wt}}}}_{i} )^{2} + (y^{{\text{m}}}_{i} - y^{{{\text{wt}}}}_{i} )^{2} + (z^{{\text{m}}}_{i} - z^{{{\text{wt}}}}_{i} )^{2} } }, $$
(7)

where x m ,y m ,z m and x wt ,y wt ,z wt are the equilibrium coordinates of the residues’ mass centres. Hence:

$$ {\text{d}}R = {\sum\limits_{i = 1}^N {\alpha _{i} s_{i} } }{\text{ or }}R_{{m }} = R_{{wt}} + {\sum\limits_{i = 1}^N {\alpha _{i} s_{i} } }. $$
(8)

Coefficients α i summarize all parameters affecting R, which do not depend on mutations.

Concerning the sum in Eq. 8, we notice that most of its terms contribute very little because of relatively small s i or/and small α i . Moreover, since α i can be positive or negative, they can cancel each other in many cases. On the other hand there must be some regions of the 50S subunit, like the PTase center, that are expected to contribute largely. In the frame of this analysis, we tried to identify the residues that exhibit the largest contributions in the sum of Eq. 8, keeping only two variables in the regression analysis. Of course this is not a rigorous analysis. Nevertheless, it is operationally a useful one, since a posteriori the results of the regression analysis are meaningful.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Papadopoulos, G., Grudinin, S., Kalpaxis, D.L. et al. Changes in the level of poly(Phe) synthesis in Escherichia coli ribosomes containing mutants of L4 ribosomal protein from Thermus thermophilus can be explained by structural changes in the peptidyltransferase center: a molecular dynamics simulation analysis. Eur Biophys J 35, 675–683 (2006). https://doi.org/10.1007/s00249-006-0076-4

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00249-006-0076-4

Keywords

Navigation