234
Views
0
CrossRef citations to date
0
Altmetric
Research Article

A quantitative structure-activity relationship study for structurally diverse HIV-1 protease inhibitors: contribution of conformational flexibility to inhibitory activity

Pages 609-615 | Received 12 Feb 2006, Accepted 20 Apr 2006, Published online: 04 Oct 2008

Abstract

In this study, we investigated by linear regression model the SAR data of the 15 HIV-1 protease inhibitors possessing structurally diverse scaffolds. First, a regression model was developed only using the enzyme-inhibitor interaction energy as a term of the model, but did not provide a good correlation with the inhibitory activity (R2 = 0.580 and Q2 = 0.500). Then, we focused on the conformational flexibility of the inhibitors which may represent the diversity of the inhibitors, and added two conformational parameters into the model, respectively: the number of rotatable bonds of ligands (ΔSrot) and the distortion energy of ligands (ΔElig). The regression model by adding ΔElig successfully improved the quality of the model (R2 = 0.771 and Q2 = 0.713) while the model with ΔSrot was unsuccessful. The prediction for a training inhibitor by the ΔElig model also showed good agreement with experimental activity. These results suggest that the conformational flexibility of HIV-1 protease inhibitors directly contributes to the enzyme inhibition.

Introduction

HIV-1 protease is one of the promising therapeutic targets of AIDS, and several protease inhibitor are widely and successfully being used for HIV/AIDS treatments Citation1-3. New inhibitors, however, are seamlessly being developed to overcome a drug-resistant virus that becomes an issue of the therapy Citation4-11.

The technology of structure-based design and/or computational chemistry has played a significant role in understanding the mechanism of HIV-1 protease inhibition and development of the protease inhibitors, e.g., identification of cyclic-urea inhibitors [Citation12]. Among the previously reported QSAR studies on HIV-1 protease inhibitors, good correlation models of the inhibitory activity were obtained using the enzyme-inhibitor interaction energy. For instance, the Merck group analyzed 33 HIV-1 protease inhibitors and obtained good correlation models of the activity with the interaction energy using the MM2X/OPTIMOL method (R2 = 0.7835 and Q2 = 0.7551 as the best correlation among the models) [Citation13]. Also, Gago and coworkers used the COMBINE methodology to analyze 49 inhibitors, and generated the highly predictive correlation model (R2 = 0.91 and Q2 = 0.81) [Citation14].

Although the significant contribution of interaction energy to the potency of the inhibition has been demonstrated by these studies, the inhibitors used for the analyses were cognate series of inhibitors in terms of chemical structure. Therefore, one may argue about the limitation of designing new compounds having structurally diverse scaffolds using this model. In fact, the known protease inhibitors so far have a variety of chemical scaffolds, e.g., from linear to cyclic scaffolds. Recently, Karplus and coworkers successfully established the correlation model by the estimation of the absolute binding free energy of enzyme-ligand interaction for such structurally diverse inhibitors (R2 = 0.83 and Q2 = 0.71)[Citation15]. This method is based on the conformational sampling by molecular dynamics and the electrostatic calculation by the Poisson or Poisson-Boltzmann equation. This model should be useful for designing a new compound with a new scaffold, but would be difficult to apply to medicinal-chemistry programs in non-computational labs.

In this study, we tried to apply a simple regression model for a set of structurally diverse protease inhibitors with the interaction energy alone initially, and then investigated the combination of other parameters to improve the regression model.

Material and methods

All forcefield-based computations were done by the MacroModel molecular modeling package with the BatchMin molecular mechanics engine [Citation16]. The MMFF94s parameter was used as the forcefield [Citation17]. The MMFF and AMBER94 [Citation18] atomic charges were loaded for inhibitor and enzyme atoms, respectively. SGI O2 R5000 workstation was used for all computations. The crystallographic structures of HIV-1 protease complexed with inhibitors were obtained from the Protein Data Bank (PDB). The PDB entry IDs of the complexes and the inhibitory activities of the inhibitors are listed in . The chemical structure of the inhibitors is shown in .

Table I.  HIV-1 protease inhibitors used in this study.

Figure 1 Chemical structures of HIV-1 protease inhibitors used a) as a training set, and b) as a test.

Figure 1 Chemical structures of HIV-1 protease inhibitors used a) as a training set, and b) as a test.

Computational results and discussion

Preparation of enzyme-inhibitor complex structures

HIV-1 protease inhibitors used in this study were selected from SAR data published in various literature reports( and ). Although they all are classified into one class, peptidic-inhibitors, the SAR data should be appropriate for this study since the set of SAR data has diversity in terms of chemical structure, and the inhibitory activity (Ki) is widely dispersed from a single picomolar to subnanomolar level. Fifteen inhibitors were used for the QSAR analysis as a training set and one inhibitor (JE2147) was used for the prediction of its activity as a test.

To build a protease structure complexed with each inhibitor for computation, all water molecules were removed from the x-ray structure except for the water molecule Wat301, which is known to be commonly conserved in HIV-1 protease-inhibitor complexes, and to be hydrogen-bonded to the flip loops of the enzyme in a dimeric form [Citation19]. For the complexes with cyclic-urea inhibitors, there is no water molecule equivalent to Wat301 since the carboxyl group of the inhibitors positionally replaces Wat301. Hydrogen atoms were then added to all heavy atoms of the enzyme and inhibitor, except for the aspartate residue at the 25th position of the enzyme (discussed below).

Determination of Asp protonation state

For the catalytically-significant aspartate residue at the 25th position, the side-chain atom of either Asp25 or Asp25′ in the dimer needs to be treated as a protonated form, i.e., one carboxyl and one carboxy anion groups in the side-chain atoms, according to known evidence [Citation20]. To determine the protonation state of aspartate, two protonated states in a complex with each inhibitor were separately prepared, i.e., one complex has a protonation on Asp25 and another has a protonation on Asp25′. Both complexes were then energy-minimized by the conjugated gradient method with the 0.01 kJ/Å-mol gradient convergent criterion. The protonation state for a complex exhibiting lower conformational energy between the two minimized complexes was chosen as the protonation state of the aspartate for the complex (). This assignment method has previously been used by Karplus's group [Citation15]. Our results showed a good agreement with their assignment except for a few complexes including AG1343, L735524, and L738317.

Table II.  Energy (kcal/mol) of the complexes with protonation at Asp25 or Asp25′a.

Calculation of enzyme-inhibitor interaction energy

For calculation of the enzyme-inhibitor interaction energy (ΔEint), the hydrogen-atom assigned complex structure was subjected to a short energy-minimization with the 0.1 kJ/Å-mol gradient convergent criterion in order to remove steric clash in the complex. For the minimization, the electrostatic energy was calculated by distance-dependant dielectric electrostatics with dielectric constant ϵ = 4r. The electrostatic non-bonded interaction was computed with a 12 Å cut-off distance and the van der Waals non-bonded interaction was computed with a 7 Å cut-off distance. Based on the short-minimized complex, ΔEint was calculated by the Equation (1). where Ecomplex is the conformational energy of the short-minimized complex; Eenzyme is the energy of the enzyme structure without the inhibitor; and Einhibitor is the conformational energy of the inhibitor alone. The ΔEint values of each complex are listed in .

Table III.  Calculated parameters for linear regression modeling.

Linear regression modeling with the interaction energy

To correlate the experimentally determined inhibitory activity (ΔGexp), linear regression equations were developed by the ΔEint value (). The squared correlation coefficient (R2) in the obtained regression model was not sufficiently high but the cross-validated correlation coefficient (Q2) was moderate (R2 = 0.580 and Q2 = 0.500). The regression model is shown as Model-1 in and calculated inhibitory activity (ΔGcalc) is listed in . The quality of the Model-1 equation was not as good as the previously reported models, indicating that the SAR data in this study are not well explained by the interaction energy alone. A major difference in the SAR data between the previous and our studies is the diversity of the chemical structure of the inhibitors; the previous studies used cognate series of inhibitors, whereas the SAR data of this study were composed of structurally diverse inhibitors. Therefore, this suggests that a different regression model is needed for structurally diverse inhibitors.

Table IV.  Statistical results of linear regression models.

Table V.  Experimental, calculated, and predicted free-energy differences.

Linear regression modeling with conformational parameters

Structurally diverse compounds likely offer a variety of conformational flexibility, e.g., linear or fixed structure. In fact, the number of rotatable bonds in the SAR data of this study widely varied from 9 to 19 (). Thus, to develop a good correlation model for the structurally diverse inhibitors, we focused on the conformational flexibility of the inhibitors and added conformational parameters into the Model-1 equation. So far, various conformational parameters to handle the conformational flexibility of ligands have been developed in virtual screening or QSAR modeling methods Citation21-27. Among them, in this study, two commonly used conformational parameters were employed to develop linear regression equations; one is a conformational entropic parameter derived from the number of rotatable bonds of ligands (ΔSrot), and the other is a conformational enthalpic parameter derived from the distortion energy of ligands (ΔElig).

ΔSrot represents the loss of conformational entropy, and was calculated by multiplying the number of rotatable bonds of a ligand (Nrot) by penalty coefficient (Crot) as shown in the Equation (2). where Crot = 0.3113 (kcal/mol/rotatable-bond) was used in this study. This value is used in Bohm's scoring function [Citation23] and estimation of docking energy in AutoDock method [Citation28]. In this study, the rotatable bonds of the inhibitors in the terminal functionalities and cyclic moieties were not counted. Since the ΔSrot value theoretically is positive or equal to zero, the following equation was applied for regression modeling so that the ΔSrot value is adopted as the penalty term in the free energy change equation.

The results of a linear regression equation with ΔSrot showed that the R2 and Q2 values were slightly improved to 0.643 and 0.567, respectively, from those of Model-1 (Model-2 in ). This fairly small improvement could be due to the limitation of correct estimation of conformational flexibility only by the number of the rotatable bonds, e.g., difficulty to estimate flexibility of the cyclic moiety.

Next, the conformational enthalpic parameter, ΔElig, was used for the regression modeling. ΔElig was calculated by the subtraction of the conformational energy of a ligand in a local minimum (Eligmin) from the conformational energy of a ligand in the bound state (Eligcomplex) as shown in Equation (3) (see ). The Eligcomplex value used was the same as the Einhibitor value defined in Equation (1). The Eligmin value was calculated by energy-minimizing the inhibitor structure, taken out from the complex, in free state.

As a result, a significant improvement of the correlation coefficients was achieved (R2 = 0.771 and Q2 = 0.713; Model-3 in ). The quality of this regression model has become comparable to that of the previously reported models including the Karplus's model. The relevance of ΔElig in the binding process has been investigated by several computational studies Citation21Citation29-32. The results obtained in this study also showed that the distortion energy of the inhibitors has a significant effect on the inhibitory activity for HIV-1 protease.

Prediction of inhibitory activity

The relevance of the obtained regression models, Model-1, Model-2, and Model-3, was examined by predicting the activity of an inhibitor, JE2147, which was excluded in the regression modeling ( and ). The best result of the prediction was obtained from the Model-3 equation which has the highest Q2 value and the lowest SPRESS value among the regression models (). The residual between the predicted and experimental activities for JE2147 was 0.902 kcal/mol.

Conclusion

In this study, the SAR data of HIV-1 protease inhibitors including structural diversity was investigated by linear regression models. We found that the linear regression model in a combination of the interaction energy with the distortion energy showed good correlation coefficients and predictive power while the interaction energy alone did not develop a good regression model. This suggests that the conformational flexibility of HIV-1 protease inhibitors directly contributes to the enzyme inhibition. In fact, in the field of medicinal chemistry, it is well known that the conformational flexibility of inhibitors significantly influences inhibitory activity, and the conformational fixation of molecules is one of the common tactics used to enhance biological activity [Citation33].

In conclusion, the regression model deduced in this study has quality as good as the previous reported models, while it is composed only of two terms in the equation and heavy computation is not needed. Thus, this model can be useful for development of a new chemical scaffold for HIV-1 protease inhibitors.

References

  • Abdel-Rahman HM, Al-Karamany GS, El-Koussi NA, Youssef AF, Kiso Y. HIV protease inhibitors: Peptidomimetic drugs and future perspectives. Curr Med Chem 2002; 9: 1905–1922
  • Randolph JT, DeGoey DA. Peptidomimetic inhibitors of HIV protease. Curr Topics Med Chem 2004; 4: 1079–1095
  • Rodriguez-Barrios F, Gago F. HIV protease inhibition: Limited recent progress and advances in understanding current pitfalls. Curr Topics Med Chem 2004; 4: 991–1007
  • Clavel F, Hance AJ. HIV drug resistance. N Engl J Med 2004; 350: 1023–1035
  • Werber Y. HIV drug market. Nat Rev Drug Discovery 2003; 2: 513–514
  • Menendez-Arias L. Targeting HIV:Antiretroviral therapy and development of drug resistance. Trends Pharmacol Sci 2002; 23: 381–388
  • de Mendoza C, Soriano V. Resistance to HIV protease inhibitors: Mechanisms and clinical consequences. Curr Drug Metab 2004; 5: 321–328
  • Fassler A, Bold G, Capraro H-G, Cozens R, Mestan J, Poncioni B, Rosel J, Tintelnot-Blomley M, Lang M. Azapeptide analogs as potent human immunodeficiency virus type-1 protease inhibitors with oral bioavailability. J Med Chem 1996; 39: 3203–3216
  • Bold G, Fassler A, Capraro H-G, Cozens R, Klimkait T, Lazdins J, Mestan J, Poncioni B, Rosel J, Stover D, Tintelnot-Blomley M, Acemoglu F, Beck W, Boss E, Eschbach M, Hurlimann T, Masso E, Roussel S, Ucci-Stoll K, Wyss D, Lang M. New aza-dipeptide analogues as potent and orally absorbed HIV-1 protease inhibitors: Candidates for clinical development. J Med Chem 1998; 41: 3387–3401
  • Raja A, Lebbos J, Kirkpatrick P. Fresh from the pipeline: Atazanavir sulphate. Nat Rev Drug Discovery 2003; 2: 857–858
  • Ekegren JK, Unge T, Safa MZ, Wallberg H, Samuelsson B, Hallberg A. A new class of HIV-1 protease inhibitors containing a tertiary alcohol in the transition-state mimicking scaffold. J Med Chem 2005; 48: 8098–8102
  • Lam PY, Jadhav PK, Eyermann CJ, Hodge CN, Ru Y, Bacheler LT, Meek JL, Otto MJ, Rayner MM, Wong YN, Chang C-H, Weber P, Jackson D, Sharpe T, Erickson-Viitanen S. Rational design of potent, bioavailable, nonpeptide cyclic ureas as HIV protease inhibitors. Science 1994; 263: 380–384
  • Holloway MK, Wai JM, Halgren TA, Fitzgerald PMD, Vacca JP, Dorsey BD, Levin RB, Thompson WJ, Chen LJ, deSolms SJ, Gaffin N, Ghosh AK, Giuliani EA, Graham SL, Guare JP, Hungate RW, Lyle TA, Sanders WM, Tucker TJ, Wiggins M, Wiscount CM, Woltersdorf OW, Young SD, Darke PL, Zugay JA. A priori prediction of activity for HIV-1 protease inhibitors employing energy minimization in the active site. J Med Chem 1995; 38: 305–317
  • Perez C, Pastor M, Ortiz AR, Gago F. Comparative binding energy analysis of hiv-1 protease inhibitors: Incorporation of solvent effects and validation as a powerful tool in receptor-based drug design. J Med Chem 1998; 41: 836–852
  • Zoete V, Michielin O, Karplus M. Protein-Ligand binding free energy estimation using molecular mechanics and continuum electrostatics. Application to HIV-1 protease inhibitors. J Comput.-Aided Mol Des 2003; 17: 861–880
  • Mohamadi F, Richards NGJ, Guida WC, Liskamp R, Lipton M, Caufield C, Chang G, Hendrickson T, Still WC. MacroModel - An integrated software system for modeling organic and bioorganic molecules using molecular mechanics. J Comput Chem 1990; 11: 440–467
  • Halgren TA. Merck molecular force field: i. basis, form, scope, parameterization and performance of MMFF94. J Comput Chem 1996; 17: 490–519
  • Cornell WD, Cieplak P, Bayly CI, Gould IR, Merz KM, Ferguson DM, Spellmeyer DC, Fox T, Caldwell JW, Kollman PA. A second generation force field for the simulation of proteins, nucleic acids, and organic molecules. J Am Chem Soc 1995; 117: 5179–5197
  • Wlodawer A, Erickson JW. Structure-Based inhibitors of HIV-1 protease. Annu Rev Biochem 1993; 62: 543–585
  • Piana S, Carloni P. Conformational flexibility of the catalytic asp dyad in HIV-1 protease: An ab initio study on the free enzyme. Proteins 2000; 39: 26–36
  • Pettersson I, Liljefors T. Structure-Activity relationships for apomorphine congeners. Conformational energies vs. biological activities. J Comput.-Aided Mol Des 1987; 1: 143–152
  • Lopez de Cornpadre RL, Pearlstein RA, Hopfinger AJ, Seydel JK. A quantitative structure-activity relationship analysis of some 4-aminodiphenyl sulfone antibacterial agents using linear free energy and molecular modeling methods. J Med Chem 1987; 30: 900–906
  • Bohm HJ. The development of a simple empirical scoring function to estimate the binding constant for a protein-ligand complex of known three-dimensional structure. J Comput.-Aided Mol Des 1994; 8: 243–256
  • Tokarski JS, Hopfinger AJ. Prediction of ligand-receptor binding thermodynamics by free energy force field (FEFF) 3D-QSAR analysis: Application to a set of peptidometic renin inhibitors. J Chem Inf Comput Sci 1997; 37: 792–811
  • Eldridge MD, Murray CW, Auton TR, Paolini GV, Mee RP. Empirical scoring functions: I. the development of a fast empirical scoring function to estimate the binding affinity of ligands in receptor complexes. J Comput.-Aided Mol Des 1997; 11: 425–445
  • Wang R, Liu L, Lai L, Tang Y. SCORE: A new empirical method for estimating the binding affinity of a protein-ligand complex. J Mol Model 1998; 4: 379–394
  • Friesner RA, Banks JL, Murphy RB, Halgren TA, Klicic JJ, Mainz DT, Repasky MP, Knoll EH, Shelley M, Perry JK, Shaw DE, Francis P, Shenkin PS. Glide: A new approach for rapid, accurate docking and scoring. 1. Method and assessment of docking accuracy. J Med Chem 2004; 47: 1739–1749
  • Morris GM, Goodsell DS, Halliday RS, Huey R, Hart WE, Belew RK, Olson AJ. Automated docking using a lamarckian genetic algorithm and empirical binding free energy function. J Comp Chem 1998; 19: 1639–1662
  • Nicklaus MC, Wang S, Driscoll JS, Milne GW. Conformational changes of small molecules binding to proteins. Bioorg Med Chem 1995; 3: 411–428
  • Bostrom J, Norrby PO, Liljefors T. Conformational energy penalties of protein-bound ligands. J Comput.-Aided Mol Des 1998; 12: 383–396
  • Vieth M, Hirst JD, Brooks CL, 3rd. Do active site conformations of small ligands correspond to low free-energy solution structures?. J Comput.-Aided Mol Des 1998; 12: 563–572
  • Perola E, Charifson PS. Conformational analysis of drug-like molecules bound to proteins: An extensive study of ligand reorganization upon binding. J Med Chem 2004; 47: 2499–2510
  • Fairlie DP, Tyndall JDA, Reld RC, Wong AK, Abbenante G, Scanlon MJ, March DR, Bergman DA, Chai CLL, Burkett BA. Conformational selection of inhibitors and substrates by proteolytic enzymes: Implications for drug design and polypeptide processing. J Med Chem 2000; 43: 1271–1281
  • Hosur MV, Bhat TN, Kempf DJ, Baldwin ET, Liu B, Gulnik S, Wideburg NE, Norbeck DW, Appelt K, Erickson JW. Influence of stereochemistry on activity and binding modes for c2 symmetry-based diol inhibitors of HIV-1 protease. J Am Chem Soc 1994; 116: 847–855
  • Kaldor SW, Kalish VJ, Davies JF, 2nd, Shetty BV, Fritz JE, Appelt K, Burgess JA, Campanale KM, Chirgadze NY, Clawson DK, Dressman BA, Hatch SD, Khalil DA, Kosa MB, Lubbehusen PP, Muesing MA, Patick AK, Reich SH, Su KS, Tatlock JH. Viracept (nelfinavir mesylate, AG1343): A potent, orally bioavailable inhibitor of HIV-1 protease. J Med Chem 1997; 40: 3979–3985
  • Backbro K, Lowgren S, Osterlund K, Atepo J, Unge T. Unexpected binding mode of a cyclic sulfamide HIV-1 protease inhibitor. J Med Chem 1997; 40: 898–902
  • Jhoti H, Singh OMP, Weir MP, Cooke R, Murray-Rust P, Wonacott A. X-ray crystallographic studies of a series of penicillin-derived asymmetric inhibitors of HIV-1 protease. Biochemistry 1994; 33: 8417–8427
  • Baldwin ET, Bhat TN, Gulnik S, Liu B, Topol IA, Kiso Y, Mimoto T, Mitsuya H, Erickson JW. Structure of HIV-1 protease with KNI-272, a tight-binding transition-state Analog Containing Allophenylnorstatine. Structure 1995; 3: 581–590
  • Chen Z, Li Y, Chen E, Hall DL, Darke PL, Culberson C, Shafer JA, Kuo LC. Crystal structure at 1.9-Å resolution of human immunodeficiency virus (HIV) II protease complexed with L-735,524, an orally bioavailable inhibitor of the HIV proteases. J Biol Chem 1994; 269: 26344–26348
  • Munshi S, Chen Z, Li Y, Olsen DB, Fraley ME, Hungate RW, Kuo LC. Rapid X-ray diffraction analysis of HIV-1 protease-inhibitor complexes: Inhibitor exchange in single crystals of the bound enzyme. Acta Crystallogr D Biol Crystallogr 1998; 54: 1053–1060
  • Hoog SS, Zhao B, Winborne E, Fisher S, Green DW, DesJarlais RL, Newlander KA, Callahan JF, Moore ML, Huffman WF, Abdel-Meguid SS. A check on rational drug design: Crystal structure of a complex of human immunodeficiency virus type 1 protease with a novel gamma-turn mimetic inhibitor. J Med Chem 1995; 38: 3246–3252
  • Thompson SK, Murthy KHM, Zhao B, Winborne E, Green DW, Fisher SM, DesJarlais RL, Tomaszek TA, Jr, Meek TD, Gleason JG, Abdel-Meguid SS. Rational design, synthesis, and crystallographic analysis of a hydroxyethylene-based HIV-1 protease inhibitor containing a heterocyclic P1′-P2′ amide bond isostere. J Med Chem 1994; 37: 3100–3107
  • Hong L, Treharne A, Hartsuck JA, Foundling S, Tang J. Crystal structures of complexes of a peptidic inhibitor with wild-type and two mutant HIV-1 proteases. Biochemistry 1996; 35: 10627–10633
  • Reiling KK, Endres NF, Dauber DS, Craik CS, Stroud RM. Anisotropic dynamics of the JE-2147-HIV protease complex: Drug resistance and thermodynamic binding mode examined in a 1.09 Å structure. Biochemistry 2002; 41: 4582–4594

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.