Skip to main content
Log in

Revisiting Dosing Regimen Using Pharmacokinetic/Pharmacodynamic Mathematical Modeling: Densification and Intensification of Combination Cancer Therapy

  • Original Research Article
  • Published:
Clinical Pharmacokinetics Aims and scope Submit manuscript

Abstract

Controlling effects of drugs administered in combination is particularly challenging with a densified regimen because of life-threatening hematological toxicities. We have developed a mathematical model to optimize drug dosing regimens and to redesign the dose intensification–dose escalation process, using densified cycles of combined anticancer drugs. A generic mathematical model was developed to describe the main components of the real process, including pharmacokinetics, safety and efficacy pharmacodynamics, and non-hematological toxicity risk. This model allowed for computing the distribution of the total drug amount of each drug in combination, for each escalation dose level, in order to minimize the average tumor mass for each cycle. This was achieved while complying with absolute neutrophil count clinical constraints and without exceeding a fixed risk of non-hematological dose-limiting toxicity. The innovative part of this work was the development of densifying and intensifying designs in a unified procedure. This model enabled us to determine the appropriate regimen in a pilot phase I/II study in metastatic breast patients for a 2-week-cycle treatment of docetaxel plus epirubicin doublet, and to propose a new dose-ranging process. In addition to the present application, this method can be further used to achieve optimization of any combination therapy, thus improving the efficacy versus toxicity balance of such a regimen.

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

Similar content being viewed by others

References

  1. Morales S, Lorenzo A, Ramos M, Ballesteros P, Mendez M, Almanza C, et al. Docetaxel plus epirubicin is a highly active, well-tolerated, first-line chemotherapy for metastatic breast cancer: results of a large, multicentre phase II study. Cancer Chemother Pharmacol. 2004;53(1):75–81.

    Article  CAS  PubMed  Google Scholar 

  2. Nishimura R, Rai Y, Matsuo F, Anan K, Hara S, Imamura S, et al. Neoadjuvant epirubicin/docetaxel (ET) concomitant chemotherapy for primary breast cancer with tumor diameter >=3.1 cm: results of the Kyushu ET therapy phase II trial. Anticancer Res. 2012;32(8):3259–65.

    CAS  PubMed  Google Scholar 

  3. Citron ML, Berry DA, Cirrincione C, Hudis C, Winer EP, Gradishar WJ, et al. Randomized trial of dose-dense versus conventionally scheduled and sequential versus concurrent combination chemotherapy as postoperative adjuvant treatment of node-positive primary breast cancer: first report of intergroup trial C9741/Cancer and Leukemia Group B Trial 9741. J Clin Oncol. 2003;21(8):1431–9.

    Article  CAS  PubMed  Google Scholar 

  4. Cottu PH, Extra JM, Espie M, Marolleau JP, DeRoquancourt A, Makke J, et al. High-dose sequential epirubicin and cyclophosphamide with peripheral blood stem cell support for advanced breast cancer: results of a phase II study. Br J Cancer. 2001;85(9):1240–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Friedrichs K, Holzel F, Janicke F. Combination of taxanes and anthracyclines in first-line chemotherapy of metastatic breast cancer: an interim report. Eur J Cancer. 2002;38(13):1730–8.

    Article  CAS  PubMed  Google Scholar 

  6. Piccart-Gebhart MJ. Mathematics and oncology: a match for life? J Clin Oncol. 2003;21(8):1425–8.

    Article  CAS  PubMed  Google Scholar 

  7. Salminen E, Korpela J, Varpula M, Asola R, Varjo P, Pyrhonen S, et al. Epirubicin/docetaxel regimen in progressive breast cancer-a phase II study. Anticancer Drugs. 2002;13(9):925–9.

    Article  CAS  PubMed  Google Scholar 

  8. Iliadis A, Barbolosi D. Optimizing drug regimens in cancer chemotherapy by an efficacy-toxicity mathematical model. Comput Biomed Res. 2000;33(3):211–26.

    Article  CAS  PubMed  Google Scholar 

  9. Swan GW. Role of optimal control theory in cancer chemotherapy. Math Biosci. 1990;101(2):237–84.

    Article  CAS  PubMed  Google Scholar 

  10. Martin R, Teo KL. Optimal control of drug administration in cancer chemotherapy. 1st ed. Singapore: World Scientific; 1994.

    Google Scholar 

  11. Klumpen HJ, Samer CF, Mathijssen RH, Schellens JH, Gurney H. Moving towards dose individualization of tyrosine kinase inhibitors. Cancer Treat Rev. 2011;37(4):251–60.

    Article  PubMed  Google Scholar 

  12. Zandvliet AS, Schellens JH, Beijnen JH, Huitema AD. Population pharmacokinetics and pharmacodynamics for treatment optimization in clinical oncology. Clin Pharmacokinet. 2008;47(8):487–513.

    Article  CAS  PubMed  Google Scholar 

  13. McClish DK, Roberts JD. Phase I studies of weekly administration of cytotoxic agents: implications of a mathematical model. Invest New Drugs. 2003;21(3):299–308.

    Article  CAS  PubMed  Google Scholar 

  14. Rombout F, Aarons L, Karlsson M, Man A, Mentré F, Nygren P, et al. Modelling and simulation in the development and use of anti-cancer agents: an underused tool? J Pharmacokinet Pharmacodyn. 2004;31(6):419–40.

    Article  PubMed  Google Scholar 

  15. Braun TM, Thall PF, Nguyen H, DeLima M. Simultaneously optimizing dose and schedule of a new cytotoxic agent. Clin Trials. 2007;4(2):113–24.

    Article  PubMed  Google Scholar 

  16. Meille C, Gentet JC, Barbolosi D, André N, Doz F, Iliadis A. New adaptive method for phase I trials in oncology. Clin Pharmacol Ther. 2008;83(6):873–81.

    Article  CAS  PubMed  Google Scholar 

  17. MATLAB. High-performance numeric computation and visualization software. 7.0 ed. Natick: The Math Works; 2004.

  18. Viens P, Roche H, Kerbrat P, Fumoleau P, Guastalla JP, Delozier T. Epirubicin-docetaxel combination in first-line chemotherapy for patients with metastatic breast cancer: final results of a dose-finding and efficacy study. Am J Clin Oncol. 2001;24(4):328–35.

    Article  CAS  PubMed  Google Scholar 

  19. De Jonge ME, van den Bongard HJ, Huitema AD, Mathot RA, Rosing H, Baas P, et al. Bayesian pharmacokinetically guided dosing of paclitaxel in patients with non-small cell lung cancer. Clin Cancer Res. 2004;10(7):2237–44.

    Article  PubMed  Google Scholar 

  20. Salinger DH, McCune JS, Ren AG, Shen DD, Slattery JT, Phillips B, et al. Real-time dose adjustment of cyclophosphamide in a preparative regimen for hematopoietic cell transplant: a Bayesian pharmacokinetic approach. Clin Cancer Res. 2006;12(16):4888–98.

    Article  CAS  PubMed  Google Scholar 

  21. Engels FK, Loos WJ, van der Bol JM, De Bruijn P, Mathijssen RH, Verweij J, et al. Therapeutic drug monitoring for the individualization of docetaxel dosing: a randomized pharmacokinetic study. Clin Cancer Res. 2011;17(2):353–62.

    Article  CAS  PubMed  Google Scholar 

  22. Sandstrom M, Lindman H, Nygren P, Lidbrink E, Bergh J, Karlsson MO. Model describing the relationship between pharmacokinetics and hematologic toxicity of the epirubicin-docetaxel regimen in breast cancer patients. J Clin Oncol. 2005;23(3):413–21.

    Article  CAS  PubMed  Google Scholar 

  23. Vainas O, Ariad S, Amir O, Mermershtain W, Vainstein V, Kleiman M, et al. Personalising docetaxel and G-CSF schedules in cancer patients by a clinically validated computational model. Br J Cancer. 2012;107(5):814–22.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Bissett D, Setanoians A, Cassidy J, Graham MA, Chadwick GA, Wilson P, et al. Phase I and pharmacokinetic study of taxotere (RP 56976) administered as a 24-hour infusion. Cancer Res. 1993;53(3):523–7.

    CAS  PubMed  Google Scholar 

  25. Bruno R, Vivier N, Vergniol JC, De Phillips SL, Montay G, Sheiner LB. A population pharmacokinetic model for docetaxel (Taxotere): model building and validation. J Pharmacokin Biopharm. 1996;24(2):153–72.

    Article  CAS  Google Scholar 

  26. Goncalves A, Viret F, Ciccolini J, Genre D, Gravis G, Giovanini M, et al. Phase I and pharmacokinetic study of escalating dose of docetaxel administered with granulocyte colony-stimulating factor support in adult advanced solid tumors. Clin Cancer Res. 2003;9(1):102–8.

    CAS  PubMed  Google Scholar 

  27. Vrignaud P, Eghbali H, Hoerni B, Iliadis A, Robert J. Pharmacokinetics and metabolism of epirubicin during repetitive courses of administration in hodgkin’s patients. Eur J Cancer Clin Oncol. 1985;21(11):1307–13.

    Article  CAS  PubMed  Google Scholar 

  28. Venturini M, Michelotti A, Papaldo P, DelMastro L, Bergaglio M, Lionetto R, et al. Identification of the highest dose of docetaxel associable with active doses of epirubicin. Results from a dose-finding study in advanced breast cancer patients. Ann Oncol. 2001;12(8):1097–106.

    Article  CAS  PubMed  Google Scholar 

  29. Meille C, Iliadis A, Barbolosi D, Frances N, Freyer G. An interface model for dosage adjustment connects hematotoxicity to pharmacokinetics. J Pharmacokinet Pharmacodyn. 2008;35(6):619–33.

    Article  CAS  PubMed  Google Scholar 

  30. Carmo-Pereira J, Costal FO, Miles DW, Henriques E, Richards MA, Rubens RD. High-dose epirubicin as primary chemotherapy in advanced breast carcinoma: a phase II study. Cancer Chemother Pharmacol. 1991;27(5):394–6.

    Article  CAS  PubMed  Google Scholar 

  31. Extra JM, Rousseau F, Bruno R, Clavel M, LeBail N, Marty M. Phase I and pharmacokinetic study of Taxotere (RP 56976; NSC 628503) given as a short intravenous infusion. Cancer Res. 1993;53(5):1037–42.

    CAS  PubMed  Google Scholar 

  32. Fabi A, Papaldo P, Pino MS, Ferretti G, Carlini P, Pacetti U, et al. Epirubicin plus docetaxel in metastatic breast cancer: escalating dose does not improve efficacy. A phase II study. Anticancer Res. 2004;24(3b):1963–8.

    CAS  PubMed  Google Scholar 

  33. Feld R, Wierzbicki R, Walde PLD, Shepherd FA, Evans WK, Gupta S, et al. Phase I–II study of high-dose epirubicin in advanced non-small-cell lung cancer. J Clin Oncol. 1992;10(2):297–303.

    CAS  PubMed  Google Scholar 

  34. Friberg LE, Henningsson A, Maas H, Nguyen L, Karlsson MO. Model of chemotherapy-induced myelosuppression with parameter consistency across drugs. J Clin Oncol. 2002;20(24):4713–21.

    Article  PubMed  Google Scholar 

  35. Laurence V, Merabet FB, Cottu PH, Cuvier C, Espie M, Marty M. Dose-dense docetaxel and epirubicine combination in advanced breast cancer: preliminary results. Ann Oncol. 2000;11(4):28.

    Google Scholar 

  36. Milla-Santos A, Milla L, Rallo L, Solano V. High-dose epirubicin plus docetaxel at standard dose with lenograstim support as first-line therapy in advanced breast cancer. Am J Clin Oncol. 2001;24(2):138–42.

    Article  CAS  PubMed  Google Scholar 

  37. Miller DK, Munshi N, Loesch D, Einhorn LE, Sledge GW. A phase II trial of high dose epirubicin in patients with advanced breast carcinoma. Cancer. 2000;88(2):375–80.

    Article  CAS  PubMed  Google Scholar 

  38. Paciucci PA, Raptis G, Bleiweiss I, Weltz C, Lehrer D, Gurry R. Neo-adjuvant therapy with dose-dense docetaxel plus short-term filgrastim rescue for locally advanced breast cancer. Anticancer Drugs. 2002;13(8):791–5.

    Article  CAS  PubMed  Google Scholar 

  39. Pagani O, Sessa C, Nole F, Crivellari D, Lombardi D, Thurlimann B, et al. Epidoxorubicin and docetaxel as first-line chemotherapy in patients with advanced breast cancer: a multicentric phase I–II study. Ann Oncol. 2000;11(8):985–91.

    Article  CAS  PubMed  Google Scholar 

  40. Polyzos A, Tsavaris N, Kosmas C, Gogas H, Vadiaka M, Markopoulos C, et al. Docetaxel and epirubicin supported by granulocyte colony-stimulating factor first-line in advanced breast cancer. Anticancer Res. 2003;23(3C):2917–23.

    CAS  PubMed  Google Scholar 

  41. Tomiak E, Piccart MJ, Kerger J, Lips S, Awada A, De Valeriola D, et al. Phase I study of docetaxel administered as a 1-hour intravenous infusion on a weekly basis. J Clin Oncol. 1994;12(7):1458–67.

    CAS  PubMed  Google Scholar 

  42. Ostby I, Kvalheim G, Rusten LS, Grottum P. Mathematical modeling of granulocyte reconstitution after high-dose chemotherapy with stem cell support: effect of post-transplant G-CSF treatment. J Theor Biol. 2004;231(1):69–83.

    Article  CAS  PubMed  Google Scholar 

  43. Roskos LK, Lum P, Lockbaum P, Schwab G, Yang BB. Pharmacokinetic/pharmacodynamic modeling of pegfilgrastim in healthy subjects. J Clin Pharmacol. 2006;46(7):747–57.

    Article  CAS  PubMed  Google Scholar 

  44. Matis JH, Kiffe TR, Parthasarathy PR. On the cumulants of population size for the stochastic power law logistic model. Theor Popul Biol. 1998;53(1):16–29.

    Article  CAS  PubMed  Google Scholar 

  45. Burstein HJ, Manola J, Younger J, Parker LM, Bunnell CA, Scheib R, et al. Docetaxel administered on a weekly basis for metastatic breast cancer. J Clin Oncol. 2000;18(6):1212–9.

    CAS  PubMed  Google Scholar 

  46. Palmeri L, Vaglica M, Palmeri S. Weekly docetaxel in the treatment of metastatic breast cancer. Ther Clin Risk Manag. 2008;4(5):1047–59.

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgments

The authors would like to thank R. Maraval-Gaget, study coordinator at the Centre Hospitalier Lyon Sud, and Dr. B. Tranchand for their involvement in the MODEL1 study, as well as Prof. J. Guitton for the analytical assays. We also thank the Chugai Pharmaceutical Company for sharing unpublished data on file for G-CSF.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Athanassios Iliadis.

Ethics declarations

Conflicts of interest

Christophe Meille, Dominique Barbolosi, Joseph Ciccolini, Gilles Freyer, and Athanassios Iliadis have no conflicts of interest to declare.

Funding

No funding has been received for the conduct of this study and/or preparation of this manuscript.

Appendix

Appendix

1.1 Pharmacokinetic Modeling

The PKs of the two drugs were described by an L—compartment model:

$$ \begin{array}{*{20}l} {\dot{c}_{1} \left( t \right) = - k_{e} \cdot c_{1} \left( t \right) - \sum\limits_{s = 2}^{L} {k_{1s} \cdot \left[ {c_{1} \left( t \right) - c_{s} \left( t \right)} \right]} + {{u\left( t \right)} \mathord{\left/ {\vphantom {{u\left( t \right)} {V_{1} }}} \right. \kern-0pt} {V_{1} }}} \hfill & {\quad c_{1} \left( 0 \right) = 0} \hfill & {} \hfill \\ {\dot{c}_{s} \left( t \right) = k_{s1} \cdot \left[ {c_{1} \left( t \right) - c_{s} \left( t \right)} \right]} \hfill & {\quad c_{s} \left( 0 \right) = 0} \hfill & {\quad s = 2:L} \hfill \\ \end{array} $$
(1)

where \( c_{1} \left( t \right) \) and \( c_{s} \left( t \right) \) are concentrations in the first and \( s - th \) peripheral compartment, respectively. The function \( u\left( t \right) \) describes the administration protocol corresponding to intermittent intravenous infusions:

$$ u\left( t \right) = \sum\limits_{j = 1}^{{n_{\text{D}} }} {\sum\limits_{i = 1}^{{n_{\text{A}} }} {\frac{{d_{ij} }}{{T_{0} }} \cdot \left[ {H\left( {t - t_{ij} } \right) - H\left( {t - t_{ij} - T_{0} } \right)} \right]} } $$
(2)

with

$$ t_{ij} = t_{0} + \tau_{1} \cdot \left( {i - 1} \right) + \tau_{2} \cdot \left( {j - 1} \right), $$

where index \( i \) is the administration number (\( n_{\text{A}} \) administrations per day), index \( j \) is the day number (\( n_{\text{D}} \) days of treatment per cycle), \( H\left( t \right) \) is the Heaviside step function (equal to 1 for \( t \ge 0 \) and 0 otherwise), \( T_{0} \) is the infusion duration, \( t_{0} \) is a time lag, and \( \tau_{1} \) and \( \tau_{2} \) are periods of repeating the administration within a day and repeating the days of treatment per cycle, respectively. The \( n_{\text{A}} \times n_{\text{D}} \) drug doses \( d_{ij} \) were compiled in vector form \( \underline{d} \). The schedule characteristics were also compiled in the vector form \( \underline{t}^{ * } = \left[ {\begin{array}{*{20}c} {T_{0} } & {t_{0} } & {\tau_{1} } & {\tau_{2} } \\ \end{array} } \right]^{T} \). The input function \( u\left( t \right) \) could then be stated as \( u\left( {t,\underline{d} ,\underline{t}^{ * } } \right) \) since depending on \( t \), \( \underline{d} \) and \( \underline{t}^{ * } \). To specify PKs for docetaxel and epirubicin implied in the combination, doses and concentrations were superscripted \( \underline{d}^{\left( D \right)} \), \( c^{\left( D \right)} \left( t \right) \) and \( \underline{d}^{\left( E \right)} \), \( c^{\left( E \right)} \left( t \right) \), respectively.

To prevent DLT events specific to long infusions [24], the two drugs were administered according to the following time schedule, as in the work of Viens et al. [18]: epirubicin was administered first (\( t_{0}^{\left( E \right)} = 0{\text{ h}} \)) as a 15-min (\( T_{0}^{\left( E \right)} = 0.25{\text{ h}} \)) constant rate infusion, and 1 h later (\( t_{0}^{\left( D \right)} = 1{\text{ h}} \)) docetaxel was administered as a 1-h (\( T_{0}^{\left( D \right)} = 1{\text{ h}} \)) constant rate infusion. This scheme, separated by 2-h discontinuation intervals (\( \tau_{1} = 2{\text{ h}} \)), was allowed to be repeated three times a day (\( n_{\text{A}} = 3 \)) for the first 3 days of the cycle (\( \begin{array}{*{20}c} {\tau_{2} = 1{\text{ d}}} & {n_{\text{D}} = 3} \\ \end{array} \)). These data completed the schedule component \( \underline{t}^{ * } \) in the input function \( u\left( {t,\underline{d} ,\underline{t}^{ * } } \right) \) (Eq. 2). Therefore, the number of doses \( \underline{d} = d_{ij} \) to be optimized was \( n_{A} \times n_{D} \times \left( {2{\text{ drugs}}} \right) = 18 \).

In population analyses involving these drugs either as single agents [2527] or in combination [22], three-compartment models (\( L = 3 \)) were used to describe the docetaxel and epirubicin kinetic profiles. Assuming no PK interaction [28], these models were also used to describe PKs in the present combination. Typical values of the two drugs are presented in Table 1.

1.2 PK/PD Interface Modeling

The interface model first introduced by Meille et al. [29] has been used to dynamically connect PKs and PDs. The interface model receives as input the PK levels of the administered drug in the central compartment, \( c_{1} \left( t \right) \), and supplies as output the exposure variable, \( y\left( t \right) \), linked to the PD response. This model was described by the following non-linear differential equation:

$$ \begin{array}{*{20}l} {\dot{y}\left( t \right) = - \alpha \cdot \exp \left[ { - y\left( t \right)} \right] \cdot y\left( t \right) + \beta \cdot \left[ {c_{1} \left( t \right) - \zeta } \right] \cdot H\left[ {c_{1} \left( t \right) - \zeta } \right]} \hfill & {} \hfill & {y\left( 0 \right) = 0} \hfill \\ \end{array} $$
(3)

where parameters \( \alpha \), \( \beta \), and \( \zeta \) are positive adjustable parameters. Exposures and parameters were superscripted by \( \left( D \right) \) and \( \left( E \right) \) for docetaxel and epirubicin, respectively.

1.3 Dynamic Safety Modeling

The dynamic model is semi-mechanistic and describes the toxicity induced by anticancer drugs on the hematopoietic chain. Because of the drugs used and the background knowledge from other studies [1, 5], only neutropenia was considered in the dynamic safety model. Initially proposed by Meille et al. [29], this model was slightly modified to incorporate the cytotoxic effect of drugs in combination and the concomitant administration of G-CSF. The dynamic safety model is described by the equations:

$$ \begin{array}{*{20}l} {\dot{w}_{1} \left( t \right) = \gamma \cdot {\lambda \mathord{\left/ {\vphantom {\lambda \omega }} \right. \kern-0pt} \omega } \cdot w_{0} \cdot \varPhi \left[ {w\left( t \right),\varphi } \right] - \gamma \cdot w_{1} \left( t \right) - N\left[ {y\left( t \right),\nu } \right] \cdot w_{1} \left( t \right)} \hfill & {\quad w_{1} \left( 0 \right) = {\lambda \mathord{\left/ {\vphantom {\lambda \omega }} \right. \kern-0pt} \omega } \cdot w_{0} } \hfill \\ {\dot{w}_{2} \left( t \right) = \gamma \cdot \left[ {w_{1} \left( t \right) - w_{2} \left( t \right)} \right] - N\left[ {y\left( t \right),\nu } \right] \cdot w_{2} \left( t \right)} \hfill & {\quad w_{2} \left( 0 \right) = {\lambda \mathord{\left/ {\vphantom {\lambda \omega }} \right. \kern-0pt} \omega } \cdot w_{0} } \hfill \\ {\dot{w}_{3} \left( t \right) = \omega \cdot w_{2} \left( {t - \tau } \right) - \lambda \cdot w_{3} \left( t \right)} \hfill & {\quad w_{3} \left( { - \tau \le t \le 0} \right) = w_{0} } \hfill \\ {\dot{w}\left( t \right) = \lambda \cdot \left\{ { \, M\left[ {z\left( t \right),\mu } \right] \cdot w_{3} \left( t \right) - w\left( t \right) \, } \right\}} \hfill & {\quad w\left( 0 \right) = w_{0}} \hfill \\ \end{array} $$
(4)

States \( w_{1} \left( t \right) \) and \( w_{2} \left( t \right) \) are two forms of cells corresponding to two consecutive steps of the maturation chain. Both disappear with loss rate \( \gamma \). States \( w_{3} \left( t \right) \) and \( w\left( t \right) \) represent circulating ANC, which disappear from the blood with rate constant \( \lambda \). Because of the maturation times in the hematopoietic chain, a change in the progenitor levels resulted in a change in ANC \( w_{3} \left( t \right) \) only after a time \( \tau \). The rate constant \( \omega \) is not structurally identifiable and therefore was set at \( \omega = 1{\text{ day}}^{ - 1} \).

The production rate of cells from the progenitor cells was assumed to be regulated by a homeostatic mechanism \( \varPhi \left[ {w\left( t \right),\varphi } \right] \), controlled by the mature neutrophil levels, \( w\left( t \right) \) with parameter \( \varphi \):

$$ \varPhi \left[ {w\left( t \right),\varphi } \right] = \left[ {{{w\left( t \right)} \mathord{\left/ {\vphantom {{w\left( t \right)} {w_{0} }}} \right. \kern-0pt} {w_{0} }}} \right]^{ - \varphi } $$
(5)

Circulating in the central compartment, anticancer drug levels \( c_{1} \left( t \right) \) were assumed to result in exposure \( y\left( t \right) \) that disturbs the neutrophil lineage by a direct hematological toxicity mechanism with parameters \( \nu \):

$$ N\left[ {y\left( t \right),\nu } \right] = \nu^{\left( D \right)} \cdot y^{\left( D \right)} \left( t \right) + \nu^{\left( E \right)} \cdot y^{\left( E \right)} \left( t \right) + \nu_{0} \cdot y^{\left( D \right)} \left( t \right) \cdot y^{\left( E \right)} \left( t \right) $$
(6)

Finally, the stimulation mechanism of G-CSF (4) is represented by the function \( M\left[ {z\left( t \right),\mu } \right] \) with parameters \( \mu \).

Typical parameter values were obtained by modeling the ANC profiles from the phase I study of Viens et al. [18]. This study was conducted on 65 patients receiving the drug combination every 3 weeks without G-CSF support. Hematological toxicity profiles from the above dataset were used to estimate parameters involved in the interface model and the dynamic model. For example, Fig. 2 illustrates the fitted model of a given profile. The averages of estimated parameters for all available profiles were considered as typical values (shown in Table 1). The simulated ANC profiles using these typical values were in agreement with profiles reported in the literature by several investigations [18, 24, 26, 3041].

Fig. 2
figure 2

ANC simulated profile fitted to experimental data for one patient receiving combination chemotherapy. ANC profile (solid line) and the associated 95 % confidence bounds (dashed lines) for three consecutive cycles for one patient receiving 75 mg/m2 of docetaxel and 90 mg/m2 of epirubicin from the phase I study of Viens et al. [18]. Full circles represent the experimental data. ANC absolute neutrophil count

1.4 PK/PD of Granulocyte Colony-Stimulating Factor

The dynamic safety model (4) includes the effect of concomitant administration of G-CSF growth factors [42, 43] administered by subcutaneous route. The PKs of G-CSF were described by the following compartment model:

$$ \begin{array}{*{20}c} {\dot{z}_{1} \left( t \right) = - h_{F} \cdot z_{3} \left( t \right) \cdot z_{1} \left( t \right) - h_{12} \cdot \left[ {z_{1} \left( t \right) - z_{2} \left( t \right)} \right] + {{\upsilon \left( t \right)} \mathord{\left/ {\vphantom {{\upsilon \left( t \right)} {V_{Z1} }}} \right. \kern-0pt} {V_{Z1} }}}&{z_{1} \left( 0 \right) = 0} \\ {\dot{z}_{2} \left( t \right) = h_{21} \cdot \left[ {z_{1} \left( t \right) - z_{2} \left( t \right)} \right]}&{z_{2} \left( 0 \right) = 0} \\ {\dot{z}_{3} \left( t \right) = w\left( t \right) - h_{0} \cdot z_{3} \left( t \right) - h_{F} \cdot z_{3} \left( t \right) \cdot z_{1} \left( t \right) + h_{B} \cdot z_{4} \left( t \right)} & {z_{3} \left( 0 \right) = {{w_{0} } \mathord{\left/ {\vphantom {{w_{0} } {h_{0} }}} \right. \kern-0pt} {h_{0} }}} \\ {\dot{z}_{4} \left( t \right) = h_{F} \cdot z_{3} \left( t \right) \cdot z_{1} \left( t \right) - h_{B} \cdot z_{4} \left( t \right)} & {z_{4} \left( 0 \right) = 0} \\ \end{array} $$
(7)

where \( z_{1} \left( t \right) \) and \( z_{2} \left( t \right) \) are concentrations in the central and peripheral compartment, respectively. The G-CSF elimination rate \( h_{F} \cdot z_{3} \left( t \right) \) was controlled by the \( z_{3} \left( t \right) \) state of the third differential equation. This equation uses the mature ANC levels \( w\left( t \right) \) as input, a constant rate \( h_{0} \) elimination, and an enzymatic-like mechanism describing the association of \( z_{3} \left( t \right) \) with \( z_{1} \left( t \right) \). This association yields a complex \( z_{4} \left( t \right) \) with rate \( h_{F} \), which is dissociated to \( z_{3} \left( t \right) \) and \( z_{1} \left( t \right) \) with rate \( h_{B} \). The distribution volume of the central compartment was denoted \( V_{Z1} \). The input function \( \upsilon \left( t \right) \) was a fractional Weibull-type absorption with parameters \( g_{1} \) and \( g_{2} \), and was given by the integral form:

$$ \int\limits_{0}^{t} {\upsilon \left( {t'} \right)dt'} = \sum\limits_{k = 1}^{{n_{G} }} {\delta_{k} \cdot \left[ {1 - \exp \left[ { - \left[ {g_{1} \cdot \left( {t - \bar{t}_{k} } \right)} \right]^{{g_{2} }} } \right]} \right]} $$
(8)

implying \( n_{G} \) G-CSF doses \( \delta_{k} \) administered at times \( \bar{t}_{k} \). Finally, the stimulation mechanism (4) was:

$$ M\left[ {z\left( t \right),\mu } \right] = 1 + \mu_{0} \cdot \left[ { \, 1 - \exp \left[ { - \left( {{{z_{1} \left( t \right)} \mathord{\left/ {\vphantom {{z_{1} \left( t \right)} {\mu_{1} }}} \right. \kern-0pt} {\mu_{1} }}} \right)^{{\mu_{2} }} } \right] \, } \right] $$
(9)

Therefore, the chain \( w\left( t \right) \to \left[ {h_{F} \cdot z_{3} \left( t \right)} \right] \to z_{1} \left( t \right) \to M\left[ {z\left( t \right),\mu } \right] \) resulted in an homeostatic mechanism on \( w\left( t \right) \) levels, externally controlled by the G-CSF administration scheme \( \upsilon \left( t \right) \); when \( w\left( t \right) \) decreases or increases, G-CSF levels are high or low because its clearance is increased or decreased, respectively.

Typical PK/PD parameter values for G-CSF were obtained from unpublished data on file (Chugai Pharmaceutical Co.). The data came from two phase I studies, where the granocyte was administered as single or repeated doses in 40 healthy volunteers. For both studies, several blood samples were drawn and G-CSF concentrations and ANC were evaluated. By pooling these raw data, a single average profile was obtained for G-CSF and ANC. The typical values of the G-CSF model were then estimated (Table 2). Figure 3 illustrates the fit of the G-CSF model on the naively pooled profiles.

Fig. 3
figure 3

G-CSF and ANC simulated profiles fitted to experimental data obtained in healthy volunteers. G-CSF (upper panel) and ANC (lower panel) experimental median profiles (full circles) from the data on file (Chugai Pharmaceutical Co.), fitted by the mathematical model (solid lines) and the associated 95 % confidence bounds (dashed lines). From time zero to day 10, the single-dose administration, and from day 10 to day 18, the repeated dose administration of G-CSF. G-CSF granulocyte colony-stimulating factors, ANC absolute neutrophil count

1.5 Hematological Constraints

The following clinical constraints on acceptable ANC levels were applied to the dynamic model:

  1. 1.

    “ANC must be maintained higher than the fixed level \( W_{\text{D}} \)”, translated by:

    $$ w\left( t \right) \ge W_{\text{D}} $$
    (10)
  2. 2.

    “Neutropenia should not exceed the period of time \( T_{\text{U}} \)”. To prevent the ANC from staying too long below a fixed level \( W_{\text{U}} \), the time \( t_{\text{U}} \) over which the ANC remains lower than \( W_{\text{U}} \) should comply with:

    $$ t_{\text{U}} \left[ {w\left( t \right) \le W_{\text{U}} } \right] \le T_{\text{U}} $$
    (11)

    with \( W_{\text{U}} > W_{\text{D}} \) usually. Chronometer \( t_{\text{U}} \left( t \right) \) was evaluated from the following differential equation:

    $$ \begin{array}{*{20}l} {\frac{{{\text{d}}t_{\text{U}} \left( t \right)}}{{{\text{d}}t}} = H\left[ {W_{\text{U}} - w\left( t \right)} \right]} \hfill & {t_{\text{U}} \left( 0 \right) = 0} \hfill \\ \end{array} $$
    (12)
  3. 3.

    “ANC levels at the end of the cycle, just before the next administration, should be higher than the minimum level \( W_{0} \)”:

    $$ w\left( T \right) \ge W_{0} $$
    (13)

1.6 Modeling Antitumor Efficacy

A disturbed Gompertzian tumor growth model described tumor fate in the presence of anticancer agents [44]:

$$ \begin{array}{*{20}c} {\frac{{{\text{d}}n\left( t \right)}}{{{\text{d}}t}} = \rho \cdot n\left( t \right) \cdot \ln \left[ {{\theta \mathord{\left/ {\vphantom {\theta {n\left( t \right)}}} \right. \kern-0pt} {n\left( t \right)}}} \right] - \kappa \cdot f\left( {c_{L}^{\left( D \right)} ,\;c_{L}^{\left( E \right)} } \right) \cdot n\left( t \right)} & {\quad n\left( 0 \right) = n_{0} } \\ \end{array} $$
(14)

where \( n\left( t \right) \) is the number of tumor cells at time \( t \), and \( n_{0} \) the initial tumor size. Parameters \( \rho \) and \( \theta \) are the tumor proliferation rate and the largest tumor mass, respectively. In the cell-loss term, constant \( \kappa \) and term \( f\left( {c_{L}^{\left( D \right)} ,c_{L}^{\left( E \right)} } \right) \) represent the global potency and effective concentration of combined drugs in the tumor site, respectively. The effective concentration term has a Weibull-like structure:

$$ f\left( {c_{L}^{\left( D \right)} ,c_{L}^{\left( E \right)} } \right) = 1 - \exp \left[ { - \left( {\underline{q}^{T} \cdot B \cdot \underline{q} } \right)^{{{\xi \mathord{\left/ {\vphantom {\xi 2}} \right. \kern-0pt} 2}}} } \right] $$
(15)

with \( {q}^{\left({D}\right)} = {\left( {c_{L}^{\left( D \right)} - C_{MIN}^{\left( D \right)} } \right) \cdot H\left( {c_{L}^{\left( D \right)} - C_{MIN}^{\left( D \right)} } \right)}, \, {q}^{\left({E}\right)} = {\left( {c_{L}^{\left( E \right)} - C_{MIN}^{\left(E \right)} } \right) \cdot H\left( {c_{L}^{\left( E \right)} - C_{MIN}^{\left( E \right)} } \right)}\) and \(\underline{q} = {\left[ {q^{\left( D \right)}\, q^{\left( E \right)} } \right]}^T\). In this formulation, \( \xi \) is a shape parameter, the array \( B \) weights the relative potency of combined drugs, and concentrations \( C_{MIN} \) are thresholds below which no tumor cells were killed. Concentrations \( c_{L} \left( t \right) \) in the deepest PK compartment were considered as the effective drug concentrations. As \( n\left( t \right) \) depends on concentration of drugs, it also depends on the administration protocol and could therefore be stated by \( n\left( {t,\underline{d} ,\underline{t}^{ * } } \right) \).

Typical values of these parameters are reported in Table 1. The values of \( \rho \) and \( \theta \) correspond to a doubling time of 2 months and a maximum tumor weight of 1 kg in human, respectively. The initial tumor size \( n_{0} \) was set according to the evaluation of tumor mass before cancer therapy, usually near 30 g.

The parameters involved in \( f\left( {c_{L}^{\left( D \right)} ,c_{L}^{\left( E \right)} } \right) \) were obtained from in vitro cytotoxicity experiments. These experiments performed with the MCF7 human breast cancer line generated data on the impact of docetaxel and epirubicin, used either alone or as a combination, on cancer cell proliferation. Tumor cells were incubated with 60 crossed concentration pairs of docetaxel (from \( 10^{ - 3} \) to 30 nM) and epirubicin (from \( 10^{ - 3} \) to 100 µM); the relative viability of cells was assessed at 72 h. The fit of the Weibull-like structure (15) to these data led to calculating \( B \), \( \xi \), and \( C_{MIN} \). These values are reported in Table 1. The off-diagonal element in \( B \) corresponds to a correlation coefficient equal to 0.6, indicating synergistic effect. Global potency \( \kappa \) was calibrated by fitting the simulated treatment responses by the above models to the responses reported by previously published clinical work [26, 30, 32, 33, 35, 36, 38, 40, 45, 46].

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Meille, C., Barbolosi, D., Ciccolini, J. et al. Revisiting Dosing Regimen Using Pharmacokinetic/Pharmacodynamic Mathematical Modeling: Densification and Intensification of Combination Cancer Therapy. Clin Pharmacokinet 55, 1015–1025 (2016). https://doi.org/10.1007/s40262-016-0374-7

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s40262-016-0374-7

Keywords

Navigation