Skip to main content
Log in

Optimal Therapy Scheduling Based on a Pair of Collaterally Sensitive Drugs

  • Original Article
  • Published:
Bulletin of Mathematical Biology Aims and scope Submit manuscript

Abstract

Despite major strides in the treatment of cancer, the development of drug resistance remains a major hurdle. One strategy which has been proposed to address this is the sequential application of drug therapies where resistance to one drug induces sensitivity to another drug, a concept called collateral sensitivity. The optimal timing of drug switching in these situations, however, remains unknown. To study this, we developed a dynamical model of sequential therapy on heterogeneous tumors comprised of resistant and sensitive cells. A pair of drugs (DrugA, DrugB) are utilized and are periodically switched during therapy. Assuming resistant cells to one drug are collaterally sensitive to the opposing drug, we classified cancer cells into two groups, \(A_\mathrm{R}\) and \(B_\mathrm{R}\), each of which is a subpopulation of cells resistant to the indicated drug and concurrently sensitive to the other, and we subsequently explored the resulting population dynamics. Specifically, based on a system of ordinary differential equations for \(A_\mathrm{R}\) and \(B_\mathrm{R}\), we determined that the optimal treatment strategy consists of two stages: an initial stage in which a chosen effective drug is utilized until a specific time point, T, and a second stage in which drugs are switched repeatedly, during which each drug is used for a relative duration (i.e., \(f \Delta t\)-long for DrugA and \((1-f) \Delta t\)-long for DrugB with \(0 \le f \le 1\) and \(\Delta t \ge 0\)). We prove that the optimal duration of the initial stage, in which the first drug is administered, T, is shorter than the period in which it remains effective in decreasing the total population, contrary to current clinical intuition. We further analyzed the relationship between population makeup, \(\mathcal {A/B} = A_\mathrm{R}/B_\mathrm{R}\), and the effect of each drug. We determine a critical ratio, which we term \(\mathcal {(A/B)}^{*}\), at which the two drugs are equally effective. As the first stage of the optimal strategy is applied, \(\mathcal {A/B}\) changes monotonically to \(\mathcal {(A/B)}^{*}\) and then, during the second stage, remains at \(\mathcal {(A/B)}^{*}\) thereafter. Beyond our analytic results, we explored an individual-based stochastic model and presented the distribution of extinction times for the classes of solutions found. Taken together, our results suggest opportunities to improve therapy scheduling in clinical oncology.

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
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11

Similar content being viewed by others

References

  • Amirouchene-Angelozzi N, Swanton C, Bardelli A (2017) Tumor evolution as a therapeutic target. Cancer Discov 7(8):805–817

    Article  Google Scholar 

  • Atuegwu NC, Arlinghaus LR, Li X, Chakravarthy AB, Abramson VG, Sanders ME, Yankeelov TE (2013) Parameterizing the logistic model of tumor growth by dw-mri and dce-mri data to predict treatment response and changes in breast cancer cellularity during neoadjuvant chemotherapy. Transl Oncol 6(3):256–264

    Article  Google Scholar 

  • Berry SJ, Coffey DS, Walsh PC, Ewing LL (1984) The development of human benign prostatic hyperplasia with age. J Urol 132(3):474–479

    Article  Google Scholar 

  • Boston EA, Gaffney EA (2011) The influence of toxicity constraints in models of chemotherapeutic protocol escalation. Math Med Biol J IMA 28(4):357–384

    Article  MathSciNet  MATH  Google Scholar 

  • Chen J-H, Kuo Y-H, Luh HP (2013) Optimal policies of non-cross-resistant chemotherapy on goldie and coldmans cancer model. Math Biosci 245(2):282–298

    Article  MathSciNet  MATH  Google Scholar 

  • Coldman A, Goldie J (1983) A model for the resistance of tumor cells to cancer chemotherapeutic agents. Math Biosci 65(2):291–307

    Article  MATH  Google Scholar 

  • Dhawan A, Nichol D, Kinose F, Abazeed ME, Marusyk A, Haura EB, Scott JG (2017) Collateral sensitivity networks reveal evolutionary instability and novel treatment strategies in ALK mutated non-small cell lung cancer. Sci Rep 7:1232

    Article  Google Scholar 

  • Gaffney E (2004) The application of mathematical modelling to aspects of adjuvant chemotherapy scheduling. J Math Biol 48(4):375–422

    Article  MathSciNet  MATH  Google Scholar 

  • Gaffney E (2005) The mathematical modelling of adjuvant chemotherapy scheduling: incorporating the effects of protocol rest phases and pharmacokinetics. Bull Math Biol 67(3):563–611

    Article  MathSciNet  MATH  Google Scholar 

  • Gerlee P (2013) The model muddle: in search of tumor growth laws. Cancer Res 73(8):2407–2411

    Article  Google Scholar 

  • Gillespie DT (1976) A general method for numerically simulating the stochastic time evolution of coupled chemical reactions. J Comput Phys 22(4):403–434

    Article  MathSciNet  Google Scholar 

  • Gillies RJ, Verduzco D, Gatenby RA (2012) Evolutionary dynamics unifies carcinogenesis and cancer therapy. Nat Rev Cancer 12(7):487

    Article  Google Scholar 

  • Goldie J (1982) Rationale for the use of alternating non-cross-resistant chemotherapy. Cancer Treat Rep 66:439–449

    Google Scholar 

  • Goldie J, Coldman A (1979) A mathematic model for relating the drug sensitivity of tumors to their spontaneous mutation rate. Cancer Treat Rep 63(11–12):1727–1733

    Google Scholar 

  • Goldie J, Coldman A (1983) Quantitative model for multiple levels of drug resistance in clinical tumors. Cancer Treat Rep 67(10):923–931

    Google Scholar 

  • Goldie JH, Coldman AJ (2009) Drug resistance in cancer: mechanisms and models. Cambridge University Press, Cambridge

    MATH  Google Scholar 

  • Holohan C, Van Schaeybroeck S, Longley DB, Johnston PG (2013) Cancer drug resistance: an evolving paradigm. Nat Rev Cancer 13(10):714

    Article  Google Scholar 

  • Hutchison DJ (1963) Cross resistance and collateral sensitivity studies in cancer chemotherapy. Adv Cancer Res 7:235–350

    Article  Google Scholar 

  • Imamovic L, Sommer MO (2013) Use of collateral sensitivity networks to design drug cycling protocols that avoid resistance development. Sci Transl Med 5(204):204ra132–204ra132

    Article  Google Scholar 

  • Jackson TL, Byrne HM (2000) A mathematical model to study the effects of drug resistance and vasculature on the response of solid tumors to chemotherapy. Math Biosci 164(1):17–38

    Article  MathSciNet  MATH  Google Scholar 

  • Jonsson VD, Blakely CM, Lin L, Asthana S, Matni N, Olivas V, Pazarentzos E, Gubens MA, Bastian BC, Taylor BS et al (2017) Novel computational method for predicting polytherapy switching strategies to overcome tumor heterogeneity and evolution. Sci Rep 7:44206

    Article  Google Scholar 

  • Katouli AA, Komarova NL (2011) The worst drug rule revisited: mathematical modeling of cyclic cancer treatments. Bull Math Biol 73(3):549–584

    Article  MathSciNet  MATH  Google Scholar 

  • Kaznatcheev A, Vander Velde R, Scott JG, Basanta D (2017) Cancer treatment scheduling and dynamic heterogeneity in social dilemmas of tumour acidity and vasculature. Br J Cancer 116(6):785

    Article  Google Scholar 

  • Komarova NL, Wodarz D (2003) Evolutionary dynamics of mutator phenotypes in cancer. Cancer Res 63(20):6635–6642

    Google Scholar 

  • Komarova NL, Wodarz D (2005) Drug resistance in cancer: principles of emergence and prevention. Proc Natl Acad Sci USA 102(27):9714–9719

    Article  Google Scholar 

  • Legler JM, Feuer EJ, Potosky AL, Merrill RM, Kramer BS (1998) The role of prostate-specific antigen (psa) testing patterns in the recent prostate cancer incidence declinein the united states. Cancer Causes Control 9(5):519–527

    Article  Google Scholar 

  • Marusyk A, Almendro V, Polyak K (2012) Intra-tumour heterogeneity: a looking glass for cancer? Nat Rev Cancer 12(5):323

    Article  Google Scholar 

  • Monzon FA, Ogino S, Hammond MEH, Halling KC, Bloom KJ, Nikiforova MN (2009) The role of kras mutation testing in the management of patients with metastatic colorectal cancer. Arch Pathol Lab Med 133(10):1600–1606

    Google Scholar 

  • Nichol D, Jeavons P, Fletcher AG, Bonomo RA, Maini PK, Paul JL, Gatenby RA, Anderson AR, Scott JG (2015) Steering evolution with sequential therapy to prevent the emergence of bacterial antibiotic resistance. PLoS Comput Biol 11(9):e1004493

    Article  Google Scholar 

  • Nichol D, Rutter J, Bryant C, Jeavons P, Anderson A, Bonomo R, Scott J (2017) Collateral sensitivity is contingent on the repeatability of evolution, bioRxiv, pp 185892

  • Palmer AC, Sorger PK (2017) Combination cancer therapy can confer benefit via patient-to-patient variability without drug additivity or synergy. Cell 171(7):1678–1691

    Article  Google Scholar 

  • Pisco AO, Brock A, Zhou J, Moor A, Mojtahedi M, Jackson D, Huang S (2013) Non-darwinian dynamics in therapy-induced cancer drug resistance. Nat Commun 4:2467

    Article  Google Scholar 

  • Rejniak KA (2005) A single-cell approach in modeling the dynamics of tumor microregions. Math Biosci Eng 2(3):643–655

    Article  MathSciNet  MATH  Google Scholar 

  • Scheel C, Weinberg RA (2011) Phenotypic plasticity and epithelial-mesenchymal transitions in cancer and normal stem cells? Int J Cancer 129(10):2310–2314

    Article  Google Scholar 

  • Scott J, Marusyk A (2017) Somatic clonal evolution: a selection-centric perspective. Biochimica et Biophysica Acta (BBA) Rev Cancer 1867(2):139–150

    Article  Google Scholar 

  • Scott JG, Fletcher AG, Anderson AR, Maini PK (2016) Spatial metrics of tumour vascular organisation predict radiation efficacy in a computational model. PLoS Comput Biol 12(1):e1004712

    Article  Google Scholar 

  • Swanson KR, Rockne RC, Claridge J, Chaplain MA, Alvord EC, Anderson AR (2011) Quantifying the role of angiogenesis in malignant progression of gliomas: in silico modeling integrates imaging and histology. Cancer Res 71(24):7366–7375

    Article  Google Scholar 

  • Thomas A, El Rouby S, Reed JC, Krajewski S, Silber R, Potmesil M, Newcomb EW (1996) Drug-induced apoptosis in b-cell chronic lymphocytic leukemia: relationship between p53 gene mutation and bcl-2/bax proteins in drug resistance. Oncogene 12(5):1055–1062

    Google Scholar 

  • Tomasetti C, Levy D (2010) An elementary approach to modeling drug resistance in cancer. Math Biosci Eng MBE 7(4):905

    Article  MathSciNet  MATH  Google Scholar 

  • Werner B, Scott JG, Sottoriva A, Anderson AR, Traulsen A, Altrock PM (2016) The cancer stem cell fraction in hierarchically organized tumors can be estimated using mathematical modeling and patient-specific treatment trajectories. Cancer Res 76(7):1705–1713

    Article  Google Scholar 

  • Wilson WH, Teruya-Feldstein J, Fest T, Harris C, Steinberg SM, Jaffe ES, Raffeld M (1997) Relationship of p53, bcl-2, and tumor proliferation to clinical drug resistance in non-hodgkin9s lymphomas. Blood 89(2):601–609

    Google Scholar 

  • Zhao B, Sedlak JC, Srinivas R, Creixell P, Pritchard JR, Tidor B, Lauffenburger DA, Hemann MT (2016) Exploiting temporal collateral sensitivity in tumor clonal evolution. Cell 165(1):234–246

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jacob G. Scott.

Appendices

Appendix A: Derivations of explicit expressions

1.1 A.1: Details of Equations (4), (5), (6), (7), (9), (10) and (12)

  1. 1.

    \(T_\mathrm{max}\): Equation (4)

\(T_\mathrm{max}\) is a minimum point of \(C_p(t)\) (from (3)). Therefore,

$$\begin{aligned}&\ C_p'(T_\mathrm{max})=0 \\&\quad \Leftrightarrow \ -(g-s)\left( \displaystyle \frac{r-s}{g+r-s}C_\mathrm{S}^0 \right) e^{-(g-s) \ T_\mathrm{max}} \\&\quad + r \left( \displaystyle \frac{g \ (C_\mathrm{S}^0 + C_\mathrm{R}^0) + (r - s) \ C_\mathrm{R}^0}{g+r-s} \right) e^{r \ T_\mathrm{max}} = 0 \\&\quad \Leftrightarrow \ (g-s)\left( \displaystyle \frac{r-s}{g+r-s}C_\mathrm{S}^0 \right) e^{-(g-s) \ T_\mathrm{max}} \\&\quad = r \left( \displaystyle \frac{g \ (C_\mathrm{S}^0 + C_\mathrm{R}^0) + (r - s) \ C_\mathrm{R}^0}{g+r-s} \right) e^{r \ T_\mathrm{max}} \\&\quad \Leftrightarrow \ (g-s)(r-s)C_\mathrm{S}^0 \ e^{-(g-s) \ T_\mathrm{max}} = r \left( (g \ (C_\mathrm{S}^0 + C_\mathrm{R}^0) + (r - s) \ C_\mathrm{R}^0) \right) e^{r \ T_\mathrm{max}} \\&\quad \Leftrightarrow \ e^{(g+r-s)T_\mathrm{max}}=\displaystyle \frac{(g-s)(r-s)C_\mathrm{S}^0}{r \left( g \ (C_\mathrm{S}^0 + C_\mathrm{R}^0) + (r - s) \ C_\mathrm{R}^0 \right) } \\&\quad \Leftrightarrow \ e^{(g+r-s)T_\mathrm{max}}=\displaystyle \frac{(g-s)(r-s)}{r (g(\mathcal {(R/S)}_0+1) + (r-s)\mathcal {(R/S)}_0)} \\&\quad \Leftrightarrow \ T_\mathrm{max}=\frac{\ln \left[ \displaystyle \frac{(g-s)(r-s)}{r (g(\mathcal {(R/S)}_0+1) + (r-s)\mathcal {(R/S)}_0)} \right] }{g+r-s} \end{aligned}$$
  1. 2.

    \(T_\mathrm{min}\): Equation (5)

Let us consider the case of drug switch with DrugA being the “pre-switch” drug and DrugB being the “post-switch” drug. If, at a specific time point \(t_1\), cell population is decreasing faster by continuing DrugA-therapy than by changing drug to DrugB,

$$\begin{aligned} C_\mathrm{P}'(0 \ |p_A,\{B_\mathrm{R}(t_1),A_\mathrm{R}(t_1)\}) > C_\mathrm{P}'(0 \ |p_B,\{A_\mathrm{R}(t_1),B_\mathrm{R}(t_1)\}), \end{aligned}$$

from Eq. (3) where \(\left( \begin{array}{l} B_\mathrm{R}(t_1) \\ A_\mathrm{R}(t_1) \end{array} \right) = \left( \begin{array}{ll} e^{-(g_A - s_A) \ t_1} &{}\quad 0 \\ \displaystyle \frac{g_A \ (e^{r_A \ t_1} - e^{-(g_A - s_A) \ t_1})}{g_A+r_A-s_A} &{}\quad e^{r_A \ t_1} \end{array} \right) \left( \begin{array}{l} B_\mathrm{R}(0) \\ A_\mathrm{R}(0) \end{array} \right) \) evaluated from Eq. (2). Then,

$$\begin{aligned}&\ C_\mathrm{P}'(0 \ |p_A,\{B_\mathrm{R}(t_1),A_\mathrm{R}(t_1)\})> C_\mathrm{P}'(0 \ |p_B,\{A_\mathrm{R}(t_1),B_\mathrm{R}(t_1)\}) \\&\quad \Leftrightarrow \ -(g_A-s_A)\left( \displaystyle \frac{r_A-s_A}{g_A+r_A-s_A}B_\mathrm{R}(t_1) \right) \\&\quad + r_A \left( \displaystyle \frac{g_A \ (A_\mathrm{R}(t_1) + B_\mathrm{R}(t_1)) + (r_A - s_A) \ A_\mathrm{R}(t_1)}{g_A+r_A-s_A} \right) \\&\quad> -(g_B-s_B)\left( \displaystyle \frac{r_B-s_B}{g_B+r_B-s_B}A_\mathrm{R}(t_1) \right) \\&\quad + r_B \left( \displaystyle \frac{g_B \ (A_\mathrm{R}(t_1) + B_\mathrm{R}(t_1)) + (r_B - s_B) \ B_\mathrm{R}(t_1)}{g_B+r_B-s_B} \right) \\&\quad \Leftrightarrow \ r_A \ A_\mathrm{R}(t_1) + s_A \ B_\mathrm{R}(t_1)> r_B \ B_\mathrm{R}(t_1) + s_B \ A_\mathrm{R}(t_1) \\&\quad \Leftrightarrow \ \displaystyle \frac{A_\mathrm{R}(t_1)}{B_\mathrm{R}(t_1)}> \displaystyle \frac{r_B - s_A}{r_A - s_B} \\&\quad \Leftrightarrow \ \displaystyle \frac{\displaystyle \frac{g_A \ (e^{r_A \ t_1} - e^{-(g_A - s_A) \ t_1})}{g_A+r_A-s_A} \ B_\mathrm{R}(0)+ e^{r_A \ t_1} \ A_\mathrm{R}(0)}{e^{-(g_A - s_A) \ t_1} \ B_\mathrm{R}(0)}> \displaystyle \frac{r_B - s_A}{r_A - s_B} \\&\quad \Leftrightarrow \ \displaystyle \frac{g_A \ (e^{(g_A+r_A-s_A) \ t_1} - 1)}{g_A+r_A-s_A} + e^{(g_A+r_A-s_A) \ t_1} \ \mathcal {(A/B)}_0> \displaystyle \frac{r_B - s_A}{r_A - s_B} \\&\quad \Leftrightarrow \ e^{(g_A+r_A-s_A) \ t_1} \left( \displaystyle \frac{g_A}{g_A+r_A-s_A} + \mathcal {(A/B)}_0 \right) > \displaystyle \frac{r_B - s_A}{r_A - s_B} + \displaystyle \frac{g_A}{g_A+r_A-s_A} \\&\quad \Leftrightarrow \ t_1 < \frac{\ln \left[ \displaystyle \frac{(r_A-s_A)(r_B-s_A)+g_A(r_A+r_B-s_A-s_B)}{(r_A-s_B) (g_A + (g_A+r_A-s_A)\mathcal {(A/B)}_0)} \right] }{g_A+r_A-s_A} \\&\quad \left( = T_\mathrm{min}(\{s_A,r_A,g_A\},\{s_B,r_B\},\mathcal {(A/B)}_0) \right) . \end{aligned}$$

Similarly,

$$\begin{aligned} t_1>T_\mathrm{min}(\{s_A,r_A,g_A\},\{s_B,r_B\},\mathcal {(A/B)}_0), \end{aligned}$$

if and only if the population is dropping faster using DrugB than by continuing to use DrugA, and

$$\begin{aligned} t_1=T_\mathrm{min}(\{s_A,r_A,g_A\},\{s_B,r_B\},\mathcal {(A/B)}_0), \end{aligned}$$

if and only if the population is dropping at an equal rate with either drug.

The general form of \(T_\mathrm{min}\) is

$$\begin{aligned}&T_\mathrm{min}(\{s_1,r_1,g_1\},\{s_2,r_2\},\mathcal {(R/S)}_0)\\&\quad =\displaystyle \frac{\displaystyle \ln \left[ \displaystyle \frac{(r_1-s_1)(r_2-s_1)+g_1(r_1+r_2-s_1-s_2)}{(r_1-s_2) (g_1 + (g_1+r_1-s_1)\mathcal {(R/S)}_0)} \right] }{g_1+r_1-s_1}, \end{aligned}$$

where the parameters of “pre-switch” and “post-switch” drugs are \(\{s_1,r_1,g_1\}\) and \(\{s_2,r_2,g_2\}\), respectively, and initial population makeup, \(\mathcal {(R/S)}_0\), is the resistant cell population divided by the sensitive cell population for the “pre-switch” drug.

  1. 3.

    \(T_\mathrm{gap}\): Equations (6)–(7)

$$\begin{aligned}&T_\mathrm{gap}(\{s_1,r_1,g_1\},\{s_2,r_2\}) = T_\mathrm{max}(\{s_1,r_1,g_1\},\mathcal {(R/S)}_0)\\&\qquad -T_\mathrm{min}(\{s_1,r_1,g_1\},\{s_2,r_2\},\mathcal {(R/S)}_0) \\&\quad = \frac{\ln \left[ \displaystyle \frac{(g_1-s_1)(r_1-s_1)}{r_1 (g(\mathcal {(R/S)}_0+1) + (r_1-s_1)\mathcal {(R/S)}_0)} \right] }{g_1+r_1-s_1} \\&\qquad - \frac{\ln \left[ \displaystyle \frac{(r_1-s_1)(r_2-s_1)+g_1(r_1+r_2-s_1-s_2)}{(r_1-s_2) (g_1 + (g_1+r_1-s_1)\mathcal {(R/S)}_0)} \right] }{g_1+r_1-s_1} \\&\quad = \frac{\ln \left[ \displaystyle \frac{(g_1-s_1)(r_1-s_1)}{r_1 (g(\mathcal {(R/S)}_0+1) + (r_1-s_1)\mathcal {(R/S)}_0)} \ \displaystyle \frac{(r_1-s_2) (g_1 + (g_1+r_1-s_1)\mathcal {(R/S)}_0)}{(r_1-s_1)(r_2-s_1)+g_1(r_1+r_2-s_1-s_2)} \right] }{g_1+r_1-s_1} \\&\quad =\frac{\ln \left[ \displaystyle \frac{(g_1-s_1)(r_1-s_1)(r_1-s_2)}{r_1((r_1-s_1)(r_2-s_1)+g_1(r_1+r_2-s_1-s_2))} \right] }{g_1+r_1-s_1} \end{aligned}$$

And,

$$\begin{aligned}&T_\mathrm{min}(\{s_1,r_1,g_1\},\{s_2,r_2\},\mathcal {(R/S)}_0)< T_\mathrm{max}(\{s_1,r_1,g_1\},\mathcal {(R/S)}_0) \\&\quad \Leftrightarrow T_\mathrm{gap}(\{s_1,r_1,g_1\},\{s_2,r_2\})> 0 \\&\quad \Leftrightarrow \frac{\ln \left[ \displaystyle \frac{(g_1-s_1)(r_1-s_1)(r_1-s_2)}{r_1((r_1-s_1)(r_2-s_1)+g_1(r_1+r_2-s_1-s_2))} \right] }{g_1+r_1-s_1}> 0 \\&\quad \Leftrightarrow \ln \left[ \displaystyle \frac{(g_1-s_1)(r_1-s_1)(r_1-s_2)}{r_1((r_1-s_1)(r_2-s_1)+g_1(r_1+r_2-s_1-s_2))} \right]> 0 \\&\quad \Leftrightarrow \displaystyle \frac{(g_1-s_1)(r_1-s_1)(r_1-s_2)}{r_1((r_1-s_1)(r_2-s_1)+g_1(r_1+r_2-s_1-s_2))}> 1 \\&\quad \Leftrightarrow (g_1-s_1)(r_1-s_1)(r_1-s_2)> r_1((r_1-s_1)(r_2-s_1)+g_1(r_1+r_2-s_1-s_2)) \\&\quad \Leftrightarrow g_1 \ s_1 \ s_2 - s_1^2 \ s_2 + r_1 \ s_1 \ s_2> g_1 \ r_1 \ r_2 - s_1 \ r_1 \ r_2 + r_1^2 \ r_2 \\&\quad \Leftrightarrow (g_1 + r_1 - s_1)(s_1 \ s_2 - r_1 \ r_2)> 0 \\&\quad \Leftrightarrow r_1 \ r_2 - s_1 \ s_2 >0 \quad \Leftrightarrow \quad r_1 \ r_2 < s_1 \ s_2 \end{aligned}$$

Similarly, \(T_\mathrm{gap}=0\) if and only if \(r_1 \ r_2 = s_1 \ s_2\), and \(T_\mathrm{gap}<0\) if and only if \(r_1 \ r_2 > s_1 \ s_2\).

  1. 4.

    \(\mathcal {A/B}\) at \(T_\mathrm{max}\) and \(T_\mathrm{min}\): Equations (9)–(10).

It is clear that

$$\begin{aligned} \mathcal {A/B}(T_\mathrm{min}^A)=\mathcal {A/B}(T_\mathrm{min}^B)=\displaystyle \frac{r_B-s_A}{r_A-s_B}, \end{aligned}$$

and

$$\begin{aligned} \mathcal {A/B}(T_\mathrm{max}^A)=\displaystyle \frac{-s_A}{r_A} \quad \text {and} \quad \mathcal {A/B}(T_\mathrm{max}^B)=\displaystyle \frac{r_B}{-s_B} \end{aligned}$$

by the expressions of \(A_\mathrm{R}(t)\), \(B_\mathrm{R}(t)\), \(T_\mathrm{max}\) and \(T_\mathrm{min}\) from Eqs. (2), (5) and (4).

Otherwise, it can be proved more simply using the concept of \(T_\mathrm{min}\) and \(T_\mathrm{max}\). Since \(C_\mathrm{S}'(t)+C_\mathrm{R}'(t)=s \ C_\mathrm{S}(t) + r \ C_\mathrm{R}(t)\), from the differential system (1), the derivatives of \(A_\mathrm{R}(t)+B_\mathrm{R}(t)\) are \(s_A \ B_\mathrm{R}(t) + r_A \ A_\mathrm{R}(t)\) and \(s_B \ A_\mathrm{R}(t) + r_B \ B_\mathrm{R}(t)\) under DrugA and DrugB, respectively. At \(T_\mathrm{min}\) (whether it is \(T_\mathrm{min}^A\) or \(T_\mathrm{min}^B\)), the derivatives of total populations are equivalent either under DrugA or under DrugB. Then,

$$\begin{aligned}&s_A \ B_R(T_\mathrm{min}) + r_A \ A_R(T_\mathrm{min}) = s_B \ A_R(T_\mathrm{min}) + r_B \ B_R(T_\mathrm{min}) \\&\displaystyle \frac{A_R(T_\mathrm{min})}{B_R(T_\mathrm{min})}=\displaystyle \frac{r_B-s_A}{r_A-s_B} \\&\mathcal {A/B}(T_\mathrm{min})=\displaystyle \frac{r_B-s_A}{r_A-s_B} \end{aligned}$$

Therefore,

$$\begin{aligned} \mathcal {A/B}(T_\mathrm{min}^A)=\mathcal {A/B}(T_\mathrm{min}^B)=\displaystyle \frac{r_B-s_A}{r_A-s_B}, \end{aligned}$$

Under DrugA at \(T_\mathrm{max}^A\), \(A_R'(t)+B_R'(t)=0\). Therefore,

$$\begin{aligned}&s_A \ B_R(T_\mathrm{max}^A) + r_A \ A_R(T_\mathrm{max}^A)=0 \\&\mathcal {A/B}(T_\mathrm{max}^A)=\displaystyle \frac{-s_A}{r_A}. \end{aligned}$$

Similarly, \(\mathcal {A/B}(T_\mathrm{max}^B)=\displaystyle \frac{r_B}{-s_B}\).

  1. 5.

    \(k^*\): Equation (12)

The sizes of the subpopulations after \(\Delta t\)-long therapy with DrugA started from initial population makeup of \(\mathcal {A/B}(0)=\mathcal {(A/B)}^*\) are

$$\begin{aligned} \left( \begin{array}{l} B_\mathrm{R}(\Delta t) \\ A_\mathrm{R}(\Delta t) \end{array} \right) = \left( \begin{array}{ll} e^{-(g_A - s_A) \ \Delta t} &{}\quad 0 \\ \displaystyle \frac{g_A \ (e^{r_A \ \Delta t} - e^{-(g_A - s_A) \ \Delta t})}{g_A+r_A-s_A} &{}\quad e^{r_A \ \Delta t} \end{array} \right) \left( \begin{array}{l} K \\ K \ \mathcal {(A/B)}^* \end{array} \right) \end{aligned}$$

derived from Eq. (2), with some constant K scaling population size. Then, the population makeup at the \(\Delta t\) and its derivative in terms of \(\Delta t\) are

$$\begin{aligned} \mathcal {(A/B)}_{\Delta t}:=\displaystyle \frac{A_\mathrm{R}(\Delta t)}{B_\mathrm{R}(\Delta t)}=\displaystyle \frac{g_A \ (e^{(g_A+r_A-s_A) \ \Delta t} - 1)}{g_A+r_A-s_A} + e^{(g_A+r_A-s_A) \ \Delta t} \mathcal {(A/B)}^*\\ \displaystyle \frac{d \ (\mathcal {(A/B)}_{\Delta t})}{d \ (\Delta t)}= g_A \ e^{(g_A+r_A-s_A) \ \Delta t} + (g_A+r_A-s_A) e^{(g_A+r_A-s_A) \ \Delta t} \mathcal {(A/B)}^* \end{aligned}$$

The time taken from \(t=\Delta t\) to reach back to the time of \(\mathcal {A/B}(t)=\mathcal {(A/B)}^*\) given DrugB is

$$\begin{aligned}&T_\mathrm{min}(\{s_B,r_B,g_B\},\{s_A,r_A\},1/\mathcal {(A/B)}_{\Delta t}) \\&\quad = \displaystyle \frac{\displaystyle \ln \left[ \displaystyle \frac{(r_B-s_B)(r_A-s_B)+g_B(r_B+r_A-s_B-s_A)}{(r_B-s_A) (g_B + (g_B+r_B-s_B)/\mathcal {(A/B)}_{\Delta t})} \right] }{g_B+r_B-s_B} \end{aligned}$$

from Eq. (5).

Then, the relative ratio between the periods of DrugA and DrugB, \(k'\), illustrated in Fig. 9, and its limit, \(k^*\), can be derived using:

$$\begin{aligned} k'&= \displaystyle \frac{\Delta t}{T_\mathrm{min}(\{s_B,r_B,g_B\},\{s_A,r_A\},1/\mathcal {(A/B)}_{\Delta t})} \\ k^*&= \displaystyle \lim _{\Delta t \rightarrow 0} k' = \displaystyle \lim _{\Delta t \rightarrow 0} \displaystyle \frac{(g_B+r_B-s_B) \ \Delta t}{\displaystyle \ln \left[ \displaystyle \frac{(r_B-s_B)(r_A-s_B)+g_B(r_B+r_A-s_B-s_A)}{(r_B-s_A) (g_B + (g_B+r_B-s_B)/\mathcal {(A/B)}_{\Delta t})} \right] } \\&= \displaystyle \lim _{\Delta t \rightarrow 0} \displaystyle \frac{(g_B+r_B-s_B) \ \Delta t}{-\displaystyle \ln \left[ \displaystyle g_B + (g_B+r_B-s_B)/\mathcal {(A/B)}_{\Delta t} \right] + \displaystyle \ln K} \\&\quad \text { with } K= \displaystyle \frac{(r_B-s_B)(r_A-s_B)+g_B(r_B+r_A-s_B-s_A)}{r_B-s_A} \\&= \displaystyle \lim _{\Delta t \rightarrow 0} \displaystyle \frac{g_B+r_B-s_B}{-\displaystyle \frac{d}{d \ (\Delta t)} \displaystyle \ln \left[ g_B + (g_B+r_B-s_B)/\mathcal {(A/B)}_{\Delta t} \right] } \\&\quad \text {by L'Hospital's rule} \\&= \displaystyle \lim _{\Delta t \rightarrow 0} \displaystyle \frac{g_B+r_B-s_B}{-\displaystyle \frac{(g_B+r_B-s_B)(-(\mathcal {(A/B)}_{\Delta t})^{-2})}{g_B + (g_B+r_B-s_B)/\mathcal {(A/B)}_{\Delta t}} \displaystyle \frac{d \ (\mathcal {(A/B)}_{\Delta t})}{d \ (\Delta t)}} \\&= \displaystyle \frac{g_B+r_B-s_B}{\displaystyle \frac{(g_B+r_B-s_B)/(\mathcal {(A/B)}^*)^2}{g_B + (g_B+r_B-s_B)/\mathcal {(A/B)}^*} (g_A+(g_A+r_A-s_A)\mathcal {(A/B)}^*)} \\&\quad \text { since, } \displaystyle \lim _{\Delta t \rightarrow 0} \mathcal {(A/B)}_{\Delta t} = \mathcal {(A/B)}^* \\&\quad \text { and } \displaystyle \lim _{\Delta t \rightarrow 0} \displaystyle \frac{d \ (\mathcal {(A/B)}_{\Delta t})}{d \ (\Delta t)}= g_A + (g_A+r_A-s_A) \mathcal {(A/B)}^* \\&= \displaystyle \frac{g_B \ \mathcal {(A/B)}^* + (g_B+r_B-s_B)}{g_A / \mathcal {(A/B)}^* + (g_A+r_A-s_A)} \\&= \displaystyle \frac{(r_A-s_B)((r_A-s_A)(r_B-s_A)+g_A(r_A+r_B-s_A-s_B))}{(r_B-s_A)((r_B-s_B)(r_A-s_B)+g_B(r_A+r_B-s_A-s_B))} \\&\quad \text { since, } \mathcal {(A/B)}^*=\displaystyle \frac{r_B-s_A}{r_A-s_B} \end{aligned}$$

1.2 A.2: Differential system of instantaneous drug switch

The goal of this section is to derive the simple differential equations of \(V=\{A_\mathrm{R},B_\mathrm{R}\}\) under instantaneous drug switch (Theorem A.8). For the sake of convenience, we want to use matrix operations and equations based on the vectors and matrices defined below.

Definition

\(\mathbb {D}_A:=\left( \begin{array}{ll} r_A &{} g_A \\ 0 &{} s_A - g_A \end{array} \right) \) , \(\mathbb {D}_B:=\left( \begin{array}{ll} s_B - g_B &{} 0 \\ g_B &{} r_B \end{array} \right) \), \(V(t):=\left( \begin{array}{l} A_\mathrm{R}(t) \\ B_\mathrm{R}(t) \end{array} \right) \),

$$\begin{aligned}&\mathbb {M}_A(t):=\left( \begin{array}{ll} e^{r_A \ t} &{} \displaystyle \frac{g_A \ (e^{r_A \ t} - e^{-(g_A-s_A) \ t})}{g_A+r_A-s_A} \\ 0 &{} e^{-(g_A - s_A) \ t} \end{array} \right) ,\\&\mathbb {M}_B(t):=\left( \begin{array}{ll} e^{-(g_B - s_B) \ t} &{} 0 \\ \displaystyle \frac{g_B \ (e^{r_B \ t} - e^{-(g_B-s_B) \ t})}{g_B+r_B-s_B} &{} e^{r_B \ t} \end{array} \right) , \\&\mathbb {A}_\epsilon :=\mathbb {M}_A(f \ \epsilon ), \mathbb {B}_\epsilon :=\mathbb {M}_B((1-f)\epsilon ), \\&\min \left[ V(t_1), V(t_2), \cdots , V(t_n)\right] :=\left( \begin{array}{l} \min \left[ A_\mathrm{R}(t_1), A_\mathrm{R}(t_2), \cdots , A_\mathrm{R}(t_n)\right] \\ \min \left[ B_\mathrm{R}(t_1), B_\mathrm{R}(t_2), \cdots , B_\mathrm{R}(t_n)\right] \end{array} \right) , \\&\max \left[ V(t_1), V(t_2), \cdots , V(t_n)\right] :=\left( \begin{array}{l} \max \left[ A_\mathrm{R}(t_1), A_\mathrm{R}(t_2), \cdots , A_\mathrm{R}(t_n)\right] \\ \max \left[ B_\mathrm{R}(t_1), B_\mathrm{R}(t_2), \cdots , B_\mathrm{R}(t_n)\right] \end{array} \right) . \end{aligned}$$

Proposition A.1

Using Drug A therapy:

$$\begin{aligned} V'(t) = \mathbb {D}_A \ V(t), \ V(t_0+\Delta t) = \mathbb {M}_A(\Delta t) \ V(t_0). \end{aligned}$$

Using Drug B therapy:

$$\begin{aligned} V'(t) = \mathbb {D}_B \ V(t), \ V(t_0+\Delta t) = \mathbb {M}_B(\Delta t) \ V(t_0). \end{aligned}$$

Proposition A.2

Both \(A_\mathrm{R}\) and \(B_\mathrm{R}\) are monotonic functions under either therapy. In the presence of Drug A, \(A_\mathrm{R}\) is increasing, and \(B_\mathrm{R}\) is decreasing. And, in the presence of Drug B, \(A_\mathrm{R}\) is decreasing, and \(B_\mathrm{R}\) is increasing.

Proposition A.3

\(\left. {\mathbb {A}_\epsilon } \right| _{\epsilon =0} = \left. {\mathbb {B}_\epsilon } \right| _{\epsilon =0} = I_2\) for all \(0 \le f \le 1\)

Proposition A.4

\(\left. {\displaystyle \frac{d}{d\epsilon } \mathbb {A}_\epsilon } \right| _{\epsilon =0} = f \ \mathbb {D}_A\), \(\left. {\displaystyle \frac{d}{d\epsilon } \mathbb {B}_\epsilon } \right| _{\epsilon =0} = (1-f) \mathbb {D}_B\) for all \(0 \le f \le 1\)

Lemma A.5

\(\displaystyle \lim _{\epsilon \rightarrow 0} \displaystyle \frac{\mathbb {B}_\epsilon \mathbb {A}_\epsilon -I_2}{\epsilon }= f \ \mathbb {D}_A + (1-f) \mathbb {D}_B\) for all \(0 \le f \le 1\)

Proof

$$\begin{aligned} \displaystyle \lim _{\epsilon \rightarrow 0} \displaystyle \frac{\mathbb {B}_\epsilon \mathbb {A}_\epsilon -I_2}{\epsilon }&=\displaystyle \lim _{\epsilon \rightarrow 0} \displaystyle \frac{\frac{d}{d\epsilon } (\mathbb {B}_\epsilon \mathbb {A}_\epsilon -I_2)}{\frac{d}{d\epsilon } \epsilon }&\text {(by L'Hospital's Rule)} \\&= \displaystyle \lim _{\epsilon \rightarrow 0} \displaystyle \frac{\frac{d\mathbb {B}_\epsilon }{d\epsilon } \mathbb {A}_\epsilon + \mathbb {B}_\epsilon \frac{d\mathbb {A}_\epsilon }{d\epsilon }}{1}&\\&=f \ \mathbb {D}_A + (1-f) \mathbb {D}_B&(\text {by Propositions}~ \hbox {A}.3-\hbox {A}.4) \end{aligned}$$

\(\square \)

Lemma A.6

\(\displaystyle \lim _{\epsilon \rightarrow 0} \displaystyle \frac{(\mathbb {B}_\epsilon \mathbb {A}_\epsilon )^n-I_2}{n \ \epsilon }= f \ \mathbb {D}_A + (1-f) \mathbb {D}_B\) for any positive integer, n, and for all \(0 \le f \le 1\)

Proof

Let \(F(n) :=\displaystyle \lim _{\epsilon \rightarrow 0} \displaystyle \frac{(\mathbb {B}_\epsilon \mathbb {A}_\epsilon )^n-I_2}{n \ \epsilon }\) and \(L :=f \ \mathbb {D}_A + (1-f) \mathbb {D}_B\).

Then, we need to prove that \(F(n) = L\) for \(n = 1,2,3,\ldots \)

If \(n=1\),

$$\begin{aligned} F(n)&= F(1) = L&(\text {by Lemma}~ \hbox {A}.5) \end{aligned}$$

Otherwise, if \(n \ge 2\) and \(F(m) = L\) for all \(1 \le m \le n-1\),

$$\begin{aligned} F(n)&= \displaystyle \lim _{\epsilon \rightarrow 0} \displaystyle \frac{(\mathbb {B}_\epsilon \mathbb {A}_\epsilon )^n-I_2}{n \ \epsilon } \\&=\displaystyle \lim _{\epsilon \rightarrow 0} \displaystyle \frac{((\mathbb {B}_\epsilon \mathbb {A}_\epsilon )^{n-1}-I_2) (\mathbb {B}_\epsilon \mathbb {A}_\epsilon )+(\mathbb {B}_\epsilon \mathbb {A}_\epsilon -I_2)}{n \ \epsilon } \\&= \displaystyle \frac{n-1}{n} \lim _{\epsilon \rightarrow 0} \displaystyle \frac{((\mathbb {B}_\epsilon \mathbb {A}_\epsilon )^{n-1}-I_2) (\mathbb {B}_\epsilon \mathbb {A}_\epsilon )}{(n-1) \ \epsilon } + \frac{1}{n}\displaystyle \lim _{\epsilon \rightarrow 0} \displaystyle \frac{\mathbb {B}_\epsilon \mathbb {A}_\epsilon -I_2}{\epsilon } \\&= \frac{n-1}{n}F(n-1) + \frac{1}{n}F(1) \\&= \frac{n-1}{n}L + \frac{1}{n}L \quad \qquad \text {(by the inductive assumption)} \\&= L \end{aligned}$$

Therefore, proved. \(\square \)

Lemma A.7

\(\displaystyle \lim _{\epsilon \rightarrow 0} \displaystyle \frac{\mathbb {A}_\epsilon (\mathbb {B}_\epsilon \mathbb {A}_\epsilon )^n-I_2}{(n+f) \ \epsilon }= \frac{(n+1)f}{n+f} \ \mathbb {D}_A + \frac{n(1-f)}{n+f} \mathbb {D}_B\) for any positive integer, n, and for all \(0 \le f \le 1\)

Proof

Using mathematical induction, if \(n=1\),

$$\begin{aligned}&\displaystyle \lim _{\epsilon \rightarrow 0} \displaystyle \frac{\mathbb {A}_\epsilon (\mathbb {B}_\epsilon \mathbb {A}_\epsilon )-I_2}{(1+f) \ \epsilon } \\&\quad = \displaystyle \frac{1}{1+f} \lim _{\epsilon \rightarrow 0} \displaystyle \frac{\mathbb {A}_\epsilon (\mathbb {B}_\epsilon \mathbb {A}_\epsilon -I_2)+(\mathbb {A}_\epsilon -I_2)}{\epsilon } \\&\quad = \displaystyle \frac{1}{1+f} \left[ \lim _{\epsilon \rightarrow 0} \mathbb {A}_\epsilon \lim _{\epsilon \rightarrow 0} \displaystyle \frac{\mathbb {B}_\epsilon \mathbb {A}_\epsilon -I_2}{\epsilon } + \lim _{\epsilon \rightarrow 0} \displaystyle \frac{\mathbb {A}_\epsilon -I_2}{\epsilon } \right] \\&\quad = \displaystyle \frac{1}{1+f} \left[ I_2 (f \ \mathbb {D}_A + (1-f) \mathbb {D}_B) + \left. {\displaystyle \frac{d}{d\epsilon } \mathbb {A}_\epsilon } \right| _{\epsilon =0} \right] ~ (\text {by Proposition}~\hbox {A}.3~\text {and Lemma}~\hbox {A}.5) \\&\quad = \displaystyle \frac{1}{1+f} \left[ (f \ \mathbb {D}_A + (1-f) \mathbb {D}_B) + f \ \mathbb {D}_A \right] \qquad (\text {by Proposition}~ \hbox {A}.4) \\&\quad = \displaystyle \frac{2 \ f}{1+f} \mathbb {D}_A + \displaystyle \frac{1-f}{1+f} \mathbb {D}_B \quad \qquad \text {The equality is true for} \ n=1 \end{aligned}$$

If \(n \ge 2\), and the equality works for all integers \(1 \le m \le n-1\),

$$\begin{aligned}&\displaystyle \lim _{\epsilon \rightarrow 0} \displaystyle \frac{\mathbb {A}_\epsilon (\mathbb {B}_\epsilon \mathbb {A}_\epsilon )^n-I_2}{(n+f) \ \epsilon }&\\&\quad = \frac{1}{n+f} \left[ \displaystyle \lim _{\epsilon \rightarrow 0} \displaystyle \frac{(\mathbb {A}_\epsilon (\mathbb {B}_\epsilon \mathbb {A}_\epsilon )^{n-1}-I_2)(\mathbb {B}_\epsilon \mathbb {A}_\epsilon )+(\mathbb {B}_\epsilon \mathbb {A}_\epsilon -I_2)}{\epsilon } \right]&\\&\quad = \frac{1}{n+f} \left[ ((n-1)+f) \displaystyle \lim _{\epsilon \rightarrow 0} \displaystyle \frac{(\mathbb {A}_\epsilon (\mathbb {B}_\epsilon \mathbb {A}_\epsilon )^{n-1}-I_2)}{((n-1)+f)\epsilon } \displaystyle \lim _{\epsilon \rightarrow 0} (\mathbb {B}_\epsilon \mathbb {A}_\epsilon ) + \displaystyle \lim _{\epsilon \rightarrow 0} \displaystyle \frac{\mathbb {B}_\epsilon \mathbb {A}_\epsilon -I_2}{\epsilon } \right]&\\&\quad = \frac{1}{n+f} \left[ ((n-1)+f) \left( \frac{n \ f}{(n-1)+f} \ \mathbb {D}_A + \frac{(n-1)(1-f)}{(n-1)+f} \mathbb {D}_B \right) (I_2 \ I_2) \right.&\\&\quad \left. + (f \ \mathbb {D}_A + (1-f) \mathbb {D}_B) \right]&\\&\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ (\text {by the inductive assumption and Proposition}~\hbox {A}.3~\text {and Lemma}~\hbox {A}.5) \\&\quad = \frac{(n+1)f}{n+f} \ \mathbb {D}_A + \frac{n(1-f)}{n+f} \mathbb {D}_B \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \text {(The equality is true for} \ n \ge 2 \text {)} \end{aligned}$$

Therefore, proved. \(\square \)

Theorem A.8

If Drug A and Drug B are prescribed in turn with a relative intensity of f and \(1-f\), and are switched instantaneously, V obeys

$$\begin{aligned} \displaystyle \frac{\mathrm{d}V}{\mathrm{d}t} = (f \ \mathbb {D}_A + (1-f)\mathbb {D}_B)V \end{aligned}$$

Proof

For any time point \(t_0\), let us define \(V_\epsilon (t)\) as a vector-valued function of \(A_\mathrm{R}(t)\) and \(B_\mathrm{R}(t)\) describing the cell population dynamics under a periodic therapy starting at \(t_0\) with DrugA assigned at \(t_0+m \ \epsilon \le t < t_0+(m+f)\epsilon \) and DrugB at \(t_0+(m+f)\epsilon \le t < t_0+(m+1)\epsilon \) for \(m=0,1,2,3,...\). Then, by Proposition A.1 and the definitions of \(\mathbb {A}\) and \(\mathbb {B}\),

$$\begin{aligned} V_\epsilon (t_0+m \ \epsilon )&= (\mathbb {B}_\epsilon \mathbb {A}_\epsilon )^m \ V(t_0),\ \ \ \ \ V_\epsilon (t_0+(m+f)\epsilon ) = \mathbb {A}_\epsilon (\mathbb {B}_\epsilon \mathbb {A}_\epsilon )^m \ V(t_0) \end{aligned}$$
(*1)

where \(V(t_0)=\left( \begin{array}{l} A_\mathrm{R}(t_0) \\ B_\mathrm{R}(t_0) \end{array} \right) \). And, \(V_0(t)\) represents instantaneous drug switching.

For any \(\Delta t>0\) and any positive integer n, there exists \(\epsilon =\epsilon (n,\Delta t)\) such that

$$\begin{aligned} \frac{\Delta t}{n+1}< \epsilon \le \frac{\Delta t}{n} \ \ \ \ \ \ \ \ \ \ \text {or} \ \ \ \ \ \ \ \ \ \ 1 \le \displaystyle \frac{\Delta t}{n \ \epsilon } < 1 + \displaystyle \frac{1}{n}. \end{aligned}$$

Then by the squeeze theorem,

$$\begin{aligned} \lim _{\Delta t \rightarrow 0} \epsilon (n, \Delta t) = 0 \text { for any positive integer}~ n,~\text {and } \lim _{n \rightarrow \infty } \displaystyle \frac{\Delta t}{n \ \epsilon (n, \Delta t)} = 1 \text { for any}~ \Delta t > 0. \end{aligned}$$
(*2)

For such \(\Delta t\), n and \(\epsilon (n,\Delta t)\), \(V_\epsilon (t_0+\Delta t)\) is bounded, since local extrema can occur only when drugs are switched by Proposition A.2. That is,

$$\begin{aligned}&\min \left[ V_\epsilon (t_0+n \ \epsilon ), V_\epsilon (t_0+(n+f)\epsilon ), V_\epsilon (t_0+(n+1)\epsilon )\right] \ \le \ V_\epsilon (t_0+\Delta t)&\\&\quad \le \ \max \left[ V_\epsilon (t_0+n \ \epsilon ), V_\epsilon (t_0+(n+f)\epsilon ), V_\epsilon (t_0+(n+1)\epsilon )\right] , \end{aligned}$$
(*3)

Also,

$$\begin{aligned}&\displaystyle \lim _{\Delta t \rightarrow 0} \displaystyle \frac{\lim _{n \rightarrow \infty }V_{\epsilon (n,\Delta t)}(t_0+n \ \epsilon (n,\Delta t))-V(t_0)}{\Delta t} \\&\quad = \displaystyle \lim _{\Delta t \rightarrow 0} \ \displaystyle \lim _{n \rightarrow \infty } \displaystyle \frac{(\mathbb {B}_\epsilon \mathbb {A}_\epsilon )^n-I_2}{\Delta t} V(t_0)\displaystyle&(\text {by} (*1)) \\&\quad = \displaystyle \frac{\lim _{\Delta t \rightarrow 0} \lim _{n \rightarrow \infty } \left[ (\mathbb {B}_\epsilon \mathbb {A}_\epsilon )^n-I_2 \right] /(n \ \epsilon )}{\lim _{\Delta t \rightarrow 0} \lim _{n \rightarrow \infty } \Delta t / (n \ \epsilon )} V(t_0)&\\&\quad = \displaystyle \frac{\lim _{n \rightarrow \infty } \left[ \lim _{\Delta t \rightarrow 0} \left[ (\mathbb {B}_\epsilon \mathbb {A}_\epsilon )^n-I_2 \right] /(n \ \epsilon ) \right] }{\lim _{\Delta t \rightarrow 0} \left[ \lim _{n \rightarrow \infty } \Delta t / (n \ \epsilon ) \right] } V(t_0)&\\&\quad = \displaystyle \frac{\lim _{n \rightarrow \infty } \left[ \lim _{\epsilon \rightarrow 0} \left[ (\mathbb {B}_\epsilon \mathbb {A}_\epsilon )^n-I_2 \right] /(n \ \epsilon ) \right] }{\lim _{\Delta t \rightarrow 0} 1} V(t_0)&(\text {by}~ (*2)) \\&\quad = \lim _{n \rightarrow \infty } \left[ f \ \mathbb {D}_A + (1-f) \mathbb {D}_B \right] V(t_0)&(\text {by Lemma}~\hbox {A}.6) \\&\quad = (f \ \mathbb {D}_A + (1-f) \mathbb {D}_B) V(t_0). \end{aligned}$$
(*4)

And,

$$\begin{aligned}&\displaystyle \lim _{\Delta t \rightarrow 0} \displaystyle \frac{\lim _{n \rightarrow \infty }V_{\epsilon (n,\Delta t)}(t_0 + (n+f) \ \epsilon (n,\Delta t))-V(t_0)}{\Delta t} \\&\quad = \displaystyle \lim _{\Delta t \rightarrow 0} \ \displaystyle \lim _{n \rightarrow \infty } \displaystyle \frac{\mathbb {A}_\epsilon (\mathbb {B}_\epsilon \mathbb {A}_\epsilon )^n-I_2}{\Delta t} V(t_0)\displaystyle&(\text {by}~ (*1)) \\&\quad = \displaystyle \frac{\lim _{\Delta t \rightarrow 0} \lim _{n \rightarrow \infty } \left[ \mathbb {A}_\epsilon (\mathbb {B}_\epsilon \mathbb {A}_\epsilon )^n-I_2 \right] /((n+f) \ \epsilon )}{\lim _{\Delta t \rightarrow 0} \lim _{n \rightarrow \infty } \Delta t / ((n+f) \ \epsilon )} V(t_0)&\\&\quad = \displaystyle \frac{\lim _{n \rightarrow \infty } \left[ \lim _{\Delta t \rightarrow 0} \left[ (\mathbb {B}_\epsilon \mathbb {A}_\epsilon )^n-I_2 \right] /((n+f) \ \epsilon ) \right] }{\lim _{\Delta t \rightarrow 0} \left[ \lim _{n \rightarrow \infty } (\Delta t / (n \ \epsilon ))(n/(n+f)) \right] } V(t_0)&\\&\quad = \displaystyle \frac{\lim _{n \rightarrow \infty } \left[ \lim _{\epsilon \rightarrow 0} \left[ (\mathbb {B}_\epsilon \mathbb {A}_\epsilon )^n-I_2 \right] /((n+f) \ \epsilon ) \right] }{\lim _{\Delta t \rightarrow 0} 1} V(t_0)&(\text {by}~ (*2)) \\&\quad = \lim _{n \rightarrow \infty } \left[ \frac{(n+1)f}{n+f} \ \mathbb {D}_A + \frac{n(1-f)}{n+f} \mathbb {D}_B \right] V(t_0)&(\text {by Lemma}~ \hbox {A}.7) \\&\quad = (f \ \mathbb {D}_A + (1-f) \mathbb {D}_B) V(t_0) \end{aligned}$$
(*5)

Similar to (*4),

$$\begin{aligned} \displaystyle \lim _{\Delta t \rightarrow 0} \displaystyle \frac{\lim _{n \rightarrow \infty }V_{\epsilon (n,\Delta t)}(t_0+(n+1) \ \epsilon (n,\Delta t))-V(t_0)}{\Delta t}&= (f \ \mathbb {D}_A + (1-f) \mathbb {D}_B) V(t_0) \end{aligned}$$
(*6)

By (*4)–(*6),

$$\begin{aligned}&\min \left[ \displaystyle \lim _{\Delta t \rightarrow 0} \displaystyle \frac{\lim _{n \rightarrow \infty }V_\epsilon (t_0+n \ \epsilon )-V(t_0)}{\Delta t}, \displaystyle \lim _{\Delta t \rightarrow 0} \displaystyle \frac{\lim _{n \rightarrow \infty }V_\epsilon (t_0+(n+f) \ \epsilon )-V(t_0)}{\Delta t} \right. , \\&\quad \left. \displaystyle \lim _{\Delta t \rightarrow 0} \displaystyle \frac{\lim _{n \rightarrow \infty }V_\epsilon (t_0+(n+1) \ \epsilon )-V(t_0)}{\Delta t} \right] = \max \left[ \displaystyle \lim _{\Delta t \rightarrow 0} \displaystyle \frac{\lim _{n \rightarrow \infty }V_\epsilon (t_0+n \ \epsilon )-V(t_0)}{\Delta t}, \right.&\\&\quad \left. \displaystyle \lim _{\Delta t \rightarrow 0} \displaystyle \lim _{\Delta t \rightarrow 0} \displaystyle \frac{\lim _{n \rightarrow \infty }V_\epsilon (t_0+(n+f) \ \epsilon )-V(t_0)}{\Delta t}, \displaystyle \lim _{\Delta t \rightarrow 0} \displaystyle \frac{\lim _{n \rightarrow \infty }V_\epsilon (t_0+(n+1) \ \epsilon )-V(t_0)}{\Delta t} \right]&\\&\quad = (f \ \mathbb {D}_A + (1-f) \mathbb {D}_B) V(t_0) \end{aligned}$$
(*7)

Then, by (*3), (*7) and the squeeze theorem,

$$\begin{aligned} \displaystyle \left. \frac{\mathrm{d}}{\mathrm{d}t} V_0 \right| _{t=t_0}= \displaystyle \lim _{\Delta t \rightarrow 0} \displaystyle \frac{\lim _{n \rightarrow \infty }V_\epsilon (t_0 + \Delta t)-V(t_0)}{\Delta t} = (f \ \mathbb {D}_A + (1-f) \mathbb {D}_B) V(t_0) \end{aligned}$$

Therefore,

$$\begin{aligned} \displaystyle \frac{\mathrm{d}V}{\mathrm{d}t} = (f \ \mathbb {D}_A + (1-f)\mathbb {D}_B)V \end{aligned}$$

\(\square \)

1.3 A.3: Population dynamics with the optimal regimen

In this section, we want to write the differential equations of \(V=\{A_\mathrm{R},B_\mathrm{R}\}\) under the optimal control strategy described in Section 3.3. Based on “Appendix A.2” and a couple of lemma/theorem, we will reach to a concise form of a differential system described at Theorem A.11.

Lemma A.9

\(\left\{ \displaystyle \frac{r_A r_B - s_A s_B}{r_A+r_B-s_A-s_B},\left( \begin{array}{l} \mathcal {(A/B)}^* \\ 1 \end{array} \right) \right\} \) is an eigenpair of \(f^* \ \mathbb {D}_A + (1-f^*)\mathbb {D}_B\) with \(\mathcal {(A/B)}^*\) and \(f^*=k^*/(1+k^*)\) defined by Eqs. (9) and (12).

Proof

Let \(\mathbb {D}^*:=f^* \ \mathbb {D}_A + (1-f^*)\mathbb {D}_B\), and \(\lambda =\displaystyle \frac{r_A r_B - s_A s_B}{r_A+r_B-s_A-s_B}\). Then,

$$\begin{aligned} \mathbb {D}^*-\lambda \ I_2 = C_1 \left( \begin{array}{l} C_2 \ U^T \\ C_3 \ U^T \end{array} \right) , \end{aligned}$$

where \(U=\left( \begin{array}{l} 1 \\ -\mathcal {(A/B)}^* \end{array} \right) \) along with

$$\begin{aligned}&C_1=-(g_A(r_A-s_B)+g_B(r_B-s_A)\\&\quad +(r_B-s_A)(r_A-s_B))(r_A+r_B-s_A-s_B)/(r_A-s_B), \\&C_2=g_A((r_A-s_B)(r_B-s_B)+g_B(r_A+r_B-s_A-s_B), \\&C_3=-g_B((r_B-s_A)(r_A-s_A)+g_A(r_A+r_B-s_A-s_B)). \end{aligned}$$

Since \(U^T \ V = 0\) where \(V=((r_B-s_A)/(r_A-s_B),1)^T\), \((\lambda ,V)\) is an eigenpair of \(\mathbb {D}^*\). \(\square \)

Theorem A.10

In Stage 2 of the optimal strategy, both \(A_\mathrm{R}\) and \(B_\mathrm{R}\) change with a constant net proliferation rate,

$$\begin{aligned}\lambda = \displaystyle \frac{r_A r_B - s_A s_B}{r_A + r_B - s_A - s_B}. \end{aligned}$$

Proof

Without a loss of generality, let us prove it only when \(\mathcal {A/B}(0)<\mathcal {(A/B)}^*\).

If \(\mathcal {A/B}(0)<\mathcal {(A/B)}^*\), DrugA has a better effect initially. So following the optimal therapy scheduling, DrugA is assigned alone at the beginning as long as \(T_\mathrm{min}^A=T_\mathrm{min}(p_A,p_B,\mathcal {A/B}(0))\) (Stage 1), and then, Stage 2 starts at \(T_\mathrm{min}^A\) with initial condition

$$\begin{aligned} V(T_\mathrm{min}^A)=\mathbb {M}_A(T_\mathrm{min}^A)V(0)=C \left( \begin{array}{l} \mathcal {(A/B)}^* \\ 1 \end{array} \right) \quad \quad \qquad \qquad \quad \quad \quad \quad \text {(**1)} \end{aligned}$$

where \(C=\displaystyle \frac{P(0)}{1+\mathcal {A/B}(0)} \left( \displaystyle \frac{(r_A-s_A)(r_B-s_A)+g_A(r_A+r_B-s_A-s_B)}{(r_A-s_B)(g_A+\mathcal {A/B}(0)(g_A+r_A-s_A))}\right) ^{-\frac{g_A-s_A}{g_A+r_A-s_A}}\).

By Theorem A.8, in Stage 2, V(t) obeys

$$\begin{aligned} \displaystyle \frac{\mathrm{d}V}{\mathrm{d}t} = \mathbb {D}^* V,~\hbox {where}~\mathbb {D}^* = f^* \mathbb {D}_A + (1-f^*)\mathbb {D}_B \quad \quad \qquad \qquad \quad \quad \quad \quad \text {(**2)} \end{aligned}$$

By Lemma A.9, \(V(T_\mathrm{min}^A)\) is an eigenvector of \(\mathbb {D}^*\) with the corresponding eigenvalue, \(\lambda \). Then, the solution of (**2) with the initial value (**1) is

$$\begin{aligned} V(t+T_\mathrm{min}^A)=e^{\lambda \ t}V(T_\mathrm{min}^A). \end{aligned}$$

\(\square \)

Theorem A.11

With optimal therapy utilizing DrugA and DrugB, V obeys the following equations and solutions.

If \(\mathcal {A/B}(0)<\mathcal {(A/B)}^*\),

$$\begin{aligned} \displaystyle \frac{\mathrm{d}V}{\mathrm{d}t} = \left\{ \begin{array}{ll} \mathbb {D}_A V &{} \text {if }\quad 0 \le t \le T_\mathrm{min}^A\\ \lambda \ V &{} \text {if }\quad t> T_\mathrm{min}^A \end{array} \right. ~\hbox {and}~ V(t) = \left\{ \begin{array}{ll} \mathbb {M}_A(t) V(0) &{} \text {if }\quad 0 \le t \le T_\mathrm{min}^A\\ e^{\lambda \ (t-T_\mathrm{min}^A)} V(T_\mathrm{min}^A) &{} \text {if }\quad t > T_\mathrm{min}^A \end{array} \right. \end{aligned}$$

Similarly, if \(\mathcal {A/B}(0) \ge \mathcal {(A/B)}^*\),

$$\begin{aligned} \displaystyle \frac{\mathrm{d}V}{{rm d}t} = \left\{ \begin{array}{ll} \mathbb {D}_B V &{} \text {if }\quad 0 \le t \le T_\mathrm{min}^B\\ \lambda \ V &{} \text {if }\quad t> T_\mathrm{min}^B \end{array} \right. ~\hbox {and}~ V(t) = \left\{ \begin{array}{ll} \mathbb {M}_B(t) V(0) &{} \text {if }\quad 0 \le t \le T_\mathrm{min}^B\\ e^{\lambda \ (t-T_\mathrm{min}^B)} V(T_\mathrm{min}^B) &{} \text {if }\quad t > T_\mathrm{min}^B \end{array} \right. \end{aligned}$$

Proof

Straightforward, by Theorem A.10 \(\square \)

Appendix B: Sensitivity analysis on optimal scheduling

The two determinant quantities of optimal control scheduling are (i) the duration of the first stage (\(T_\mathrm{min}^1\)), and (ii) the relative intensity between two drugs in the second stage (\(k^*\)). Here, we show sensitivity analysis on the quantities related to them, \(T_\mathrm{gap}\) and \(f^*\), over a range of (scaled) model parameters. Additionally over the same range, we studied how much our \(T_\mathrm{min}\)-based optimal scheme is better than the \(T_\mathrm{max}\)-based scheme evaluated by the integral in Eq. (13).

  1. 1.

    Sensitivity analysis of \(T_\mathrm{gap}\)

    Using \(g_1\), we non-dimentionalize all the values, like

$$\begin{aligned} \{\overline{s_1},\overline{r_1}|\overline{s_2},\overline{r_2}\} := \displaystyle \frac{1}{g_1} \{s_1,r_1|s_2,r_2\} \quad \hbox {and} \quad \overline{T_\mathrm{gap}}:= g_1 \ T_\mathrm{gap} \end{aligned}$$

then,

$$\begin{aligned} \overline{T_\mathrm{gap}}(\{\overline{s_1},\overline{r_1}\},\{\overline{s_2},\overline{r_2}\})&:= \frac{\ln \left[ \displaystyle \frac{(1-\overline{s_1})(\overline{r_1}-\overline{s_1})(\overline{r_1}-\overline{s_2})}{\overline{r_1}((\overline{r_1}-\overline{s_1})(\overline{r_2}-\overline{s_1})+(\overline{r_1}+\overline{r_2}-\overline{s_1}-\overline{s_2}))} \right] }{1+\overline{r_1}-\overline{s_1}} \end{aligned}$$

In general, cells mutate slower than they proliferate , so we ran sensitivity analysis on \(T_\mathrm{gap}\) for all \(a \gg 1\) for \({a \in \{-\overline{s_1}, -\overline{s_2}, \overline{r_1}, \overline{r_2}\}}\). Figure 12 shows \(T_\mathrm{gap}\) over the range of \({20 \le -\overline{s_1}, -\overline{s_2}, \overline{r_1}, \overline{r_2} \le 100}\). So, under the assumption that \(g_1 \ll \min \{-s_1,-s_2,r_1,r_2\}\),

$$\begin{aligned} T_\mathrm{gap}(\{s_1,r_1\},\{s_2,r_2\}) \approx \displaystyle \frac{\ln \left[ \displaystyle \frac{-s_1(r_1-s_2)}{r_1(r_2-s_1)} \right] }{r_1-s_1}, \end{aligned}$$

which approximates the contour curves in Fig. 12.

Fig. 12
figure 12

(Colour figure Online) Contour maps of \(T_\mathrm{gap}\) over ranges of \(10 \le a \le 100\) for \({a \in \{-\overline{s_1},-\overline{s_2},\overline{r_1},\overline{r_2}\}=\{-s_1,-s_2,r_1,r_2\}/g_1}\) and \(r_1 r_2 < s_1 s_2\) (Condition (6)). As \(-s_2\) decreases and/or \(r_2\) increases, the optimal switching timing to the second drug is delayed (\(T_\mathrm{min}\uparrow \) and \(T_\mathrm{gap}\downarrow \)). As \(r_1\) increases, \(T_\mathrm{gap}\) decreases. Also, \(T_\mathrm{gap}\) and \(s_1\) have a non-monotonic relationship as shown on the graphs

  1. 2.

    Sensitivity analysis of \(f^*\)

Regarding the regulated intensities among the two drugs, \(k^*\), we assumed that \(g_1 \approx g_2 := g\), similarly assuming that they are both much smaller than \(\{-s_1,-s_2,r_1,r_2\}\). Then, we normalized all the parameters with the unit of g, like

$$\begin{aligned} {\{\overline{s_1},\overline{r_1}|\overline{s_2},\overline{r_2}\} := \displaystyle \frac{1}{g} \{s_1,r_1|s_2,r_2\}.} \end{aligned}$$

\(k^*\) can be rewritten in terms of the dimensionless parameters.

$$\begin{aligned} {k^*(\{\overline{s_1},\overline{r_1}\},\{\overline{s_2},\overline{r_2}\}) = \displaystyle \frac{(\overline{r_1}-\overline{s_2})((\overline{r_1}-\overline{s_1})(\overline{r_2}-\overline{s_1})+(\overline{r_1}+\overline{r_2}-\overline{s_1}-\overline{s_2}))}{(\overline{r_2}-\overline{s_1})((\overline{r_2}-\overline{s_2})(\overline{r_1}-\overline{s_2})+(\overline{r_1}+\overline{r_2}-\overline{s_1}-\overline{s_2}))}} \end{aligned}$$

In this sensitivity analysis, we use

$$\begin{aligned} f^*:=\displaystyle \frac{k^*}{1+k^*}, \end{aligned}$$

which represents intensity fraction of the initially better drug out of the total therapy. We evaluated \(f^*\) over the same ranges of \(\{s_1,s_2,r_1,r_2\}\), like the previous exercise (see Fig. 13) over the range \(\max \{g_1,g_2\} \ll \min \{-s_1,-s_2,r_1,r_2\}\), so \(k^*\) and \(f^*\) can be approximated by the simpler forms:

$$\begin{aligned} k^* \approx \displaystyle \frac{r_1-s_1}{r_2-s_2} \quad \hbox {and} \quad f^* \approx \displaystyle \frac{r_1-s_1}{r_1+r_2-s_1-s_2} \end{aligned}$$
Fig. 13
figure 13

(Colour figure Online) Contour maps of \(f^*\) over ranges of \(10 \le a \le 100\) for \({a \in \{-\overline{s_1},-\overline{s_2},\overline{r_1},\overline{r_2}\}=\{-s_1,-s_2,r_1,r_2\}/g}\) and \(r_1 r_2 < s_1 s_2\) (Condition 6). \(k^*\) (or \(f^*\)) increases, as \(r_1\) and/or \(-s_1\) decreases and/or as \(r_2\) and/or \(-s_2\) increases

  1. 3.

    Sensitivity analysis of Integral (13)

To study the sensitivity of the advantage of using the optimal control defined by Integral (13), we assumed that \(g_1 \approx g_2 \approx g = 0.001\). Then, similar to the previous studies, we explored the sensitivity of the normalized parameters in terms of g, that is (Fig. 14):

$$\begin{aligned} {\{\overline{s_1},\overline{r_1}|\overline{s_2},\overline{r_2}\} := \displaystyle \frac{1}{g} \{s_1,r_1|s_2,r_2\}.} \end{aligned}$$
Fig. 14
figure 14

(Colour figure Online) Contour maps of the measured advantageous effect of the optimal therapy defined by the integration (13) over ranges of \(10 \le a \le 100\) for \({a \in \{-\overline{s_1},-\overline{s_2},\overline{r_1},\overline{r_2}\}}\) and \(r_1 r_2 < s_1 s_2\) (Condition (6)) Here, \({\{-\overline{s_1},-\overline{s_2},\overline{r_1},\overline{r_2}\}=\{-s_1,-s_2,r_1,r_2\}/g}\) and \(g=0.001\). The measured effect increases as \(r_1\), \(r_2\) decreases and/or \(-s_1\) increases

Appendix C: Clinical implementation of instantaneous switch in the optimal strategy

In clinical practice, the instantaneous drug switch which we suggest in the second stage of the optimal treatment scheduling is not implementable. Therefore, we compared similar schedules to the optimal case. In the “similar” schedules, the first stage, using an initial drug, remained the same as the optimal schedule. However the second part, where we previously used an instantaneous switch (with \(\Delta t = 0\)), was modified to use a fast switch (\(\Delta t \gtrsim 0\)). Figure 15a, b shows how instantaneous switching (\(\Delta t=0\)) and fast switching (multiple choices of \(\Delta t \gtrsim 0\)) compare in terms of population size using different drug parameters. As expected, the smaller \(\Delta t\) is , the closer to the ideal case. And, a choice of a reasonably small \(\Delta t\) (like 1 day or 3 days) results in an outcome quite close to the optimal scenario.

We repeated this exercise with \(k^*\) (from Eq. 12) instead of \(k(\Delta t)\) modulated by \(\Delta t\) (Fig. 15c, d). Only small differences are observed between Fig. 15a, b and c, d which justifies the general usefulness of \(k^*\) independent of \(\Delta t\).

Fig. 15
figure 15

Graphs showing regular drug switching in Stage 2 with different \(\{\Delta t, k(\Delta t,p_A,p_B)\}\): \(\Delta t=1\) day (blue), \(\Delta t=4\) days (red), \(\Delta t=7\) days (green), and \(\Delta t=10\) days (magenta). Parameters/conditions: \(p_A=\{-0.18,0.008,0.00075\}\)/day, \(p_B=\{-0.9,0.016,0.00125\}\)/day and \(\{A_\mathrm{R}^0,B_\mathrm{R}^0\}=\{0.1,0.9\}\), a Total population histories, \(C_\mathrm{P}^n\) for \(n \in \{1,4,7,10\}\) days, b Differences between the optimal population history \(C_\mathrm{P}^*\) (i.e., when \(\Delta t=0\)) and each case with positive \(\Delta t\) (i.e., \(C_\mathrm{P}^n-C_\mathrm{P}^*\)). The inserts interesting ranges. c and d are equivalent with (a) and (b) except that \(k^*(p_A,p_B)\}\) has been used instead of \(k(\Delta t,p_A,p_B)\}\) (Color figure online)

Appendix D: Stochastic simulation codes

The computational code written in Python will be provided at Github (https://github.com/nryoon12/Optimal-Therapy-Scheduling-Based-on-a-Pair-of-Collaterally-Sensitive-Drugs).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Yoon, N., Vander Velde, R., Marusyk, A. et al. Optimal Therapy Scheduling Based on a Pair of Collaterally Sensitive Drugs. Bull Math Biol 80, 1776–1809 (2018). https://doi.org/10.1007/s11538-018-0434-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11538-018-0434-2

Keywords

Navigation