Skip to main content
Log in

A pharmacokinetic and pharmacodynamic analysis of drug forgiveness

  • Original Paper
  • Published:
Journal of Pharmacokinetics and Pharmacodynamics Aims and scope Submit manuscript

Abstract

Nonadherence to medication is a major public health problem. To combat nonadherence, some clinicians have suggested using “forgiving” drugs, which maintain efficacy in spite of delayed or missed doses. What pharmacokinetic (PK) and pharmacodynamic (PD) factors make a drug forgiving? In this paper, we address this question by analyzing a linear PK/PD model for a patient with imperfect adherence. We assume that the drug effect is far from maximal and consider direct effect, effect compartment (biophase), and indirect response PD models. We prove that the average drug effect relative to the clinically desired effect is simply the fraction of prescribed doses actually taken by the patient. Hence, under these assumptions, drug forgiveness cannot be defined in terms of the average effect. We argue that forgiveness should instead be understood in terms of effect fluctuations. We prove that the rates of PK absorption, PK elimination, and PD elimination are exactly equivalent for determining effect fluctuations. We prove all the aforementioned results for any pattern of nonadherence, including late doses, missed doses, drug holidays, extra doses, etc. To obtain quantitative estimates of effect fluctuations, we consider a simple statistical pattern of nonadherence and analytically calculate the coefficient of variation of effect. We further show how effect fluctuations can be reduced by taking an extra “make up” dose following a missed dose if any one of the aforementioned PK/PD rates is sufficiently slow. We illustrate some of our results for a nonlinear indirect response model of metformin.

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

Similar content being viewed by others

References

  1. Vrijens B, De Geest S, Hughes DA, Przemyslaw K, Demonceau J, Ruppar T, Dobbels F, Fargher E, Morrison V, Lewek P et al (2012) A new taxonomy for describing and defining adherence to medications. Br J Clin Pharmacol 73(5):691–705

    Article  PubMed  PubMed Central  Google Scholar 

  2. Osterberg L, Blaschke T (2005) Adherence to medication. N Engl J Med 353(5):487–497

    Article  CAS  PubMed  Google Scholar 

  3. Eduardo S et al (2003) Adherence to long-term therapies: evidence for action. World Health Organization, Geneva

    Google Scholar 

  4. Brian HR, Pauline MH, Amit G, Patty M (2002) Interventions for helping patients to follow prescriptions for medications. Cochrane Database Syst Rev. https://doi.org/10.1002/14651858.CD000011

    Article  Google Scholar 

  5. Urquhart J (1998) Pharmacodynamics of variable patient compliance: implications for pharmaceutical value. Adv Drug Deliv Rev 33(3):207–219

    Article  CAS  PubMed  Google Scholar 

  6. Assawasuwannakit P, Braund R, Duffull SB (2015) Quantification of the forgiveness of drugs to imperfect adherence. CPT: Pharmacomet Syst Pharmacol 4(3):204–211

    Google Scholar 

  7. Assawasuwannakit P, Braund R, Duffull SB (2016) A framework for quantifying the influence of adherence and dose individualization. Clin Pharmacol Ther 99(4):452–459

    Article  CAS  PubMed  Google Scholar 

  8. Pellock JM, Brittain ST (2016) Use of computer simulations to test the concept of dose forgiveness in the era of extended-release (XR) drugs. Epilepsy Behav 55:21–23

    Article  PubMed  Google Scholar 

  9. Morrison A, Stauffer ME, Kaufman AS (2017) Relationship between adherence rate threshold and drug ‘forgiveness’. Clin Pharmacokinet 56(12):1435–1440

    Article  CAS  PubMed  Google Scholar 

  10. Véronique D (2011) Drug forgiveness and interpatient pharmacokinetic variability in tuberculosis. J Infect Dis 204:1827–1829

    Article  Google Scholar 

  11. Lowy A, Munk VC, Ong SH, Burnier M, Vrijens B, Tousset EP, Urquhart J (2011) Effects on blood pressure and cardiovascular risk of variations in patients? adherence to prescribed antihypertensive drugs: role of duration of drug action. Int J Clin Pract 65(1):41–53

    Article  CAS  PubMed  Google Scholar 

  12. Nony P, Boissel J-P (2002) Use of sensitivity functions to characterise and compare the forgiveness of drugs. Clin Pharmacokinet 41(5):371–380

    Article  CAS  PubMed  Google Scholar 

  13. Gohore DG, Fenneteau F, Barrière O, Li J, Nekka F et al (2010) Rational drug delineation: a global sensitivity approach based on therapeutic tolerability to deviations in execution. Pharmacol Pharm 1(02):42

    Article  CAS  Google Scholar 

  14. Osterberg LG, Urquhart J, Blaschke TF (2010) Understanding forgiveness: minding and mining the gaps between pharmacokinetics and therapeutics. Clin Pharmacol Ther 88(4):457–459

    Article  CAS  PubMed  Google Scholar 

  15. Boissel J-P, Nony P (2002) Using pharmacokinetic-pharmacodynamic relationships to predict the effect of poor compliance. Clin Pharmacokinet 41(1):1–6

    Article  CAS  PubMed  Google Scholar 

  16. Melanie AF, Marilyn EM, Donald EM (2012) Mechanism-based pharmacodynamic modeling. Computational toxicology. Springer, New York, pp 583–600

    Google Scholar 

  17. Hong Y, Rohatagi S, Habtemariam B, Walker JR, Schwartz SL, Mager DE (2008) Population exposure-response modeling of metformin in patients with type 2 diabetes mellitus. J Clin Pharmacol 48(6):696–707

    Article  CAS  PubMed  Google Scholar 

  18. Bender CM, Orszag SA (2013) Advanced mathematical methods for scientists and engineers I: Asymptotic methods and perturbation theory. Springer, New York

    Google Scholar 

  19. Counterman ED, Lawley SD (2021) What should patients do if they miss a dose of medication? A theoretical approach. J Pharmacokinet Pharmacodynam

  20. Counterman ED, Lawley SD (2022) Designing drug regimens that mitigate nonadherence. Bull Math Biol 84(1):1–36

    Article  Google Scholar 

  21. Stauffer ME, Hutson P, Kaufman AS, Morrison A (2017) The adherence rate threshold is drug specific. Drugs in R&D 17(4):645–653

    Article  CAS  Google Scholar 

  22. Garnett WR, McLean AM, Zhang Y, Clausen S, Tulloch SJ (2003) Simulation of the effect of patient nonadherence on plasma concentrations of carbamazepine from twice-daily extended-release capsules. Curr Med Res Opin 19(6):519–525

    Article  CAS  PubMed  Google Scholar 

  23. Reed RC, Dutta S (2004) Predicted serum valproic acid concentrations in patients missing and replacing a dose of extended-release divalproex sodium. Am J Health Syst Pharm 61(21):2284–2289

    Article  CAS  PubMed  Google Scholar 

  24. Dutta S, Reed RC (2006) Effect of delayed and/or missed enteric-coated divalproex doses on valproic acid concentrations: simulation and dose replacement recommendations for the clinician 1. J Clin Pharm Ther 31(4):321–329

    Article  CAS  PubMed  Google Scholar 

  25. Ding J, Zhang Y, Jiao Z, Wang Y (2012) The effect of poor compliance on the pharmacokinetics of carbamazepine and its epoxide metabolite using monte carlo simulation. Acta Pharmacol Sin 33(11):1431–1440

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Chen C, Wright J, Gidal B, Messenheimer J (2013) Assessing impact of real-world dosing irregularities with lamotrigine extended-release and immediate-release formulations by pharmacokinetic simulation. Ther Drug Monit 35(2):188–193

    Article  CAS  PubMed  Google Scholar 

  27. Gidal Barry E, Oneeb M, Jim F, Ziad H, Haichen Y, Jin Z, Randi F, Antonio L (2014) The practical impact of altered dosing on perampanel plasma concentrations: pharmacokinetic modeling from clinical studies. Epilepsy Behav 35:6–12

    Article  CAS  PubMed  Google Scholar 

  28. Brittain ST, Wheless JW (2015) Pharmacokinetic simulations of topiramate plasma concentrations following dosing irregularities with extended-release vs immediate-release formulations. Epilepsy Behav 52:31-36D

    Article  PubMed  Google Scholar 

  29. Sunkaraneni S, Blum D, Ludwig E, Chudasama V, Fiedler-Kelly J, Marvanova M, Bainbridge J, Phillips L (2018) Population pharmacokinetic evaluation and missed-dose simulations for eslicarbazepine acetate monotherapy in patients with partial-onset seizures. Clin Pharmacol Drug Dev 7(3):287–297

    Article  CAS  PubMed  Google Scholar 

  30. Hard ML, Wehr AY, Sadler BM, Mills RJ, Moltke L (2018) Population pharmacokinetic analysis and model-based simulations of aripiprazole for a 1-day initiation regimen for the long-acting antipsychotic aripiprazole lauroxil. Eur J Drug Metab Pharmacokinet 43(4):461–469

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Elkomy MH (2020) Changing the drug delivery system: Does it add to non-compliance ramifications control? A simulation study on the pharmacokinetics and pharmacodynamics of atypical antipsychotic drug. Pharmaceutics 12(4):297

    Article  PubMed Central  Google Scholar 

  32. Jia-qin G, Guo Y, Jiao Z, Ding J, Li G-F (2020) How to handle delayed or missed doses: a population pharmacokinetic perspective. Eur J Drug Metab Pharmacokinet 45(2):163–172

    Article  CAS  Google Scholar 

  33. Li J, Nekka F (2007) A pharmacokinetic formalism explicitly integrating the patient drug compliance. J Pharmacokinet Pharmacodyn 34(1):115–139

    Article  PubMed  Google Scholar 

  34. Lévy-Véhel P-E, Lévy-Véhel J (2013) Variability and singularity arising from poor compliance in a pharmacokinetic model I: the multi-IV case. J Pharmacokinet Pharmacodyn 40(1):15–39

    Article  PubMed  CAS  Google Scholar 

  35. Fermín LJ, Lévy-Véhel J (2017) Variability and singularity arising from poor compliance in a pharmacokinetic model II: the multi-oral case. J Math Biol 74(4):809–841

    Article  PubMed  Google Scholar 

  36. Jie M (2017) Stochastic modeling of random drug taking processes and the use of singular perturbation methods in pharmacokinetics. Ph.D. Thesis, The University of Utah

  37. Hof F, Bridge LJ (2021) Exact solutions and equi-dosing regimen regions for multi-dose pharmacokinetics models with transit compartments. J Pharmacokinet Pharmacodyn 48(1):99–131

    Article  CAS  PubMed  Google Scholar 

  38. Savic RM, Jonker DM, Kerbusch T, Karlsson MO (2007) Implementation of a transit compartment model for describing drug absorption in pharmacokinetic studies. J Pharmacokinet Pharmacodyn 34(5):711–726

    Article  CAS  PubMed  Google Scholar 

  39. Vrijens B, Vincze G, Kristanto P, Urquhart J, Burnier M (2008) Adherence to prescribed antihypertensive drug treatments: longitudinal study of electronically compiled dosing histories. BMJ 336(7653):1114–1117

    Article  PubMed  PubMed Central  Google Scholar 

  40. Hammer SM, Vaida F, Bennett KK, Holohan MK, Sheiner L, Eron JJ, Wheat LJ, Mitsuyasu RT, Gulick RM, Valentine FT et al (2002) Dual vs single protease inhibitor therapy following antiretroviral treatment failure: a randomized trial. JAMA 288(2):169–180

    Article  CAS  PubMed  Google Scholar 

  41. Sun J, Nagaraj HN, Reynolds NR (2007) Discrete stochastic models for compliance analysis based on an AIDS clinical trial group (ACTG) study. Biom J 49(5):731–741

    Article  PubMed  Google Scholar 

  42. Crauel H (2001) Random point attractors versus random set attractors. J Lond Math Soc (2) 63(2):413–427

    Article  Google Scholar 

  43. Mattingly JC (1999) Ergodicity of \(2\)D Navier–Stokes equations with random forcing and large viscosity. Commun Math Phys 206(2):273–288

    Article  Google Scholar 

  44. Schmalfuß B (1996) A random fixed point theorem based on Lyapunov exponents. Random Comput Dynam 4(4):257–268

    Google Scholar 

  45. Lawley SD, Mattingly JC, Reed MC (2015) Stochastic switching in infinite dimensions with applications to random parabolic PDE. SIAM J Math Anal 47(4):3035–3063

    Article  Google Scholar 

  46. Lawley SD, Keener JP (2019) Electrodiffusive flux through a stochastically gated ion channel. SIAM J Appl Math 79(2):551–571

    Article  Google Scholar 

  47. Norris JR (1998) Markov chains. Statistical & probabilistic mathematics. Cambridge University Press, Cambridge

    Google Scholar 

  48. Durrett R (2010) Probability: theory and examples, 4th edn. Cambridge University Press, Cambridge

    Book  Google Scholar 

Download references

Funding

SDL was supported by the National Science Foundation (Grant Nos. DMS-1944574 and DMS-1814832).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Sean D. Lawley.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix

Appendix

In this appendix, we present the mathematical proofs and derivations of some results in the main text.

Proof of formula for E(t) in (21)

To verify (21), note first that (21) and the definitions in (22) imply that \(E(0)=0\), as desired. Next, the definition of E(t) in (18) and the ODE for \(c_{\text {e}}\) in (6) imply that E(t) satisfies

$$\begin{aligned} \frac{\text {d}}{\text {d}t}E =\frac{E_{\max }}{\text {EC}_{\text {50}}}k_{ \text {p}\text {e}} {c_{ \text {p}}}-k_{\text {e}0} E. \end{aligned}$$
(49)

Differentiating (21) with respect to t and using the value of \(c_{ \text {p}}\) in (4) shows that (21) indeed satisfies (49), which verifies (21).□

Proof that \(\langle E \rangle =\mu \langle E^{\mathrm{perf}} \rangle\)

To prove \(\langle E \rangle =\mu \langle E^{\mathrm{perf}} \rangle\), we let \(k>0\) and consider the large T behavior of

$$\begin{aligned} F(T,k):=\frac{1}{T}\int _{0}^{T}\sum _{n:t_{n}\le t}e^{-k(t-t_{n})}f_{n}\,\text {d}t. \end{aligned}$$

Interchanging the sum and integral yields

$$\begin{aligned} F(T,k)&=\frac{1}{T}\sum _{n:t_{n}\le T}\int _{t_{n}}^{T}e^{-k(t-t_{n})}f_{n}\,\text {d}t\nonumber \\&=\frac{1}{k}\frac{1}{T}\sum _{n:t_{n}\le T}f_{n} -\frac{1}{k}\frac{1}{T}\sum _{n:t_{n}\le T}e^{-k(T-t_{n})}f_{n}. \end{aligned}$$
(50)

By (30), the first term in (50) has the large T limit,

$$\begin{aligned} \lim _{T\rightarrow \infty }\frac{1}{k}\frac{1}{T}\sum _{n:t_{n}\le T}f_{n} =\mu \frac{1}{k\tau }. \end{aligned}$$
(51)

We claim that the second term in (50) vanishes as \(T\rightarrow \infty\). To see this, let \(\varepsilon \in (0,1)\) and observe that

$$\begin{aligned}&\frac{1}{T}\sum _{n:t_{n}\le T}e^{-k(T-t_{n})}f_{n}\\&\quad =\frac{1}{T}\sum _{n:t_{n}\le (1-\varepsilon )T}e^{-k(T-t_{n})}f_{n} +\frac{1}{T}\sum _{n:(1-\varepsilon )T<t_{n}\le T}e^{-k(T-t_{n})}f_{n}\\&\quad \le e^{-\varepsilon kT}\frac{1}{T}\sum _{n:t_{n}\le (1-\varepsilon )T}f_{n} +\frac{1}{T}\sum _{n:(1-\varepsilon )T<t_{n}\le T}f_{n}. \end{aligned}$$

Since (30) implies that

$$\begin{aligned} \lim _{T\rightarrow \infty }e^{-\varepsilon kT}\frac{1}{T}\sum _{n:t_{n}\le (1-\varepsilon )T}f_{n} =0, \end{aligned}$$

it follows that

$$\begin{aligned} \lim _{T\rightarrow \infty }\frac{1}{T}\sum _{n:t_{n}\le T}e^{-k(T-t_{n})}f_{n} \le \lim _{T\rightarrow \infty }\frac{1}{T}\sum _{n:(1-\varepsilon )T<t_{n}\le T}f_{n}. \end{aligned}$$
(52)

To estimate the upper bound in (52), let \(\varepsilon _{2}>0\) and observe that (30) implies that if T is sufficiently large, then

$$\begin{aligned} \frac{1}{T}\sum _{n:t_{n}\le T(1-\varepsilon )}f_{n} +\frac{1}{T}\sum _{n:T(1-\varepsilon )<t_{n}\le T}f_{n} \le \mu /\tau +\varepsilon _{2}, \end{aligned}$$
(53)

and

$$\begin{aligned} \mu /\tau -\varepsilon _{2} \le \frac{1}{T(1-\varepsilon )}\sum _{n:t_{n}\le T(1-\varepsilon )}f_{n}. \end{aligned}$$
(54)

Combining (53)–(54) yields

$$\begin{aligned} \lim _{T\rightarrow \infty }\frac{1}{T}\sum _{n:T(1-\varepsilon )<t_{n}\le T}f_{n}&\le \mu /\tau +\varepsilon _{2} -(1-\varepsilon )(\mu /\tau -\varepsilon _{2})\\&=2\varepsilon _{2}+\varepsilon \mu /\tau -\varepsilon \varepsilon _{2}. \end{aligned}$$

Since \(\varepsilon \in (0,1)\) and \(\varepsilon _{2}>0\) are arbitrary, we conclude from (52) that

$$\begin{aligned} \lim _{T\rightarrow \infty }\frac{1}{T}\sum _{n:t_{n}\le T}e^{-k(T-t_{n})}f_{n}=0, \end{aligned}$$

and therefore (50) and (51) imply

$$\begin{aligned} \lim _{T\rightarrow \infty }F(T,k) =\mu \frac{1}{k\tau }. \end{aligned}$$
(55)

Finally, using the expression for E(t) in (25), the values of \(f_{n}\) and \(t_{n}\) in (26) for perfect adherence, and the result in (55), we conclude that

$$\begin{aligned} \langle E \rangle =\mu \frac{E_{\max }}{\text {EC}_{\text {50}}}\frac{DF}{V_{}}\Big ( \frac{b_{ \text {p}0}}{k_{ \text {p}0}\tau } +\frac{b_{ \text {a} \text {p}}}{k_{ \text {a} \text {p}}\tau } +\frac{b_{\text {e}0}}{k_{\text {e}0}\tau } \Big ) =\mu \langle E^{\mathrm{perf}} \rangle . \end{aligned}$$

Derivation of \(\text {CV}(E^{\text {single}})\)

Using the model of nonadherence introduced in “Coefficient of variation estimate”, applying (33) gives

$$\begin{aligned}&\psi (t) :=\frac{E(t)}{\mu \langle E^{\mathrm{perf}} \rangle } =\frac{1}{\mu }\sum _{n=0}^{{\lfloor }{t/\tau }{\rfloor }}\Big (a_{ \text {p}0}e^{-k_{ \text {p}0}(t-n\tau )}\\&\quad +\,a_{ \text {a} \text {p}}e^{-k_{ \text {a} \text {p}}(t-n\tau )}+a_{\text {e}0}e^{-k_{\text {e}0}(t-n\tau )}\Big )f_{n}, \end{aligned}$$

where the coefficients \(a_{ \text {p}0}\), \(a_{ \text {a} \text {p}}\), and \(a_{\text {e}0}\) are defined in (34), \({\lfloor }{t/\tau }{\rfloor }\) denotes the largest integer less than or equal to \(t/\tau\), and \(\{f_{n}\}_{n\in \mathbb {Z}}\) are as in (37). Note that we now take the sequence \(\{f_{n}\}_{n\in \mathbb {Z}}\) to be bi-infinite (i.e. \(n\in \mathbb {Z}\)), which is convenient for the analysis below.

For \(t\in [0,\tau ]\), define

$$\begin{aligned} \Psi (t) :=\frac{1}{\mu }(a_{ \text {p}0}e^{-k_{ \text {p}0}t}A_{ \text {p}0} +a_{ \text {a} \text {p}}e^{-k_{ \text {a} \text {p}}t}A_{ \text {a} \text {p}} +a_{\text {e}0}e^{-k_{\text {e}0}t}A_{\text {e}0}), \end{aligned}$$
(56)

where

$$\begin{aligned} A_{ij} :=\sum _{n=0}^{\infty }(e^{-k_{ij}\tau })^{n}f_{-n},\quad \text {for }i,j\in \{ \text {a}, \text {p},0\}. \end{aligned}$$
(57)

Note that \(A_{ij}\) converges almost surely by the Weierstrass M-test since \(|f_{n}|\) is bounded by a deterministic constant for all \(n\in \mathbb {Z}\) (namely, \(|f_{n}|\le 1\) in this case). Random variables of the form in (57) are often referred to as random pullback attractors [42,43,44,45,46].

It then follows from (32) and Theorem 1 in [20] that

$$\begin{aligned} \begin{aligned} \text {CV}(E^{\text {single}})&=\sqrt{\lim _{T\rightarrow \infty }\frac{1}{T}\int _{0}^{T}\Big (\psi (t)-1\Big )^{2}\,\text {d}t}\\&=\sqrt{\frac{1}{\tau }\int _{0}^{\tau }\mathbb {E}\Big [\big (\Psi (t)-1\big )^{2}\Big ]\,\text {d}t}, \end{aligned} \end{aligned}$$
(58)

where \(\mathbb {E}\) denotes mathematical expectation. Using the definition of \(\Psi (t)\) in (56), equation (58) implies that \(\text {CV}(E^{\text {single}})\) requires calculating expectations of the form

$$\begin{aligned} \mathbb {E}[A_{ij}] \quad \text {and}\quad \mathbb {E}[A_{ij}A_{lm}],\quad \text {for }i,j,l,m\in \{ \text {a}, \text {p},0\}. \end{aligned}$$
(59)

These expectations are obtained immediately from Corollary 3 in [20]. In particular, we have that

$$\begin{aligned} \mathbb {E}[A_{ij}]&=\frac{p}{1-e^{-k_{ij}\tau }},\\ \mathbb {E}[A_{ij}A_{lm}]&=\frac{p}{1-e^{-k_{ij}\tau }e^{-k_{lm}\tau }}\\&\quad +\,p^{2}\Big (\frac{1}{(1-e^{-k_{ij}\tau })(1-e^{-k_{lm}\tau })}\\&\quad -\frac{1}{1-e^{-k_{ij}\tau }e^{-k_{lm}\tau }}\Big ). \end{aligned}$$

Plugging these formulas into (58) and performing the integration yields for the formula for \(\text {CV}(E^{\text {single}})\) in (39).

Derivation of \(\mu ^{\text {double}}\)

For the double dose protocol in ““Make up” doses reduce variation for slow PK or PD”, it is immediate that \(\{f_{n}\}_{n\in \mathbb {Z}}\) is an irreducible, time-homogeneous Markov chain with the following transition probabilities,

$$\begin{aligned} \mathbb {P}(f_{n+1}=0\,|\,f_{n}=i)&=1-p\quad \text {if }i\in \{0,1,2\},\\ \mathbb {P}(f_{n+1}=1\,|\,f_{n}=i)&={\left\{ \begin{array}{ll} 0 &{} \text {if }i=0,\\ p &{} \text {if }i=1,\\ p &{} \text {if }i=2, \end{array}\right. }\\ \mathbb {P}(f_{n+1}=2\,|\,f_{n}=i)&={\left\{ \begin{array}{ll} p &{} \text {if }i=0,\\ 0 &{} \text {if }i=1,\\ 0 &{} \text {if }i=2. \end{array}\right. } \end{aligned}$$

Using standard Markov chain results (for example, see section 1.7 in [47]), a quick calculation shows that the stationary distribution of \(\{f_{n}\}_{n\in \mathbb {Z}}\) is

$$\begin{aligned} \mathbb {P}(f_{n}=i) ={\left\{ \begin{array}{ll} 1-p &{} \text {if }i=0,\\ p^{2} &{} \text {if }i=1,\\ p(1-p) &{} \text {if }i=2. \end{array}\right. } \end{aligned}$$

Using (30), Birkhoff’s ergodic theorem (for example, see section 7.2 in [48]) thus implies that the long-term fraction of doses taken for the double dose protocol is

$$\begin{aligned} \mu ^{\text {double}} =\lim _{N\rightarrow \infty }\frac{1}{N}\sum _{n=0}^{N-1}f_{n} =\sum _{i=0}^{2}i\mathbb {P}(f_{n}=i) =p+p(1-p). \end{aligned}$$

Derivation of \(\text {CV}(E^{\text {double}})\)

The derivation of the formula for \(\text {CV}(E^{\text {double}})\) in (47) is identical to the derivation presented in “Derivation of \(\text {CV}(E^{\text {single}})\)” above except the formulas for the expectations in (59) are given by the formulas for the double dose protocol in Corollary 3 in [20].

Maximum plasma concentration for double dose protocol

Fig. 8
figure 8

Contour plot of the factor \(\varphi\) in (61) which bounds the maximum increase in plasma concentration caused by following the double dose protocol (see (60))

Theorem 5 in [20] yields the following upper bound for the maximum possible plasma compartment concentration obtained by following the double dose protocol,

$$\begin{aligned} c_{ \text {p}}^{\text {double}}(t) \le \max _{s\ge 0}c_{ \text {p}}^{\mathrm{perf}}(s) +\varphi \langle c_{ \text {p}}^{\mathrm{perf}} \rangle , \end{aligned}$$
(60)

where \(\varphi\) is the following dimensionless factor,

$$\begin{aligned} \varphi :=\tau (k_{ \text {a} \text {p}})^{\frac{k_{ \text {p}0}}{k_{ \text {p}0}-k_{ \text {a} \text {p}}}}(k_{ \text {p}0})^{\frac{k_{ \text {a} \text {p}}}{k_{ \text {a} \text {p}}-k_{ \text {p}0}}}, \end{aligned}$$
(61)

and \(\langle c_{ \text {p}}^{\mathrm{perf}} \rangle =\frac{DF}{V}\frac{1}{\tau }\frac{1}{k_{ \text {p}0}}\) is the long-term average plasma concentration for perfect adherence. The upper bound in (60) is valid for any time \(t\ge 0\) and any sequence of remembering or forgetting \(\{\xi _{n}\}_{n\ge 0}\) for the adherence model in ““Make up” doses reduce variation for slow PK or PD” above. The first term in the righthand side of (60) is the maximum plasma concentration obtained by a patient with perfect adherence. The second term in the righthand side of (60) bounds the maximum possible increase in plasma concentration caused by the double dose protocol.

In Fig. 8, we plot the factor \(\varphi\) in (61). This plot shows that \(\varphi \ll 1\) if \(k_{ \text {a} \text {p}}\tau \ll 1\) and/or \(k_{ \text {p}0}\tau \ll 1\). This means that if the PK absorption rate is slow compared to the dosing interval (\(k_{ \text {a} \text {p}}\tau \ll 1\)) and/or the PK elimination rate is slow compared to the dosing interval (\(k_{ \text {p}0}\tau \ll 1\)), then the double dose protocol can at most cause the plasma concentration to rise only slightly above the concentration for perfect adherence.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

McAllister, N.P., Lawley, S.D. A pharmacokinetic and pharmacodynamic analysis of drug forgiveness. J Pharmacokinet Pharmacodyn 49, 363–379 (2022). https://doi.org/10.1007/s10928-022-09808-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10928-022-09808-w

Keywords

Navigation