Skip to main content
Log in

Should patients skip late doses of medication? A pharmacokinetic perspective

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

Abstract

Missed doses, late doses, and other dosing irregularities are major barriers to effective pharmacotherapy, especially for the treatment of chronic conditions. What should a patient do if they did not take their last dose at the prescribed time? Should they take it late or skip it? In this paper, we investigate the pharmacokinetic effects of taking a late dose. We consider a single compartment model with linear absorption and elimination for a patient instructed to take doses at regular time intervals. We suppose that the patient forgets to take a dose and then realizes some time later and must decide what remedial steps to take. Using mathematical analysis, we derive several metrics which quantify the effects of taking the dose late. The metrics involve the difference between the drug concentration time courses for the case that the dose is taken late and the case that the dose is taken on time. In particular, the metrics are the integral of the absolute difference over all time, the maximum of the difference, and the maximum of the integral of the difference over any single dosing interval. We apply these general mathematical formulas to levothyroxine, atorvastatin, and immediate release and extended release formulations of lamotrigine. We further show how population variability can be immediately incorporated into these results. Finally, we use this analysis to propose general principles and strategies for dealing with dosing irregularities.

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

    Article  CAS  PubMed  Google Scholar 

  2. Coleman CI, Limone B, Sobieraj DM, Lee S, Roberts MS, Kaur R, Alam T (2012) Dosing frequency and medication adherence in chronic disease. J Manag Care Pharm 18(7):527–539

    Article  PubMed  Google Scholar 

  3. Burnier M (2019) Is there a threshold for medication adherence? lessons learnt from electronic monitoring of drug adherence. Front Pharmacol 9:1540

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  4. Thier SL, Yu-Isenberg KS, Leas BF, Cantrell CR, DeBussey S, Goldfarb NI, Nash DB (2008) In chronic disease, nationwide data show poor adherence by patients to medication and by physicians to guidelines. Manag Care 17(2):48–52

    PubMed  Google Scholar 

  5. Paschal AM, Rush SE, Sadler T (2014) Factors associated with medication adherence in patients with epilepsy and recommendations for improvement. Epilepsy Behav 31:346–350

    Article  PubMed  Google Scholar 

  6. Cramer JA, Glassman M, Vincent R (2002) The relationship between poor medication compliance and seizures. Epilepsy Behav 3(4):338–342

    Article  PubMed  Google Scholar 

  7. Howard J, Wildman K, Blain J, Wills S, Brown D (1999) The importance of drug information from a patient perspective. J Soc Admin Pharm 16(3–4):115–126

    Google Scholar 

  8. Gilbert A, Roughead L, Sansom L et al (2002) I’ve missed a dose; what should I do? Austral Prescrib 25(1):16–17

    Article  Google Scholar 

  9. Abdullah A, Hughes DA (2020) What should patients do if they miss a dose? A systematic review of patient information leaflets and summaries of product characteristics. Eur J Clin Pharmacol 1:1–10

    Google Scholar 

  10. Sankar AP, Nevedal DC, Neufeld Stewart, Luborsky MR (2007) What is a missed dose? implications for construct validity and patient adherence. AIDS Care 19(6):775–780

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Millum J, Grady C (2013) The ethics of placebo-controlled trials: methodological justifications. Contemp Clin Trials 36(2):510–514

    Article  PubMed  Google Scholar 

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

    Article  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  14. Gibaldi M, Perrier D (1982) Pharmacokinetics. Marcelly Dekker, 2nd edn. Wiley, New York

    Book  Google Scholar 

  15. Bauer LA (2015) Clinical pharmacokinetic equations and calculations. McGraw-Hill Medical, New York

    Google Scholar 

  16. Web app at https://eclark24.shinyapps.io/LateDoseMetrics/

  17. Lexicomp. Levothyroxine. http://online.lexi.com. Accessed 12 Jan 2022

  18. Colucci P, Seng YC, Murray D, Salvatore B (2013) A review of the pharmacokinetics of levothyroxine for the treatment of hypothyroidism. Eur Endocrinol 9(1):40

    Article  PubMed  PubMed Central  Google Scholar 

  19. Demiana WF, Urien S, Carreau V, Chauvenet M, Jean SH, Giral P, Bruckert E, Lechat P (2006) Use of an indirect effect model to describe the ldl cholesterol-lowering effect by statins in hypercholesterolaemic patients. Fund Clin Pharmacol 20(3):321–330

    Article  CAS  Google Scholar 

  20. 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. Therapeut Drug Monit 35(2):188–193

    Article  CAS  Google Scholar 

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

    Google Scholar 

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

    Article  CAS  Google Scholar 

  23. 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 

  24. 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 

  25. Véronique D (2011) Drug forgiveness and interpatient pharmacokinetic variability in tuberculosis. J Infect Dis 204:1

    Google Scholar 

  26. 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 

  27. 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 

  28. McAllister NP, Lawley SD (2022) A pharmacokinetic and pharmacodynamic analysis of drug forgiveness. Submitted

  29. Brent GA, Koenig RJ (2017) Thyroid and antithyroid drugs. McGraw-Hill Education, New York

    Google Scholar 

  30. Clinic Mayo (2022) Levothyroxine. https://www.mayoclinic.org/drugs-supplements/levothyroxine-oral-route/proper-use/drg-20072133. Accessed 12 Jan 2022

  31. RxList. Synthroid. https://www.rxlist.com/synthroid-drug/patient-images-side-effects.htm. Accessed 12 Jan 2022

  32. Drugs.com. Levothyroxine. https://www.drugs.com/levothyroxine.html. Accessed 12 Jan 2022

  33. National Health Service (UK). Levothyroxine. https://www.nhs.uk/medicines/levothyroxine/. Accessed 12 Jan 2022

  34. Jacqueline J, Bianco AC, Bauer AJ, Burman KD, Cappola AR, Celi FS, Cooper DS, Kim BW, Peeters RP, Rosenthal MS et al (2014) Guidelines for the treatment of hypothyroidism: prepared by the American thyroid association task force on thyroid hormone replacement. Thyroid 24(12):1670–1751

    Article  Google Scholar 

  35. Lennernäs H (2003) Clinical pharmacokinetics of atorvastatin. Clin Pharmacokinet 42(13):1141–1160

    Article  PubMed  Google Scholar 

  36. April V, Cynthia S, Judith D (2017) Davis’s drug guide for nurses, 15th edn. New York, F.A, Davis Company

    Google Scholar 

  37. Healthline. Dosage for lipitor: what you need to know. https://www.healthline.com/health/drugs/lipitor-dosage. Accessed 12 Jan 2022

  38. Mould DR, Upton RN (2013) Basic concepts in population modeling, simulation, and model-based drug development part 2: introduction to pharmacokinetic modeling methods. CPT 2(4):1–14

    Google Scholar 

  39. Qiu Y, Zhou D (2011) Understanding design and development of modified release solid oral dosage forms. J Valid Technol 17(2):23

    Google Scholar 

  40. Logan JD (2013) Applied mathematics. Wiley, New York

    Google Scholar 

  41. Garnett WR, McLean AM, Yuxin Z, Susan C, Tulloch SJ (2033) 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

  42. 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 

  43. 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 Therapeut 31(4):321–329

    Article  CAS  Google Scholar 

  44. 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 

  45. Gidal BE, Majid O, Ferry J, Hussein Z, Yang H, Zhu J, Fain R, Laurenza A (2014) The practical impact of altered dosing on perampanel plasma concentrations: pharmacokinetic modeling from clinical studies. Epilepsy Behav 35:6–12

    Article  PubMed  Google Scholar 

  46. 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 

  47. Hard ML, Wehr AY, Sadler BM, Mills RJ, von 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 

  48. 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 

  49. Counterman ED, Lawley SD (2021) What should patients do if they miss a dose of medication? A theoretical approach. J Pharmacokinet Pharmacodyn 48:873

    Article  PubMed  Google Scholar 

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

    Article  Google Scholar 

  51. 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 

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

    Article  Google Scholar 

  53. McAllister NP, Lawley SD (2022) A pharmacokinetic and pharmacodynamic analysis of drug forgiveness. J Pharmacokinet Pharmacodyn 1:1–17

    Google Scholar 

  54. Felmlee MA, Morris ME, Mager DE (2012) Mechanism-based pharmacodynamic modeling. Computational toxicology. Springer, Cham, pp 583–600

    Chapter  Google Scholar 

Download references

Funding

SDL was supported by the National Science Foundation (Grant Nos. CAREER 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 collect details of the mathematical analysis. We also present a fourth metric in Sect. 5.3 for studying the effects of a late dose.

Derivation of mathematical formulas

To calculate \(\rho ^{\text {skip}}\) in (15), we note that

$$\begin{aligned} |c^{\text {skip}}(t)-c(t)| =c_{0}(t), \end{aligned}$$
(45)

where \(c_{0}(t)\) is defined in (5). Integrating (45) yields (16).

To calculate \(\rho _{-}\) and \(\rho _{+}\) defined in (13)-(14), we first use that (10) and (8) imply that \(\rho _{-}\) and \(\rho _{+}\) can be written in terms of \(\delta (t,d)\) in (8) as

$$\begin{aligned} \rho _{-} =\frac{-1}{\tau {\langle }c^{\text {perf}}{\rangle }}\int _{0}^{{{s_{0}}}}\delta (t,d)\,\text {d}t,\quad \rho _{+} =\frac{1}{\tau {\langle }c^{\text {perf}}{\rangle }}\int _{{{s_{0}}}}^{\infty }\delta (t,d)\,\text {d}t, \end{aligned}$$
(46)

where \({{s_{0}}}>d>0\) is such that \(\delta (t,d)<0\) for all \(t\in (0,{{s_{0}}})\) and \(\delta (t,d)>0\) for all \(t>{{s_{0}}}\). Solving for \({{s_{0}}}\) yields

$$\begin{aligned} {{s_{0}}} =\frac{1}{k_{\text {a}}-k_{\text {e}}}\ln \Big (\frac{e^{k_{\text {a}}d}-1}{e^{k_{\text {e}}d}-1}\Big )>d >0. \end{aligned}$$
(47)

Integrating (46) using (8) and (47) yields (17), which then yields (19) by (12).

To calculate \(\nu\) defined in (24), we first use (10) to obtain that \(\nu\) can be written in terms of \(\delta\) as

$$\begin{aligned} \nu =\frac{1}{{\langle }c^{\text {perf}}{\rangle }}\max _{t\in {\mathbb {R}}}\delta (t,d). \end{aligned}$$
(48)

A simple calculus exercise yields that \(\delta (t,d)\) is maximized at \(t={{s_{1}}}\), where

$$\begin{aligned} {{s_{1}}} =\frac{1}{k_{\text {a}}-k_{\text {e}}}\log \Big [\frac{(e^{k_{\text {a}}d}-1)k_{\text {a}}}{(e^{k_{\text {e}}d}-1)k_{\text {e}}}\Big ]. \end{aligned}$$
(49)

Plugging (49) into (48) yields the formula for \(\nu\) in (26).

To calculate \(\gamma\) in (25), we first use (10) to obtain that \(\gamma\) can be written in terms of \(\delta\) as

$$\begin{aligned} \gamma =\frac{1}{{\langle }c^{\text {perf}}{\rangle }}\max _{t\in {\mathbb {R}}}\frac{1}{\tau }\int _{t}^{t+\tau }\delta (s,d)\,\text {d}s. \end{aligned}$$
(50)

Integrating (50) and performing a simple calculus exercise yields that the maximum in (50) occurs at \(t={{s_{2}}}\), where

$$\begin{aligned} {{s_{2}}} =\frac{1}{k_{\text {a}}-k_{\text {e}}}\log \Big (\frac{(e^{k_{\text {a}}\tau }-1) (e^{k_{\text {a}}d}-1)}{(e^{k_{\text {e}}\tau }-1) (e^{k_{\text {e}}d}-1)}\Big )-\tau . \end{aligned}$$
(51)

Plugging (51) into (50) yields the formula for \(\gamma\) in (27).

Monotonicity

We now prove (32) and (33). Applying the chain rule to (19) yields

$$\begin{aligned} \begin{aligned} \frac{\partial }{\partial k_{\text {a}}}{{{\rho }}}&=2f_{x}(k_{\text {a}}d,k_{\text {e}}d)d,\quad \frac{\partial }{\partial k_{\text {e}}}{{{\rho }}} =2f_{y}(k_{\text {a}}d,k_{\text {e}}d)d,\\ \frac{\partial }{\partial d}{{{\rho }}}&=2f_{x}(k_{\text {a}}d,k_{\text {e}}d)k_{\text {a}}+2f_{x}(k_{\text {a}}d,k_{\text {e}}d)k_{\text {e}}, \end{aligned} \end{aligned}$$
(52)

where \(f_{x}\) and \(f_{y}\) denote the partial derivatives of f in (18) with respect to x and y, respectively. Specifically,

$$\begin{aligned} f_{x}(x,y) =\frac{y (\frac{e^y-1}{e^x-1})^{\frac{x}{x-y}} \left( e^x (y-x)-\left( e^x-1\right) \log \left( e^y-1\right) +\left( e^x-1\right) \log \left( e^x-1\right) \right) }{(x-y)^2}, \end{aligned}$$
(53)

and the formula for \(f_{y}(x,y)\) is obtained from (53) upon swapping x and y. Though (53) is a complicated expression, it is easy to plot as a function of \(x>0\) and \(y>0\) to obtain that

$$\begin{aligned} f_{x}(x,y)>0,\quad f_{y}(x,y)>0. \end{aligned}$$
(54)

Hence, (54) and the expressions in (52) yield (32).

The sign of the partial derivatives of \(\gamma\) in (33) follow immediately from (27), the chain rule, and (54).

To obtain the sign of the partial derivatives of \(\nu\) in (33), we first note that the partial derivative of g in (28) with respect to x is

$$\begin{aligned} g_{x}(x,y) =\Big (\frac{y (x \log (x)+y-x \log (y)-x)}{(x-y)^2}\Big )\left( \frac{y}{x}\right) ^{\frac{x}{x-y}}, \end{aligned}$$
(55)

and the formula for \(g_{y}(x,y)\) is obtained from (55) upon swapping x and y. Though (55) is a complicated expression, it is easy to plot as a function of \(x>0\) and \(y>0\) to obtain that

$$\begin{aligned} g_{x}(x,y)>0,\quad g_{y}(x,y)>0. \end{aligned}$$
(56)

Hence, the sign of the partial derivatives of \(\nu\) in (33) follow immediately from using (26), the chain rule, and (54) and (56).

Maximum concentration

In the main text of the paper, the metric \(\nu\) measures the maximum amount that \(c^{d}(t)\) can rise above c(t). To obtain the value for the maximum of \(c^{d}(t)\) rather than how the \(c^{d}(t)\) rises above c(t), we must make assumptions about the patient’s adherence before and after the late dose. For simplicity, we assume that the patient has perfect adherence for a long time before and after the late dose. Letting \(\{c^{\text {perf},d}(t)\}_{t\in {\mathbb {R}}}\) denote this concentration time course, we prove below that

$$\begin{aligned} \max _{t\in {\mathbb {R}}}c^{\text {perf},d}(t) =\max _{j\in \{j^{*},j^{*}+1\}}\big (\delta (j\tau +t_{j},d)+c^{\text {perf}}(t_{j})\big ), \end{aligned}$$
(57)

where \(\delta\) is defined in (8) and

$$\begin{aligned} j^{*}&={\Big \lfloor {\frac{1}{k_{\text {a}}\tau -k_{\text {e}}\tau }\log \Big (\frac{(e^{k_{\text {a}}\tau }-1) (e^{k_{\text {a}}d}-1)}{(e^{k_{\text {e}}\tau }-1) (e^{k_{\text {e}}d}-1)}\Big )-1}\Big \rfloor }, \end{aligned}$$
(58)
$$\begin{aligned} t_{j}&=\frac{1}{k_{\text {a}}-k_{\text {e}}}\log \Big (\frac{k_{\text {a}}(e^{k_{\text {e}}\tau }-1) [(e^{k_{\text {a}}\tau }-1) (e^{k_{\text {a}}d}-1)+e^{(j +1)k_{\text {a}}\tau }] e^{j (k_{\text {e}}-k_{\text {a}})\tau }}{k_{\text {e}}(e^{k_{\text {a}}\tau }-1) [(e^{k_{\text {e}}\tau }-1) (e^{k_{\text {e}}d}-1)+e^{(j+1) k_{\text {e}}\tau }]}\Big ). \end{aligned}$$
(59)

In (58), we use the floor function notation, in which \({\lfloor {x}\rfloor }\) denotes the largest integer less than or equal to x.

While the complicated formulas in (57)-(59) do not offer much intuition, they can be easily plotted to investigate how the maximum concentration depends on the various parameters. In Fig. 8, we plot

$$\begin{aligned} \theta :=\frac{\max _{t\in {\mathbb {R}}}c^{\text {perf},d}(t)-{\langle }c^{\text {perf}}{\rangle }}{{\langle }c^{\text {perf}}{\rangle }}, \end{aligned}$$
(60)

which is a dimensionless measure of how far \(c^{\text {perf},d}(t)\) rises above \({\langle }c^{\text {perf}}{\rangle }\), relative to \({\langle }c^{\text {perf}}{\rangle }\). The top left panel in Fig. 8 is for no delay (\(d=0\)) to show how the concentration time courses rises above the average \({\langle }c^{\text {perf}}{\rangle }\) for perfect adherence. In this case of no delay, \(\theta\) has a simpler formula which we denote by \(\theta ^{\text {perf}}\),

$$\begin{aligned} \theta ^{\text {perf}} =&\lim _{d\rightarrow 0}\theta =\frac{k_{\text {a}}k_{\text {e}}\tau }{k_{\text {a}}-k_{\text {e}}} \Bigg (\frac{\left( \frac{k_{\text {e}}(e^{k_{\text {a}}\tau }-1)}{k_{\text {a}}(e^{k_{\text {e}}\tau }-1)}\right) ^{\frac{k_{\text {e}}}{k_{\text {a}}-k_{\text {e}}}}}{e^{k_{\text {e}}\tau }-1}\nonumber \\&-\frac{\left( \frac{k_{\text {e}}(e^{k_{\text {a}}\tau }-1)}{k_{\text {a}}(e^{k_{\text {e}}\tau }-1)}\right) ^{\frac{k_{\text {a}}}{k_{\text {a}}-k_{\text {e}}}}}{e^{k_{\text {a}}\tau }-1}\Bigg )-1. \end{aligned}$$
(61)

The bottom right panel shows that even in the extreme case of a double dose (\(d=\tau\)), the drug concentration rises only slightly above the average if the absorption and/or elimination rate is sufficiently slow compared to \(1/\tau\). Conversely, this plot shows that the drug concentration can rise far above the average if both the absorption and elimination rate is sufficiently fast compared to \(1/\tau\). Analogous to Figs. 2 and 4, the letter markers in Fig. 8 are for the drugs in Table 1.

Fig. 8
figure 8

Maximum relative increase in concentration \(\theta\) in (60) for different values of the delay \(d\in [0,\tau ]\). Note that the top left panel is for \(d=0\) (see (61)). The letters L, A, I and X correspond to different drugs. See the text for details

To obtain the formula for \(\max _{t\in {\mathbb {R}}}c^{\text {perf},d}(t)\) in (57), we first note that

$$\begin{aligned} c^{\text {perf},d}(t) =c^{\text {perf}}(t)+c^{\text {perf},d}(t)-c^{\text {perf}}(t) =c^{\text {perf}}(t)+\delta (t,d), \end{aligned}$$
(62)

where \(c^{\text {perf}}(t)\) is defined in (6). Now, \(\delta (t,d)\) is strictly increasing for \(t\in ({{s_{0}}},{{s_{1}}})\) and strictly decreasing for \(t\in ({{s_{1}}},\infty )\). It therefore follows that \({{s_{2}}}\in ({{s_{0}}},{{s_{1}}})\) is such that \(\delta ({{t_{1}}},d)>\delta (t_{0},d)\) if \({{t_{1}}}\in ({{s_{2}}},{{s_{2}}}+\tau )\) and \(t_{0}\notin ({{s_{2}}},{{s_{2}}}+\tau )\). Since \(c^{\text {perf}}(t)\) is periodic with period \(\tau\), it follows from (62) that the maximum of \(c^{\text {perf},d}(t)\) must occur at some \(t\in [{{s_{2}}},{{s_{2}}}+\tau ]\). Now, since \(j^{*}\) in (58) satisfies \(j^{*}={\lfloor {{{s_{2}}}/\tau }\rfloor }\), it follows that the maximum of \(c^{\text {perf},d}(t)\) must occur at some \(t\in [j^{*}\tau ,(j^{*}+2)\tau ]\) since \([{{s_{2}}},{{s_{2}}}+\tau ]\subset [j^{*}\tau ,(j^{*}+2)\tau ]\). If \(t=j\tau +s\) for some fixed integer \(j\in {\mathbb {Z}}\) and some time \(s\in [0,\tau ]\), then it follows immediately from (6) that

$$\begin{aligned} c^{\text {perf}}(j\tau +s)&=\frac{DF}{V}\frac{k_{\text {a}}}{k_{\text {a}}-k_{\text {e}}}\nonumber \\&\qquad \big [e^{-k_{\text {e}}s}/(1-e^{-k_{\text {e}}\tau })-e^{-k_{\text {a}}s}/(1-e^{-k_{\text {a}}\tau })\big ]. \end{aligned}$$
(63)

Hence, (62) implies \(c^{\text {perf},d}(j\tau +s)=C_{j}^{d}(s)\) where \(C_{j}^{d}(s)\) is defined to be

$$\begin{aligned} C_{j}^{d}(s)&:=\frac{DF}{V}\frac{k_{\text {a}}}{k_{\text {a}}-k_{\text {e}}}\nonumber \\&\quad \big [e^{-k_{\text {e}}s}/(1-e^{-k_{\text {e}}\tau })-e^{-k_{\text {a}}s}/(1-e^{-k_{\text {a}}\tau })\big ]\nonumber \\&\qquad +\delta (j\tau +s,d). \end{aligned}$$
(64)

Differentiating (64) with respect to s shows that \(\frac{\text {d}}{\text {d}s}C_{j}^{d}(s)=0\) if \(s=t_{j}\) where \(t_{j}\) is defined in (59). Hence, we have obtained (57).

To obtain \(\max _{t\in {\mathbb {R}}}c^{\text {perf}}(t)\), we merely differentiate (63) with respect to s to find the maximum. Plugging the resulting expression into the definition of \(\theta\) in (60) yields the formula for \(\theta ^{\text {perf}}\) in (61).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Clark, E.D., Lawley, S.D. Should patients skip late doses of medication? A pharmacokinetic perspective. J Pharmacokinet Pharmacodyn 49, 429–444 (2022). https://doi.org/10.1007/s10928-022-09812-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10928-022-09812-0

Keywords

Navigation