Parameter | Description | Linezolid | Penicillin G | Methicillin |
α | Elimination rate of S | 39.7 | 0.1204 | 32.4 |
q | Fraction of S that acquires resistance | 4×10−3 | 0.3992 | 0.1234 |
Citation: Hermann Mena, Lena-Maria Pfurtscheller, Jhoana P. Romero-Leiton. Random perturbations in a mathematical model of bacterial resistance: Analysis and optimal control[J]. Mathematical Biosciences and Engineering, 2020, 17(5): 4477-4499. doi: 10.3934/mbe.2020247
[1] | Xiaoxiao Yan, Zhong Zhao, Yuanxian Hui, Jingen Yang . Dynamic analysis of a bacterial resistance model with impulsive state feedback control. Mathematical Biosciences and Engineering, 2023, 20(12): 20422-20436. doi: 10.3934/mbe.2023903 |
[2] | Jing Jia, Yanfeng Zhao, Zhong Zhao, Bing Liu, Xinyu Song, Yuanxian Hui . Dynamics of a within-host drug resistance model with impulsive state feedback control. Mathematical Biosciences and Engineering, 2023, 20(2): 2219-2231. doi: 10.3934/mbe.2023103 |
[3] | Edgar Alberto Vega Noguera, Simeón Casanova Trujillo, Eduardo Ibargüen-Mondragón . A within-host model on the interactions of sensitive and resistant Helicobacter pylori to antibiotic therapy considering immune response. Mathematical Biosciences and Engineering, 2025, 22(1): 185-224. doi: 10.3934/mbe.2025009 |
[4] | Qimin Huang, Mary Ann Horn, Shigui Ruan . Modeling the effect of antibiotic exposure on the transmission of methicillin-resistant Staphylococcus aureus in hospitals with environmental contamination. Mathematical Biosciences and Engineering, 2019, 16(5): 3641-3673. doi: 10.3934/mbe.2019181 |
[5] | Zhun Han, Hal L. Smith . Bacteriophage-resistant and bacteriophage-sensitive bacteria in a chemostat. Mathematical Biosciences and Engineering, 2012, 9(4): 737-765. doi: 10.3934/mbe.2012.9.737 |
[6] | Michele L. Joyner, Cammey C. Manning, Brandi N. Canter . Modeling the effects of introducing a new antibiotic in a hospital setting: A case study. Mathematical Biosciences and Engineering, 2012, 9(3): 601-625. doi: 10.3934/mbe.2012.9.601 |
[7] | Xiaxia Kang, Jie Yan, Fan Huang, Ling Yang . On the mechanism of antibiotic resistance and fecal microbiota transplantation. Mathematical Biosciences and Engineering, 2019, 16(6): 7057-7084. doi: 10.3934/mbe.2019354 |
[8] | Avner Friedman, Najat Ziyadi, Khalid Boushaba . A model of drug resistance with infection by health care workers. Mathematical Biosciences and Engineering, 2010, 7(4): 779-792. doi: 10.3934/mbe.2010.7.779 |
[9] | Jhoana P. Romero-Leiton, Kernel Prieto, Daniela Reyes-Gonzalez, Ayari Fuentes-Hernandez . Optimal control and Bayes inference applied to complex microbial communities. Mathematical Biosciences and Engineering, 2022, 19(7): 6860-6882. doi: 10.3934/mbe.2022323 |
[10] | Mudassar Imran, Hal L. Smith . A model of optimal dosing of antibiotic treatment in biofilm. Mathematical Biosciences and Engineering, 2014, 11(3): 547-571. doi: 10.3934/mbe.2014.11.547 |
Common infections such as pneumonia, urinary infections and post–surgical infections are examples of areas where antibiotics play an important role. Historically, antibiotics have been developed to combat harmful bacteria, but not all bacteria are harmful to health, for instance, in the human organism, there are beneficial bacteria such as those that inhabit the intestines and play an important role in the digestion process [1]. To date, diseases caused by bacteria can be treated with antibiotics as long as they are not resistant to these drugs. The increase in resistance rates is a major problem in medicine because a simple infection can be lethal, also it leads to a rise of social–economical burden due to increased health care costs.
The previous scenario generates the need to research in this area, in order to create strategies that allow to understand the evolution and propagation of bacterial resistance to antibiotics and in some way help to mitigate its evolution. From mathematics, the bacterial resistance phenomenon has been studied using a deterministic and stochastic approach, although deterministic work has dominated strongly on stochastic work [2]. From the deterministic setting, we can highlight the works focused on propagation and transmission of bacterial resistance [3,4,5,6,7,8], identifying the responsible factors of the bacterial resistance [9], examining the bacteria behaviour under the use of different antibiotic treatments [8,10,11,12], optimizing the use of antibiotics [13], helping to design viable control strategies [7], modelling the acquisition of resistance from external sources [14], among others. From a stochastic approach, the works of Philipsen [1] and D'Agata et al. [10] who studied the evolution process of bacterial resistance and the impact of minimizing antibiotic treatment duration are highlighted.
Although the study of deterministic mathematical models provides important results in terms of epidemiological thresholds such as the basic reproduction number, they have some limitations, since it is quite difficult to predict the future behaviour of a biological system with precision [15]. This is due to the fact of the deterministic models do not incorporate the effect of a fluctuating environment (for example, intensity of sunlight, temperature, precipitation, among others), which yields to consider non–constant parameters in the model that oscillate around a certain average value. Therefore, the stochastic mathematical modelling of biological systems is more realistic causing a greater interest in researchers dedicated to answering questions such as: what is the probability that there is an outbreak of a certain disease? How long will a disease likely persist (with or without intervention) [15]?
The stochastic approach for a deterministic mathematical model can be introduced by including noises in the parameters involved in the model, see e.g., [16,17,18,19,20,21]. Nevertheless, there are not many results about introduction of control strategies for stochastic models. As for this topic, we can name the works given on the references [22,23,24,25].
In this work, we will first study the effects caused in the existence and stability of equilibrium solutions properties with the introduction of random perturbations to some parameters of the deterministic mathematical model proposed by Romero et al. [3], which after nondimensionalization is given by the following system of ordinary differential equations (ODEs):
{dS(t)dt=βsS[1−(S+R)]−αS−μsSdR(t)dt=βrR[1−(S+R)]+qαS−μrR. | (1.1) |
In the above mathematical model the variables S and R denote the number of sensitive and resistant bacteria population to antibiotics, respectively. The parameters βs and βr represent the growth rates of sensitive and resistant bacteria, respectively, being βr≤βs [3], and μs and μr the natural death rates of sensitive and resistant bacteria, respectively. The term qαS represent the number of sensitive bacteria that acquire resistance due to contact with the antibiotic, where α is the effectiveness rate of antibiotic and q is the the proportion of bacteria that mutate. A complete description of the model (1.1) can be found in Table 2 in [3]. The qualitative analysis of the previous system was done in terms of certain thresholds representing the basic reproduction numbers of resistant and sensitive bacteria. The existence and global stability conditions of the three equilibrium solutions can be found on page 66 of [3]. Secondly, we will analyse the effects of prophylaxis (such as patient education campaigns) as a control strategy, both in the deterministic and stochastic formulation, to reduce the spread of resistant bacteria. Finally, we will do numerical experiments to validate our theoretical results.
Parameter | Description | Linezolid | Penicillin G | Methicillin |
α | Elimination rate of S | 39.7 | 0.1204 | 32.4 |
q | Fraction of S that acquires resistance | 4×10−3 | 0.3992 | 0.1234 |
This paper is organized as follows: first, we formulate the stochastic mathematical model and obtain conditions for the stability of equilibrium solutions in terms of the stochastic basic reproduction numbers. Then, we formulate and analyse the optimal control problems and investigate the influence of prophylaxis on the reduction of resistant bacteria. Finally, numerical simulations in both cases using data from [3] for bacteria of the genus Staphylococus aeureus are performed to confirm the theoretical results.
In this section we formulate the stochastic version of the mathematical model (1.1), considering the growth rates of bacteria as random rates according to [26]
~βs=βs+ρ1˙W1(t)~βr=βr+ρ2˙W2(t), |
where ρ1 and ρ2∈R and W1(t), W2(t) are independent Brownian motions. This implies that the growth rates of sensitive and resistant bacteria are equal to an average reproduction rates plus an additional time dependent term that follow a normal distribution with a mean of zero. Thus, an stochastic version of (1.1), is given by
{dS=[βsS(1−(S+R))−(μs+α)S]dt+ρ1S(1−(S+R))dW1dR=[βrR(1−(S+R))+qαS−μrR]dt+ρ2R(1−(S+R))dW2. | (2.1) |
Through the rest of this paper, let (Ω,F,P) be a complete probability space with a filtration {Ft}t≥0 satisfying the usual conditions (i.e., it is increasing and right continuous while F0 contains all P-null sets). Now, we stablish a set of biological interest for the system (2.1) as follows
Δ:={(S,R)∈R2:S>0,R>0,0<S+R<1}. | (2.2) |
Theorem 2.1. If (S0,R0)∈Δ is an initial condition and q∈(0,1], then P[(St,Rt)∈Δ]=1. In other words, either Δ is a set of positive invariants or Δ is a positively invariant set.
Proof. Let us define the following functions defined on Δ:
f1(S,R)=βsS(1−(S+R))−(α+μS)S,g1(S,R)=ρ1S(1−(S+R)),f2(S,R)=βrR(1−(S+R)))+qαS−μrR,g2(S,R)=ρ2R(1−(S+R)). |
The Lyapunov operator associated to (2.1) is given by
L=∂∂t+f1∂∂S+f2∂∂R+12g21∂2∂S2+12g22∂2∂R2. | (2.3) |
Let V:Δ→[0,+∞) be the function defined by
V(S,R)=−lnS−lnR−ln(1−(S+R)). | (2.4) |
We claim that LV is bounded above on Δ. In order to prove this, we define
L1:=∂∂t+f1∂∂S+f2∂∂R,L2:=12g21∂2∂S2+12g22∂2∂R2. |
Note that L=L1+L2. Now,
L1V=−βs(1−(S+R))+(μs+α)+βsS−(μs+α)S1−(S+R)−βr(1−(S+R))−qαSR+μr+βrR+qαS1−(S+R)−μrR1−(S+R). |
As 0<q≤1, R,S∈(0,1) and 1−(S+R)>0, this yields to
L1V≤μs+α+βs+μr+βr. | (2.5) |
Similar calculations show that
L2V≤ρ21+ρ22. | (2.6) |
It follows from (2.5) and (2.6) that there is a positive constant C such that LV≤C on Δ. Continuing with the proof, for each k∈N, we define
Δk={(S,R)∈R2:S>1k,R>1kandS+R<1−1k}. |
Note that Δk↗Δ as k→+∞. Defining
τ:=inf{t≥0:S(t)=0∨R(t)=0∨S(t)+R(t)=1}, |
and for each k∈N
τk:=inf{t≥0:S(t)=1k∨R(t)=1k∨S(t)+R(t)=1−1k}. |
Therefore τk↗τ almost surely as k→+∞.
To prove that P[(S(t),R(t))∈Δ]=1 it is equivalent to show that P[τ=+∞]=1. If we assume that P[τ=+∞]<1, then there exists η>0 such that P[τ<∞]>η. As limT→+∞P[τ<T]=P[τ<∞], there exists T>0 such that
P[τ<T]>η. | (2.7) |
As τk converges almost surely to τ, there exists k0∈N such that for every k≥k0 it holds that
P[τk<T]>η. |
Itôs formula yields to
dV=LV+∂V∂SdW1+∂V∂RdW2≤C+∂V∂SdW1+∂V∂RdW2, |
where we used that LV≤C on Δ. Therefore,
V(S(τk∧T),R(τk∧T))≤V(S(0),R(0))+CT+∫τk∧T0∂V∂SdW1+∫τk∧T0∂V∂RdW2, |
and integrating both sides of the above inequality yields to
E[V(S(τk∧T),R(τk∧T))]≤V(S(0),R(0))+CT. |
As
E[V(S(τk∧T),R(τk∧T))]≥E[V(S(τk∧T),R(τk∧T))1{τk<T}]≥−ln(1/k)P[τk<T], |
we obtain
P[τk<T]≤−CT+V(S(0),R(0))ln(1/k). |
Taking the limit k→∞, we get that P[τ<T]=0 which contradicts (2.7). Therefore, it holds that P[τ=+∞]=1, which completes the proof.
Now, based on the below definition, we will determine conditions for the stability of the equilibrium solutions of the stochastic system (2.1).
Definition 2.2 ([27]). Consider the general n–dimensional stochastic system
dX(t)=f(t,X(t))dt+g(t,X(t))dW, | (2.8) |
with initial condition X(0)=X0 and its solution is denoted by X(t,X0). Assume that f(t,0)=g(t,0)=0 for all t≥0, so X=0 is an equilibrium of (2.8). Then, the equilibrium X=0 is said to be
(ⅰ) almost surely exponentially stable if for all X0∈Rn,
lim supt→∞1tln|X(t,X0)|<0a.s.; |
(ⅱ) pth moment exponentially stable if there is a pair of positive constants C1, C2 such that for all X0∈Rn,
E[|X(t,X0)|p]≤C1|X0|pe−C2t,ont≥0. |
When p=2, it is usually said to be exponentially stable in mean square and the equilibrium X=0 is globally asymptotically stable.
Let us note that E0:=(0,0) is the trivial equilibrium of the system (2.1). Conditions for exponential moment stability of E0 in terms of Lyapunov function are given in the following lemma.
Lemma 2.3. ([27, Theorem 3]). Suppose that there exists a function V(t,x)∈C1,2(R×Rn) satisfying the inequalities
K1|x|p≤V(t,x)≤K2|x|pandLV(t,x)≤−K3|x|p,Ki>0,p>0,i∈{1,2,3}. | (2.9) |
Then the equilibrium of the system (2.1) is pth moment exponentially stable.
Theorem 2.4. If the conditions
−βs+(α+μs)−p−12ρ21>0and−βr+μr−p−12ρ22>0 | (2.10) |
hold, then the equilibrium E0 of the system (2.1) is pth moment exponentially stable.
Proof. Let us consider the Lyapunov function
V(S,R)=λ1Sp+λ2Rp, |
where λi, i=1,2 are positive constants. The first inequality of (2.9) is naturally fulfilled. Next, we show the second inequality of (2.9) and apply the Lyapunov operator (2.3) to V.
LV=pλ1[βs(1−(S+R))−(α+μs)+p−12ρ21(1−(S+R))2]Sp+pλ2qαSRp−1+pλ2[βr(1−(S+R))−μr+p−12ρ22(1−(S+R))2]Rp≤[λ1(pβs−p(α+μs)+p(p−1)2ρ21)+λ2qαε1−p]Sp+λ2[pβr−pμr+p(p−1)2ρ22+qα(p−1)ε]Rp=−[λ1(−pβs+p(α+μs)−p(p−1)2ρ21)−λ2qαε1−p]Sp−λ2[−pβr+pμr−p(p−1)2ρ22−qα(p−1)ε]Rp, |
where we used a variant of Young's inequality
xp−1y≤p−1pεxp+1pε1−pyp,x,y>0, |
as well as 0<S+R<1. Under the conditions (2.10) we can choose ε small enough and λ1, λ2>0 such that the coefficients of Sp and Rp are negative. This completes the proof.
From Theorem 2.4 we derive conditions such that the equilibrium E0 is globally asymptotically stable.
Lemma 2.5. If −βs+(α+μs)−12ρ21>0 and −βr+μr−12ρ22>0, the equilibrium E0 of the system (2.1) is globally asymptotically stable.
Proof. Taking p=2 in (2.10) yields immediately to the desired result.
Next, we want to derive similar basic reproduction numbers for the resistant and sensitive bacteria, respectively as in the deterministic formulation. Recall, for the unperturbed model, those thresholds are given by
Rr=βrμrandS0=βsα+μs, |
and depending on these values, there are three equilibrium solutions, E0 where both bacteria die out, E1 where only resistant bacteria survives, and the equilibrium of coexistence E2. As for the perturbed models these thresholds differ, see e.g., [28], we will study the behaviour of the stochastic model.
Theorem 2.6. Let (S,R)∈Δ be the solution of (2.1). The following holds:
If either
SS0:=βs−12ρ21α+μs<1,andρ21≤βs, | (2.11) |
or
ρ21>max{β2s2(α+μs),βs}, | (2.12) |
and either
RSr:=βr−12ρ22μr<1,andρ22≤βr, | (2.13) |
or
ρ22>max{β2r2μr,βr}, | (2.14) |
then the solution of the system (2.1) has the following property:
lim supt→∞logS(t)t≤−a<0a.s.,lim supt→∞logR(t)t≤−b<0a.s., |
namely it tends to zero exponentially almost surely, where a,b>0 are positive constants. In other words, both the sensitive and the resistant bacteria die out with probability one.
Proof. The proof follows closely the one given in [28]. By Itô's formula, it holds that
dlnS=[βs(1−(S+R))−(α+μs)−12ρ21(1−(S+R))2]dt+12ρ1(1−(S+R))dW1. |
As (S,R)∈Δ, we have that S+R∈(0,1) and using the large number theorem for martingales [29], we get
lim supt→∞1t∫t0(1−(S+R))dW1=0a.s.. |
Moreover the function f:(0,1)→R defined by
f(x)=βsx−ρ212x2−(α+μs)=−ρ212(x−βsρ21)2−(α+μs)+β2s2ρ21 |
has its maximum value at x=1 if ρ21<βs. This yields to
lim supt→∞1tlnS≤(−ρ212+βs−(α+μs))<0⟺SS0<1. |
On the other hand, if (2.12) holds, then
f(x)=−ρ212(x−βsρ21)2−(α+μs)+β2s2ρ21≤−(α+μs)+β2s2ρ21<0, |
and therefore, also
lim supt→∞1tlnS<0. |
Next we show that also R tends to zero exponentially almost surely by using condition (2.13). Without loss of generality it holds that 0≤S≤εR, as R∈Δ and S tends to zero exponentially a.s.. Then, Itô's formula gives
dlnR≤[βr(1−(S+R))−μr+qαε−12ρ22(1−(S+R))2]dt+12ρ2(1−(S+R))dW2. |
Using a similar argument as above, we can define a function ˜f:(0,1)→R,x↦βrx−μr−12ρ22x2, and its maximum value is obtained at x=1 if ρ22≤βr. Taking the limit t→∞, ε→0 yields to
lim supt→∞1tlnR≤−12ρ22+βr−μr<0⟺RSr<1. |
The same assertion can be shown when using condition (2.14). This follows analogously and will be omitted here.
Remark 2.7. The conditions (2.11) and (2.13) in Theorem 2.6 state that the bacteria will extinct if RSr<1 and SS0<1 and the white noise is not too large or if the white noise is large enough such that the conditions (2.12) and (2.14) are fulfilled. Moreover, the thresholds compared to the unperturbed system differ by the noise parameters ρ1, ρ2, and the conditions for the stochastic model are weaker than for the deterministic system. This will be also illustrated in more detail in Subsection 2.2.
The remaining part of this subsection deals with the persistence of bacteria.
Theorem 2.8. Let (S,R)∈Δ be the solution of (2.1). Then,
(i) If RSr>1 and (2.11) or (2.12) are fulfilled, then the sensitive bacteria tend to zero exponentially and the resistant bacteria satisfy
lim inft→∞1t∫t0R(s)ds=:lim inft→∞⟨R⟩≥RSr−1RSr>0, | (2.15) |
i.e. the resistant bacteria are persistent in mean.
(ii) If SS0>1, Rr<SS0, RSr<SS0 and μr(SS0−Rr)>SS0−S0SS0−1qαSS0, then both the sensitive and resistant bacteria are persistent in mean and they satisfy
lim inft→∞⟨R⟩≥qα(SS0−1)(qα+μr)(SS0−μrμr+qαRSr(SS0−S0+1))>0,lim inft→∞⟨S⟩≥1SS0(SS0−S0)qα+(SS0−1)μr(1−RrSS0)qα+μr(1−RrSS0)>0. |
Proof. First we show (i). Since either (2.11) or (2.12) hold, it follows from Theorem 2.6 that limt→∞S(t)=0 a.s.. Moreover, applying Itô's formula to V=lnR yields
dlnR=[βr(1−(S+R))−μr+qαSR−ρ222(1−(S+R))2]dt+ρ2(1−(S+R))dW2. |
Thus,
lnR(t)−lnR(0)t≥βr−(βr−ρ222)⟨S⟩−(βr−ρ222)⟨R⟩−μr+qα⟨SR⟩−ρ222+ρ2t∫t0(1−(S+R))dW2, |
where we used that 0<R<1 and 0<S+R<1 and in particular for 0≤S≤εR it holds that
⟨R⟩≥1βr−ρ222(βr−μr−ρ222+qαε−(βr−ρ222)ε)+ρ2t∫t0(1−(S+R))dW2+1tlnR(0)−1tlnR(t). |
The large number theorem for martingales [29] states that
limt→∞ρ2t∫t0(1−(S+R))dW2=0,limt→∞lnR(0)t=0, |
and taking the limit t→∞ and ε→0, we obtain
lim inft→∞⟨R⟩≥RSr−1RSr>0⟺RSr>1. |
Next, we show assertion (ii). Itô's formula implies that
dlnS=[βs(1−(S+R))−(α+μs)−ρ212(1−(S+R))2]dt+ρ1(1−(S+R))dW1, |
and in particular,
⟨S⟩≥1βs−ρ212(βs−(α+μs)−ρ212−(βs−ρ212)⟨R⟩)+φ1(t),=(SS0−1SS0)−⟨R⟩+φ1(t), | (2.16) |
where we used the definition of SS0 and φ1 is given by
φ1(t)=ρ1t∫t0(1−(S+R))dW1−lnS(t)−lnS(0)t. |
Similarly we can show that
⟨S⟩≤1βs−ρ212(βs−(α+μs)−(βs−ρ212)⟨R⟩)+φ1(t),=S0−1SS0−⟨R⟩+φ1(t). | (2.17) |
Integration of the second equation of (2.1) leads to
R(t)−R(0)t=(βr⟨R(1−(S+R))⟩−μr⟨R⟩+qα⟨S⟩)dt+ρ2t∫t0R(1−(S+R))dW2≥((βr−ρ222)⟨R(1−(S+R))⟩−μr⟨R⟩+qα⟨S⟩)dt+ρ2t∫t0R(1−(S+R))dW2, |
where we used that 0<S+R<1. To this end, inserting the inequalities (2.16) and (2.17) yields to
R(t)−R(0)t≥((βr−ρ222)⟨R⟩−(βr−ρ222)S0−1SS0⟨R⟩−(μr+qα)⟨R⟩+SS0−1SS0qα)dt+ρ2t∫t0R(1−(S+R))dW2+(qα−βr+ρ222)φ1(t), |
and therefore,
⟨R⟩≥qαSS0−1SS0qα+μr−(βr−ρ222)(1−S0−1SS0)+φ(t)=qα(SS0−1)(qα+μr)(SS0−μrμr+qαRSr(SS0−S0+1))+φ(t), | (2.18) |
where φ is defined by
φ(t)=ρ2t∫t0R(1−(S+R))dW2+(qα−βr+ρ222)φ1(t)−R(t)−R(0)t. |
Using again the large number theorem of martingales, we obtain that
limt→∞φ(t)=0. |
Thus, it holds that
lim inft→∞⟨R⟩≥qα(SS0−1)(qα+μr)(SS0−μrμr+qαRSr(SS0−S0+1))>0, |
if the conditions of (ⅱ) are fulfilled. Similarly one could show that
lim supt→∞⟨R⟩≤qα(S0−1)(qα+μr)(SS0−μrRrqα+μr). | (2.19) |
Finally, we have to show that also lim inft→∞⟨S⟩>0. Due to (2.16) and (2.19), it holds that
lim inft→∞⟨S⟩≥SS0−1SS0−lim supt→∞⟨R⟩≥SS0−1SS0−qα(S0−1)(qα+μr)(SS0−μrRrμr+qα)=1SS0((SS0−1)−qα(S0−1)qα+μr−βrSS0)=SS0−1SS0μr(1−RrSS0)qα+μr(1−RrSS0)+SS0−S0SS0qαqα+μr(1−RrSS0). |
This concludes the proof.
In this subsection, we do some numerical experiments to corroborate our theoretical results. The values of the dimensionless parameters were obtained from [3] and are given in Table 1. According to [3], the unit of time before nondimensionalization corresponds to days.
Parameter | Definition | Value |
βs | Growth rate of S | 0.4 |
βr | Growth rate of R | 0.1 |
μs | Natural death rate of S | 0.2 |
μr | Natural death rate of R | 0.5 |
ρ1 | Intensity of noise for βs | 0.5 |
ρ2 | Intensity of noise for βr | 0.5 |
Recall the parameter α represents the elimination rate of bacteria by antibiotic which is considered to vary depending on the antibiotic supplied: Linezolid, Penicillin G or Methicillin. Table 2 shows the values of these variations.
The previous section tells us, that the trivial equilibrium E0 of the system (2.1) exists if SS0<1 and ρ21<βs and RSr<1 and ρ22<βr, respectively or if the noise parameters are large compared to the other model parameters. The goal of this subsection is to confirm the theoretical results with numerical simulations. In the following, we approximate the mean of the solution of the stochastic system by performing a Monte–Carlo simulation using 5000 different paths and compare it with the deterministic results.
As a first example, we take the values from Table 1 and Linezolid as antibiotic. In this case it holds that Rr=0.2, S0=0.01 and RSr=−0.05 and SS0=0.069 respectively. Thus, using the results of Table 2 from [3], we should obtain the equilibrium E0 see Figure 1).
We see that both the resistant and the sensitive bacteria tend to zero for the deterministic model and this is also the case for the stochastic model.
As a next example, we use Methicillin and change the natural death rate of R to 0.09. Then, for the deterministic model, we have Rr=1.11 and S0=0.0123 but ρ22>βr respectively. The theoretical analysis of the previous section tells that in the stochastic case both bacteria dies out while in the deterministic case, the resistant bacteria survives. However, if we decrease ρ2 to 0.1, we obtain that RSr>1 and the behaviour of the stochastic system coincides with the deterministic one. Therefore, in Figure 2 the mean of the solution of the stochastic model is plotted for both, ρ2=0.5 and ρ2=0.1 respectively.
The theoretical results for the equilibrium E1 coincide with the numerical results as can be seen in Figure 2. For both models only the sensitive bacteria goes extinct if we use ρ2=0.1, however, for ρ2=0.1 both bacteria die out.
Finally, Penicillin G is taken as antibiotic. Then, S0>1 but for the stochastic setting we have SS0<1. Thus, we need to change ρ1 to 0.3, which leads to SS0>1 and both, the sensitive and resistant bacteria should be present, which is confirmed by Figure 3.
The numerical simulations of equations (1.1) and (2.1) which can be seen in Figures 1–3 confirm the theoretical results derived in Section 2 and are in line with the deterministic results from [3].
From a mathematical approach only a small number of studies have established analytical results for the optimal control of an infectious disease under drug resistance, see e.g., [11,30,31,32,33]. Here, we do not consider a population–model for individuals infected with a bacterial disease, but we consider a simple model where sensitive and resistant bacteria interact. Thus, in this section we study the optimal control problem for the deterministic model (1.1) which motivates the stochastic case (2.1). We first analyse the setting, show the existence of the control and also present some numerical simulations.
In order to formulate the deterministic optimal control problem, the following additional hypothesis in the model (1.1) are considered: the population of resistant bacteria are perturbed by prophylaxis at a rate (1−η(t))qα, being η(t) the control by prophylaxis, which assumes values between 0 and 1, where η=0 is assumed if the prophylaxis is ineffective and η=1 if it is completely effective, that is, there is no mutation due to antibiotics. Thus, we have that the control variable η(t) provides information about the amount of patients that must be educated.
The main goal will be to minimize the number of resistant bacteria. For this purpose, the following cost function is considered
J(η)=∫T0(aR+12bη2)dt. | (3.1) |
In above function a represents the social cost, which depends on the number of resistant bacteria due to mutations, and the function 12bη2 defines the absolute cost associated to the control strategy, such as implementation, ordering, distribution, among others.
Taking x=(S,R) and the above considerations, the following optimal control problem is formulated
{J(η)=min∫T0(aR+12bη2)dtdSdt=βsS[1−(S+R)]−αS−μsSdRdt=βrR[1−(S+R)]+(1−η)qαS−μrRx(0)=(ˉS,ˉR)=x0x(T)=(Sf,Rf)=x1. | (3.2) |
In our control problem, we assume an initial time t0=0, a final time T fixed representing the implementation time of the control strategy, and also we assume free dynamic variables x1 at the final time, while the coordinates of the initial condition x0 are the coordinates of a non–trivial equilibrium of the system (1.1). Additionally, we assume that control variable is in an appropriate set of admissible controls, namely U={η(t):η(t)is Lebesgue measurable and0≤η(t)≤1,t∈[0,T]}.
In the following, we prove the existence and uniqueness of the control η. To this end, using arguments similar to [34], we have to show that the following properties have to be fulfilled:
(ⅰ) The set of all solutions to the state equations from (3.2) with corresponding control functions in U is not empty.
(ⅱ) The right side of the state equation from (3.2) is continuous, bounded above by a sum of bounded control and state, and can be written as a linear function of the control variables with coefficients dependent on time and the state variables.
(ⅲ) The integrand in the cost function aR+12bη2 is convex on U and additionally satisfies
aR+12bη2≥k1|η|δ−k2wherek1,k2>0,δ>1. |
Assumption (ⅰ) is fulfilled by Theorem 1 from [4]. Rewriting the right side of the state system as ˙x=f(x)+g(x)η immediately leads to Assumption (ⅱ). Moreover, the integrand in the cost function is clearly convex and we have that
aR+12bη2≥−aR+12bη2≥−k2+k1|η|2, |
such that Assumption (ⅲ) is fulfilled. Thus, there exists an optimal control.
Now, the Pontryagin principle for bounded controls [35] is used to calculate the optimal control of (3.2). To this end, we observe that the Hamiltonian is given by
H=aR+12bη2+λ1[βsS(1−(S+R))−αS−μsS]+λ2[βrR(1−(S+R))+qαS−μrR], | (3.3) |
where λi, i=1,2 are the adjoint variables which determine the adjoint system. The adjoint system and the formulation (3.2) define the optimal system. The main result of this section is summarized in the following theorem.
Theorem 3.1. For (3.2) there exists a corresponding optimal solution (S⋆(t),R⋆(t)) that minimize J(η) in [0,T]. Moreover, there exits an adjoint function λ(t)=(λ1(t),λ2(t)) such that
{dλ1dt=λ1(2βsS+βsR+μs+α−βs)−λ2[βrR+(1−η(t))qα]dλ2dt=−a+λ1βsS+λ2(2βrR+βrS+μr−βr) | (3.4) |
with transversality conditions λi(t)=0 for i=1,2 which satisfy
η⋆=min{max{0,λ2qαSb},1}. |
Proof. The Pontryagin Principle applied to (3.2) guarantees the existence of adjoint variables λi, i=1,2 that satisfy
˙λ1=−∂H∂S,λ1(T)=0,˙λ2=−∂H∂R,λ2(T)=0,H=maxη∈UH. |
Replacing the derivatives of H with respect to S and R in above equations we obtain the system (3.4). The optimality condition for the Hamiltonian is given by ∂H∂η⋆=0, or equivalently bη⋆−λ2qαS=0, from where η⋆=λ2qαSb. In consequence, the optimal control η⋆ is given by
η⋆=min{max{0,λ2qαSb},1}. |
In this subsection, some numerical simulations for the problem (3.2) are performed to observe the effects of prophylaxis as control strategy. To this end, we use the forward–backward sweep method described on [36]. The implementation time of control is approximately 10 days and the values of the relative weights are a=b=0.002 as it is suggested in [37]. Here we omit the case of Linezolid as the bacteria die out even without control strategy.
In Figure 4 we observe the control effectiveness. From the first moment the Methicillin–resistant bacteria population R is controlled while without control R grows. On the other hand, η starts at the limit of 90% during the first moment and then decreases rapidly until reaching its lower level of 0% before the first day.
Figure 5 shows the case of Penicillin G–resistant bacteria. Despite the cost and effort generated by controlling the resistance with the prophylaxis strategy, control can not be achieved. As less resistant bacteria survive, the costs of the prophylaxis, have to be much lower than the social costs, to achieve an effective control.
As discussed in the previous sections, introducing random perturbations to the deterministic system will lead to different behaviour of the bacteria. Thus, we will study in this subsection a stochastic optimal control problem where the state equations are given by
{dS=[βsS(1−(S+R))−(μs+α)S]dt+ρ1S(1−(S+R))dW1dR=[βrR(1−(S+R))+(1−η)qαS−μrR]dt+ρ2R(1−(S+R))dW2. | (3.5) |
The solutions (S,R) are random variables, thus we need to take the expected value of the cost function (3.1), which leads in minimizing
J(η)=E[∫T0(aR+12bη2)dt]. |
As the diffusion does not contain the control variable, the Hamiltonian H is defined by (see e.g. [38])
H=aR+12bη2+p1[βsS(1−(S+R))−(μs+α)S]+p2[βrR(1−(S+R))+(1−η)qαS−μrR]+q1[ρ1S(1−(S+R))]+q2[ρ2R(1−(S+R))], |
where the pairs (p1,q1) and (p2,q2) solve the backward stochastic differential equations
{dp1(t)=−[(βs(1−(S+R)−S)−(μs+α))p1+(−βrR+(1−η)qα)p2+(ρ1(1−(S+R)−S)q1−ρ2yq2)]dt+q1dW1dp2(t)=−a−[−βsSp1+(βr(1−(S+R)−R)−μr)p2−ρ1Sq1+ρ2(1−(S+R)−R)q2]dt+q2dW2p1(T)=0p2(T)=0. | (3.6) |
Moreover, as in the deterministic case, it holds that ∂H∂η⋆=0, or equivalently
η⋆=min{max{0,p2qαSb},1}. | (3.7) |
The main difficulty in the numerical simulations is the approximation of the stochastic forward–backward differential equation (3.5)–(3.6). Due to the four–step scheme [39], these equations are related to the solution of quasi–linear partial differential equations which we solve by the probabilistic approach presented by [40]. The implementation time of control strategies is again 10 days and the values of the relative weights are a=b=0.002. The parameters in the equations are the same as in Subsection 2.2, and as in Subsection 3.2, only the cases of Methicillin Penicillin G resistant bacteria are considered. Moreover, from the study in Section 2, we choose for the Methicillin simulation ρ2=0.1 and for the Penicillin simulation ρ1=0.3, respectively, as for greater values of ρ1 and ρ2 both bacteria die out and no control strategy is necessary.
We observe a similar behaviour as in the deterministic case comparing with Figures 4–5. Again, Methicillin-resistant bacteria is controlled effectively by η and the resistant bacteria without control grows much faster than the controlled one (see Figure 6). However, repeating the experiment with Penicillin G-resistant bacteria, the uncontrolled and the controlled resistant bacteria behave similarly (Figure 7) and a lower value of b is needed in order to have an effective control strategy.
This paper presented a mathematical study describing the dynamical behaviour of bacterial resistance to antibiotics model with perturbed reproduction rates. Our purpose was based on analysing this behaviour using both a deterministic model and a stochastic one. The deterministic case was studied in [3]. Concerning the stochastic model, we obtained sufficient conditions for the stability of the trivial equilibrium E0 in the probability sense by using a suitable Lyapunov function and other techniques of stochastic analysis. The investigation of this stochastic model revealed that the stochastic stability of E0 depends on the magnitude of the intensity of noise. Moreover, we determined the stochastic reproduction numbers for sensitive and resistant bacteria and we gave conditions for the cases that only resistant bacteria survive and that both bacteria persist.
The main result of this work was based on the formulation of optimal control problems in both cases (deterministic and stochastic). The control strategy used was prophylaxis, that is, control by patient education campaigns and an adequate supply of antibiotics. Sufficient conditions were derived to show existence and uniqueness of the control problem. The numerical simulations of the deterministic control problem were performed by using data from bacterial cells of the genus Staphylococcus under three types of antibiotics supply: Penicillin G (low efficacy), Methicillin (medium efficacy) and Linezolid (high efficiency). The numerical results revealed that when bacteria are treated with Linezolid they are naturally eradicated, when they are treated with Methicillin the sensitive bacteria are eliminated and a persistence of resistant bacteria is generated, which can be reduced incorporating a prophylactic control strategy. The most interesting case occurred when bacteria are treated with Penicillin G, as there is a persistence of sensitive and resistant bacteria that cannot be controlled with the same costs for prophylaxis. Therefore, these results suggested that one has to lower the costs for the prophylaxis in comparison to the social costs to obtain an effective control strategy.
We want to thank to anonymous referees for their valuable suggestions and comments that helped us to improve this manuscript. This work was supported by a scholarship of the Vizerektorat für Forschung, University of Innsbruck and by the Austrian Science Fund (FWF)–project id: P27926.
No potential conflict of interest was reported by the authors.
[1] | K. R. Philipsen, Nonlinear Stochastic Modelling of Antimicrobial resistance in Bacterial Populations, PhD thesis, Technical University of Denmark (DTU), 2010. |
[2] | I. Nåsell, Stochastic models of some endemic infections, Math. Biosci., 179 (2002), 1-19. |
[3] | J. P. Romero-Leiton, E. Ibargüen-Mondragón, L. Esteva, Un modelo matemático sobre bacterias sensibles y resistentes a antibióticos, Matemáticas: Enseñanza Universitaria, 19 (2011), 55-73. |
[4] | J. P. Romero-Leiton, E. Ibargüen-Mondragón, Sobre la resistencia bacteriana hacia antibióticos de acción bactericida y bacteriostática, Revista Integración, 32 (2014), 101-116. |
[5] | E. Ibargüen-Mondragón, S. Mosquera, M. Cerón, E. M. Burbano-Rosero, S. P. Hidalgo-Bonilla, L. Esteva, J. P. Romero-Leiton, Mathematical modeling on bacterial resistance to multiple antibiotics caused by spontaneous mutations, Biosystems, 117 (2014), 60-67. |
[6] | E. Ibargüen-Mondragón, J. P. Romero-Leiton, L. Esteva, E. Burbano, Mathematical modeling of bacterial resistance to antibiotics by mutations and plasmids, J. Biol. Systems, 24 (2016), 129-146. |
[7] | R. A. Weinstein, M. J. Bonten, D. J. Austin, M. Lipsitch, Understanding the spread of antibiotic resistant pathogens in hospitals: mathematical models as tools for control, Clin. Infect. Dis., 33 (2001), 1739-1746. |
[8] | J. Alavez-Ramirez, J. R. A. Castellanos, L. Esteva, J. A. Flores, J. L. Fuentes-Allen, G. GarcíaRamos, G. Gómez, J. López-Estrada, Within-host population dynamics of antibiotic-resistant M. tuberculosis, Math. Med. Biol., 24 (2007), 35-56. |
[9] | D. Austin, R. Anderson, Studies of antibiotic resistance within the patient, hospitals and the community using simple mathematical models, Philos. Trans. R. Soc. Lond. B. Biol. Sci., 354 (1999), 721-738. |
[10] | E. M. D'Agata, P. Magal, D. Olivier, S. Ruan, G. F. Webb, Modeling antibiotic resistance in hospitals: the impact of minimizing treatment duration, J. Theor. Biol., 249 (2007), 487-499. |
[11] | S. Bonhoeffer, M. Lipsitch, B. R. Levin, Evaluating treatment protocols to prevent antibiotic resistance, Proc. Natl. Acad. Sci. USA, 94 (1997), 12106-12111. |
[12] | M. Bootsma, M. van der Horst, T. Guryeva, B. Ter Kuile, O. Diekmann, Modeling non-inherited antibiotic resistance, Bull. Math. Biol., 74 (2012), 1691-1705. |
[13] | E. Massad, M. N. Burattini, F. A. B. Coutinho, An optimization model for antibiotic use, Appl. Math. Comput., 201 (2008), 161-167. |
[14] | F. Hellweger, X. Ruan, S. Sanchez, A simple model of tetracycline antibiotic resistance in the aquatic environment (with application to the Poudre river), Int. J. Environ. Res. Public Health, 8 (2011), 480-497. |
[15] | T. Saha, M. Bandyopadhyay, Dynamical analysis of a delayed ratio-dependent prey-predator model within fluctuating environment, Appl. Math. Comput., 196 (2008), 458-478. |
[16] | S. B. Mortensen, S. Klim, B. Dammann, N. R. Kristensen, H. Madsen, R. V. Overgaard, A Matlab framework for estimation of NLME models using stochastic differential equations, J. Pharmacokinet. Pharmacodyn., 34 (2007), 623-642. |
[17] | S. Klim, S. B. Mortensen, N. R. Kristensen, R. V. Overgaard, H. Madsen, Population stochastic modelling (PSM)-an R package for mixed-effects models based on stochastic differential equations, Comput. Methods Programs Biomed., 94 (2009), 279-289. |
[18] | M. W. Pedersen, D. Righton, U. H. Thygesen, K. H. Andersen, H. Madsen, Geolocation of north sea cod (gadus morhua) using hidden markov models and behavioural switching, Can. J. Fish. Aquat. Sci., 65 (2008), 2367-2377. |
[19] | J. L. Jacobsen, H. Madsen, P. Harremoës, A stochastic model for two-station hydraulics exhibiting transient impact, Water Sci. Technol., 36 (1997), 19-26. |
[20] | H. Jonsdottir, H. Madsen, O. P. Palsson, Parameter estimation in stochastic rainfall-runoff models, J. Hydrol., 326 (2006), 379-393. |
[21] | J. U.-M. Nielsen, Price-quality competition in the exports of the central and eastern european countries, Intereconomics, 35 (2000), 94-101. |
[22] | S. U. Acikgoz, U. M. Diwekar, Blood glucose regulation with stochastic optimal control for insulin-dependent diabetic patients, Chem. Eng. Sci., 65 (2010), 1227-1236. |
[23] | P. Grandits, R. M. Kovacevic, V. M. Veliov, Optimal control and the value of information for a stochastic epidemiological SIS-model, J. Math. Anal. Appl., 476 (2019), 665 - 695. |
[24] | P. J. Witbooi, G. E. Muller, G. J. Van Schalkwyk, Vaccination control in a stochastic SVIR epidemic model, Comput. Math. Methods Med., 2015. |
[25] | R. Aboulaich, A. Darouichi, I. Elmouki, A. Jraifi, A stochastic optimal control model for BCG immunotherapy in superficial bladder cancer, Math. Model. Nat. Phenom., 12 (2017), 99-119. |
[26] | J. H. Brown, J. F. Gillooly, A. P. Allen, V. M. Savage, G. B. West, Toward a metabolic theory of ecology, Ecology, 85 (2004), 1771-1789. |
[27] | A. Lahrouz, L. Omari, D. Kiouach, Global analysis of a deterministic and stochastic nonlinear SIRS epidemic model. |
[28] | Y. Zhao, D. Jiang, X. Mao, A. Gray, The threshold of a stochastic SIRS epidemic model in a population with varying size, Discrete Contin. Dyn. Syst. Ser. B, 20 (2015). 1277-1295, |
[29] | X. Mao, Stochastic differential equations and applications, Elsevier, 2007. |
[30] | N. Chehrazi, L. Cipriano, E. Enns, Dynamics of drug resistance: Optimal control of an infectious disease, Available at SSRN, URL http://dx.doi.org/10.2139/ssrn.2927549. |
[31] | N. I. Stilianakis, A. S. Perelson, F. G. Hayden, Emergence of drug resistance during an influenza epidemic: insights from a mathematical model, J. Infect. Dis., 177 (1998), 863-873. |
[32] | K. Leung, M. Lipsitch, K. Y. Yuen, J. T. Wu, Monitoring the fitness of antiviral-resistant influenza strains during an epidemic: A mathematical modelling study, Lancet Infect. Dis., 17 (2017), 339- 347. |
[33] | B. Petrie, R. Barden, B. Kasprzyk-Hordern, A review on emerging contaminants in wastewaters and the environment: Current knowledge, understudied areas and recommendations for future monitoring, Water Res., 72 (2015), 3-27. |
[34] | A. Permatasari, R. Tjahjana, T. Udjiani, Existence and characterization of optimal control in mathematics model of diabetics population, in J. Phys. Conf. Ser., vol. 983, 2018, 1-6. |
[35] | L. S. Pontryagin, V. G. Boltyanskii, R. V. Gamkrelidze, E. F. Mishchenko, The Mathematical Theory of Optimal Processes., New York and London: Interscience Publisher, 1962. |
[36] | S. Lenhart, J. T. Workman, Optimal control applied to biological models, CRC Press, 2007. |
[37] | J. P. Romero-Leiton, E. Ibargüen-Mondragón, Stability analysis and optimal control intervention strategies of a malaria mathematical model, Appl. Sci., 21 (2019), 184-217. |
[38] | J. Yong, X. Y. Zhou, Stochastic controls: Hamiltonian systems and HJB equations, vol. 43, Springer Science & Business Media, 1999. |
[39] | J. Ma, P. Protter, J. Yong, Solving forward-backward stochastic differential equations explicitly-a four step scheme, Probab. Theory Related Fields, 98 (1994), 339-359. |
[40] | G. Milstein, M. V. Tretyakov, Numerical algorithms for forward-backward stochastic differential equations, SIAM J. Sci. Comput., 28 (2006), 561-582. |
1. | Jhoana P. Romero-Leiton, Kernel Prieto, Daniela Reyes-Gonzalez, Ayari Fuentes-Hernandez, Optimal control and Bayes inference applied to complex microbial communities, 2022, 19, 1551-0018, 6860, 10.3934/mbe.2022323 | |
2. | Jing Jia, Yanfeng Zhao, Zhong Zhao, Bing Liu, Xinyu Song, Yuanxian Hui, Dynamics of a within-host drug resistance model with impulsive state feedback control, 2022, 20, 1551-0018, 2219, 10.3934/mbe.2023103 | |
3. | Sida Kang, Tianhao Liu, Hongyu Liu, Yuhan Hu, Xilin Hou, Sanket Kaushik, Dynamic analysis and optimal control of stochastic information cross-dissemination and variation model with random parametric perturbations, 2024, 19, 1932-6203, e0303300, 10.1371/journal.pone.0303300 | |
4. | Sha He, Sanyi Tang, Qimin Zhang, Libin Rong, Robert A. Cheke, Modelling optimal control of air pollution to reduce respiratory diseases, 2023, 458, 00963003, 128223, 10.1016/j.amc.2023.128223 | |
5. | Sida Kang, Xilin Hou, Yuhan Hu, Hongyu Liu, Dynamic analysis and optimal control of a stochastic information spreading model considering super-spreader and implicit exposer with random parametric perturbations, 2023, 11, 2296-424X, 10.3389/fphy.2023.1194804 | |
6. | Gerasimos Rigatos, Masoud Abbaszadeh, Pierluigi Siano, Mohammed Al-Numay, Farouk Zouari, A Nonlinear Optimal Control Approach for Bacterial Infections Under Antibiotics Resistance, 2024, 37, 1009-6124, 2293, 10.1007/s11424-024-3566-5 | |
7. | Zainab Dere, N.G. Cogan, Bhargav R. Karamched, Optimal control strategies for mitigating antibiotic resistance: Integrating virus dynamics for enhanced intervention design, 2025, 00255564, 109464, 10.1016/j.mbs.2025.109464 |
Parameter | Description | Linezolid | Penicillin G | Methicillin |
α | Elimination rate of S | 39.7 | 0.1204 | 32.4 |
q | Fraction of S that acquires resistance | 4×10−3 | 0.3992 | 0.1234 |
Parameter | Definition | Value |
βs | Growth rate of S | 0.4 |
βr | Growth rate of R | 0.1 |
μs | Natural death rate of S | 0.2 |
μr | Natural death rate of R | 0.5 |
ρ1 | Intensity of noise for βs | 0.5 |
ρ2 | Intensity of noise for βr | 0.5 |
Parameter | Description | Linezolid | Penicillin G | Methicillin |
α | Elimination rate of S | 39.7 | 0.1204 | 32.4 |
q | Fraction of S that acquires resistance | 4×10−3 | 0.3992 | 0.1234 |
Parameter | Definition | Value |
βs | Growth rate of S | 0.4 |
βr | Growth rate of R | 0.1 |
μs | Natural death rate of S | 0.2 |
μr | Natural death rate of R | 0.5 |
ρ1 | Intensity of noise for βs | 0.5 |
ρ2 | Intensity of noise for βr | 0.5 |