1,280
Views
7
CrossRef citations to date
0
Altmetric
Articles

Modeling the seasonality of Methicillin-resistant Staphylococcus aureus infections in hospitals with environmental contamination

, , &
Pages 99-122 | Received 30 Apr 2018, Accepted 05 Aug 2018, Published online: 22 Aug 2018

ABSTRACT

A deterministic mathematical model with periodic antibiotic prescribing rate is constructed to study the seasonality of Methicillin-resistant Staphylococcus aureus (MRSA) infections taking antibiotic exposure and environmental contamination into consideration. The basic reproduction number R0 for the periodic model is calculated under the assumption that there are only uncolonized patients with antibiotic exposure at admission. Sensitivity analysis of R0 with respect to some essential parameters is performed. It is shown that the infection would go to extinction if the basic reproduction number is less than unity and would persist if it is greater than unity. Numerical simulations indicate that environmental cleaning is the most important intervention to control the infection, which emphasizes the effect of environmental contamination in MRSA infections. It is also important to highlight the importance of effective antimicrobial stewardship programmes, increase active screening at admission and subsequent isolation of positive cases, and treat patients quickly and efficiently.

1. Introduction

Methicillin-resistant Staphylococcus aureus (MRSA), a type of staph bacteria first discovered in 1961, is one of the most common causes of hospital-acquired infections. As a considerable threat to global public health, MRSA causes many hard-to-treat infections such as serious skin infections, brain abscess (central nervous system infection), endophthalmitis, pneumonia (lung infection), and bloodstream infections. We usually treat staph bacteria with antibiotics, however, as antibiotics are abused to be prescribed to inhibit these kinds of bacteria infections, so far MRSA has been resistant to many common antibiotics such as methicillin, oxacillin, penicillin, and amoxicillin. Based on a Centers for Disease Control and Prevention (CDC) report [Citation5], 30–50% of antibiotics patients accepted in hospitals are unnecessary or inappropriate. Even though some antibiotics still work, MRSA is constantly adapting, which makes researchers difficult to keep developing new antibiotics. Hence whether a patient has antibiotic exposure or not is kind of important for his or her treatment. In fact, many studies observe that patients with antibiotic exposure tend to be more likely to be colonized by MRSA, which results in a lengthier duration in hospitals, a higher chance of failed treatment, a more expensive cost, a larger shedding rate of bacteria to environment, and even a higher mortality of death [Citation6,Citation10,Citation23,Citation24]. Hence it is necessary to consider antibiotic exposure and use of antibiotics in hospitals as influential factors in the transmission of MRSA.

Furthermore, in recent decades seasonal variation of MRSA infections in the hospital settings has been widely observed, especially in surgical wound, skin and soft tissue, urine, and respiratory tract in young children [Citation12,Citation14,Citation17,Citation18,Citation21,Citation22]. Reasons for this seasonal variation of MRSA infections in hospital are very complicated and still controversial. Previous studies believe that the seasonality involves temperature variation, insect bites, seasonal influenza, community-associated MRSA (CA-MRSA) infection, school season, seasonal community antibiotic use, which may result in a seasonal pattern of antibiotic prescriptions in hospitals. Especially, in the work of Sun et al. [Citation22], seasonality in the prescription data was found (see Figure ). Moreover, they performed a seasonal decomposition analysis for the MRSA isolates and found out that both fluoroquinolone prescriptions and the percentage of MRSA isolates that were resistant to ciprofloxacin peaked in the winter. Similar results were found for both the percentage of MRSA isolates resistant to clindamycin and macrolide/lincosamide prescriptions (see Figure ). Though this does not totally reflect the antibiotic usage in hospitals, [Citation21,Citation22] indicate that the usage of antibiotics in hospitals should also fluctuate seasonally. We believe that the seasonal pattern of antibiotic consumption implies the seasonal circulation of pathogens, which could induce seasonal risk of infections for hospitalized patients due to their weakened immune systems, and thus results in seasonal antibiotic consumption rates among hospitalized patients. Therefore, in our model we assume the antibiotic prescribing rate as a periodic function depending on time t, which has a period of 365 days and represents that antibiotic prescribing rate increases starting at the beginning of August, gains a peak in winter and then decreases starting at the beginning of February according to the data shown in Figure .

Figure 1. Number of prescriptions for antibiotic drug classes, by month. Source: IMS Health, Xponent, 1999–2007. Abbreviation: TMP/Sulfra, trimethoprim/sulfamethoxazole [Citation22].

Figure 1. Number of prescriptions for antibiotic drug classes, by month. Source: IMS Health, Xponent, 1999–2007. Abbreviation: TMP/Sulfra, trimethoprim/sulfamethoxazole [Citation22].

Figure 2. (a) Seasonal pattern of fluoroquinolone prescriptions and MRSA isolates resistant to ciprofloxacin; Mean monthly seasonal variation for fluoroquinolone prescriptions and MRSA isolates resistant to clindamycin for inpatient, outpatient and combined isolates as calculated by STL method. Prescription data source: IMS Health, Xponent, 1999–2007; Resistance data source: The Surveillance Network (TSN) Database-USA (Focus Diagnostics, Herndon, VA, USA) and (b) Seasonal pattern of macrolide and lincosamide prescriptions and MRSA isolates resistant to ciprofloxacin; Mean monthly seasonal variation for macrolide and lincosamide prescriptions and MRSA isolates resistant to clindamycin for inpatient, outpatient and combined isolates as calculated by STL method. Prescription data source: IMS Health, Xponent, 1999–2007; Resistance data source: The Surveillance Network (TSN) Database-USA (Focus Diagnostics, Herndon, VA, USA) [Citation22].

Figure 2. (a) Seasonal pattern of fluoroquinolone prescriptions and MRSA isolates resistant to ciprofloxacin; Mean monthly seasonal variation for fluoroquinolone prescriptions and MRSA isolates resistant to clindamycin for inpatient, outpatient and combined isolates as calculated by STL method. Prescription data source: IMS Health, Xponent, 1999–2007; Resistance data source: The Surveillance Network (TSN) Database-USA (Focus Diagnostics, Herndon, VA, USA) and (b) Seasonal pattern of macrolide and lincosamide prescriptions and MRSA isolates resistant to ciprofloxacin; Mean monthly seasonal variation for macrolide and lincosamide prescriptions and MRSA isolates resistant to clindamycin for inpatient, outpatient and combined isolates as calculated by STL method. Prescription data source: IMS Health, Xponent, 1999–2007; Resistance data source: The Surveillance Network (TSN) Database-USA (Focus Diagnostics, Herndon, VA, USA) [Citation22].

Mathematical modelling, as a powerful tool in quantifying the complex and numerous factors, has been widely developed to explore the transmission of MRSA [Citation1,Citation3,Citation6-9,Citation11,Citation25,Citation26,Citation28]. In their work, the direct patient-healthcare worker transmission is shown to be an essential factor in the transmission of MRSA in hospitals, as well as the indirect transmission via environmental contamination based on the observation that MRSA has the ability to be alive for days, weeks or even months on environmental surfaces in healthcare facilities, doors, and gowns. To the best of our knowledge, no model has been developed to address the seasonality in the transmission of MRSA. Studying seasonality of MRSA infections is helpful in developing efficient control programmes, lowering the long-term health risks, and distributing public resources.

We organize the paper as follows. We develop a periodic mathematical model to describe a comprehensive transmission of MRSA in Section 2. In Section 3, boundedness and positivity of solutions, the basic reproduction number, the extinction and uniform persistence of infections are analyzed. Simulations and discussion of the model behaviours and sensitive analysis of the basic reproduction number are given in the last two sections.

2. The periodic deterministic model

In order to describe the seasonal transmission of MRSA by a mathematical model, we first denote the patients, health-care workers (HCWS) and free-living bacteria in the environment as the following seven compartments [Citation16]:

  • Pu(t)=number of uncolonized patients without antibiotic exposure at time t.

  • PuA(t)=number of uncolonized patients with antibiotic exposure at time t.

  • Pc(t)=number of colonized patients without antibiotic exposure at time t.

  • PcA(t)=number of colonized patients with antibiotic exposure at time t.

  • Hu(t)number of uncontaminated health care workers at time t.

  • Hc(t)=number of contaminated health care workers at time t.

  • Be(t)=number of the free-living bacteria in the environment at time t.

The flowchart describing the transmission dynamics of MRSA in hospitals among these seven compartments is given in Figure . Our periodic deterministic model is derived based on the following assumptions and descriptions:
  1. We assume that a patient would have antibiotic exposure if he or she has received antibiotics within the month at admission or is currently receiving antibiotic treatment in the hospital.

  2. Based on the seasonal pattern of antibiotic usage found in Sun et al. [Citation22], we use a periodic function ϵ(t)=ϵ0(1+ϵ1sin((2π/365)(t240))) to describe the antibiotic prescription rate in the hospital. ϵ(t) has a period of 365 days, and represents that antibiotic prescription rate increases starting at the beginning of August, gains a peak in winter and then decreases starting at the beginning of February according to the data shown in Figures  and . ϵ0 is the baseline antibiotic prescription rate and ϵ1 is the magnitude of change.

  3. The free-living bacteria are uniformly distributed in the environment.

  4. The total number of patients in a unit is a constant Np. That is equivalent to say that patients are admitted at a total rate Ω(t)=γuPu+γcPc+γuAPuA+γcAPcA, where γu,γuA,γc, and γcA are the corresponding discharge rates of patients from these four compartments. We also denote θu,θuA,θc,θcA as the corresponding proportion of patients Pu,PuA,Pc,PcA at admission. It was estimated that the fraction of patients with antibiotic exposure of new admission to be 0.38, i.e. θuA+θcA=0.38 [Citation6,Citation15].

  5. The total number of health-care workers is a constant Nh.

  6. We assume that the bacterial reproduction cannot occur due to lack of proper condition in the hospital, even though the free-living bacteria are able to survive in the environment for a long time. As a result, shedding from colonized patients is one of the key transmission of bacteria to contaminate environment υpPc+υpAPcA. υp, and υpA are the shedding rate of bacteria from patients without or with antibiotic exposure, respectively. In addition, when the contaminated HCWs touch the environmental surfaces such as door handles, health facilities, bedding, they leave bacteria there υhHc, which is another way to contaminate the environment. Of course, hospitals always have a standard cleaning rate or disinfection rate γb.

  7. We assume that there is no contact between patients, which means that if an uncolonized patient without antibiotic exposure becomes colonized without antibiotic exposure, he/she either contacts contaminated HCWs at rate αpβp(1η)PuHc or touches the contaminated environment at rate κpPuBe. A similar process happens when an uncolonized patient with antibiotic exposure becomes colonized with antibiotic exposure at rate αpβpA(1η)PuAHc+κpAPuABe. αp is the contact rate per day, βp and βpA are the chance of colonization per contact for uncolonized patients without or with antibiotic exposure, respectively, η is the hand hygiene compliance, κp and κpA are the chance of colonization by touching contaminated environment for uncolonized patients without or with antibiotic exposure, respectively. Besides, when an uncolonized patient without antibiotic exposure is accepting an antibiotic treatment for other diseases, which is the reason for his/her hospitalization, he/she then becomes uncolonized patients with antibiotic exposure ϵ(t)Pu. That is similar to how a colonized patient without antibiotic exposure becomes colonized with antibiotic exposure ϵ(t)Pc. Note that an uncolonized patient without antibiotic exposure cannot move to the colonized with antibiotic exposure in one step.

  8. An uncontaminated HCW becomes contaminated when he/she contacts colonized patients or touches contaminated environmental surfaces at rate αpβh(1η)PcHu+αpβhA(1η)PcAHu, where βh,βhA are chance of contamination per contact with Pc or PcA, respectively. HCWs have a decontaminated rate μc to move from contaminated state to uncontaminated state.

  9. Antibiotics can not only kill the bad bacteria that make patients sick, but also commensal bacteria, this may disturb the balance of patients' commensal microbiota. Then patients with prior antibiotic exposure are more likely to have a higher colonization rate of MRSA, so we assume βpAβp. In addition, by previous studies [Citation6,Citation13], we estimate that uncolonized patients with antibiotic exposure are 1.67 times more vulnerable than uncolonized patients without antibiotic exposure, i.e. βpA=1.67×βp. Moreover, patients with recent antibiotic exposure are more likely to experience adverse effects or new health problems (such as gastrointestinal symptoms, respiratory infections, skin rashes, and so on); and a higher probability of treatment failure because of the acquisition of resistance elements, which lead to lengthier hospital stays. So we assume that γcA1γc1γuA1γu1. We also follow the assumptions and parametrizions from [Citation6,Citation13,Citation25] for the values of βh, βhA, υp, υpA, υh, κp, κpA, κh as shown in Table .

Figure 3. Transmission flowchart of MRSA among patients, health-care works and environment.

Figure 3. Transmission flowchart of MRSA among patients, health-care works and environment.

Table 1. Parameters and descriptions.

Detailed parameter values are given in Table . We hence formulate the periodic mathematical model as follows: (1) dPudt=θuΩ(t)αpβp(1η)PuHcκpPuBeγuPuϵ(t)Pu,dPcdt=θcΩ(t)+αpβp(1η)PuHc+κpPuBeγcPcϵ(t)Pc,dPuAdt=θuAΩ(t)αpβpA(1η)PuAHcκpAPuABeγuAPuA+ϵ(t)Pu,dPcAdt=θcAΩ(t)+αpβpA(1η)PuAHc+κpAPuABeγcAPcA+ϵ(t)Pc,dHudt=αpβh(1η)PcHuαpβhA(1η)PcAHuκhHuBe+μcHc,dHcdt=αpβh(1η)PcHu+αpβhA(1η)PcAHu+κhHuBeμcHc,dBedt=υpPc+υpAPcA+υhHcγbBe(1) with initial conditions Pu(0)=Pu0, PuA(0)=PuA0, Pc(0)=Pc0, PcA(0)=PcA0, Hu(0)=Hu0, Hc(0)=Hc0, Be(0)=Be0, where Ω(t)=(γuPu+γcPc+γuAPuA+γcAPcA) and ϵ(t)=ϵ0(1+ϵ1sin((2π/365)(t240))).

3. Mathematical analysis

3.1. Basic reproduction number

The basic reproduction number R0 for the periodic deterministic model (Equation1) is constructed according to the definition in Bacae¨r and Guenaoui [Citation2] and follow the general calculation procedure in Wang and Zhao [Citation27]. When θu=0, θc=0, and θcA=0, that is only uncolonized patients with antibiotic exposure are admitted into hospital, the infection-free infection (IFE) is defined as E0=(Pu,Pc,PuA,PcA,Hu,Hc,Be)=(0,0,Np,0,Nh,0,0), We can rewrite the variables of periodic ODE system (Equation1) as a vector E0=(Pc,PcA,Hc,Be,Pu,PuA,Hu)=(0,0,0,0,0,Np,Nh). Following the general calculation procedure in Wang and Zhao [Citation27], we have F=αpβp(1η)PuHc+κpPuBeαpβpA(1η)PuAHc+κpAPuABeαpβh(1η)PcHu+αpβhA(1η)PcAHu+κhHuBe0000,V=γcPc+ϵ(t)PcθcΩγcAPcA[ϵ(t)Pc+θcAΩ]μcHcγbBe(υpPc+υpAPcA+υhHc)αpβp(1η)PuHc+κpPuBe+γuPu+ϵ(t)PuθuΩαpβpA(1η)PuAHc+κpAPuABe+γuAPuA[ϵ(t)Pu+θuAΩ]αpβh(1η)PcHu+αpβhA(1η)PcAHu+κhHuBeμcHc, where ϵ(t)=ϵ0(1+ϵ1sin((2π/365)(t240))) and Ω(t)=(γuPu+γcPc+γuAPuA+γcAPcA). We also have V=γcPc+ϵ(t)PcγcAPcAμcHcγbBeαpβp(1η)PuHc+κpPuBe+γuPu+ϵ(t)PuαpβpA(1η)PuAHc+κpAPuABe+γuAPuAαpβh(1η)PcHu+αpβhA(1η)PcAHu+κhHuBe,V+=θcΩϵ(t)Pc+θcAΩ0υpPc+υpAPcA+υhHcθuΩϵ(t)Pu+θuAΩμcHc. So we derive that F(t)=000000αpβpA(1η)NpκpANpαpβh(1η)NhαpβhA(1η)Nh0κhNh0000,V(t)=γc+ϵ(t)000ϵ(t)γcA0000μc0υpυpAυhγb, and M(t)=γuϵ(t)00γu+ϵ(t)00000. Let Y(t,s), ts be the evolution operator of the system (2) dydt=V(t)y.(2) That is, for each sR, the 4×4 matrix Y(t,s) satisfies ddtY(t,s)=V(t)Y(t,s),ts,Y(s,s)=I, where I is the 4×4 identity matrix. In order to characterize R0, we consider the following linear ω-periodic system (3) dwdt=V(t)+F(t)λω,tR+(3) with parameter λ(0,). Let W(t,s,λ), ts, be the evolution operator of the system (Equation3) on R4. Clearly,ΦFV=W(t,0,1),t0.

According to the method in Wang and Zhao [Citation27], we let φ be ω-periodic in s and the initial distribution of infectious individuals. So F(s)φ(s) is the rate of new infections produced by the infected individuals who were introduced at time s. When ts, Y(t,s)F(s)φ(s) gives the distribution of those infected individuals who were newly infected by φ(s) and remain in the infected compartments at time t. Naturally, tY(t,s)F(s)φ(s)ds=0Y(t,ta)F(ta)φ(ta)da is the distribution of accumulative new infections at time t produced by all those infected individuals φ(s) introduced at time previous to t.

Let Cω be the ordered Banach space of all ω-periodic functions from R to R4, which is equipped with the maximum norm and the positive cone Cω+:={φCω:φ(t)0,tR+}. Then we can define a linear operator L:CωCω by (Lφ)(t)0Y(t,ta)F(t,ta)φ(ta)da,tR+, φCω, L is called the next infection operator and the spectral radius of L is defined as the basic reproduction number R0:=ρ(L) for the periodic epidemic model. In order to determine the threshold dynamics, we use Theorems 2.1 and 2.2 in Wang and Zhao [Citation27]. First of all, we need to verify the seven assuptions in the theorems.

(A1)–(A5) The first five conditions can be easily verified by observing F, V+ andV.

(A6) ρ(ΦM(ω))<1, where ρ(ΦM(ω)) is the spectral radius of ΦM(ω). ΦM(t) is the monodromy matrix of the linear ω-periodic system dq/dt=M(t)q with M=γuϵ(t)00γu+ϵ(t)00000. Hence, we have, ΦM(t)=00eγu+ϵ(t)dt1212eγu+ϵ(t)dt0120. It is obvious that ρ(ΦM(t))<1, since eγu+ϵ(t)dt<1 based on γu+ϵ(t)>0 in our parameter setting.

(A7) ρ(ΦV(ω))<1, where ΦV(t) is the monodromy matrix of the linear ω-periodic sysstem dy/dt=V(t)y with V=γcϵ(t)000ϵ(t)γcA0000μc0υpυpAυhγb. Hence, we have, ΦV(t)=eγu+ϵ(t)dt000c1eγcAt00c20eμct0c3υpAγbγcAeγcAtυhγbμceμcteγbt, where c1,c2 and c3 are no need to be calculated, even though they can be calculated. Since it is a lower triangular matrix with all elements in diagonal are less than one,ρ(ΦV(ω))<1.

Hence, all assumptions (A1)–(A7) hold, So by (ii) in Theorems 2.1 and 2.2 in Wang and Zhao [Citation27], we have the following results.

Lemma 3.1

R0=λ is the unique solution of ρ(W(ω,0,λ))=1, where W(t,s,λ),ts, is the evolution operator of system (Equation3).

Theorem 3.2

If R0<1, then the infection-free equilibrium E0 is locally asymptotically stable; If R0>1, then E0 is unstable.

Lemma 3.3

For the basic reproduction number R0, we have

  1. R0=1 if and only of ρ(ΦFV(ω))=1.

  2. R0>1 if and only of ρ(ΦFV(ω))>1.

  3. R0<1 if and only of ρ(ΦFV(ω))<1.

Remark 3.4

If ρ(ΦFV(ω))<1, then the disease-free equilibrium E0 is locally asymptotically stable; If ρ(ΦFV(ω))>1, then E0 is unstable.

In order to characteristic R0, we consider F(t)λV(t)=(γc+ϵ(t))000ϵ(t)γcAαpβpA(1η)NpλκpANpλαpβh(1η)NhλαpβhA(1η)NhλμcκhNhλυpυpAυhγb, where ϵ(t)=ϵ0(1+ϵ1sin((2π/365)(t240))). We want to calculate the monodromy matrix of the system (4) dxdt=F(t)λV(t)x.(4) By observing the matrix F(t)/λV(t), we can see that x1(t) can be solved directly.

When x1(t)=0, we have x˙2(t)x˙3(t)x˙4(t)=Ax2(t)x3(t)x4(t), where A=γcAαpβpA(1η)NpλκpANpλαpβhA(1η)NhλμcκhNhλυpAυhγb is a constant matrix.

When x1(t)=eγu+ϵ(t), we have x˙2(t)x˙3(t)x˙4(t)=Ax2(t)x3(t)x4(t)+f(t), where f(t)=ϵ(t)eγu+ϵ(t)αpβh(1η)Nhλeγu+ϵ(t)υpeγu+ϵ(t). According to the results in Chapter 1 of Perko [Citation19], we are able to find the monodromy matrix of the system (Equation4). However, the high-dimension of the matrix F/λV makes the analytical solution for R0 complicated. Hence, we derive R0 numerically in next section.

3.2. Extinction of infection

Based on the biological background of the model (Equation1), we consider solutions of model (Equation1) with nonnegative initial values: Pu00,PuA00,Pc00,PcA00,Hu00,Hc00,Be00.

Lemma 3.5

If Pu0,PuA0,Pc0,PcA0,Hu0,Hc0,Be00, i.e. the initial values are nonnegative, then the solution of model (Equation1) is nonnegative for all t0 and ultimately bounded. In particular, if Pu0,PuA0,Pc0,PcA0,Hu0,Hc0,Be0>0, i.e. the initial values are positive, then the solutions of model (Equation1) is also positive for all t0.

Proof.

According to the continuous dependence of solutions with respect to initial values, we only need to prove that when the initial values are positive, i.e. Pu0,PuA0,Pc0,PcA0,Hu0,Hc0,Be0>0, the solution of model (Equation1) is also positive for all t0. Let m(t)=min{Pu(t),PuA(t),Pc(t),PcA(t),Hu(t),Hc(t),Be(t)},t>0. By the assumption that the initial values are positive, we clearly have, m(0)>0. So we assume that there exists a t1>0 such that m(t1)=0 and m(t)>0 for all t[ 0,t1).

If m(t1)=Pu(t1), from the first equation of model (Equation1), it follows that dPu/dt(αpβp(1η)Hc(t)+κpBe(t)+γu+ϵ(t))Pu for all t[ 0,t1). Since Hc(t),Be(t)>0,ϵ(t)=ϵ0(1+ϵ1sin((2π/365)(t240)))>0 for all t[ 0,t1), we have 0=Pu(t1)Pu0exp0t1(αpβp(1η)Hc(s)+κpBe(s)+γu+ϵ(s))ds>0, which leads to a contradiction. We can get similar contradictions in the other cases. Hence, the solutions remain in the positive cone if the initial conditions are in the positive cone R7.

Next, denote M(t)=Pu(t)+PuA(t)+Pc(t)+PcA(t)+Hu(t)+Hc(t)+Be(t). Then dM(t)dt=dBe(t)dt=υpPc+υpAPcA+υhHcγbBeυpNp+υpANp+υhNhγbBe(t), where Np=Pu(t)+PuA(t)+Pc(t)+PcA(t) and Nh=Hu(t)+Hc(t), which implies that Be(t)(υpNp+υpANp+υhNh)γb(1eγbt)+Be0eγbt. So Be(t) is bounded by a fixed number K=(υpNp+υpANp+υhNh)γb+Be0. Let N=Np+Nh+K, we have Pu(t)+PuA(t)+Pc(t)+PcA(t)+Hu(t)+Hc(t)+Be(t)N. Thus, the solution is ultimately bounded. This completes the proof.

Remark 3.6

Denote G:={(Pu,PuA,Pc,PcA,Hu,Hc,Be)R+7:Pu+PuA+Pc+PcA+Hu+Hc+BeN)}, Lemma 3.5 implies that G is positively invariant set with respect to solutions of model (Equation1).

Theorem 3.7

If R0<1, then the infection-free equilibrium E0=(0,0,Np,0,Nh,0,0) is globally asymptotically stable.

Proof.

According to Theorem 3.2, E0 is locally asymptotically stable when R0<1. According to Lemma 3.3, we know that R0<1 is equivalent to ρ(ΦFV(ω))<1, where FV is the defined as F(t)V(t)=γcϵ(t)000ϵ(t)γcAαpβpA(1η)NpκpANpαpβh(1η)NhαpβhA(1η)NhμcκhNhυpυpAυhγb. By the continuity, we can always find a small enough positive constant δ such that ρ(ΦFV+δN(ω))<1, where N(t)=00αpβp(1η)κp00αpβpA(1η)κpA00000000. Now we try to prove the global attractivity of the disease-free equilibrium E0. By the non-negativity of solutions and the assumption that θu=θc=θcA=0,θuA=1, we have the following result from the first equation of the model (Equation1): dPudtγuPuϵ(t)Pu. Note that ϵ(t)=ϵ0(1+ϵ1sin((2π/365)(t240)))>0,t. That is, δ>0, there exists t1>0, such that Pu(t)δ,tt1. Similarly, by the third equation of the model (Equation1) and the fact that γu=max{γu,γc,γcA}, we get dPuAdtγu(NpPuA)+γuAPuAγuAPuA+ϵ(t)(NpPuA), that is, dPuAdt(γu+ϵ(t))Np(γu+ϵ(t))PuA. Then δ>0, there exists t2>0, such that PuA(t)Np+δ,tt2. Let T=max{t1,t2}, If t>T, since θu=θc=θcA=0,θuA=1, then (5) Pc(t)αpβp(1η)δHc+κpδBeγcPcϵ(t)Pc,PcA(t)αpβpA(1η)(Np+δ)Hc+κpA(Np+δ)BeγcAPcA+ϵ(t)Pc,Hc(t)αpβh(1η)NhPc+αpβhA(1η)NhPcA+κhNhBeμcHc,Be(t)υpPc+υpAPcA+υhHcγbBe.(5) Considering the following auxiliary system: (6) P~c(t)=αpβp(1η)δH~c+κpδB~eγcP~cϵ(t)P~c,P~cA(t)=αpβpA(1η)(Np+δ)H~c+κpA(Np+δ)B~eγcAP~cA+ϵ(t)P~c,H~c(t)=αpβh(1η)NhP~c+αpβhA(1η)NhP~cA+κhNhB~eμcH~c,B~e(t)=υpP~c+υpAP~cA+υhH~cγbB~e,(6) which can be written as, (7) dx(t)dt=(F(t)V(t)+δN(t))x(t),x(t)=(P~c(t),P~cA(t),H~c(t),B~e(t))T.(7) Hence, there exists a positive ω-periodic function f(t)=(f1(t),f2(t),f3(t),f4(t))T such that x(t)=eμtf(t) is a solution of system (Equation7) where μ=1ωlnρ(ΦFV+δN(ω)), according to the Lemma 2.1 in Zhang and Zhao [Citation29]. Note that ρ(ΦFV+δN(ω))<1, which implies that lnρ(ΦFV+δN(ω))<0, that is to say, μ<0. Then limtx(t)=0. Let S(t)=(Pc(t),PcA(t),Hc(t),Be(t))T, by comparison principle, we have limtS(t)=0, which is equivalent to say that limtPc=0,limtPcA=0,limtHc=0,limtBe=0. Therefore, E0 is globally attractive when R0<1. This completes the proof.

3.3. Persistence of infection

Finally, we prove that the model is uniformly persistent, which implies the the persistence of MRSA infections.

Theorem 3.8

If R0>1, then model (Equation1) is uniformly persistent.

Proof.

We follow the persistence theory of nonautonomous models given in Zhao [Citation30] to discuss the uniform persistence of model (Equation1). We first define X={(Pu,Pc,PuA,PcA,Hu,Hc,Be):Pu0,Pc0,PuA0,PcA0,Hu0,Hc0,Be0}, X0={(Pu,Pc,PuA,PcA,Hu,Hc,Be)X:Pc>0,PcA>0,Hc>0,Be>0}, X0=XX0. Note that both X and X0 are positively invariant with respect to system (Equation1), and X0 is relatively closed in X. Since our model (Equation1) is ω-periodic (ω=365 days), the Poincaré map associated with our model (Equation1) P:XX is defined by P(x0)=φ(ω,x0),x0X, where x0=(Pu(0),Pc(0),PuA(0),PcA(0),Hu(0),Hc(0),Be(0)) and φ(t,x0) is the unique solution of model (Equation1) with initial values φ(0,x0)=x0. Note that a continuous mapping f:XX is said to be compact if f maps any bounded set to a precompact set in X [Citation30]. According to Lemma 3.5, the Poincaré map P is compact and point dissipative on X, which implies that there exists a global attractor by Theorem 1.1.3 in [Citation30].

Define M={x0X0:Pn(x0)X0, n=1,2,}, where x0=(Pu(0),Pc(0),PuA(0),PcA(0),Hu(0),Hc(0),Be(0)). We want to verify that M={(Pu,0,PuA,0,Hu,0,0):Pu0,PuA0,Hu=Nh}. We first verify that M{(Pu,0,PuA,0,Hu,0,0):Pu0,PuA0,Hu=Nh}, which is equivalent to verify that if x0{(Pu,0,PuA,0,Hu,0,0):Pu0,PuA0,Hu=Np}, then x0M. For any point x0=(Pu(0),Pc(0),PuA(0),PcA(0),Hu(0),Hc(0),Be(0)), we suppose that x0{(Pu,0,PuA,0,Hu,0,0):Pu0,PuA0,Hu=Np}, that is to say one of Pc(0),PcA(0),Hc(0),Be(0) is not zero. Without loss of generality, we suppose that Pc(0)>0,PcA(0)=0,Hc(0)=0,Be(0)=0. By the fourth, sixth and seventh equations of model (Equation1), note that ϵ(t)>0t, we have dPcA(0)dtϵ(0)Pc(0)>0;rdHc(0)dtαpβhNpPc(0)>0;dBe(0)dtκpPc(0)>0. Thus, there exists δ0>0, if 0<t<δ0, then Pc(t)>0,PcA(t)>0,Hc(t)>0,Be(t)>0, which implies that x0X0. Other cases (Pc(0)>0, or PcA(0)>0, or Hc(0)>0) can be proved in the similar way. Thus x0M. That is to say, for any x0{(Pu,0,PuA,0,Hu,0,0):Pu0,PuA0,Hu=Nh}, x0M. So M{(Pu,0,PuA,0,Hu,0,0):Pu0,PuA0,Hu=Nh}.

Obviously we have {(Pu,0,PuA,0,Hu,0,0):Pu0,PuA0,Hu=Nh}M, since if x0=(Pu(0),0,PuA(0),0,Np,0,0), then the solutions (Pu(t),Pc(t),PuA(t),PcA(t),Hu(t),Hc(t),Be(t))(Pu(t),0,PuA(t),0,Nh,0,0) where Pu(t)>0,PuA>0. Therefore M={(Pu,0,PuA,0,Hu,0,0):Pu0,PuA0,Hu=Nh}. There is only one equilibrium E0=(0,0,Np,0,Nh,0,0) in M, so x0Mω(x0)=E0. Therefore E0 is a compact and isolated invariant sets inX0.

Let x0=(Pu(0),Pc(0),PuA(0),PcA(0),Hu(0),Hc(0),Be(0))X0 be any initial value. Next we claim that there exist a positive constant δ such that (8) lim supnPn(x0)E0δ.(8) Suppose that claim (Equation8) is not true, i.e. for any δ>0, lim supnPn(x0)E0δ for some x0X0. That is to say, there exists a big enough n1>0, for all n>n1, Pn(x0)E0δ Followed by the continuity of solution φ(t,x0) with respect to the initial values, we know >0, there exists a δ>0 such that if x0E0δ, then φ(t,x0)φ(t,E0)<, t[0,ω]. Hence we obtain φ(t,Pn(x0))φ(t,E0)< for all n>n1 andt[0,ω].

Now for any big enough t0, we can rewrite t=nω+tˆ, where n=[t/ω] is the greatest integer less than or equal to t/ω and tˆ[0,ω]. We can always choose t big enough to make sure that n>n1. Hence for big enough t , we have φ(t,x0)φ(t,E0)=φ(tˆ,Pn(x0))φ(tˆ,E0)<. It follows that 0Pu(t),c(t), NpPuA(t)Np+, PcA(t), NhHu(t)Nh+, Hc(t), Be(t), for any t big enough. Thus for t big enough, we have (9) Pc(t)γcPcϵ(t)Pc,PcA(t)αpβpA(1η)(Np)Hc+κpA(Np)BeγcAPcA+ϵ(t)Pc,Hc(t)αpβh(1η)(Nh)Pc+αpβhA(1η)(Nh)PcA+κh(Nh)BeμcHc,Be(t)υpPc+υpAPcA+υhHcγbBe.(9) Consider the following auxiliary system: (10) P~c(t)=γcPc~ϵ(t)Pc~,P~cA(t)=αpβpA(1η)(Np)Hc~+κpA(Np)Be~γcAPcA~+ϵ(t)Pc~,H~c(t)=αpβh(1η)(Nh)Pc~+αpβhA(1η)(Nh)PcA~+κh(Nh)Be~μcHc~,B~e(t)=υpPc~+υpAPcA~+υhHc~γbBe~,(10) which can be written as, (11) dx(t)dt=(F(t)V(t)N(t))x(t),  x(t)=(P~c(t),P~cA(t),H~c(t),B~e(t))T,(11) where F(t)V(t)=γcϵ(t)000ϵ(t)γcAαpβpA(1η)NpκpANpαpβh(1η)NhαpβhA(1η)NhμcκhNhυpυpAυhγb,N(t)=000000αpβpA(1η)κpAαpβh(1η)αpβhA(1η)0κh0000. Hence there exists a positive ω-periodic function g(t)=(g1(t),g2(t),g3(t),g4(t))T such that x(t)=eμtg(t) is a solution of system (Equation11) where μ=1ωlnρ(ΦFVN(ω)), according to the Lemma 2.1 in Zhang and Zhao [Citation29]. Note that ρ(ΦFVN(ω))>1, which implies that lnρ(ΦFVN(ω))>0, that is to say, μ>0. Then limtx(t)=. Let J(t)=(Pc(t),PcA(t),Hc(t),Be(t))T, by comparison principle, we have limtJ(t)=, which is equivalent to say that limtPc=,limtPcA=,limtHc=,limtBe=. The claim implies that E0 is an isolated invariant set in X and Ws(E0)X0=. Therefore the Poincar map P is uniformly persistent with respect to (X0,X0) if R0>1 by Theorems 1.3.1 and 3.1.1 in [Citation30]. This completes the proof.

4. Numerical simulations

The deterministic model with periodic transmission rate is simulated for 1000 days with initial values (Pu0,PuA0,Pc0,PcA0,Hu0,Hc0,Be0)=(4,6,7,6,17,6,1000) and detailed parameter values in Table . The simulated solutions of the model are periodic as shown in Figure . Based on the seasonal pattern of antibiotic usage observed in Sun et al. [Citation22], we assume that antibiotic prescription rate in hospital increases starting at the beginning of August, gains a peak in winter and then decreases starting at the beginning of February according to the data shown in Figures  and , which results in the similar pattern of colonized patients with antibiotic exposure in Figures  and (a), but with a lag about 15-days. We suggest that there may be a temporal correlation between antibiotic use and resistance. Figures  and  tell us that the prevalence of colonized patients with antibiotic exposure has periodic phenomenon between about 34% and 39% and the prevalence of colonized patients without antibiotic exposure is between 4% and 6%. While when there is no admission of colonized patients, i.e. θc=θcA=0, Figure  implies that the prevalence of colonized patients with antibiotic exposure reduces to between 20% and 23% and the prevalence of colonized patients without antibiotic exposure is between 3% and 5%. This means that detection and isolation of MRSA colonized patients on admission may be a useful intervention to control the hospital infection. While when only uncolonized patients without antibiotic exposure are admitted to hospital, Figure  indicates that the prevalence of colonized patients with antibiotic exposure is between 12% and 15% and the prevalence of colonized patients without antibiotic exposure is between 3% and 4%. We suggest that in order to control the infection in hospital, it is important to increase the public education about how to use antibiotics properly in community.

Figure 4. Solutions of uncolonized patients without or with antibiotic exposure (Pu(t),PuA) and colonized patients without or with antibiotic exposure (Pc(t),PcA) of the model (Equation1) with initial values (Pu0,PuA0,Pc0,PcA0,Hu0,Hc0,Be0)=(4,6,7,6,17,6,1000). Parameters are given in Table .

Figure 4. Solutions of uncolonized patients without or with antibiotic exposure (Pu(t),PuA) and colonized patients without or with antibiotic exposure (Pc(t),PcA) of the model (Equation1(1) dPudt=θuΩ(t)−αpβp(1−η)PuHc−κpPuBe−γuPu−ϵ(t)Pu,dPcdt=θcΩ(t)+αpβp(1−η)PuHc+κpPuBe−γcPc−ϵ(t)Pc,dPuAdt=θuAΩ(t)−αpβpA(1−η)PuAHc−κpAPuABe−γuAPuA+ϵ(t)Pu,dPcAdt=θcAΩ(t)+αpβpA(1−η)PuAHc+κpAPuABe−γcAPcA+ϵ(t)Pc,dHudt=−αpβh(1−η)PcHu−αpβhA(1−η)PcAHu−κhHuBe+μcHc,dHcdt=αpβh(1−η)PcHu+αpβhA(1−η)PcAHu+κhHuBe−μcHc,dBedt=υpPc+υpAPcA+υhHc−γbBe(1) ) with initial values (Pu0,PuA0,Pc0,PcA0,Hu0,Hc0,Be0)=(4,6,7,6,17,6,1000). Parameters are given in Table 1.

Figure 5. (a) Prevalence of colonized patients with or without antibiotic exposure of model (Equation1) with initial values (Pu0,PuA0,Pc0,PcA0,Hu0,Hc0,Be0)=(4,6,7,6,17,6,1000). Parameters are given in Table . Compared with antibiotic prescribing rate and (b) The free-living bacterial load in the environment.

Figure 5. (a) Prevalence of colonized patients with or without antibiotic exposure of model (Equation1(1) dPudt=θuΩ(t)−αpβp(1−η)PuHc−κpPuBe−γuPu−ϵ(t)Pu,dPcdt=θcΩ(t)+αpβp(1−η)PuHc+κpPuBe−γcPc−ϵ(t)Pc,dPuAdt=θuAΩ(t)−αpβpA(1−η)PuAHc−κpAPuABe−γuAPuA+ϵ(t)Pu,dPcAdt=θcAΩ(t)+αpβpA(1−η)PuAHc+κpAPuABe−γcAPcA+ϵ(t)Pc,dHudt=−αpβh(1−η)PcHu−αpβhA(1−η)PcAHu−κhHuBe+μcHc,dHcdt=αpβh(1−η)PcHu+αpβhA(1−η)PcAHu+κhHuBe−μcHc,dBedt=υpPc+υpAPcA+υhHc−γbBe(1) ) with initial values (Pu0,PuA0,Pc0,PcA0,Hu0,Hc0,Be0)=(4,6,7,6,17,6,1000). Parameters are given in Table 1. Compared with antibiotic prescribing rate and (b) The free-living bacterial load in the environment.

Figure 6. (a) Prevalence of colonized patients with or without antibiotic exposure of the model (Equation1) with initial values (Pu0,PuA0,Pc0,PcA0,Hu0,Hc0,Be0)=(4,6,7,6,17,6,1000), θu=0.62,θuA=0.38,θc=0,θcA=0 and other parameter values given in Table ; (b) The free-living bacterial load in the environment.

Figure 6. (a) Prevalence of colonized patients with or without antibiotic exposure of the model (Equation1(1) dPudt=θuΩ(t)−αpβp(1−η)PuHc−κpPuBe−γuPu−ϵ(t)Pu,dPcdt=θcΩ(t)+αpβp(1−η)PuHc+κpPuBe−γcPc−ϵ(t)Pc,dPuAdt=θuAΩ(t)−αpβpA(1−η)PuAHc−κpAPuABe−γuAPuA+ϵ(t)Pu,dPcAdt=θcAΩ(t)+αpβpA(1−η)PuAHc+κpAPuABe−γcAPcA+ϵ(t)Pc,dHudt=−αpβh(1−η)PcHu−αpβhA(1−η)PcAHu−κhHuBe+μcHc,dHcdt=αpβh(1−η)PcHu+αpβhA(1−η)PcAHu+κhHuBe−μcHc,dBedt=υpPc+υpAPcA+υhHc−γbBe(1) ) with initial values (Pu0,PuA0,Pc0,PcA0,Hu0,Hc0,Be0)=(4,6,7,6,17,6,1000), θu=0.62,θuA=0.38,θc=0,θcA=0 and other parameter values given in Table 1; (b) The free-living bacterial load in the environment.

Figure 7. (a) Prevalence of colonozied patients with or without antibiotic exposure of modified model (Equation1) with initial values (Pu0,PuA0,Pc0,PcA0,Hu0,Hc0,Be0)=(4,6,7,6,17,6,1000), θu=1,θuA=0,θc=0,θcA=0 and other parameter values given in Table  and (b) The free-living bacterial load in the environment.

Figure 7. (a) Prevalence of colonozied patients with or without antibiotic exposure of modified model (Equation1(1) dPudt=θuΩ(t)−αpβp(1−η)PuHc−κpPuBe−γuPu−ϵ(t)Pu,dPcdt=θcΩ(t)+αpβp(1−η)PuHc+κpPuBe−γcPc−ϵ(t)Pc,dPuAdt=θuAΩ(t)−αpβpA(1−η)PuAHc−κpAPuABe−γuAPuA+ϵ(t)Pu,dPcAdt=θcAΩ(t)+αpβpA(1−η)PuAHc+κpAPuABe−γcAPcA+ϵ(t)Pc,dHudt=−αpβh(1−η)PcHu−αpβhA(1−η)PcAHu−κhHuBe+μcHc,dHcdt=αpβh(1−η)PcHu+αpβhA(1−η)PcAHu+κhHuBe−μcHc,dBedt=υpPc+υpAPcA+υhHc−γbBe(1) ) with initial values (Pu0,PuA0,Pc0,PcA0,Hu0,Hc0,Be0)=(4,6,7,6,17,6,1000), θu=1,θuA=0,θc=0,θcA=0 and other parameter values given in Table 1 and (b) The free-living bacterial load in the environment.

Based on the calculation procedure about the basic reproduction number discussed above, we calculate the basic reproduction number R0 to be 1.476 with the parameter values in Table . By Theorem 3.8, we conclude that the infection will persist with the baseline parameter values. In Figure , we perform some sensitivity analysis to explore the effect of the following parameters on changing the basic reproduction number R0: (a) The cleaning/disinfection rate of environment γb; (b) Shedding rate of bacteria from colonized patients with antibiotic exposure to environment υpA; (c) The discharge rate of colonized patients with antibiotic exposure γcA; (d) The hand hygiene compliance with HCWs η; (e) The contact rate between patients and HCWs αp; (f) The decontaminated rate of HCWs μc. Figure (a) shows that increasing the environmental cleaning/disinfection rate γb from 0.6 to 1 can reduce the basic reproduction number from 1.705 to 1.065, which is the most efficient intervention. Since we assume that the free-living bacteria have no proper condition to reproduce themselves, shedding bacteria from colonized patients is a crucial factor in environmental contamination, which is verified in Figure (b) where if the shedding rate of colonized patients with antibiotic exposure υpA is below 300, the basic reproduction number can be below 1. This again emphasizes the importance of environmental cleaning. Figure (c) indicates that the discharge rate (the inverse of stay in hospital) of colonized patients with antibiotic exposure γcA greatly increase the basic reproduction number especially when they have a lengthier stay than 18 days (baseline value i.e. 0.0551 ). However, it is hard to treat colonized patients with antibiotic exposure efficiently and quickly since they have resistance to many common antibiotics, which usually leads to a lengthier stay to make the situation worse. Hence how to make an efficient and right treatment plan for colonized patients with antibiotic exposure is a challenge and also a key to control the infection. In Figure (d), it seems that the hand hygiene compliance of HCWs (from the baseline value 0.4 to 1) make little difference to change the basic reproduction number, which is a little surprising, since the hand hygiene is always thought to be an important intervention. We think that this is because the direct transmission through HCWs is well-known so hospitals have paid enough attention to the hand hygiene of HCWs while the indirect transmission through contaminated environment lacks our surveillance and is more important than we thought. That is why the environmental cleaning γb and the shedding rate γcA affect greatly the basic reproduction number in our sensitivity analysis Figure (a,b). Hence, we believe that it is necessary to strengthen the surveillance of environmental cleaning with feedback to cleaning team, and try to use more efficient cleaning products. Figure (e,f) imply how the contact rate αp and decontaminated rate of HCWs μc affect the basic reproduction number.

Figure 8. Effects of parameters on the basic reproduction number R0: (a) γb, (b) υpA, (c) γcA, (d) η, (e) αp and (f) μc. Other parameters values are given in Tabel .

Figure 8. Effects of parameters on the basic reproduction number R0: (a) γb, (b) υpA, (c) γcA, (d) η, (e) αp and (f) μc. Other parameters values are given in Tabel 1.

5. Discussion

We presented a comprehensive mathematical model with periodic transmission rate to study MRSA infections in hospitals, including key factors such as environmental contamination and antibiotic exposure. Both the direct transmission via HCWs and the indirect transmission via free-living bacteria in the environment were taken into account. Inspired by the work of Sun et al. [Citation22], we modelled the antibiotic prescribing rate as a periodic function depending on time t in the transmission of MRSA, i.e. ϵ(t)=ϵ0(1+ϵ1sin((2π/365)(t240))), which has a period of one year (365 days) and implies that antibiotic prescribing rate increases starting at the beginning of August, gains a peak in winter and then decreases starting at the beginning of February according to the data shown in Figures  and . Based on the definition in Bacae¨r and Guenaoui [Citation2] and the calculation procedure in Wang and Zhao [Citation27], we deduced the basic reproduction number R0 for the periodic deterministic model and carried out some mathematical analysis to prove that the infection would go to extinction if the basic reproduction number is less than unity and would persist if it is greater than unity. On the basis of parameter values given in Table , the basic reproduction number is estimated to be 1.476, which implies that MRSA infections persist in hospitals. Our simulations suggest that the prevalence of colonized patients with antibiotic exposure has periodic phenomenon between about 34% and 39% and the prevalence of colonized patients without antibiotic exposure is between 4% and 6% in Figures  and . In addition, since we observe a lag about 15 days between the pattern of colonized patients with antibiotic exposure and antibiotic prescription rate in Figure , we suggest that there may be a temporal correlation between antibiotic use and resistance. By controlling the proportion of patients from four compartments on admission, Figures  and  imply that the prevalence of colonized patients with or without antibiotic exposure would reduce greatly if only uncolonized patients without antibiotic exposure are admitted. This means that detection and isolation of MRSA colonized patients on admission may be a useful intervention to control the hospital infection, and also strengthens the importance to increase the public education about how to use antibiotics properly at community.

It follows from the sensitivity analysis that the basic reproduction number is sensitive to the cleaning/disinfection rate of environment γb, shedding rate of bacteria from colonized patients with antibiotic exposure to environment υpA, and the discharge rate of colonized patients with antibiotic exposure γcA. In particular, environmental cleaning is the most important intervention to control the infection according to our sensitivity analysis. Figure (a) shows that increasing the environmental cleaning/disinfection rate γb from 0.6 to 1 reduces the basic reproduction number from 1.705 to 1.065. Besides, if the shedding rate of colonized patients with antibiotic exposure υpA is below 300, the basic reproduction number can be below 1 (Figure (b)). Because the free-living bacteria have no proper condition to reproduce themselves in hospitals, shedding bacteria from colonized patients becomes a key factor in transmission of MRSA. This also indirectly shows the impact of environmental cleaning. We also found that if colonized patients with antibiotic exposure stay hospitals more than 18 days on average, the basic reproduction number increases dramatically. However, colonized patients with antibiotic exposure usually have resistance to many common antibiotics, which makes it harder and longer to treat them. So how to make an efficient and right treatment plan for colonized patients with antibiotic exposure is a challenge to control the infection. We also observed that the hand hygiene compliance of HCWs change little on the basic reproduction number. We guess the reason is that hospitals have paid enough attention to the hand hygiene of HCWs while still lack attention on the indirect transmission via contaminated environment that maybe is much more important than we thought. This again explains why the environmental cleaning γb and the shedding rate γcA affect greatly the basic reproduction number in our sensitivity analysis.

Hence, in order to control the infection, we believe it is necessary to strengthen the surveillance of environmental cleaning with feedback to cleaning team, try to use more efficient cleaning products, highlight the necessary of effective antimicrobial stewardship programmes, increase active screening on admission and subsequent isolation of positive cases, and treat patients quickly and efficiently. Nevertheless, a comprehensive cost-effectiveness analysis of control policy is needed in future work.

Our model emphasizes the importance of incorporating the indirect transmission via free-living bacteria in the environment, where they are assumed to be uniformly distributed. However, bacterial density varies in rooms of hospitals [Citation4], so future work should take the environmental heterogeneity into consideration.

Acknowledgements

We would like to thank the two anonymous reviewers for their helpful comments and suggestions which helped us to improve the paper.

Disclosure statement

No potential conflict of interest was reported by the authors.

Additional information

Funding

This research was partially supported by National Science Foundation ( DMS-1412454) and the University of Miami Provost's Research Award; Directorate for Mathematical and Physical Sciences.

References

  • D.J. Austin and R.M. 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(1384) (1999), pp. 721–738. doi: 10.1098/rstb.1999.0425
  • N. Bacaër and S. Guernaoui, The epidemic threshold of vector-borne diseases with seasonality, J. Math. Biol. 53(3) (2006), pp. 421–436. doi: 10.1007/s00285-006-0015-0
  • J.M. Boyce, G. Potter-Bynoe, C. Chenevert and T. King, Environmental contamination due to Methicillin-resistant staphylococcus aureus possible infection control implications, Infect. Control Hosp. Epidemiol. 18(9) (1997), pp. 622–627. doi: 10.1086/502213
  • C. Browne and G.F. Webb, A nosocomial epidemic model with infection of patients due to contaminated rooms, Discrete Continuous Dyn. Syst. Ser. B 12 (2015), pp. 761–787.
  • Center for Disease Control and Prevention, Antibiotic/antimicrobial Resistance, last updated March 29, 2018. https://www.cdc.gov/drugresistance/index.html.
  • F. Chamchod, S. Ruan, Modeling Methicillin-resistant staphylococcus aureus in hospitals: Transmission dynamics, antibiotic usage and its history. Theor. Biol. Med. Model. 9(1) (2012), 25. doi: 10.1186/1742-4682-9-25.
  • E.M.C. D'Agata, M.A. Horn, S. Ruan … J.R. Wares, Efficacy of infection control interventions in reducing the spread of multidrug-resistant organisms in the hospital setting. PLoS One 7(2) (2012), e30170. doi: 10.1371/journal.pone.0030170.
  • E.M.C. D'Agata, P. Magal, D. Olivier, S. Ruan and G.F. Webb, Modeling antibiotic resistance in hospitals: The impact of minimizing treatment duration, J. Theor. Biol. 249(3) (2007), pp. 487–499. doi: 10.1016/j.jtbi.2007.08.011
  • E.M.C D'agata, G.F. Webb, M.A. Horn, R.C. Moellering and S. Ruan, Modeling the invasion of community-acquired Methicillin-resistant staphylococcus aureus into hospitals, Clin. Infect. Dis. 48(3) (2009), pp. 274–284. doi: 10.1086/595844
  • S.J. Dancer, How antibiotics can make us sick: The less obvious adverse effects of antimicrobial chemotherapy, Lancet Infect. Dis. 4(10) (2004), pp. 611–619. doi: 10.1016/S1473-3099(04)01145-4
  • S.J. Dancer, Importance of the environment in Meticillin-resistant staphylococcus aureus acquisition: The case for hospital cleaning, Lancet Infect. Dis. 8(2) (2008), pp. 101–113. doi: 10.1016/S1473-3099(07)70241-4
  • T. Delorme, A. Garcia and P. Nasr, A longitudinal analysis of Methicillin-resistant and sensitive staphylococcus aureus incidence in respect to specimen source, patient location, and temperature variation, Int. J. Infect. Dis. 54 (2017), pp. 50–57. doi: 10.1016/j.ijid.2016.11.405
  • T.N. Doan, D.C.M. Kong, C. Marshall, C.M.J. Kirkpatrick and E.S. McBryde, Modeling the impact of interventions against acinetobacter baumannii transmission in intensive care units, Virulence 7(2) (2016), pp. 141–152. doi: 10.1080/21505594.2015.1076615
  • M.R. Eber, M. Shardell, M.L. Schweizer … E.N. Perencevich, Seasonal and temperature-associated increases in gram-negative bacterial bloodstream infections among hospitalized patients. PLoS One 6(9) (2011), e25298. doi: 10.1371/journal.pone.0025298.
  • J.T. Fishbain, J.C. Lee, H.D. Nguyen, J.A. Mikita, C.P. Mikita, C.F.T Uyehara and D.R. Hospenthal, Nosocomial transmission of Methicillin-resistant staphylococcus aureus: A blinded study to establish baseline acquisition rates, Infect. Control Hosp. Epidemiol. 24(6) (2003), pp. 415–421. doi: 10.1086/502224
  • Q. Huang, M.A. Horn and S. Ruan, Modeling the effect of antibiotic exposure on the transmission of Methicillin-resistant staphylococcus aureus in hospitals with environmental contamination. preprint
  • S. Leekha, D.J. Diekema and E.N. Perencevich, Seasonality of staphylococcal infections, Clin. Microbiol. Infect. 18(10) (2012), pp. 927–933. doi: 10.1111/j.1469-0691.2012.03955.x
  • L.A. Mermel, J.T. Machan, S. Parenteau, Seasonality of MRSA infections. PLoS One 6(3) (2011), e17925. doi: 10.1371/journal.pone.0017925.
  • L. Perko, Differential Equations and Dynamical Systems, 3rd ed., Springer, New York, 2000.
  • R.E. Polk, C. Fox, A. Mahoney, J. Letcavage and C. MacDougall, Measurement of adult antibacterial drug use in 130 US hospitals: Comparison of defined daily dose and days of therapy, Clin. Infect. Dis.44(5) (2007), pp. 664–670. doi: 10.1086/511640
  • R.E. Polk, C.K. Johnson, D. McClish, R.P. Wenzel and M.B. Edmond, Predicting hospital rates of Fluoroquinolone-resistant pseudomonas aeruginosa from fluoroquinolone use in US hospitals and their surrounding communities, Clin. Infect. Dis. 39(4) (2004), pp. 497–503. doi: 10.1086/422647
  • L. Sun, E.Y. Klein and R. Laxminarayan, Seasonality and temporal correlation between community antibiotic use and resistance in the United States, Clin. Infect. Dis. 55(5) (2012), pp. 687–694. doi: 10.1093/cid/cis509
  • E. Tacconelli, Antimicrobial use: Risk driver of multidrug resistant microorganisms in healthcare settings, Curr. Opin. Infect. Dis. 22(4) (2009), pp. 352–358. doi: 10.1097/QCO.0b013e32832d52e0
  • E. Tacconelli, G. De Angelis, M.A. Cataldo, E. Pozzi and R. Cauda, Does antibiotic exposure increase the risk of Methicillin-resistant staphylococcus aureus (MRSA) isolation? A systematic review and meta-analysis, J. Antimicrob. Chemother. 61(1) (2007), pp. 26–38. doi: 10.1093/jac/dkm416
  • J. Wang, L. Wang, P. Magal, Y. Wang, S. Zhuo, X. Lu and S. Ruan, Modelling the transmission dynamics of Meticillin-resistant staphylococcus aureus in Beijing Tongren Hospital, J. Hosp. Infect.79(4) (2011), pp. 302–308. doi: 10.1016/j.jhin.2011.08.019
  • L. Wang, S. Ruan, Modeling nosocomial infections of Methicillin-resistant staphylococcus aureus with environment contamination. Sci. Rep. 7 (2017), 580. doi: 10.1038/s41598-017-00261-1.
  • W. Wang and X.-Q. Zhao, Threshold dynamics for compartmental epidemic models in periodic environments, J. Dyn. Differ. Equ. 20(3) (2008), pp. 699–717. doi: 10.1007/s10884-008-9111-8
  • X. Wang, Y. Xiao, J. Wang and X. Lu, A mathematical model of effects of environmental contamination and presence of volunteers on hospital infections in China, J. Theor. Biol. 293 (2012), pp. 161–173. doi: 10.1016/j.jtbi.2011.10.009
  • F. Zhang and X.-Q. Zhao, A periodic epidemic model in a patchy environment, J. Math. Anal. Appl.325(1) (2007), pp. 496–516. doi: 10.1016/j.jmaa.2006.01.085
  • X.-Q. Zhao, Dynamical Systems in Population Biology, Springer, New York, 2003.