Processing math: 100%
Research article Special Issues

Optimal control and Bayes inference applied to complex microbial communities


  • Interactions between species are essential in ecosystems, but sometimes competition dominates over mutualism. The transition between mutualism-competition can have several implications and consequences, and it has hardly been studied in experimental settings. This work studies the mutualism between cross-feeding bacteria in strains that supply an essential amino acid for their mutualistic partner when both strains are exposed to antimicrobials. When the strains are free of antimicrobials, we found that, depending on the amount of amino acids freely available in the environment, the strains can exhibit extinction, mutualism, or competition. The availability of resources modulates the behavior of both species. When the strains are exposed to antimicrobials, the population dynamics depend on the proportion of bacteria resistant to the antimicrobial, finding that the extinction of both strains is eminent for low levels of the resource. In contrast, competition between both strains continues for high levels of the resource. An optimal control problem was then formulated to reduce the proportion of resistant bacteria, which showed that under cooperation, both strains (sensitive and resistant) are immediately controlled, while under competition, only the density of one of the strains is decreased. In contrast, its mutualist partner with control is increased. Finally, using our experimental data, we did parameters estimation in order to fit our mathematical model to the experimental data.

    Citation: Jhoana P. Romero-Leiton, Kernel Prieto, Daniela Reyes-Gonzalez, Ayari Fuentes-Hernandez. Optimal control and Bayes inference applied to complex microbial communities[J]. Mathematical Biosciences and Engineering, 2022, 19(7): 6860-6882. doi: 10.3934/mbe.2022323

    Related Papers:

    [1] Hermann Mena, Lena-Maria Pfurtscheller, Jhoana P. Romero-Leiton . Random perturbations in a mathematical model of bacterial resistance: Analysis and optimal control. Mathematical Biosciences and Engineering, 2020, 17(5): 4477-4499. doi: 10.3934/mbe.2020247
    [2] Colette Calmelet, John Hotchkiss, Philip Crooke . A mathematical model for antibiotic control of bacteria in peritoneal dialysis associated peritonitis. Mathematical Biosciences and Engineering, 2014, 11(6): 1449-1464. doi: 10.3934/mbe.2014.11.1449
    [3] 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
    [4] 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
    [5] Harry J. Dudley, Zhiyong Jason Ren, David M. Bortz . Competitive exclusion in a DAE model for microbial electrolysis cells. Mathematical Biosciences and Engineering, 2020, 17(5): 6217-6239. doi: 10.3934/mbe.2020329
    [6] 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
    [7] Christopher Botelho, Jude Dzevela Kong, Mentor Ali Ber Lucien, Zhisheng Shuai, Hao Wang . A mathematical model for Vibrio-phage interactions. Mathematical Biosciences and Engineering, 2021, 18(3): 2688-2712. doi: 10.3934/mbe.2021137
    [8] Miller Cerón Gómez, Eduardo Ibarguen Mondragon, Eddy Lopez Molano, Arsenio Hidalgo-Troya, Maria A. Mármol-Martínez, Deisy Lorena Guerrero-Ceballos, Mario A. Pantoja, Camilo Paz-García, Jenny Gómez-Arrieta, Mariela Burbano-Rosero . Mathematical model of interaction Escherichia coli and Coliphages. Mathematical Biosciences and Engineering, 2023, 20(6): 9712-9727. doi: 10.3934/mbe.2023426
    [9] 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
    [10] Eduardo Ibargüen-Mondragón, Lourdes Esteva, Edith Mariela Burbano-Rosero . Mathematical model for the growth of Mycobacterium tuberculosis in the granuloma. Mathematical Biosciences and Engineering, 2018, 15(2): 407-428. doi: 10.3934/mbe.2018018
  • Interactions between species are essential in ecosystems, but sometimes competition dominates over mutualism. The transition between mutualism-competition can have several implications and consequences, and it has hardly been studied in experimental settings. This work studies the mutualism between cross-feeding bacteria in strains that supply an essential amino acid for their mutualistic partner when both strains are exposed to antimicrobials. When the strains are free of antimicrobials, we found that, depending on the amount of amino acids freely available in the environment, the strains can exhibit extinction, mutualism, or competition. The availability of resources modulates the behavior of both species. When the strains are exposed to antimicrobials, the population dynamics depend on the proportion of bacteria resistant to the antimicrobial, finding that the extinction of both strains is eminent for low levels of the resource. In contrast, competition between both strains continues for high levels of the resource. An optimal control problem was then formulated to reduce the proportion of resistant bacteria, which showed that under cooperation, both strains (sensitive and resistant) are immediately controlled, while under competition, only the density of one of the strains is decreased. In contrast, its mutualist partner with control is increased. Finally, using our experimental data, we did parameters estimation in order to fit our mathematical model to the experimental data.



    The rapid surge of resistant strains and the void in discovering of new antimicrobials has made antimicrobial resistance one of the most critical problems in public health nowadays. There has been a great effort from the scientific community to understand the basic mechanisms by which bacteria become resistant to several antibiotics; almost all efforts have focused on the study of a single strain, and it is just in recent times that we have been looking at how the interaction between members of complex communities are shaping the susceptibility profile.

    In nature, microbes co-exist in multi-species communities with ecological dynamics driven by the complex interaction between individual strains and the environment. Syntrophic interactions, whereby bacterial cells exchange costly metabolites for the benefit of both interacting partners, are pervasive in bacterial communities as they enhance bacterial survival in hostile environments.

    Availability of resources affects the nature of the interactions between members of a microbial consortium [1]. These interactions can be determinants of the stability of the consortia. Based on two-pair interactions, there are typically three outcomes; cooperation where both members of the community are beneficial to the other one; competition where the presence of one member is detrimental to the other one and neutral interaction; where the members of the community have no interaction.

    Typically competitive interactions end up in competitive exclusion, decreasing the diversity of the consortia. For syntrophic interactions, if the competition for resources is very strong, one community member's collapse could collapse the total population.

    The effect of changes in interactions have on the control of antibiotic resistance is unknown. There are some examples where cooperation can help a community to survive in stressful environments [2]. We model a synthetic community composed of two different strains; both are auxotrophic to an amino acid, an essential metabolite. This condition makes them incapable of growing on their own in an environment without amino acids. In contrast, they can grow in a minimal media when they are together due to the leak of all the other amino acids from their partner [3]. Auxotroph cells are essential for natural communities [4], they shape the composition and can be determinant of the stability of a microbial community. For instance, they have been found in aquatic communities [5], microbial soil communities[6] and the human microbiome [7,8]. Although auxotrophies could increase the dependence of individuals in the communities, they could also increase fitness by reducing the nutrient requirements of an environment, which means that a community could support growth in a poorer context.

    In the past years, auxotrophic strains have been used to study metabolic cross-feeding interactions in particular environmental context [1,9,10]. Pair-wise interactions are useful to determine changes in the interaction and to study division of labor inside the community, such as public goods. A relevant example of public goods is the production of enzymes able to degrade antibiotics.

    Cross-feeding interactions in natural environments are important to increase diversity in the population by decreasing the fitness of the individual and increasing the fitness of the community[11]. It has been shown that these interactions are dynamic and can change with respect to the environment. Auxotrophies are a particular example of this. It has also been shown that auxotrophic phenotypes appear very fast in a population growing in a rich nutrient environment [10]. This feature can be crucial in diverse communities, like the microbiota, particularly when an antimicrobial is present, which disrupts the community members [4]. If auxotrophic phenotypes appear, having functional redundancies could help preserve general function in a community. For this reason, understanding what is the effect of an antimicrobial in a syntrophic consortium could be of great value.

    This paper explores a population-based model and control theory to understand the effect of antimicrobials in microbial communities with syntrophic interactions and how the environment minimizes resistant bacteria in the system.

    We first propose an optimal control problem based on Hoek et al. in [1] of an auxotrophic consortium of two bacteria competing for the same resources in the presence of a gradient of antibiotic. Then we described our experimental data and performed numerical experiments. Finally, using Bayesian Inference, we estimate some interesting mathematical model parameters using our experimental data. Numerical experiments validate all theoretical results.

    In order to better understand the different types of interaction between cross-feeding bacteria, we formulate a similar mathematical model to the proposed by Hoek et al. In [1], two strains of yeast, one auxotrophic to tryptophan (trp-) and the other auxotrophic to leucine (leu-) interact in an environment with different resource availability. In this formulation, we assume that strain trp- (X) and strain leu- (Y) are divided into susceptible and resistant bacteria (Xs, Xr and Ys, Yr, respectively). We assume both strains interact with a bacterial infection in an individual's body. We are interested in modeling the dynamics of bacterial resistance acquisition through antimicrobial resistance induced by antibiotic-mediated selection. Thus, the following hypothesis is considered for the strain X (similar assumptions can be considered for the strain Y): sensitive bacteria grows at a rate proportional to resistant bacteria (if a is the growth rate of sensitive bacteria, then aˉγx is the growth rate of resistant bacteria) following the logistic equation with a carrying capacity K. ˉF is the amount of supplemented amino acids, ˉκ represents an effective Monod constant, and β is a parameter that quantifies the asymmetry of benefit that each bacteria receives from its partner. The density of sensitive bacteria become resistant by antibiotic-mediated selection is given through the term ϵqλXs, where ϵ is the efficacy of the antimicrobial, λ is the antimicrobial supply rate, q is the proportion of mutation (therefore (1q) represents the proportion of elimination). Both population turnover at a constant dilution rate ˉδ. Thus, the per capita growth rate of both bacteria is adjusted by the mutualistic partner as well as the supplemented amino acids:

    ˙Xs=ˉγx(Xs+Xr)(Ys+Yr+ˉFYs+Yr+ˉF+ˉκ)(1Xs+Xr+Ys+YrK)ϵqλXsϵ(1q)λXsˉδXs˙Xr=aˉγx(Xs+Xr)(Ys+Yr+ˉFYs+Yr+ˉF+ˉκ)(1Xs+Xr+Ys+YrK)+ϵqλXsˉδXr˙Ys=ˉγy(Ys+Yr)(β(Xs+Xr)+ˉFβ(Xs+Xr)+ˉF+ˉκ)(1Xs+Xr+Ys+YrK)ϵqλYsϵ(1q)λYsˉδYs˙Yr=aˉγy(Ys+Yr)(β(Xs+Xr)+ˉFβ(Xs+Xr)+ˉF+ˉκ)(1Xs+Xr+Ys+YrK)+ϵqλYsˉδYr. (2.1)

    A complete description of the parameters involved in Model (2.1) can be found on Table 1.

    Table 1.  Definition and dimension of the parameters involved in Model (2.1).
    Parameter Definition Unit
    a Cost of the resistance Dimensionless
    ˉγx Growth rate of strain X 1/time
    ˉγy Growth rate of strain Y 1/time
    β Asymmetry constant Dimensionless
    ˉF Amount of supplementary amino acid Population
    ˉκ Monod constant Population
    ϵ Antimicrobial efficacy 1/time
    λ Supply concentration of the antimicrobial Dimensionless
    q Mutation proportion Dimensionless
    ˉδ Dilution rate 1/time
    K Carry capacity Dimensionless

     | Show Table
    DownLoad: CSV

    Now, we define μμ=(μ1(t),μ2(t)) as a control variable associated with the mutations. Thus, both strains of susceptible bacteria mutate at a rate (1μ1(t))ϵqλXs and (1μ2(t))ϵqλYs, respectively, where for i=1,2, μi(t)[0,1] (μi=0 represents no efficacy of the control, while μi=1 indicates that the use of the control is completely effective). Thus, we have that each control variable μi(t) provides information about the amount of bacteria that must not mutate at time t. To minimize the number of resistant bacteria, we define the cost function:

    J[μμ]=T0(ˉc1Xr+ˉc2Yr+12μ1(t)2+d212μ2(t)2)dt,

    where the parameters c1 and c2 represent social cost and the parameters d1 and d2 represent relative weights associated to the controls. Additionally, we define the boundary conditions:

    X(0)=(Xs0,Xr0,Ys0,Yr0)=X0X(T)=(Xsf,Xrf,Ysf,Yyf)=Xf,

    and we assume that the final time T is a fixed implementation time of the control strategies, the final state Xf is variable, and the initial state X0 is a coexistence equilibrium of System (2.1). We suppose that each control is in the set of Lebesgue measurable functions with 0μ(t)1, t[0,T] (U called the set of admissible controls).

    Let us define ω as a rescaling parameter with dimension 1/population×time. To obtain a dimensionless formulation of System (2.1), the state variables have been divided by the carrying capacity of the environment, the parameters and the time have also been adjusted to be dimensionless as follows

    Sx=XsK,Rx=XrK,Sy=YsK,Ry=YrKandτ=ωKt,

    and

    c1=ˉc1K,c2=ˉc2K,F=ˉFK,κ=ˉκK,γx=ˉγxωK,γy=ˉγyωKc=ϵλωKandδ=ˉδωK.

    Thus, our optimal control problem can be written in the following dimensionless form

    {minJ[μμ]=T0(c1Rx+c2Ry+d112μ1(t)2+d212μ2(t)2)dtdSxdτ=γx(Sx+Rx)(Sy+Ry+FSy+Ry+F+κ)[1(Sx+Rx+Sy+Ry)](1μ1(t))qcSx(1q)cSxδSxdRxdτ=aγx(Sx+Rx)(Sy+Ry+FSy+Ry+F+κ)[1(Sx+Rx+Sy+Ry)]+(1μ1(t))qcSxδRxdSydτ=γy(Sy+Ry)(β(Sx+Rx)+Fβ(Sx+Rx)+F+κ)[1(Sx+Rx+Sy+Ry)](1μ2(t))qcSy(1q)cSyδSydRydτ=aγy(Sy+Ry)(β(Sx+Rx)+Fβ(Sx+Rx)+F+κ)[1(Sx+Rx+Sy+Ry)]+(1μ2(t))qcSyδRyX(0)=(Sx0,Rx0.Sy0.Ry0)=X0X(T)=(Sxf,Rxf.Syf.Ryf)=Xf. (2.2)

    Let us set as X=(Sx,Rx,Sy,Ry) the vector of states, Z=(z1,z2,z3,z4) the vector of adjoint variables and f0(t,X,μμ) the integrand of the cost function. We will use the Pontryagin principle for bounded controls to compute the optimal control of Problem (2.2). First of all, following the classical results given in the references [12,13] it is easy to verify the existence of optimal controls (it is enough to check that the properties (i)–(iii) of Section 3.4 from reference [14] are fulfilled). Now, the Hamiltonian H associated with the Problem (2.2) is defined by H(t,X(t),μ(t),Z(t))=f0(t,X,μ)+Z(t)f(X,t,μ), that is

    H=c1Rx+c2Ry+d112μ21+d212μ22+z1[γx(Sx+Rx)(Sy+Ry+FSy+Ry+F+κ)[1(Sx+Rx+Sy+Ry)](1μ1)qcSx(1q)cSxδSx]+z2[aγx(Sx+Rx)(Sy+Ry+FSy+Ry+F+κ)[1(Sx+Rx+Sy+Ry)]+(1μ1)qcSxδRx]+z3[γy(Sy+Ry)(β(Sx+Rx)+Fβ(Sx+Rx)+F+κ)[1(Sx+Rx+Sy+Ry)](1μ2)qcSy(1q)cSyδSy]+z4[aγy(Sy+Ry)(β(Sx+Rx)+Fβ(Sx+Rx)+F+κ)[1(Sx+Rx+Sy+Ry)]+(1μ2)qcSyδRy]. (2.3)

    The adjoint system and state equations define the optimal system. The following theorem summarizes the main result of this section.

    Theorem 2.1. There is an optimal solution X(t) that minimize J[μμ] in [0,T]. Moreover, there exits a vector of adjoint variables Z such that

    ˙z1=γx((Sy+Ry+F)(1+2Sx+Sy+2Rx+Ry)(Sy+Ry+F+κ))(z1+az2)c+γy(F2+Fκβκ+β(βR2x+κRy+2Rx(F+κ+βSx)(β(Sx+Rx)+F+κ)2)(z3+az4)+γy(Sx(2(F+κ)+βSx)+κSy)(Sy+Ry)(β(Sx+Rx)+F+κ)2)(z3+az4)+cqμ1(z2z1)c(qz2z1)+δz1˙z2=γx((Sy+Ry+F)(1+2Sx+Sy+2Rx+Ry)(Sy+Ry+F+κ))(z1+az2)+γy(F2+Fκβκ+β(βR2x+κRy+2Rx(F+κ+βSx)(β(Sx+Rx)+F+κ)2)(z3+az4)+γy(Sx(2(F+κ)+βSx)+κSy)(Sy+Ry)(β(Sx+Rx)+F+κ)2)(z3+az4)c1+δz2˙z3=γy((β(Sx+Rx)+F)(1+Sx+2Sy+Rx+2Ry)(β(Sx+Rx)+F+κ))(z3+az4)+γx(F2+(1+F)κ+κRx+R2y+κSx+2(F+κ)Sy(Sy+Ry+F+κ)2)(z1+az2)+γx(S2y+2Ry(F+κ+Sy)(Sy+Ry)(Sy+Ry+F+κ)2)(z1+az2)+cqμ2(z4z3)c(qz4z3)+δz3˙z4=γy((β(Sx+Rx)+F)(1+Sx+2Sy+Rx+2Ry)(β(Sx+Rx)+F+κ))(z3+az4)+γx(F2+(1+F)κ+κRx+R2y+κSx+2(F+κ)Sy(Sy+Ry+F+κ)2)(z1+az2)+γx(S2y+2Ry(F+κ+Sy)(Sy+Ry)(Sy+Ry+F+κ)2)(z1+az2)c2+δz4, (2.4)

    with transversality condition zi(t)=0 for i=1,2,3,4 which satisfies:

    μ1=qcSx(z2z1)d1μ2=qcSy(z4z3)d2. (2.5)

    To see a detailed proof of the previous theorem see Appendix A.

    We used Escherichia coli strains from the Keio collection containing either ΔilvA or ΔtyrA deletions [15], each strain was transformed with the plasmid pBGT-1 [16] carrying blaTEM-1 gene that confers resistance to ampicillin and a GFP fluorescent marker inducible with arabinose, pBGT-1 is a non-conjugative plasmid with around 20 copies in average per cell. Strains were grown in LB medium (Lysogeny Broth) with 40 µg/mL of kanamycin and 100 µg/mL of ampicillin (for those with plasmid) at 30℃ for 16 hours after incubation strains were washed with M9 salts and re-suspended in M9 minimal medium supplemented with glucose.

    To measure susceptibility to the antimicrobial, we performed dose-response experiments in 96-well plates. The ampicillin concentrations used were 0, 6, 15, 36, 89,220,540, 1326, 3257 and 8000 µg/mL, each well was inoculated with 20 µL of clean cells for a total of 200 µL per well and incubated in an ELx808 plate reader at 30℃ with continuous shaking, OD630 was measured every 20 minutes for 24 hours, data not shown.

    Co-cultures were performed in 96-well plates with 180 µL of LB media with 6 µg/mL of ampicillin, and 20 µL of cells with 80% of susceptible cells (ΔilvA or ΔtyrA) and 20% of resistant cells (ΔtyrA-pBGT or ΔilvA-pBGT). Two equal plates were grown in a ELx808 plate reader at 30℃ with eight replicates for condition in each plate, OD630 was measured every 20 minutes for 24 hours, one plate was used for growth rate analysis, and the second one was used to measure colony forming units (CFU), relative abundance and cell viability, every 2 hours we collected one of the replicates of the second plate and divided the sample, as we had to open the well, this procedure was destructive to the sample, to have sampled every 2 hours, we completely used one of the plates, the other one was kept in the plate reader.

    For relative abundance and cell viability, one replicate was collected every 2 hours and supplemented with 20 µL of arabinose (0.5%), incubated at 30℃ for 4 hours, and stored at 4℃ for one day to induce GFP fluorescence. After storage, 100 µL of each sample was frozen, 5 µL from a 1:1000 dilution were incubated for 2 days at 30C in selective agar with M9 minimal media, supplemented with 10.9 mg/L isoleucine and 7.15 mg/L tyrosine and 0.5% of arabinose for CFU counting, the rest of the sample was used to measured abundance of resistant cells using a CytoFLEX S cytometer (20,000 events per sample).

    For experimental data, we performed 16 replicas in total, 8 were kept at a plate reader to measure OD shown in Figure 1A, the other eight replicas were used to measured abundance of cells and viability, the experimental data measured in the flux cytometer every 2 hours splitting resistance and susceptible bacteria are shown in Figure 1B.

    Figure 1.  Experimental data from co-cultures of cells with 80% of susceptible cells (ΔilvA or ΔtyrA) and 20% of resistant cells (ΔtyrA-pBGT or ΔilvA-pBGT) grown in LB media with 6 µg/ml of ampicillin (low concentration). In A we showed the optical density of eight replicas of the co-culture in 24 hours, solid line represent the average of the 8 replicas and shadows showed standard error for each data point, measurements taken every 20 minutes. B shows the frequency of susceptible and resistant bacteria, dots represents data point every 2 hours measured in the flux cytometer, all data processed with Matlab.

    In order to performed the optimal control numerical experiments, we used the Backward-Forward Sweep Method described on Lenhart et al. [13] in Matlab interface, and also we used our experimental data given in Section 3 and some data collected in the reference [1], from two non-mating budding yeast strains of Saccharomyces cerevisiae that were designed to be deficient in the biosynthesis of an essential amino acid and also to overproduce the amino acid required by the ir partner. In such laboratory experiments, the auxotrophic strain of leucine (leu-) overproduces tryptophan, while the auxotrophic strain of tryptophan (trp-) overproduces leucine. According to reference[1], these strains have previously been shown to form a cross-feeding mutualism when grown on solid agar, with each strain losing the amino acid needed by its partner. We assume that the carrying capacity K=0.1×106 and the temporal value is ω=0.3×105. The values of the other parameters taken in this study are given in Table 2.

    Table 2.  Values of the parameters involved in Models (2.1) and (2.2). Some of them were taken from the reference [1], whereas other from our experimental data.
    Parameter Dimensional model Dimensionless model
    Growth rate of bacteria X 0.3 1
    Growth rate of bacteria Y 0.288 0.96
    Asymmetry constant 2 2
    Dilution rate 0.15 0.5
    Monod constant 12,000 0.12
    Cost of resistance 0.01 0.01
    Antimicrobial concentration λ=10 c=16.67

     | Show Table
    DownLoad: CSV

    In these experiments we consider that the amount of supplementary amino acids and the proportion of mutations are variables. The mutation portion (q) values used were 0.1, 0.5 and 0.9, parameters for the amount of amino acids used are shown in Table 3.

    Table 3.  Values for the amount of supplementary amino acids.
    Dimensional model (ˉF) Dimensionless model (F)
    1000 0.01
    5000 0.05
    150,000 1.5

     | Show Table
    DownLoad: CSV

    In Figure 2 we show some numerical experiments of sensitive bacteria (X and Y) of Model (2.2), when there are no antimicrobials, for different values of F. From this figure, we can see that when F is small enough (F = 0.01) there is the extinction of both strains. When F grows (F = 0.05) there is cooperation or mutualism between both strains, whereas, for large values of F (1.5), there is competition between both strains, being evident in the increase in the strain X and the extinction of the strain Y.

    Figure 2.  Simulations of sensitive strains X and Y of Model (2.2), when there are no antimicrobials, for different values of supplementary amino acids (F). When F is small, there is extinction, when F grows, there is mutualism, whereas, for large values of F, there is competition.

    In Figure 3, we supply antimicrobials in our system, we keep fixed the value of F = 0.01 (where the is the extinction of both bacteria), and vary the values of the mutation proportion q (0.1, 0.5 and 0.9). In this case, we can see that the sensitive strains tend to zero regardless of the value of q, whereas for large values of q, the resistant bacteria of both types (X and Y) increase in density at the beginning of the time.

    Figure 3.  Simulations of the state equations of Model (2.2) for F = 0.01 and different values of q. Here, the sensitive strains tend to zero regardless of the value of q, whereas for large values of q, the resistant bacteria of both phenotypes increase in density at the beginning of the time.

    A different scenario occurs for F = 0.05 (where there is cooperation between both strains), which can be evident in Figure 4. Here we can see that both strains become extinct only for small values of q (q = 0.5 or less), while if q increases (0.9), they coexist in density over time.

    Figure 4.  Simulations of the state equations of Model (2.2) for F = 0.05 and different values of q. Here, both strains become extinct only for small values of q (q = 0.5 or less), while if q increases, they coexist in density over time.

    Finally, for F = 1.5 (see Figure 5), we can see that the strain Y decreases no matter what value of q is taken. On the other hand, the strain X stabilizes at increasing densities as the value of q changes.

    Figure 5.  Simulations of the state equations of Model (2.2) for F = 1.5 and different values of q. The strain Y decreases no matter what value of q is taken. The strain X stabilizes at increasing densities as the value of q changes.

    To observe the effect of controlling the antimicrobial resistance in the system where the strains X and Y interact in an environment with antimicrobials, we use the Forward-Backward Sweep Method proposed by Lenhart and Workman [13]. The values of the parameters associated with Problem (2.2) are shown in Table 4. We only introduce the control variables for the cases where there is a high proportion of mutation (q = 0.9) and for the mutualism (F = 0.05) and competition (F = 1.5). In Figure 6 we show the results of control when F=0.05 and q=0.9 (where there is cooperation). In this case, with controls, both populations of bacteria are controlled immediately, maintaining the two controls at their maximum effort during the first 40 days and then rapidly decreasing to zero during the rest of the control campaign. In Figure 8 we show the results when F=1.5 and q=0.9 (when there is competition). Here, it can be seen that only the strain X is controlled from the first day of the campaign. But the strain Y increases with the control. The effort is only made for the first control at 100%, while the second control remains at zero throughout the campaign.

    Table 4.  Values of the parameters associated with the optimal control problem (2.2).
    Parameter Value
    Weights d1 0.0001
    d2 0.0001
    Social costs c1 0.0001
    c2 0.0001

     | Show Table
    DownLoad: CSV
    Figure 6.  Simulations of the Optimal Control (2.2) under mutualism (F = 0.05 and q = 0.9). With controls, both bacteria are controlled immediately.
    Figure 7.  Simulations of the Optimal Control (2.2) under competition (F = 1.5 and q = 0.9). Only the strain X is controlled from the first day of the campaign. But the strain Y increases with the control.
    Figure 8.  Simulations of the Optimal Control (2.2) under competition (F = 1.5 and q = 0.9). Only the strain X is controlled from the first day of the campaign. But the strain Y increases with the control.

    For the parameter estimation of Model (2.1) we used the experimental data described in Section 3. To simplify notation, we will drop the bar symbol on the parameters from here. The experimental data is shown in Figure 9. Since the data size is very low, i.e., we have just six point for each time series of type of bacteria, we have implemented a re-sample with an interpolation for the values of odd hours in order to increase the data size. Specifically, we used a rich media called LB, and low ampicillin concentration. The values are described in Table 5.

    Figure 9.  Experimental data: a rich media (LB) and low ampicillin concentration.
    Table 5.  Posterior mean and quantiles of all the estimated parameters of Model (2.1) using the experimental data described in Section 3.
    Parameter Mean Std Min 25% 50% 75%
    κ 6332296.7333 107726.2128 5970880.0000 6257630.0000 6332720.0000 6407300.0000
    γx 7.4709 0.0481 7.3150 7.4371 7.4713 7.5055
    γx 5.9583 0.0909 5.6493 5.9035 5.9557 6.0205
    q 0.0199 0.0008 0.0175 0.0194 0.0199 0.0205

     | Show Table
    DownLoad: CSV

    The fitting curve or estimation of the parameters of a model is considered an inverse problem. Some references of works using Bayesian inference are available in references [17,18,19,20,21]. Let F:RmRs×k, denoted by F(θ) be the forward problem, where θ are the parameters of Model (2.1) to estimate, m the number of parameter to estimate, and k the number of state variables. The inverse problem is formulated as a standard optimization problem

    minθRmF(θ)yobs2, (5.1)

    with yobs. Given yobs=(˜I,˜A), which correspond to strain tyrA- (X) and strain ilvA- (Y) are divided into susceptible and resistant bacteria (Xs, Xr and Ys, Yr, respectively), the conditional probability distribution π(θ|yobs), called the posterior distribution of θ is given by the Bayes' theorem:

    π(θ|yobs)π(yobs|θ)π(θ). (5.2)

    All the available information regarding the unknown parameter θ is codified into the prior distribution π(θ), which specifies our belief in a parameter before observing the data. All the available information regarding the way of how was obtained the measured data is codified into the likelihood distribution π(yobs|θ). This likelihood can be seen as an objective or cost function, as it punishes deviations of the model from the data. A Poisson distribution, P(y|μ), with respect to the time, is typically used to account for the discrete nature of these counts, where µ is the mean of the random variable y, i.e., E[Y]=μ. We assume independent Poisson distributed noise η, i.e., all dependency in the data is codified into the model (2.1). In other words, the positive definite noise covariance matrix η is assumed to be diagonal. The posterior distribution π(θ|yobs) given by (5.2) does not have an analytical closed form since the likelihood function, which depends on the solution of the non-linear model (2.1), does not have an explicit solution. Then, we explore the posterior distribution using the Stan Statistics package [22]. We have used the Automatic Differentiation Variational Inference method (ADVI), which is based on the automatic variational inference algorithm. Specifically, we have used the Full-Rank submethod of ADVI. We have used the interface in Python (PyStan) [22], for more details see this Github link. Table 5 shows the posterior mean, quantiles of all estimated parameters of model (2.1). Table 6 shows the prior distributions summary of the estimated parameters of Model (2.1) using the experiment data set. Gamma distributions were used for parameters κ,γx,γy and q. Uniform distribution were used for the initial conditions Sx0,Rx0,Sy0,Ry0. Figure 11 shows the joint probability density distributions of the estimated parameters within 95% (HPD) using the experimental data. The blue lines represent the medians. Figure 12 shows Model fit for sensitive and resistant bacteria of Model (2.1) using the experiment data. Blue and red dot points represent strain ilvA (X) data, i.e., the susceptible and resistant bacteria, called Sx and Rx, respectively. Orange and purple solid lines represent the median of posterior distribution of the sensitive and resistant bacteria (ilvA-pBGT (X), respectively. Shaded area represent the 95% probability bands for the expected value of sensitive (orange line), resistant bacteria (blue line). Figure 13 shows Model fit for sensitive and resistant bacteria of the model (2.1) using the experiment data. Blue and red dot points represent strain tyrA (Y) data, i.e., the susceptible and resistant bacteria, called Sy and Ry, respectively. Orange and purple solid lines represent the median of the posterior distribution of the sensitive and resistant bacteria strain tyrA-pBGT (Y), respectively. Shaded area represent the 95% probability bands for the expected value of sensitive (orange line), resistant bacteria (blue line).

    Table 6.  Prior distributions summary of the estimated parameters of model (2.1) using the experiment data set. Gamma distributions were used for parameters κ,γx,γy and q. Uniform distribution were used for the initial conditions Sx0,Rx0,Sy0,Ry0.
    Parameter Support Prior distribution
    κ [5×103,5×107] Gamma, a=2.5, b=1
    γx [0,10] Gamma, a=2.5, b=1
    γy [0,10] Gamma, a=2.5, b=1
    q [0,0.2] Gamma, a=2.5, b=1
    Sx0 [1×105,5×107] Uniform
    Rx0 [1×105,5×107] Uniform
    Sy0 [1×105,5×107] Uniform
    Ry0 [1×105,5×107] Uniform

     | Show Table
    DownLoad: CSV
    Figure 10.  Credible intervals of the parameters of the model (2.1) within 95% Highest-Posterior Density (HPD).
    Figure 11.  Joint probability density distributions of the estimated parameters within 95% (HPD) using the experimental data. The blue lines represent the medians.
    Figure 12.  Model fit for sensitive and resistant bacteria of Model (2.1) using the experiment data. Blue and red dot points represent strain ilvA (X) data, i.e., the susceptible and resistant bacteria, called Sx and Rx, respectively. Orange and purple solid lines represent the median of the posterior distribution of the sensitive and resistant bacteria (ilvA-pBGT (X), respectively. Shaded area represent the 95% probability bands for the expected value of sensitive (orange line), resistant bacteria (blue line).
    Figure 13.  Model fit for sensitive and resistant bacteria of the model (2.1) using the experiment data. Blue and red dot points represent strain tyrA (Y) data, i.e., the susceptible and resistant bacteria, called Sy and Ry, respectively. Orange and purple solid lines represent the median of the posterior distribution of the sensitive and resistant bacteria strain tyrA-pBGT (Y), respectively. Shaded area represent the 95% probability bands for the expected value of sensitive (orange line), resistant bacteria (blue line).

    In this work, we attempted to explain the ecological interactions between cross-feeding bacteria in strains that supply an essential amino acid for their mutualistic partner when they are exposed to antimicrobials. Although from the ecological point of view, the microbial interaction can occur between multiple bacteria, for illustration and ease of handling mathematical equations, we considered the interaction between two bacteria. We formulated a mathematical model using ODEs assuming that both strains interact in an environment with different resources availability. We assumed both strains are divided into susceptible and resistant. After that, we estimated the most important parameters of the ODEs model using Bayesian Inference.

    To validate our theoretical results with numerical experiments, we used our experimental data and other data available in the literature. More specifically, we used Escherichia coli strains from the Keio collection containing either ΔilvA or ΔtyrA deletions, where each strain was transformed with a plasmid that confers resistance to ampicillin. For some parameters, we also used data from Hoek et al. in [1] of auxotrophic consortia of two microbes competing for the same resources.

    The theoretical and numerical results showed that when the strains are free of the antimicrobial, depending on the amount of amino acids freely available in the environment, the strains can exhibit extinction, mutualism, or competition, the availability of resources modulates the behavior of both species. In contrast, if the strains are exposed to antimicrobials, the population dynamics depends on the proportion of bacteria that presents resistance to the antimicrobial, finding that for low levels of the resource, the two species become extinct, whereas, for high levels of the resource, competition between both strains is given.

    The optimal control results showed that both strains (sensitive and resistant) are immediately controlled under cooperation, while under competition, only the density of one of the strains decreases, whereas its mutualist partner with control increases. Finally, the growth rates of both strains (ˉγx and ˉγy), the Monod constant (ˉκ) the mutation proportion (q) and the initial conditions of Model (2.1) were estimated using Bayesian Inference and the data set described in Section 3. From Figures 12 and 13, we could observe that the fitting is not entirely accurate. This could be due to the lack of data and/or also because Model (2.1) should be adjusted.

    The results obtained with this study corroborated that the antimicrobial resistance phenomenon is a complex problem worldwide that the scientific community has extensively studied. This problem is not only related to biological aspects of microorganisms but also other aspects, including socioeconomic and governance factors of countries [23]. Even, some authors claim that a novel approach to antibiotic discovery would be based on the analysis of microbial consortia in their ecological context (see [24,25]). Others discuss the potential of microbial interactions to target and improve microbial dysbiosis as a strategy for the prevention or treatment of cancer [26]. Additionally, some of them affirm that exposure to sublethal concentrations of antimicrobials can indeed alter microbial metabolism and even change the behavior in beneficial ways, triggering reactions such as fleeing or hiding within the protective environment of a microbial aggregate [27].

    Therefore, research questions are left open. A fundamental challenge will be to generate more laboratory experiments to obtain more data to allow better adjustments to the model, and probably adjust the model assuming additional hypotheses.

    J. Romero thanks the support of Fundación Ceiba, Colombia. We thank A. San Millan for the plasmid p-BGT. D. Reyes-Gonzalez is a doctoral student in Programa de Doctorado en Ciencias Biomédicas, Universidad Nacional Autónoma de México, and received fellowship 572373 from CONACYT. A. Fuentes-Hernandez was funded by PAPIIT-UNAM (grant IN215920).

    No potential conflict of interest was reported by the authors.

    By the Pontryagin principle, we can guarantee the existence of adjoint variables zi, i=1,2,3,4 that satisfy:

    ˙zi=dzidt=Hxizi(T)=0,i=1,2,3,4H(X(t),μμ(t),Z(t),t)=maxμiUH(X(t),μμ(t),Z(t),t).

    From above, the adjoint system can be written as:

    ˙z1=HSx,z1(T)=0˙z3=HSy,z3(T)=0˙z2=HRx,z2(T)=0˙z4=HRy,z4(T)=0.

    By doing the respective calculations in the previous equations, we obtain System (2.4). Now, the optimality condition for the Hamiltonian is H/μμ, or equivalently:

     dHdμ1=qcSx(z1z2)+d1μ1dHdμ2=qcSy(z3z4)+d2μ2.

    From the above, we obtain the characterization given on (2.5). In consequence, μ1 satisfies:

    u1={1ifqcSx(z2z1)d1>0qcSx(z2z1)d1ifqcSx(z2z1)d110ifqcSx(z2z1)d1<0,

    or equivalently:

    μ1=min{max{0,qcSx(z2z1)d1},1}.

    Similar calculations can be done for μ2, and then we obtain the characterization given on Eq (2.5) which completes the proof.



    [1] T. A. Hoek, K. Axelrod, T. Biancalani, E. A. Yurtsev, J. Liu, J. Gore, Resource availability modulates the cooperative and competitive nature of a microbial cross-feeding mutualism, PLoS Biol., 14 (2016), e1002540. https://doi.org/10.1371/journal.pbio.1002540 doi: 10.1371/journal.pbio.1002540
    [2] E. Toby Kiers, T. M. Palmer, A. R. Ives, J. F. Bruno, J. L. Bronstein, Mutualisms in a changing world: an evolutionary perspective, Ecol. Lett., 13 (2010), 1459–1474. https://doi.org/10.1111/j.1461-0248.2010.01538.x doi: 10.1111/j.1461-0248.2010.01538.x
    [3] A. R. Figueiredo, R. Kümmerli, Microbial mutualism: Will you still need me, will you still feed me?, Curr. Biol., 30 (2020), R1041–R1043. https://doi.org/10.1016/j.cub.2020.07.002 doi: 10.1016/j.cub.2020.07.002
    [4] K. Zengler, L. S. Zaramela, The social network of microorganisms - how auxotrophies shape complex communities, Nat. Rev. Microbiol., 16 (2018), 383–390. https://doi.org/10.1038/s41579-018-0004-5 doi: 10.1038/s41579-018-0004-5
    [5] W. M. Johnson, H. Alexander, R. L. Bier, D. R. Miller, M. E. Muscarella, K. J. Pitz, et al., Auxotrophic interactions: a stabilizing attribute of aquatic microbial communities?, FEMS Microbiol. Ecol., 96 (2020), fiaa115. https://doi.org/10.1093/femsec/fiaa115 doi: 10.1093/femsec/fiaa115
    [6] X. Jiang, C. Zerfaß, S. Feng, R. Eichmann, M. Asally, P. Schäfer, et al., Impact of spatial organization on a novel auxotrophic interaction among soil microbes, ISME J., 12 (2018), 1443–1456. https://doi.org/10.1038/s41396-018-0095-z doi: 10.1038/s41396-018-0095-z
    [7] X. Zhu, S. Campanaro, L. Treu, R. Seshadri, N. Ivanova, P. G. Kougias, et al., Metabolic dependencies govern microbial syntrophies during methanogenesis in an anaerobic digestion ecosystem, Microbiome, 8 (2020), 22. https://doi.org/10.1186/s40168-019-0780-9 doi: 10.1186/s40168-019-0780-9
    [8] A. E. Douglas, The microbial exometabolome: ecological resource and architect of microbial communities, Philos. Trans. R. Soc. Lond. B Biol. Sci., 375 (2020), 20190250, https://doi.org/10.1098/rstb.2019.0250 doi: 10.1098/rstb.2019.0250
    [9] A. Dal Co, C. Brannon, M. Ackermann, Division of labor in bacteria, Elife, 7 (2018), e38578. https://doi.org/10.7554/eLife.38578 doi: 10.7554/eLife.38578
    [10] G. D'Souza, C. Kost, Experimental evolution of metabolic dependency in bacteria, PLoS Genet., 12 (2016), e1006364. https://doi.org/10.1371/journal.pgen.1006364 doi: 10.1371/journal.pgen.1006364
    [11] M. A. Henson, P. Phalak, Suboptimal community growth mediated through metabolite crossfeeding promotes species diversity in the gut microbiota, PLoS Comput. Biol., 14 (2018), e1006558. https://doi.org/10.1371/journal.pcbi.1006558 doi: 10.1371/journal.pcbi.1006558
    [12] W. H. Fleming, R. W. Rishel, Deterministic and stochastic optimal control, Springer Science and Business Media, 2012.
    [13] S. Lenhart, J. T. Workman, Optimal control applied to biological models, Chapman and Hall/CRC, 2007.
    [14] H. Mena, L. M. Pfurtscheller, J. P. Romero-Leiton, Random perturbations in a mathematical model of bacterial resistance: Analysis and optimal control, Math. Biosci. Eng., 17 (2020), 4477–4499, https://doi.org/10.3934/mbe.2020247 doi: 10.3934/mbe.2020247
    [15] T. Baba, T. Ara, M. Hasegawa, Y. Takai, Y. Okumura, M. Baba, et al., Construction of escherichia coli k-12 in-frame, single-gene knockout mutants: the keio collection, Mol. Syst. Biol., 2, https://doi.org/10.1038/msb4100050
    [16] A. San Millan, J. A. Escudero, D. R. Gifford, D. Mazel, R. C. MacLean, Multicopy plasmids potentiate the evolution of antibiotic resistance in bacteria, Nat. Ecol. Evol., 1 (2016), 1–8. https://doi.org/10.1038/s41559-016-0010 doi: 10.1038/s41559-016-0010
    [17] O. Stojanović, J. Leugering, G. Pipa, S. Ghozzi, A. Ullrich, A Bayesian Monte Carlo approach for predicting the spread of infectious diseases, PLoS ONE, 14 (2019), e0225838. https://doi.org/10.1371/journal.pone.0225838 doi: 10.1371/journal.pone.0225838
    [18] T. Luzyanina, G. Bocharov, Markov chain Monte Carlo parameter estimation of the ODE compartmental cell growth model, Math. Biol. Bioinfor., 13 (2018), 376–391.
    [19] G. Brown, A. Porter, J. Oleson, J. Hinman, Approximate Bayesian computation for spatial SEIR(S) epidemic models, Spat. Spatiotemporal Epidemiology, 24 (2018), 2685–2697, https://doi.org/10.1016/j.sste.2017.11.001 doi: 10.1016/j.sste.2017.11.001
    [20] E. Ibarguen-Mondragon, K. Prieto, S. Hidalgo-Bonilla, A model on bacterial resistance considering a generalized law of mass action for plasmid replication, J. Biol. Syst., 29 (2021), 375–412. https://doi.org/10.1142/S0218339021400118 doi: 10.1142/S0218339021400118
    [21] K. Prieto, J. P. Romero–Leiton, Current forecast of HIV/AIDS using Bayesian inference, Nat. Resour. Model., 34 (2021), e12332, https://doi.org/10.1111/nrm.12332 doi: 10.1111/nrm.12332
    [22] B. Carpenter, A. Gelman, D. Hoffman, B. Goodrich, M. Betancourt, M. Brubaker, et. al., Stan: A probabilistic programming language, J. Stat. Softw., 76 (2017), 1–32. https://doi.org/10.18637/jss.v076.i01 doi: 10.18637/jss.v076.i01
    [23] J. Riaño-Moreno, J. P. Romero-Leiton, K. Prieto, Contribution of governance and socioeconomic factors to the P. aeruginosa MDR in Europe, Antibiotics, 11 (2022), 212. https://doi.org/10.3390/antibiotics11020212 doi: 10.3390/antibiotics11020212
    [24] T. Netzker, M. Flak, M. K. Krespach, M. C. Stroe, J. Weber, V. Schroeckh, et. al., Microbial interactions trigger the production of antibiotics, Curr. Opin. Microbiol., 45 (2018), 117–123. https://doi.org/10.1016/j.mib.2018.04.002 doi: 10.1016/j.mib.2018.04.002
    [25] C. Zhang, P. D. Straight, Antibiotic discovery through microbial interactions, Curr. Opin. Microbiol., 51 (2019), 64–71, https://doi.org/10.1016/j.mib.2019.06.006 doi: 10.1016/j.mib.2019.06.006
    [26] T. Van Raay, E. Allen-Vercoe, Microbial interactions and interventions in colorectal cancer, Microbiol. Spectr., 5 (2017). https://doi.org/10.1128/microbiolspec.BAD-0004-2016 doi: 10.1128/microbiolspec.BAD-0004-2016
    [27] W. C. Ratcliff, R. F. Denison, Alternative actions for antibiotics, Science, 332 (2011), 547–548. https://doi.org/10.1126/science.1205970 doi: 10.1126/science.1205970
  • This article has been cited by:

    1. Jhoana P. Romero-Leiton, Alissen Peterson, Pablo Aguirre, Kamal Acharya, Bouchra Nasri, Assessing the impact of mutations and horizontal gene transfer on the antimicrobial resistance and its control: a mathematical model, 2025, 44, 2238-3603, 10.1007/s40314-024-03043-4
    2. Jhoana P. Romero-Leiton, Alissen Peterson, Pablo Aguirre, Carlos Bastidas-Caldes, Bouchra Nasri, Dynamics of AMR beyond a single bacterial strain: Revealing the existence of multiple equilibria and immune system-dependent transitions, 2025, 191, 09600779, 115912, 10.1016/j.chaos.2024.115912
  • Reader Comments
  • © 2022 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2375) PDF downloads(113) Cited by(2)

Figures and Tables

Figures(13)  /  Tables(6)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog