Review Special Issues

Data-driven biomarker analysis using computational omics approaches to assess neurodegenerative disease progression

  • Received: 29 November 2020 Accepted: 07 February 2021 Published: 22 February 2021
  • The complexity of biological systems suggests that current definitions of molecular dysfunctions are essential distinctions of a complex phenotype. This is well seen in neurodegenerative diseases (ND), such as Alzheimer's disease (AD) and Parkinson's disease (PD), multi-factorial pathologies characterized by high heterogeneity. These challenges make it necessary to understand the effectiveness of candidate biomarkers for early diagnosis, as well as to obtain a comprehensive mapping of how selective treatment alters the progression of the disorder. A large number of computational methods have been developed to explain network-based approaches by integrating individual components for modeling a complex system. In this review, high-throughput omics methodologies are presented for the identification of potent biomarkers associated with AD and PD pathogenesis as well as for monitoring the response of dysfunctional molecular pathways incorporating multilevel clinical information. In addition, principles for efficient data analysis pipelines are being discussed that can help address current limitations during the experimental process by increasing the reproducibility of benchmarking studies.

    Citation: Marios G. Krokidis, Themis P. Exarchos, Panagiotis Vlamos. Data-driven biomarker analysis using computational omics approaches to assess neurodegenerative disease progression[J]. Mathematical Biosciences and Engineering, 2021, 18(2): 1813-1832. doi: 10.3934/mbe.2021094

    Related Papers:

    [1] Agustín Gabriel Yabo, Jean-Baptiste Caillau, Jean-Luc Gouzé . Optimal bacterial resource allocation: metabolite production in continuous bioreactors. Mathematical Biosciences and Engineering, 2020, 17(6): 7074-7100. doi: 10.3934/mbe.2020364
    [2] Shahab Shamshirband, Javad Hassannataj Joloudari, Sahar Khanjani Shirkharkolaie, Sanaz Mojrian, Fatemeh Rahmani, Seyedakbar Mostafavi, Zulkefli Mansor . Game theory and evolutionary optimization approaches applied to resource allocation problems in computing environments: A survey. Mathematical Biosciences and Engineering, 2021, 18(6): 9190-9232. doi: 10.3934/mbe.2021453
    [3] Zehan Liu, Daoxin Qiu, Shengqiang Liu . A two-group epidemic model with heterogeneity in cognitive effects. Mathematical Biosciences and Engineering, 2025, 22(5): 1109-1139. doi: 10.3934/mbe.2025040
    [4] Xuyin Wang, Weiguo Liu, Lu Li, Peizhen Zhao, Ruifeng Zhang . Resource dependent scheduling with truncated learning effects. Mathematical Biosciences and Engineering, 2022, 19(6): 5957-5967. doi: 10.3934/mbe.2022278
    [5] Yanpei Liu, Yunjing Zhu, Yanru Bin, Ningning Chen . Resources allocation optimization algorithm based on the comprehensive utility in edge computing applications. Mathematical Biosciences and Engineering, 2022, 19(9): 9147-9167. doi: 10.3934/mbe.2022425
    [6] Xinru Zhou, Xinmiao Rong, Meng Fan, Josué-Antonio Nescolarde-Selvaa . Stoichiometric modeling of aboveground-belowground interaction of herbaceous plant. Mathematical Biosciences and Engineering, 2019, 16(1): 25-55. doi: 10.3934/mbe.2019002
    [7] Semu Mitiku Kassa . Three-level global resource allocation model for HIV control: A hierarchical decision system approach. Mathematical Biosciences and Engineering, 2018, 15(1): 255-273. doi: 10.3934/mbe.2018011
    [8] Yafei Li, Yuxi Liu . Multi-airport system flight slot optimization method based on absolute fairness. Mathematical Biosciences and Engineering, 2023, 20(10): 17919-17948. doi: 10.3934/mbe.2023797
    [9] Ali Gharouni, Lin Wang . Modeling the spread of bed bug infestation and optimal resource allocation for disinfestation. Mathematical Biosciences and Engineering, 2016, 13(5): 969-980. doi: 10.3934/mbe.2016025
    [10] Yu Shen, Hecheng Li . A new differential evolution using a bilevel optimization model for solving generalized multi-point dynamic aggregation problems. Mathematical Biosciences and Engineering, 2023, 20(8): 13754-13776. doi: 10.3934/mbe.2023612
  • The complexity of biological systems suggests that current definitions of molecular dysfunctions are essential distinctions of a complex phenotype. This is well seen in neurodegenerative diseases (ND), such as Alzheimer's disease (AD) and Parkinson's disease (PD), multi-factorial pathologies characterized by high heterogeneity. These challenges make it necessary to understand the effectiveness of candidate biomarkers for early diagnosis, as well as to obtain a comprehensive mapping of how selective treatment alters the progression of the disorder. A large number of computational methods have been developed to explain network-based approaches by integrating individual components for modeling a complex system. In this review, high-throughput omics methodologies are presented for the identification of potent biomarkers associated with AD and PD pathogenesis as well as for monitoring the response of dysfunctional molecular pathways incorporating multilevel clinical information. In addition, principles for efficient data analysis pipelines are being discussed that can help address current limitations during the experimental process by increasing the reproducibility of benchmarking studies.



    Hepatitis C is a liver inflammatory disease caused by a viral infection. Hepatitis C virus (HCV) is an RNA-type virus from the Flaviviridae family (genus Hepacivirus) which has a high replication rate. Hepatitis can develop into liver cancer [1] and can cause serious complications even death [2,3]. Around 60 – 85% of acute hepatitis C may develop into a chronic condition, 10–15% may develop cirrhosis, and 25% may develop liver cancer [1,2,4]. It is estimated that 58 million people have a chronic hepatitis C, with 1.5 million new infections occurring per year over the world[5]. There is no vaccine available for HCV [3,4,6,7], thus it is important to gain a better insight into the nature of this disease.

    One of the accurate tools to understand the dynamics of hepatitis C is using a mathematical model. There are several mathematical models developed for HCV for example in [8,9] studying the HCV model in population level. The HCV model incorporating treatment, therapy, or isolation were investigated in [10,11,12,13,14]. In reference [15], the optimal control strategies for HCV epidemics considering the uncertainty of the model was discussed.

    The spread of HCV at the cellular level was also investigated by several researchers e.g., [16,17,18,19,20]. In 1998, Neumann et al. studied the dynamic of HCV and the effect of antiviral. They have found that HCV is highly dynamic. An existing deterministic model cannot be applied to describe randomness in many biological factors, for instance, the random occurrence of cell infection, mutation, and apoptosis [21,22]. Furthermore, during the spread of HCV, different cells and infectious virus particles reacting in the same environment can give different effects. In other words, some uncertain factors influence the spread of HCV in body cells such as lifestyle (alcohol consumption or smoking), patient compliance level, lipid metabolism, metabolic syndrome, and body weight [23]. Motivated by this phenomenon, we are interested in extending the mathematical model given in [24] by considering uncertainty factors. A stochastic model will be derived from the deterministic model. This model provides better insight into the uncertainty and variability of the disease dynamic. Moreover, the solution of the stochastic model is in the form of distribution[25], while the solution of the deterministic model only produces one predictive value [26]. Several papers also discussed the stochastic HCV model, e.g., [27,28,29,30,31,32]. In contrast, in this paper, the noise parameter is added in the transmission rate representing the characteristic of variability and the treatment in the model is considered.

    The outline of this paper is organized as follows. In Section 2, the HCV model involving stochastic disturbances in the transmission rate are developed. In addition, the existence, uniqueness, and boundedness of the solution are established. In Section 3, we derive the extinction and persistence in mean condition. In Section 4, the numerical simulations are carried out to illustrate the analytical results. The conclusions are presented in the last section.

    Recall a deterministic mathematical model of HCV given in [24] as

    dTdt=Λδ1T(1η)βVTdIdt=(1η)βVTδ2IdVdt=(1ϵ)kIcV. (2.1)

    where T is the number of uninfected cells, I is the number of infected cells, and V is the number of free viruses. The uninfected cells are produced at rate Λ and die naturally at a constant rate δ1. Cells become infected when they are interacting with a virus with a constant rate β, and once infected, they will die at a constant rate δ2. HCV is produced by infected cells at a constant rate k and cleared at a constant rate c.

    For the deterministic model (2.1), the disease-free equilibrium point is E0=(Λδ1,0,0) and the endemic equilibrium point is E1=(δ2c(1η)(1ϵ)kβ,Λδ2δ1c(1η)(1ϵ)kβ,(1ϵ)kΛδ2cδ1(1η)β). Using the next-generation matrix method [33], the basic reproduction number of the system (2.1) is Rd0=(1ϵ)(1η)βkΛcδ1δ2.

    In this paper, we generalize system (2.1) by incorporating a stochastic noise parameter in the transmission rate as

    {dT=(Λδ1T(1η)βVT)dtσ(1η)VTdB(t)dI=((1η)βVTδ2I)dt+σ(1η)VTdB(t)dV=((1ϵ)kIcV)dt (2.2)

    where B(t) is a standard Brownian motion and σ is a real constant which is known as the intensity of noise. The value σ is the standard deviation of transmission rate data that represents the variability of the transmission rate. The other description of parameters model (2.2) is given in Table 1. Clearly, the system (2.1) is the special case of the system (2.2), where σ=0.

    Table 1.  Variables and parameters of the model.
    Variable/Parameter Description Unit Range
    T(t) concentration of uninfected liver cells cell/mL 0
    I(t) concentration of infected liver cells cell/mL 0
    V(t) concentration of free viruses virion/mL 0
    Λ rate of production cell/day×mL 0
    η the effectiveness of drug in stopping infection [0, 1]
    k rate of free virus production virion/cell×day 0
    ϵ the effectiveness of drug in blocking virus [0, 1]
    β transmission rate mL/virion × day 0
    δ1 death rate of uninfected liver cells 1/day 0
    δ2 death rate of infected liver cells 1/day 0
    c virion clearance 1/day 0

     | Show Table
    DownLoad: CSV

    In this section, the theorem of the existence, uniqueness, and boundedness of the solution system (2.2) are established. Let (Ω,F,P) be a complete probability space with a filtration {Ft}t0 satisfying the conditions that it is right continuous and F0 contains all P-null sets.

    Definition 2.1. [34]

    Let T>0, F(.,.):[0,T]×RnRn,G(.,.):[0,T]×RnRn×m be measurable function, and X(t) satisfy

    dX(t)=F(X(t))dt+G(X(t))dB(t), (2.3)

    where F(., .) and G(., .) are the coefficients of (2.3). Then the coefficients (2.3) are locally Lipschitz, if

    |F(t,X)F(t,ˉX)|+|G(t,X)G(t,ˉX)|C1|XˉX|;X,ˉXRn,t[0,T], (2.4)

    for some constant C10.

    Definition 2.2. [34]

    The coefficients of (2.3) satisfy linear growth condition, if

    |F(t,X)|+|G(t,X)|C2(1+|X|);XRn,t[0,T], (2.5)

    for some constant C2.

    Lemma 2.3. The system (2.2) with the initial condition in R3+ is uniformly ultimately bounded and belongs to the following closed and bounded positively invariant set

    Γ={(T,I,V)R3+|0<T+IΛμ,0<V(1ϵ)(kΛcμ)}R3+, (2.6)

    for every t0.

    Proof. Let a(t) be a function such that a(t)Λ. We know that N(t)=T(t)+I(t) is the concentration of uninfected and infected cells at time t. Take μ=min{δ1,δ2}, then

    dN=dT+dI,=(Λδ1Tδ2I)dt,dN(ΛμN)dt. (2.7)

    Let dN=(a(t)μN)dt, it follows that

    dN=(a(t)μN)dt,dNdt+μN=a(t),

    thus,

    N(t)=eμt(t0eμτa(τ)dτ+N0). (2.8)

    Since a(t)Λ, we obtain

    Neμt(t0eμτΛdτ+N0),=Λμ+(N0Λμ)eμt. (2.9)

    By taking a limit t of Eq (2.9), we get

    N(t)Λμ.

    Analog with the technique for solving Eq (2.7), for V(0)=V0 and I(t)<N(t)Λμ, we get

    dV(t)=[(1ϵ)kIcV]dt,[(1ϵ)(kΛμ)cV]dt,

    then

    V(t)(1ϵ)(kΛcμ)+(V0(1ϵ)(kΛμ))ect. (2.10)

    If we take limtV(t), then

    V(t)(1ϵ)(kΛcμ).

    Since the right-hand side of Eq (2.2) satisfies the Lipschitz condition, the solution exists and unique on [0,b) for some b>0. Assume that there exists t1(0,b) such that V(t1)=0 and all other variables are positive on t1(0,b). Therefore, for all t[0,t1]

    dV=((1ϵ)kIcV)dt,cVdt,

    thus

    V(t1)C3ect1>0,whereC3positiveconstant. (2.11)

    This is contradiction with V(t1)=0. Then V(t)>0. Analoguely, we get I(t)>0 and T(t)>0. Therefore, we obtain that (T(t),I(t)(0,Λμ) for all t[0,T] and 0<V(t)(1ε)(kΛcμ).

    Next, we define the necessary condition that guarantees the existence and uniqueness of time-global solution of Eq (2.2).

    Theorem 2.4. Let the coefficients of the system (2.2) satisfy the Lipschitz condition. Then for any initial value (T(0),I(0),V(0))Γ, there exists a unique time-global solution (T(t),I(t),V(t))ΓR3+,t0 with probability 1.

    Proof. According to Definition 2.1, Definition 2.2, and Theorem 5.2.1 in [34], there exists a unique global solution. However, we have that the coefficients of the system (2.2) only satisfy the Lipschitz condition [35], thus system (2.2) has a unique local solution on t[0,τe) for any initial value (T(0),I(0),V(0))Γ where τe is the explosion time (i.e. the time when the solution tends to infinity). To guarantee the solution of system (2.2) is a unique global solution, it is necessary to show that τe= [36,37].

    Let k0>0 be sufficiently large such that every component of (T(0),I(0),V(0)) is in the interval [1k0,k0]. For every kk0, we define the stopping time or the first passage time (i.e. first period when the stochastic process penetrates the barrier) as

    τk=inf{t[0,τe)|T(t)(1k,k),I(t)(1k,k)orV(t)(1k,k)}.

    Throughout this paper, we set inf=. It is known that the lower bound of R is an empty set and the largest infimum of the empty set is infinity. Since τk is increasing as k, and

    τ0τ1τ2...τkτk+1...,

    then it follows that

    τ=limkτk,

    thus ττe almost sure. Next, it is necessary to show

    limkτk=.

    We will prove by contradiction. Suppose that P(τ<)<1. If τ<, then there exist T>0 and ε(0,1) such that P{τkT}>ε,kk0. In this case, the proof technique analogue is to [37,38,39]. Define a function C2,1,Q:[0,)×R3+R+ where

    Q(T,I,V)=(T1lnT)+(I1lnI)+(V1lnV).

    Since

    y1lny0,y>0,

    then Q is a non-negative function. By using Itˆo's formula, we get

    dQ=Qtdt+QTdT+QIdI+QVdV+(12)(2QT2(dT)2)+(12)(2QI2(dI)2)+(12)(2V2(dV)2)+2QTIdTdI+2QTVdTdV+2QIVdIdV,=(11T){[Λδ1T(1η)βVT]dtσ(1η)VTdB(t)}+(11I){[(1η)βVTδ2I]dt+σ(1η)VTdB(t)}+(11V){[(1ϵ)kIcV]dt}+12(σVT)2(1η)2(1T2+1I2)(dB(t))2,{Λ+δ1+δ2+(1ϵ)kI+(1η)βV+c+12(σVT)2(1η)2(1T2+1I2)}dt+{σ(1η)VT(1T1I)}dB(t),Mdt+σ(1η)VT(1T1I)dB(t), (2.12)

    where M is a positive constant. Integrating Eq (2.12) from 0 to τkT yields

    Q(T(τkT),I(τkT),V(τkT))Q(T(0),I(0),V(0))Mt+τkT0σ(1η)VT(1T1I)dB(t). (2.13)

    Taking the expectation of both sides in Eq (2.13) leads to

    E(Q(T(τkT),I(τkT),V(τkT)))E(Q(T(0),I(0),V(0)))+E(Mt)+E(τkT0σ(1η)VT(1T1I)dB(t)),
    Q(T(0),I(0),V(0))+Mt. (2.14)

    Define Ωk={τkT} for any kk0. Then for every wΩk there are components of Q(T(τkT),I(τkT),V(τkT)) which are equal to either k or 1k, thus

    Q(T(τkT),I(τkT),V(τkT))min{k1lnk,1k1ln(1k)}. (2.15)

    Combining Eqs (2.14) and (2.15) yields

    Q(T(0),I(0),V(0))+MtE(Q(T(τkT,w),I(τkT,w),V(τkT,w))),=E(1Ωk(w).Q(T(τk,w),I(τk,w),V(τk,w))),P(Ωk)min{k1lnk,1k1ln(1k)},εmin{k1lnk,1k1ln(1k)}, (2.16)

    where 1Ωk is the indicator function. By taking a limit k of Eq (2.16), we obtain

    =Q(T(0),I(0),V(0))+Mt<.

    This is contradiction. We get τ= or P(τ=)=1. Therefore, there exists a unique time-global solution (T(t),I(t),V(t))ΓR3+,t0 with probability 1.

    In this section, the theorem of the extinction and persistence in the mean condition system (2.2) is derived.

    In this part, we investigate the almost surely exponential stability of the disease-free equilibrium point by using the suitable Lyapunov function and another technique of stochastic analysis.

    Lemma 3.1. If

    σ>β2(δ2(1ϵ)kc), (3.1)

    the disease-free equilibrium point (T,I,V)=(Λ/δ1,0,0) is almost surely exponentially stable in Γ.

    Proof. The technical proof of this theorem follows from [38,40]. Define a function Q=ln(I+V). Using Itˆo's formula, we obtain

    dQ=Qtdt+QTdT+QIdI+QVdV+12[2QT2dTdT+2QI2dIdI+2QV2dVdV]+2QTIdTdI+2QTVdTdV+2QIVdIdV,=1I+V[((1η)βVTδ2I)dt+σ(1η)VTdB]+1I+V[(1ϵ)kIcV]dt+12[1I+V)2(σ2(1η)2(VT)2)dt],={(1η)βZδ2II+VcVI+V+(1ϵ)kII+V12σ2(1η)2Z2}dt+σ(1η)ZdB,

    where Z=VTI+V. Futhermore, we have

    dQ12σ2(1η)2[Zβδ2(1η)]2dt+β22σ2dt[δ2(1ϵ)kc]dt+σ(1η)ZdB,β22σ2dt[δ2(1ϵ)kc]dt+σ(1η)ZdB. (3.2)

    Taking the integral of (3.2) and dividing both sides by t and then computing the limit superior t yield

    limtsup1tln(I(t)+V(t))limtsup1t{ln(I(0)+V(0))+β22σ2[δ2(1ϵ)kc]}+limtsupt01tσ(1η)ZdB. (3.3)

    By the strong law of large number for Martingales [41], we have

    limtsupt01tσ(1η)ZdB=0. (3.4)

    Thus, from Eqs (3.3) and (3.4), we obtain

    limtsup1tln(I(t)+V(t))limtsup1t{ln(I(0)+V(0))+β22σ2[δ2(1ϵ)kc]}<0.

    By Lemma 3.1, if the noise is increasing and satisfies the condition (2.10) HCV will die out. However, in this study, we also get that a bounded variation of infection rate could also lead to extinction. This is presented in the following theorem.

    Theorem 3.2. Let (T(0),I(0),V(0))Γ be the initial value of system (2.2) and (T(t),I(t),V(t))Γ be the corresponding solution. If

    Rs0=(1ϵ)(1η)βkΛcδ1δ2[14cδ21δ22(1ϵ)2(1η)2k2Λ2σ2]=Rd0[14cδ21δ22(1ϵ)2(1η)2k2Λ2σ2]<1

    and σ2<4δ1δ2β(1ϵ)(1η)kΛ, then HCV will be extinct almost surely, i.e.,

    limtT(t)=Λδ1,limtI(t)=0,limtV(t)=0.

    Proof. The proof technique follows from [38,40,42,43]. Consider the first equation of system (2.2), then we get

    T(t)T(0)=Λtt0(1η)βV(s)T(s)dst0δ1T(s)dst0σ(1η)V(s)T(s)dB(s),

    or

    1tt0T(s)ds=Λδ1(T(t)T(0))δ1t1δ1tt0(1η)βV(s)T(s)ds1δ1tt0σ(1η)V(s)T(s)dB(s).

    For simplicity, we define an integrable function X=1tt0X(s)ds, X(t)[0,+), thus

    T=Λδ1(T(t)T(0))δ1t1δ1(1η)βVT1δ1tt0σ(1η)V(s)T(s)dB(s). (3.5)

    Taking the limit of Eq (3.5) as t tend to infinity, we obtain

    limtT=1δ1limt[Λ(T(t)T(0))t(1η)βVT1tt0σ(1η)V(s)T(s)dB(s)],1δ1limt[Λ1tt0σ(1η)V(s)T(s)dB(s)],Λδ1.

    Next, define a function Q=ln((1ϵ)kI+δ2V). Applying Ito's formula, we obtain

    dQ=Qtdt+QTdT+QIdI+QVdV+12[2QT2dTdT+2QI2dIdI+2QV2dVdV]+2QTIdTdI+2QTVdTdV+2QIVdIdV,=(1ϵ)k(1ϵ)kI+δ2V[((1η)βVTδ2I)dt+σ(1η)VTdB]+δ2(1ϵ)kI+δ2V[(1ϵ)kIcV]dt+12[1((1ϵ)kI+δ2V)2(1ϵ)2k2(σ2(1η)2(VT)2)dt],[(1ϵ)k(1η)βTδ2c12δ2212(1ϵ)2k2σ2(1η)2T2]dt+(1ϵ)k(1ϵ)kI+δ2Vσ(1η)VTdB. (3.6)

    Integrating both sides of Eq (3.6) from 0 to t yields

    ln((1ϵ)kI+δ2V)ln((1ϵ)kI(0)+δ2V(0))+(1ϵ)kδ2(1η)βt0T(s)dsct14δ22(1ϵ)2k2σ2(1η)2t0(T(s))2ds+t0(1ϵ)k(1ϵ)kI(s)+δ2V(s)σ(1η)VTdB(s). (3.7)

    Dividing both sides of Eq (3.7) by t and taking a limit superior for t0, we get

    limtsup1tln((1ϵ)kI+δ2V)limtsup1t[ln((1ϵ)kI(0)+δ2V(0))+(1ϵ)kδ2(1η)βt0T(s)ds]climtsup14δ22(1ϵ)2k2σ2(1η)2t0(T(s))2ds+t0(1ϵ)k(1ϵ)kI(s)+δ2V(s)σ(1η)VTdB(s).

    By the strong law of large number for Martingales [41],

    limtsup1tt0(1ϵ)k(1ϵ)kI(s)+δ2V(s)σ(1η)VTdB(s)=0,

    we obtain

    limtsup1tln((1ϵ)kI+δ2V)(1ϵ)kδ1δ2(1η)βΛc14δ22(1ϵ)2k2σ2(1η)2(Λ)2(δ1)2,c{[(1ϵ)(1η)βkΛcδ1δ214cδ22δ21(1ϵ)2(1η)2k2Λ2σ2]1},c[Rs01]<0.

    Thus, limtI(t)=0 and limtV(t)=0. In other words, HCV will be extinct.

    In this subsection, we give a sufficient condition to guarantee the persistence in the mean condition.

    Definition 3.3. [31,40,44]

    System (2.2) is said to be persistent in mean if

    liminft1tt0[I(s)+V(s)]ds>0

    almost surely.

    Theorem 3.4. Let (T(0),I(0),V(0))Γ be the initial value of system (2.2) and (T(t),I(t),V(t))Γ be the corresponding solution. If

    Rs1=(1ϵ)(1η)βkΛcδ1δ2[12cδ21(1+δ2)2(1ϵ)2(1η)2k2Λ2σ2]=Rd0[12cδ21(1+δ2)2(1ϵ)2(1η)2k2Λ2σ2]>1,

    then HCV will be persistent in mean, i.e.,

    liminft1tt0{I(s)+V(s)}dsδ2/(δ2c)(1η)(1ϵ)βΛk[δ1c(Rs11)].

    Proof. Analogue to the proof technique in [40] and [44]. Integrating the system (2.2) yields

    T(t)T(0)+I(t)I(0)+V(t)V(0)=Λtδ1t0T(s)dsδ2t0I(s)ds+(1ϵ)kt0I(s)dsct0V(s)ds (3.8)

    Dividing both sides Eq (3.8) by t, we get

    T(t)T(0)t+I(t)I(0)t+V(t)V(0)t=Λδ1tt0T(s)dsδ2tt0I(s)ds+(1ϵ)ktt0I(s)dsctt0V(s)ds1tt0T(s)ds=Λδ1δ2δ1tt0I(s)ds+(1ϵ)kδ1tt0I(s)dscδ1tt0V(s)dsΦ (3.9)

    where

    Φ(t)=T(t)T(0)t+I(t)I(0)t+V(t)V(0)t.

    Define a function Q=ln[(1ϵ)kI(t)+(1+δ2)V(t)]. Applying Itˆo's formula, we obtain

    dQ=(1ϵ)k[((1η)βVTδ2I)dt+σ(1η)VTdB](1ϵ)kI+(1+δ2)V+(1+δ2)((1ϵ)kIcV)dt(1ϵ)kI+(1+δ2V)+12[(1η)2(1ϵ)2k2σ2(VT)2[(1ϵ)kI+(1+δ2)V]2dt],((1ϵ)(1η)kβTδ2c12(1η)2(1ϵ)2k2σ2T2(1+δ2)2)dt+σ(1η)(1ϵ)kVTdB(1ϵ)kI+(1+δ2)V. (3.10)

    Integrating Eq (3.10) from 0 to t and dividing both sides by t, we have

    1tln[(1ϵ)kI(t)+(1+δ2)V(t)]1tln[(1ϵ)kI(0)+(1+δ2)V(0)+1tt0(1ϵ)(1η)kβT(s)dsδ2c1tt012(1η)2(1ϵ)2k2σ2T(s)2ds(1+δ2)2+1tt0σ(1η)(1ϵ)kV(s)T(s)dB(s)(1ϵ)kI+(1+δ2)V. (3.11)

    According to Eqs (3.9) and (3.11), we get

    δ2δ1tt0I(s)ds+cδ1tt0V(s)dsδ2(1η)(1ϵ)βΛk[c(Rs11)Φ(t)Ψ(t)+M(t)] (3.12)

    where

    Rs1=(1ϵ)(1η)βkΛcδ1δ2[12cδ21(1+δ2)2(1ϵ)2(1η)2k2Λ2σ2],Ψ(t)=1tln[(1ϵ)kI(0)+(1+δ2)V(0),M(t)=1tt0σ(1η)(1ϵ)kV(s)T(s)dB(s)(1ϵ)kI+(1+δ2)V.

    Taking the limit inferior of Eq (3.12) as t0 yields

    limtinfδ2δ1tt0I(s)ds+cδ1tt0V(s)dslimtinfδ2(1η)(1ϵ)βΛk[c(Rs11)],limtinf1t[t0I(s)ds+t0V(s)ds]limtinfδ2/(δ2c)(1η)(1ϵ)βΛk[δ1c(Rs11)].

    This completes the proof.

    In this section, we carry out a numerical simulation in order to provide an interpretation of the solution. We use the Euler-Maruyama method to determine the solution of system (2.2). The discretization of system (2.2) is given as follows

    {Ti+1=Ti+(Λδ1Ti(1η)βViTi)Δtσ(1η)ViTiΔtζIi+1=Ii+((1η)βViTiδ2Ii)Δt+σ(1η)ViTiΔtζVi+1=Vi+((1ϵ)kIicVi)Δt (4.1)

    where t[t0,tN],Δt=tNt0N, and ζ is normally distributed N(0, 1).

    Three simulations are conducted. The first simulation is deterministic system. The second simulation is to represent the solution of a stochastic model in the extinction condition. The third simulation is illustrating the solution of a stochastic model when Lemma 3.1 and Theorem 3.2 are violated. The parameters and initial values of the model in each simulation are given in Table 2.

    Table 2.  Values and parameters of the model.
    Sim.1 Sim.2 Sim.3 Unit Reference
    T(0) 106 106 106 cell/mL [45]
    I(0) 105 105 105 cell/mL [45]
    V(0) 105 105 105 virion/mL [45]
    Λ 8×105 8×105 8×105 cell/day×mL [17]
    η 0.7 0.7 0.7 [17]
    k 0.9 0.2 0.9 virion/cell×day [17]
    ϵ 0.75 0.75 0.75 [17]
    β 5.4×108 5.4×108 5.4×107 mL/virion × day [17]
    δ1 0.0047 0.1 0.0047 1/day [17,46]
    δ2 0.3 0.3 0.3 1/day [17]
    c 0.8 1 0.8 1/day [17]
    σ 4×107 9×108

     | Show Table
    DownLoad: CSV

    Figure 1 shows the solution of a deterministic model (σ=0) with Rd0>1. It can be seen that the solution is tend to the endemic equilibrium point E1=(6.6×106,2.6×106,7×105) which means that HCV disease is persistent.

    Figure 1.  Simulation 1: a deterministic model of the system (2.1).

    Figure 2 illustrates the stochastic model with σ=4×107 the condition

    σ=1.6×1013=4×107>β2(δ2(1ϵ)kc)=0.76×107

    and Rs0<1 where σ2=1.6×1013<4δ1δ2β(1ϵ)(1η)kΛ=5.4×1013 satisfies Lemma 3.1 and Theorem 3.2 respectively. More precisely, when the choosen parameters satisfy Lemma 3.1 and Theorem 3.2, the point E0=(8×106,0,0) is almost exponentially stable. It can be seen from Figure 2 where the concentration of infected cells and free viruses go to zero. Therefore, HCV will go extinct almost surely.

    Figure 2.  Simulation 2: a stochastic model of the system (2.2) with σ=4×107, 10 paths.

    Figure 3 describes the condition when basic reproduction number Rs0=0.0214<1 and σ2=4.9×1015<4δ1δ2β(1ϵ)(1η)kΛ=5.4×1013. This is satisfying Theorem 3.2 and violating Lemma 3.1. That figure shows a stochastic noise impact the disease extinction.

    Figure 3.  Simulation 2: a stochastic model of the system (2.2) with σ=7×108, 10 paths.

    Figure 4 shows that if the intensity of the disturbance is smaller where σ2=8.1×1016<1.9×1012, thus Lemma 3.1 is violated. Taking the time interval [0,400] with 4000 points and 10 paths, it appears in Figure 4(a) that the concentration of uninfected cells rises to a peak around at t = 30 days, then drops sharply to around t = 45 days and then they rise slowly. The graphics in Figures 4(b), (c) show the concentration of infected cells and free virus after t = 50 days which tend to certain interval values. This shows that hepatitis C remains exist.

    Figure 4.  Simulation 3: a stochastic model of the system (2.2) with σ=9×108, 10 paths.

    Figure 5 describes the condition of the HCV extinction when Lemma 3.1 is satisfied while Theorem 3.2 is violated. It can be seen that figures show the concentration of infected cells and free virus from initial value that tends to zero for a long period of time.

    Figure 5.  Simulation 3: a stochastic model of the system (2.2) with β=2.5×108, 10 paths.

    Figure 6(a) shows the distribution value of uninfected cells in simulation 2 with a confidence interval of 95 %. After time t about 50 days the concentration of uninfected cells will go to a positive constant value (8×106). From Figure 6(b), the simulation at t = 5 is more diverse than t = 10. After time t 20 days, the concentration of infected cells will go to zero. Figure 6(c) shows the free virus concentration at the time t from 3 to 15 days is in the certain interval value. After time t 20 days, the virus concentration will go to zero. Therefore, from simulation 2, it can be concluded that stochastic disturbances can affect the behavior of the spread of hepatitis C. When conditions are met according to Lemma 3.1 and Theorem 3.2, at a certain time t hepatitis C will disappear. Mathematically, in this case, the basic reproduction number in the extinction condition is less than one.

    Figure 6.  A stochastic model of the system (2.2) with confidence interval on simulation 2.

    Figure 7 shows the confidence intervals for simulation 3 where we reduced the sigma values while preserving the other parameter values. Figure 7(a) shows the concentration of uninfected cells will reach peaks around 23 to 30 days and then decline. After 75 days the concentration will move to a certain interval value. The green lines show the mean value of the simulations. Figures 7(b), (c) show that after 75 days, the concentration of infected cells and free virus lead to a certain positive number interval value with a confidence interval of approximately 95 %. In other words, hepatitis C disease will persist.

    Figure 7.  A stochastic model of the system (2.2) with confidence interval on simulation 2.

    Lastly, we conduct a simulation for persistence in mean condition where β=5.4×107,δ1=0.0047,k=1,c=0.6,σ=9×108, and the rest of parameters and initial values state in Table 2. Figure 8 illustrates the solution of the stochastic model compared to the deterministic. Figure 9 shows the number of concentrations of healthy cells, infected cells, and free virus at a certain value which are more than zero. When t50, it appears that the concentrations of uninfected cells, infected cells and free viruses go to an equilibrium state. According to the numerical calculations, we obtain Rs1>1 and the endemic equilibrium point E1=(T,I,V)=(4.4×106,2.6×106,1.08×106). According to Theorem 3.4 hepatitis C remains to exist which are also depicted in Figure 9. Figure 10 describes the condition if the noise intensity is increasing, then the solution of model (2.2) will be strongly oscillating around the endemic equilibrium point of the model (2.1).

    Figure 8.  A stochastic model of the system (2.2) on persistence in mean simulation compared to a deterministic.
    Figure 9.  A stochastic model of the system (2.2) with σ=9×108, 10 paths on simulation persistence in mean.
    Figure 10.  A stochastic model of the system (2.2) with σ=2×107, 10 paths on simulation persistence in mean.

    Furthermore, in Figure 11, we generate a 95% confidence interval for t[0,400] with 4000 paths. Figure 11(a) describes the number of uninfected cells that tends to a consistent interval value after 75 days. Figure 11(b), (c) show the confidence intervals of infected liver cells and free viruses. In the time t span of 0 to 400 days the number of concentrations in a row increases to a peak around day 27 for infected cells and free viruses, then decreases. After 30 days, leading to a consistent value for infected cells from 1.9×106 to 3.9×106, and free viruses from 0.75×106 to 1.5×106.

    Figure 11.  A stochastic model of the system (2.2) with confidence interval on persistence in mean simulation.

    In this paper, we propose an epidemic model for HCV transmission at the cellular level incorporating statistical noise. The results extend the paper of Neumann, et.al.[24] in understanding the dynamics of a deterministic hepatitis C virus. Various types of Lyapunov functions are designed to study the persistence in mean conditions as well as the extinction of the stochastic system. Based on this model, there exists a unique time-global solution for any given positive initial value. In addition, numerical simulations are carried out to describe the solution behaviour of the model. We analyze that if the noise intensity is increasing, then the disease will go to extinction. If the basic reproduction number of the extinction condition is less than one, then hepatitis C will extinct. If the basic reproduction number of the persistence in the mean condition is more than one, then hepatitis C remains to exist. When the noise intensity increases, the solution will give a strong oscillation under the condition of persistence in the mean. It can provide some useful strategies for controlling the dynamics of that disease. In future research, we can consider an optimal control of the stochastic model for HCV, research on memory effect, and fractional derivatives [47]. Furthermore, it needs to investigate the stochastics noise for other parameters apart from infection rate. It is also important to consider a stochastic model for HCV transmission involving the response immune system.

    The first author thanks LPDP Indonesia for the financial support under the Doctoral Program Scholarship. The authors thanks to Department of Mathematics, Universitas Gadjah Mada (UGM) and Directorate for the Higher Education, Ministry of Research, Technology, and Higher Education of Indonesia for the Research Grant "Penelitian Disertasi Doktor (PDD)" UGM 2021 under grant no. 2261/UN1/DITLIT/DIT-LIT/PT/2021. The special thanks also for some colleagues for the discussions during the research.

    The authors declare that they have no competing financial interest or personal relationships that could have appeared to influence the work reported in this paper.



    [1] C. A. Lane, J. Hardy, J. M Schott, Alzheimer's disease, Eur. J. Neurol., 25 (2017), 59-70.
    [2] H. S. Kwon, S. H. Koh, Neuroinflammation in neurodegenerative disorders: the roles of microglia and astrocytes, Transl. Neurodegener., 9 (2020), 42.
    [3] C. Marogianni, M. Sokratous, E. Dardiotis, G. M. Hadjigeorgiou, D. Bogdanos, G. Xiromerisiou, Neurodegeneration and inflammation—an interesting interplay in Parkinson's disease, Int. J. Mol. Sci., 21 (2020), 8421.
    [4] S. H. Kim, M. Y. Noh, H. J. Kim, K. W. Oh, J. Park, S. Lee, et al., A therapeutic strategy for Alzheimer's disease focused on immune-inflammatory modulation, Dement Neurocogn. Disord., 18 (2019), 33-46. doi: 10.12779/dnd.2019.18.2.33
    [5] Z. Han, R. Tian, P. Ren, W. Zhou, P. Wang, M. Luo, et al., Parkinson's disease and Alzheimer's disease: a Mendelian randomization study, BMC Med. Genet., 19 (2018), 215.
    [6] A. Xie, J. Gao, L. Xu, D. Meng, Shared mechanisms of neurodegeneration in Alzheimer's disease and Parkinson's disease, Biomed. Res. Int., 2014 (2014), 648740.
    [7] J. Montaner, L. Ramiro, A. Simats, S. Tiedt, K. Makris, G. C. Jickling, et al., Multilevel omics for the discovery of biomarkers and therapeutic targets for stroke, Nat. Rev. Neurol., 16 (2020), 247-264.
    [8] W. J. Lukiw, A. Vergallo, S. Lista, H. Hampel, Y. Zhao, Biomarkers for Alzheimer's disease (AD) and the application of precision medicine, J. Pers. Med., 10 (2020), 138.
    [9] E. M. Cilento, L. Jin, T. Stewart, M. Shi, L. Sheng, J. Zhang, Mass spectrometry: A platform for biomarker discovery and validation for Alzheimer's and Parkinson's diseases, J. Neurochem., 151 (2019), 397-416. doi: 10.1111/jnc.14635
    [10] J. Verheijen, K. Sleegers, Understanding Alzheimer disease at the interface between genetics and transcriptomics, Trends Genet., 34 (2018), 434-447. doi: 10.1016/j.tig.2018.02.007
    [11] E. Cuyvers, K. Sleegers, Genetic variations underlying Alzheimer's disease: evidence from genome-wide association studies and beyond, Lancet Neurol., 15 (2016), 857-868. doi: 10.1016/S1474-4422(16)00127-7
    [12] F. L. Heppner, R. M. Ransohoff, B. Becher, Immune attack: The role of inflammation in Alzheimer disease, Nature, 16 (2015), 358-372.
    [13] Y. Yan, A. Zhao, Y. Qui, Y. Li, R. Yan, Y. Wang, et al., Genetic association of FERMT2, HLA-DRB1, CD2AP, and PTK2B polymorphisms with Alzheimer's disease risk in the southern Chinese population, Front. Aging Neurosci., 12 (2020), 16.
    [14] V. Escott-Price, C. Bellenguez, L. S. Wang, S. H. Choi, D. Harold, L. Jones, et al., Gene-wide analysis detects two new susceptibility genes for Alzheimer's disease. PLoS One, 9 (2014), e94661.
    [15] I. López González, P. Garcia-Esparcia, F. Llorens, I. Ferrer, Genetic and transcriptomic profiles of inflammation in neurodegenerative diseases: Alzheimer, Parkinson, Creutzfeldt-Jakob and Tauopathies, Int. J. Mol. Sci., 17 (2016), 206.
    [16] Z. P. Van Acker, M. Bretou, Wim Annaert, Endo-lysosomal dysregulations and late-onset Alzheimer's disease: impact of genetic risk factors, Mol. Neurodegener., 14 (2019), 20.
    [17] A. De Roeck, C. Van Broeckhoven, K. Sleegers, The role of ABCA7 in Alzheimer's disease: evidence from genomics, transcriptomics and methylomics, Acta Neuropathol., 138 (2019), 201-220.
    [18] C. Reitz, Novel susceptibility loci for Alzheimer's disease, Future Neurol., 10 (2015), 547-558. doi: 10.2217/fnl.15.42
    [19] C. Bellenguez, C. Charbonnier, B. Grenier-Boley, O. Quenez, K. Le Guennec, G. Nicolas, et al., Contribution to Alzheimer's disease risk of rare variants in TREM2, SORL1, and ABCA7 in 1779 cases and 1273 controls, Neurobiol. Aging, 59 (2017), 220.e1-220.e9.
    [20] P. Ciryam, R. Kundra, R. Freer, R. I. Morimoto, C. M. Dobson, M. Vendruscolo, A transcriptional signature of Alzheimer's disease is associated with a metastable subproteome at risk for aggregation, Proc. Natl. Acad. Sci. USA, 113 (2016), 4753-4758. doi: 10.1073/pnas.1516604113
    [21] L. V. Kalia, A. E. Lang, Parkinson's disease, Lancet, 386 (2015), 896-912.
    [22] A. Masato, N. Plotegher, D. Boassa, L. Bubacco, Impaired dopamine metabolism in Parkinson's disease pathogenesis, Mol. Neurodegener., 14 (2019), 35.
    [23] M. G. Krokidis, Identification of biomarkers associated with Parkinson's disease by gene expression profiling studies and bioinformatics analysis, AIMS Neurosci., 6 (2019), 333-345. doi: 10.3934/Neuroscience.2019.4.333
    [24] A. Grünewald, K. R. Kumar, C. M. Sue, New insights into the complex role of mitochondria in Parkinson's disease, Prog. Neurobiol., 177 (2019), 79-93.
    [25] C. A. D. Costa, W. E. Manaa, E. Duplan, F. Checler, The endoplasmic reticulum stress/unfolded protein response and their contributions to Parkinson's disease physiopathology, Cells, 9 (2020), 2495.
    [26] M. Deffains, M. H. Canron, M. Teil, Q. Li, B. Dehay, E. Bezard, et al., L-DOPA regulates α-synuclein accumulation in experimental parkinsonism, Neuropathol. Appl. Neurobiol., 2020.
    [27] L. P. Dolgacheva, A. V. Berezhnov, E. I. Fedotova, V. P. Zinchenko, A. Y. Abramov, Role of DJ-1 in the mechanism of pathogenesis of Parkinson's disease, J. Bioenerg. Biomembr., 51 (2019), 175-188. doi: 10.1007/s10863-019-09798-4
    [28] A. Urbizu, K. Beyer, Epigenetics in Lewy body diseases: Impact on gene expression, utility as a biomarker, and possibilities for therapy, Int. J. Mol. Sci., 21 (2020), 4718.
    [29] D. M. O'Hara, G. Pawar, S. K. Kalia, L. V. Kalia, LRRK2 and α-Synuclein: Distinct or synergistic players in Parkinson's disease?, Front. Neurosci., 14 (2020), 577.
    [30] T. K. Lin, K. J. Lin, K. L. Lin, C. W. Liou, S. D. Chen, Y. C. Chuang, et al., When friendship turns sour: effective communication between mitochondria and intracellular organelles in Parkinson's disease, Front. Cell. Dev. Biol., 8 (2020), 607392.
    [31] B. Wang, K. R. Stanford, M. Kundu, ER-to-Golgi trafficking and its implication in neurological diseases, Cells, 9 (2020), 408.
    [32] T. M. Axelsen, D. P. D. Woldbye, Gene therapy for Parkinson's disease, an update, J. Parkinsons Dis., 8 (2018), 195-215.
    [33] T. J. Collier, D. E. Jr. Redmond, K. Steece-Collier, J. W. Lipton, F. P. Manfredsson, Is a-synuclein loss-of-function a contributor to parkinsonian pathology? Evidence from non-human primates, Front. Neurosci., 10 (2016), 12.
    [34] C. Klein, N. Hattori, C. Marras, MDSGene: Closing data gaps in genotype-phenotype correlations of monogenic Parkinson's disease, J. Parkinsons Dis., 8 (2018), S25-S30.
    [35] J. Chi, Q. Xie, J. Jia, X. Liu, J. Sun, Y. Deng, L. Yi, Integrated analysis and identification of novel biomarkers in Parkinson's disease, Front. Aging Neurosci., 10 (2018), 178.
    [36] Y. I. Li, G. Wong, J. Humphrey, T. Raj, Prioritizing Parkinson's disease genes using population-scale transcriptomic data, Nat. Commun., 10 (2019), 994.
    [37] K. J. Karczewski, M. P. Snyder, Integrative omics for health and disease, Nat. Rev. Genet., 19 (2018), 299-310. doi: 10.1038/nrg.2018.4
    [38] B. Wang, V. Kumar, A. Olson, D. Ware, Reviving the transcriptome studies: an insight into the emergence of single-molecule transcriptome sequencing, Front. Genet., 10 (2019), 384.
    [39] F. R. Pinu, S. A. Goldansaz, J. Jaine, Translational metabolomics: current challenges and future opportunities, Metabolites, 9 (2019), 108.
    [40] J. L. Ren, A. H. Zhang, L. Kong, X. J. Wang, Advances in mass spectrometry-based metabolomics for investigation of metabolites, RSC Adv., 8 (2018), 22335.
    [41] J. M. Wilkins, E. Trushina, Application of metabolomics in Alzheimer's disease, Front. Neurol., 8 (2018), 719.
    [42] E. Trushina, T. Dutta, X. M. Persson, M. M. Mielke, R. C. Petersen, Identification of altered metabolic pathways in plasma and CSF in mild cognitive impairment and Alzheimer's disease using metabolomics, PLoS One, 8 (2013), e63644.
    [43] G. Paglia, M. Stocchero, S. Cacciatore, S. Lai, P. Angel, M. T. Alam, et al., Unbiased metabolomic investigation of Alzheimer's disease brain points to dysregulation of mitochondrial aspartate metabolism, J. Proteome Res., 15 (2016), 608-618. doi: 10.1021/acs.jproteome.5b01020
    [44] S. G. Snowden, A. A. Ebshiana, A. Hye, Y. An, O. Pletnikova, R. O'Brien, et al., Association between fatty acid metabolism in the brain and Alzheimer disease neuropathology and cognitive performance: a nontargeted metabolomic study, PLoS Med., 14 (2017), e1002266.
    [45] S. P. Guiraud, I. Montoliu, L. Da Silva, L. Dayon, A. N. Galindo, J. Corthesy, et al., High-throughput and simultaneous quantitative analysis of homo-cysteine-methionine cycle metabolites and co-factors in blood plasma and cerebrospinal fluid by isotope dilution LC-MS/MS, Anal. Bioanal. Chem., 409 (2017), 295-305. doi: 10.1007/s00216-016-0003-1
    [46] J. B. Toledo, M. Arnold, G. Kastenmuller, R. Chang, R. A. Baillie, X. Han, et al., Metabolic network failures in Alzheimer's disease-a biochemical road map, Alzheimers Dement., 13 (2017), 965-84. doi: 10.1016/j.jalz.2017.01.020
    [47] S. F. Graham, O. P. Chevallier, C. T. Elliott, C. Holscher, J. Johnston, B. McGuinness, et al., Untargeted metabolomic analysis of human plasma indicates differentially affected polyamine and l-arginine metabolism in mild cognitive impairment subjects converting to Alzheimer's disease, PLoS One, 10 (2015), e0119452.
    [48] M. Mapstone, F. Lin, M. A. Nalls, A. K. Cheema, A. B. Singleton, M. S. Fiandaca, et al., What success can teach us about failure: the plasma metabolome of older adults with superior memory and lessons for Alzheimer's disease, Neurobiol. Aging, 51 (2017), 148-55. doi: 10.1016/j.neurobiolaging.2016.11.007
    [49] M. Benson, Clinical implications of omics and systems medicine: focus on predictive and individualized treatment, J. Intern. Med., 279 (2016), 229-240. doi: 10.1111/joim.12412
    [50] C. Evans, J. Hardin, D. M. Stoebel, Selecting between-sample RNA-Seq normalization methods from the perspective of their assumptions, Brief Bioinform., 19 (2018), 776-792. doi: 10.1093/bib/bbx008
    [51] F. Maleki, K. Ovens, D. J. Hogan, A. J. Kusalik, Gene set analysis: challenges, opportunities, and future research, Front. Genet., 11 (2020), 654.
    [52] I. Ihnatova, V. Popovici, E. Budinska, A critical comparison of topology-based pathway analysis methods, PLoS One, 13 (2018), e0191154.
    [53] N. Xie, L. Zhang, W. Gao, C. Huang, P. E. Huber, X. Zhou, et al., Subpathway-GM: identification of metabolic subpathways via joint power of interesting genes and metabolites and their topologies within pathways, Signal. Transduct. Target. Ther., 5 (2020), 227.
    [54] Q. Yang, S. Wang, E. Dai, S. Zhou, D. Liu, H. Liu, et al., Pathway enrichment analysis approach based on topological structure and updated annotation of pathway, Brief Bioinform., 20 (2019), 168-177. doi: 10.1093/bib/bbx091
    [55] P. Kong, P. Lei, S. Zhang, D. Li, J. Zhao, B. Zhang, Integrated microarray analysis provided a new insight of the pathogenesis of Parkinson's disease, Neurosci. Lett., 662 (2018), 51-58. doi: 10.1016/j.neulet.2017.09.051
    [56] M. G. Krokidis, P. Vlamos, Transcriptomics in amyotrophic lateral sclerosis, Front. Biosci. (Elite Ed), 10 (2018), 103-121.
    [57] L. Bravo-Merodio, L. A. Williams, G. V. Gkoutos, A. Acharjee, Omics biomarker identification pipeline for translational medicine. J. Transl. Med., 17 (2019), 155.
    [58] M. A. Myszczynska, P. N. Ojamies, A. M. B. Lacoste, D. Neil, A. Saffari, R. Mead, et al., Applications of machine learning to diagnosis and treatment of neurodegenerative diseases, Nat. Rev. Neurol., 16 (2020), 440-456. doi: 10.1038/s41582-020-0377-8
    [59] J. Xin, X. Ren, L. Chen, Y. Wang, Identifying network biomarkers based on protein-protein interactions and expression data, BMC Med. Genomics, 8 (2015), S11.
    [60] L. Fu, K. Fu, Analysis of Parkinson's disease pathophysiology using an integrated genomics-bioinformatics approach, Pathophysiology, 22 (2015), 15-29. doi: 10.1016/j.pathophys.2014.10.002
    [61] R. Al-Ouran, Y. W. Wan, C. G. Mangleburg, T. V. Lee, K. Allison, J. M. Shulman, et al., A portal to visualize transcriptome profiles in mouse models of neurological disorders, Genes (Basel), 10 (2019), E759.
    [62] J. Kelly, R. Moyeed, C. Carroll, D. Albani, X. Li, Gene expression meta-analysis of Parkinson's disease and its relationship with Alzheimer's disease, Mol. Brain., 12 (2019), 16.
    [63] E. Mariani, F. Frabetti, A. Tarozzi, M. C. Pelleri, F. Pizzetti, R. Casadei, Meta-analysis of Parkinson's disease transcriptome data using TRAM software: whole substantia nigra tissue and single dopamine neuron differential gene expression, PLoS One, 11 (2016), e0161567.
    [64] P. Kong, P. Lei, S. Zhang, D. Li, J. Zhao, B. Zhang, Integrated microarray analysis provided a new insight of the pathogenesis of Parkinson's disease, Neurosci. Lett., 662 (2018), 51-58. doi: 10.1016/j.neulet.2017.09.051
    [65] E. Glaab, Computational systems biology approaches for Parkinson's disease, Cell Tissue Res., 373 (2018), 91-109. doi: 10.1007/s00441-017-2734-5
    [66] T. M. Nguyen, A. Shafi, T. Nguyen, S. Draghici, Identifying significantly impacted pathways: a comprehensive review and assessment, Genome Biol., 20 (2019), 203.
    [67] J. Ma, A. Shojaie, G. Michailidis, A comparative study of topology-based pathway enrichment analysis methods, BMC Bioinform., 20 (2019), 546.
    [68] C. Gao, H. Sun, T. Wang, M. Tang, N. I. Bohnen, M. L. T. M. Müller, et al., Model-based and model-free machine learning techniques for diagnostic prediction and classification of clinical outcomes in Parkinson's disease, Sci. Rep., 8 (2018), 7129.
    [69] Y. Hu, Z. Pan, Y. Hu, L. Zhang, J. Wang, Network and pathway-based analyses of genes associated with Parkinson's disease, Mol. Neurobiol., 54 (2017), 4452-4465. doi: 10.1007/s12035-016-9998-8
  • Reader Comments
  • © 2021 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(4140) PDF downloads(281) Cited by(9)

Figures and Tables

Figures(3)  /  Tables(1)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog