
Calcium (Ca2+) signaling plays a pivotal role in coordinating neural stem cell (NSC) proliferation across various cell cycle stages, regulating immediate early gene transcription, and governing processes like quiescence and cell division. Additionally, calcium signaling pathways are implicated in the initiation, progression, and therapeutic targeting of glioblastoma multiforme (GBM), particularly focusing on glioma stem cells (GSCs). Intracellular calcium levels are increased through the activation of channels, transporters, and calcium-binding proteins (CaBPs), which generate specific calcium signals characterized by spatial, temporal, and intensity profiles. Moreover, extracellular factors such as growth factors, neurotransmitters, and extracellular nucleotides modulate calcium levels to finely regulate NSC and GBM behavior. Calcium-associated proteins and ion channels like calcium release-activated (CRAC) channels and voltage-gated calcium channels play key roles in NSC proliferation and differentiation. Despite calcium's versatile and widespread role as a second messenger critical for regulating various cellular functions, the specific roles of calcium in stem cell niches, stem cell maintenance, and glioblastoma stem cells are still in early stages of exploration. This article aimed to provide a comprehensive and current understanding of the roles of calcium signaling in NSC behavior and interactions within their niche, which are critical for neurogenesis, brain repair mechanisms, and understanding age-related decline in stem cell function. Investigating the heterogeneity of GBM tumors resembling neurospheres and their similarity to neural stem cells (NSCs) highlights the critical involvement of calcium in governing cellular behaviors such as quiescence, proliferation, and migration. Furthermore, this manuscript illuminates various potential interventions targeting calcium channels and associated signaling pathways to mitigate GSC activities and hinder GBM recurrence, offering a promising avenue for developing novel therapeutic strategies against GBM.
Citation: Ola A Al-Ewaidat, Sopiko Gogia, Valiko Begiashvili, Moawiah M Naffaa. The multifaceted role of calcium signaling dynamics in neural cell proliferation and gliomagenesis[J]. AIMS Biophysics, 2024, 11(3): 296-328. doi: 10.3934/biophy.2024017
[1] | Khaphetsi Joseph Mahasa, Rachid Ouifki, Amina Eladdadi, Lisette de Pillis . A combination therapy of oncolytic viruses and chimeric antigen receptor T cells: a mathematical model proof-of-concept. Mathematical Biosciences and Engineering, 2022, 19(5): 4429-4457. doi: 10.3934/mbe.2022205 |
[2] | Elzbieta Ratajczyk, Urszula Ledzewicz, Maciej Leszczynski, Avner Friedman . The role of TNF-α inhibitor in glioma virotherapy: A mathematical model. Mathematical Biosciences and Engineering, 2017, 14(1): 305-319. doi: 10.3934/mbe.2017020 |
[3] | Taeyong Lee, Adrianne L. Jenner, Peter S. Kim, Jeehyun Lee . Application of control theory in a delayed-infection and immune-evading oncolytic virotherapy. Mathematical Biosciences and Engineering, 2020, 17(3): 2361-2383. doi: 10.3934/mbe.2020126 |
[4] | Zizi Wang, Zhiming Guo, Hal Smith . A mathematical model of oncolytic virotherapy with time delay. Mathematical Biosciences and Engineering, 2019, 16(4): 1836-1860. doi: 10.3934/mbe.2019089 |
[5] | Lu Gao, Yuanshun Tan, Jin Yang, Changcheng Xiang . Dynamic analysis of an age structure model for oncolytic virus therapy. Mathematical Biosciences and Engineering, 2023, 20(2): 3301-3323. doi: 10.3934/mbe.2023155 |
[6] | Peter S. Kim, Joseph J. Crivelli, Il-Kyu Choi, Chae-Ok Yun, Joanna R. Wares . Quantitative impact of immunomodulation versus oncolysis with cytokine-expressing virus therapeutics. Mathematical Biosciences and Engineering, 2015, 12(4): 841-858. doi: 10.3934/mbe.2015.12.841 |
[7] | Prathibha Ambegoda, Hsiu-Chuan Wei, Sophia R-J Jang . The role of immune cells in resistance to oncolytic viral therapy. Mathematical Biosciences and Engineering, 2024, 21(5): 5900-5946. doi: 10.3934/mbe.2024261 |
[8] | Jianjun Paul Tian . The replicability of oncolytic virus: Defining conditions in tumor virotherapy. Mathematical Biosciences and Engineering, 2011, 8(3): 841-860. doi: 10.3934/mbe.2011.8.841 |
[9] | Joseph Malinzi, Rachid Ouifki, Amina Eladdadi, Delfim F. M. Torres, K. A. Jane White . Enhancement of chemotherapy using oncolytic virotherapy: Mathematical and optimal control analysis. Mathematical Biosciences and Engineering, 2018, 15(6): 1435-1463. doi: 10.3934/mbe.2018066 |
[10] | Nada Almuallem, Dumitru Trucu, Raluca Eftimie . Oncolytic viral therapies and the delicate balance between virus-macrophage-tumour interactions: A mathematical approach. Mathematical Biosciences and Engineering, 2021, 18(1): 764-799. doi: 10.3934/mbe.2021041 |
Calcium (Ca2+) signaling plays a pivotal role in coordinating neural stem cell (NSC) proliferation across various cell cycle stages, regulating immediate early gene transcription, and governing processes like quiescence and cell division. Additionally, calcium signaling pathways are implicated in the initiation, progression, and therapeutic targeting of glioblastoma multiforme (GBM), particularly focusing on glioma stem cells (GSCs). Intracellular calcium levels are increased through the activation of channels, transporters, and calcium-binding proteins (CaBPs), which generate specific calcium signals characterized by spatial, temporal, and intensity profiles. Moreover, extracellular factors such as growth factors, neurotransmitters, and extracellular nucleotides modulate calcium levels to finely regulate NSC and GBM behavior. Calcium-associated proteins and ion channels like calcium release-activated (CRAC) channels and voltage-gated calcium channels play key roles in NSC proliferation and differentiation. Despite calcium's versatile and widespread role as a second messenger critical for regulating various cellular functions, the specific roles of calcium in stem cell niches, stem cell maintenance, and glioblastoma stem cells are still in early stages of exploration. This article aimed to provide a comprehensive and current understanding of the roles of calcium signaling in NSC behavior and interactions within their niche, which are critical for neurogenesis, brain repair mechanisms, and understanding age-related decline in stem cell function. Investigating the heterogeneity of GBM tumors resembling neurospheres and their similarity to neural stem cells (NSCs) highlights the critical involvement of calcium in governing cellular behaviors such as quiescence, proliferation, and migration. Furthermore, this manuscript illuminates various potential interventions targeting calcium channels and associated signaling pathways to mitigate GSC activities and hinder GBM recurrence, offering a promising avenue for developing novel therapeutic strategies against GBM.
Cancer is characterized by abnormal cell growth and several steps of genetic mutations. Glioblastoma multiforme (GBM), a type of brain tumor, is one of the most fatal human cancers, with a low survival rate attributed to its aggressive growth and rapid, pervasive brain invasion [1,2]. Despite recent progress in cancer treatment, toxicity and corresponding side effects such as cancer-related cognitive changes [3,4] are common in conventional anticancer therapies such as chemotherapy and radiation therapy. Targeted cancer therapy, a promising choice with recent advances, can be a great way of eradicating cancer cells by controlling its dynamics and players in the signaling pathways that enable tumor growth [5]. Thus, there is a need for optimal control of promotion or inhibition of these intracellular molecules in cancer treatment including chemo- and oncolytic virus (OV) therapies.
The ubiquitin–proteasome system within a cell regulates the degradation of proteins, thus controlling cellular function and maintaining homeostasis [6,7]. Therefore, proteasome inhibitors can be used to treat cancer due its inhibition of cancer cells's increased need of protein synthesis and degradation in cancer progression [6,8]. By interfering with appropriate degradation and recycling of proteins, a critical component for cell viability [9], proteasome inhibitors induce accumulation of ubiquitin-tagged proteins, resulting in endoplasmic reticulum (ER) stress [10] and, thus, apoptosis, i.e., programmed cell death, of cancer cells. Bortezomib, the first proteasome inhibitor agent in cancer treatment [11], was approved for multiple myeloma and lymphoma by Food and Drug Administration (FDA) [6]. Despite minimal undesirable side effects such as accumulation of anti-apoptotic proteins, bortezomib was shown to consistently overcome the well-known apoptotic resistance of cancer cells such as melanoma cells [12], especially when combined with other anti-cancer agents [7,13,14].
NFκB, a master regulator that controls inflammation and immune responses as well as cell proliferation, stays usually in an inactive mode by its suppressor, IκB [15]. Various regulators such as ER stress can activate NFκB by degrading IκB which induces the translocation of NFκB to the nucleus and subsequent activation of various key genes, including the anti-apoptotic gate keeper, Bcl2 and suppression of the final apoptosis gene BAX (Figure 1) [16,17,18,19,20]. While the systemic homeostasis can be maintained by this production of IκB and the feedback control loop in normal tissue [21], a poor management of NFκB by IκB, or a constitutively active NFκB state, can allow the critical resistance of tumor cells to conventional anti-cancer therapy [22,23,24,25,26]. Therefore, agents targeting BAX showed a great potential of killing cancer cells by up-regulating BAX, thus, inducing apoptosis [17,27]. While apoptosis is a typical cell death program, necroptosis is another form of cell death process that is caspase-independent and causes massive cell killing. This is executed by up-regulation of reactive oxygen species (ROS) from activated RIP1, a receptor-interacting-protein kinase [28]. RIP1, a key activator of necroptosis, is not involved in the apoptotic signaling pathway [28,29]. Up-regulation of both Bcl2 and RIP1 as well as down-regulation of caspase-8 were associated with necroptosis in glioma cancer cells [30].
An emerging treatment option with rapid advances for cancer patients is oncolytic virotherapy (OV) that uses various forms of cancer-lysing viruses as a targeted therapy [31,32,33]. OVs can target and kill cancer cells by infection and destruction of cell structure while avoiding unnecessary damages to normal cells in tumor microenvironment (TME) [34,35]. While these viruses can be eliminated by the immune system on the way to the target tumor [36], the high degree of replications within a cancer cell and boosting effect of OV-induced immune activities within TME [31,37] can compensate the loss by immune cells, which enables these viruses to spread throughout the tumor from the infected tumoral region [38]. Clinical trials for various types of OVs as well as combination therapies have been conducted and are under way to test their anti-tumor efficacy and safety [6,31,39]. Some of them were approved by FDA for clinical use, for instance, T-Vec, a modified herpes simplex virus 1 (oHSV), for melanoma patients [6,37,40]. However, virus clearance due to OV-induced immunity prevents an OV therapy from efficient anti-tumor efficacy [41]. Combination therapies including OVs were suggested as an alternative method to overcome this low efficacy and several chemotherapeutic candidates are being tested for synergistic impact on cancer cell killing [37,42].
Yoo et al. [6] demonstrated that bortezomib was not only effective in killing cancer cells but promotes synergistic anti-tumor efficacy by inducing unfolded protein response in cancer cells, which then leads to the nuclear localization of the oHSV and a significant increase in viral replication. A follow-up study [43] then showed that the OV-bortezomib combination therapy can induce necroptosis through stimulation of the secretion of various cytokines associated with inflammation. While bortezomib treatment alone initiated a typical apoptosis, the combination treatment induced necroptosis by diverting to different signaling pathways [20,44]. Kim et al. [45] developed a mathematical model of the nonlinear dynamics of complex interactions between cancer cells and immune cells in OV–bortezomib therapy based on the experimental observations including those reported in [6,43]. It was found that both injection of exogenous natural killer (NK) cells and deletion of endogenous NK cells can lead to better anti-tumor efficacy, illustrating the complex role of TME such as immune cells in regulation of dynamics of OV-tumor interaction in a combination anti-cancer therapy. In a follow-up study [46], Kim et al. investigated the intracellular signaling pathway that governs the cancer cell death mechanism in response to various levels of bortezomib with OV therapy. Anti-apoptosis status and two different cellular death mechanisms, i.e., apoptosis and necroptosis in response to bortezomib treatment, were characterized by up- or down-regulation of three key intracellular molecules (NFkB/Bcl2, BAX, RIP1) [46], which was consistent with experimental observations [6,20,43,44,45]. These works [45,46,47] showed that the treatment protocols significantly affect the anti-tumor efficacy in the combination therapy and illustrated the importance of the detailed analysis of both apoptotic and necroptotic death of cancer cells in bortezomib-OV therapies. Various types of mathematical models by other groups have been developed for oncolytic virus therapy [48,49,50,51,52,53,54,55,56,57,58] and bortezomib treatment [59,60,61] in various cancers.
In this study, we employ a modified mathematical model of bortezomib-OV cancer therapy, integrating cell-death pathways based on Kim et al. [46]. This refined model enables the design of optimally controlled treatment strategies. Our modeling framework utilizes optimal control theory, a mathematical tool that deals with complex biological systems that can be controlled by an external agent [62]. Optimal control theory has been used to determine effective administration schedules of anti-cancer agents in various cancer types [63,64,65,66,67] including brain tumors [47,68,69,70,71] and lung cancers [72,73]. For example, an optimal control method was utilized to prevent aggressive glioma cell invasion by up-regulating miR-451 [71] and controlling cell cycle [70] within cancer cells, or to optimize the anti-tumor efficacy in a nonlinear immune microenvironment [47]. Optimal control theory along mathematical models was used in various fields including epidemiology for COVID-19 [74] and MERS [75] and cancer research. In particular, optimal control scheme was used to construct cost-effective therapy protocols against cancer development, from earlier works [64,65,66] to recent works [76]. These studies investigated the optimal strategies in controlling tumor growth under various treatments (chemotherapy [64,65,66,69,76,77,78], immunotherapy [79,80,81,82], and review [63] by minimizing the tumor size and costs. In previous studies [45,46], authors investigated the nonlinear dynamics of NK cells in a combination therapy (OVs + bortezomib) as well as anti-tumor efficacy via the intracellular signaling pathway (NFκB-Bcl-2-BAX-RIP1) and invasion pattern of GBM cells in brain in response to OVs and bortezomib. GBM is a serious brain cancer with very low survival rate and major cause of death is regrowth of invasive tumor cells in different locations after surgery. In [46], Kim et al explores tumor growth and invasion patterns in GBM patients based on the injection location of bortezomib and OVs. These studies focus mainly on anti-tumor efficacy of NK therapy and invasion pattern of tumor cells in the presence of bortezomib and OVs through constant treatment protocols. These approaches present challenges in translating the findings into practical clinical applications. In particular, in order to better understand the OV-mediated cell death program and provide personalized care for each cancer patient, it is necessary to develop optimally controlled injection strategies. Due to the safety issues, injection amount of OVs is usually limited and high doses of bortezomib can cause serious side effects. For example, local or IV-administration of OVs can cause adverse reactions such as influenza-like symptoms (vomiting, myalgia, headache, fatigue, elevated body temperature, nausea) and local reactions (pain, rash, erythema, peripheral edema) [83,84]. Despite the high degrees of anti-tumor effects, bortezomib can also induce serious side effects such as nausea, diarrhea, tiredness, low platelets, fever, and low white blood cell count [85,86,87]. On the other hand, low level of OVs and bortezomib decreases the overall anti-tumor efficacy due to virus clearance and blood-brain barrier (BBB). Therefore, appropriate level of OVs and bortezomib need to be controlled for injection into patient's body in order to avoid serious side effects and to obtain maximal clinical outcomes. In our framework based on the detailed hybrid type of mathematical model, we used an optimal control theory in order to maximize the anti-tumor efficacy and minimize the side effects and costs associated with administration at clinics by controlling the intracellular cell death program. We also investigated the dynamical systems of intracellular module in terms of apoptosis, oncolysis, and necroptosis in much more detail. This analysis provides a platform of developing more efficient next generation of OVs and another drugs targeting the necroptosis signaling pathway. Hence, we aim to obtain optimal treatment regimens (including timing and dosages) for OV and bortezomib, striking a balance where cancer cell eradication is maximized, while concurrently minimizing undesirable side effects and financial expenditures. The OV/bortezomib-mediated intracellular signaling pathways of apoptosis, necroptosis, and oncolysis in cell death programs are very complex biological system itself, providing fundamentally different mechanism of eradicating tumor cells. The necroptosis pathway plays a key role in cancer therapy including OV therapy. For example, anti-cancer drugs such as doxorubicin can increase oncolytic effect significantly (> 100-fold in breast cancer) by enhancing OV replication and inducing apoptosis and necroptosis [88]. Other research groups found that a radiation therapy can induce necroptosis in cancer progression [89]. However, it is poorly understood how these cell death programs work in OV therapy in detail. Thus, better understanding of this complex cell-death program itself can shed light into developing a new, more effective anti-cancer drug targeting NFκB-Bcl2-BAX-RIP1 in an OV combination therapy by manipulating the signaling network. Our findings enable the proposal of sophisticated optimal treatment strategies in terms of controlling the precise components at the gene level within a tumor cell rather than a population level (cf. [47]). Furthermore, detailed mathematical analysis of the intracellular cell death cascade and optimally controlled injection schedule can provide a general framework for the pharmaceutical companies to find new compounds targeting signaling pathways in OV therapy. We employed an optimal control formulation to strategically reduce tumor size by triggering both necroptosis and apoptosis through modulating intracellular signaling pathways (NFκB/Bcl2, BAX, RIP1), while managing the cumulative administration costs of bortezomib and OVs. Our results provide efficient and cost-effective therapeutic management schemes for OVs and bortezomib, taking into account overall anti-tumor efficacy.
In this work, we investigate the complex interactions between tumor cells and other therapeutic agents such as bortezomib (B) and oncolytic virus (v) incorporating an intracellular pathway using a mathematical model. Tumor cells are classified as : Uninfected (x), infected (y), and necrotic tumor (n) cells. Four intracellular components involved in the cell-death program are considered: IκB (S), NFκB (F), BAX (A), and RIP1 (R). The regulatory network in Figure 2 describes the complex interactions among the cellular components and intracellular modules. In the model, we consider a tumor microenvironment where (ⅰ) initially uninfected tumor cells (x) grow (i.e., initial condition of x is always positive.) (ⅱ) uninfected tumor cells (x) become infected tumor cells (y) under intracellular oncolysis condition in response OV treatment (ⅲ) bortezomib alone can kill tumor cells under intracellular apoptosis condition (ⅳ) uninfected tumor cells (x) become infected tumor cells (y) under intracellular necroptosis condition in response both OV treatment and bortezomib, leading to necroptotic death.
Following administration, OVs navigate toward the tumor site and infiltrate the target cancer cells. Within these cells, OVs vigorously replicate generating a substantial viral population. As the infected cancer cells disintegrate, OVs are released, instigating the infection-elimination process anew.
Bortezomib alone, can induce cancer cell death through a typical programmed cell-death pathways, involving apoptotic signaling cascade. Conversely, the combined therapy of bortezomib and OV orchestrates a synergistic assault on cancer cells, harnessing necroptotic signaling pathways to achieve a heightened level of cell destruction.
The network of the model is shown in Figure 2 and the corresponding model dynamics can be described by a system of ordinary differential equations in a dimensionless form as follows:
dSdt=λBB1+αvk+v+σ1σ29σ29+σ4F2−S, | (2.1) |
dFdt=σ7+σ2σ210σ210+σ5S2−ω1(D)F, | (2.2) |
dAdt=σ8+σ3σ211σ211+σ6F2−ω2A, | (2.3) |
dRdt=σ12+σ13[oHSV]F−ω3R, | (2.4) |
dxdt=λx(1−xx0)−β1xBIapop−β2xvIoncol−β3xvInecrop, | (2.5) |
dydt=−δy+β2xvIoncol+β3xvInecrop, | (2.6) |
dndt=δy−μn, | (2.7) |
dvdt=uV+bδy(1+α1B)−γv, | (2.8) |
dBdt=uB−(μ1x+μ2y)BkB+B−μBB. | (2.9) |
Here, λB is the signaling activation rate of IκB in response to bortezomib signal B, which is inhibited by the presence of OVs. So, [oHSV] indicates a response switch function of OVs (v) in the form, [oHSV] = vk+v with a Hill-type parameter k. This IκB activity is inhibited by NFκB (F) with inhibition strength σ4, autocatalytic enhancement parameter σ1, and Hill-type parameters σ9. In a similar fashion, the inhibition processes of NFκB (F) and BAX (A) by IκB and NFκB, are represented by corresponding parameters σ5,σ2,σ10 and σ6,σ3,σ11, respectively. Here, ω1(t) may depend on drugs that enhance NFκB degradation, i.e., ω1(D)=ω†1eD where D is a concentration of a NFκB inhibitor. Here, the drug concentration D is a constant. The signaling sources and decay rates of NFκB, BAX, and RIP1 are represented by σ7,σ8,σ12 and ω1,ω2, and ω3, respectively. The third term on the right hand side of Eq (2.1) indicates the dimensionless natural decay of IκB. The same response switch function [oHSV] in Eq (2.4) is used for activation of RIP1 in the presence of OVs in the system. Population dynamics of uninfected tumor cells in Eq (2.5) consists of logistic growth with the growth rate λ and carrying capacity x0, bortezomib-induced cell killing with the apoptotic death rate β1, infection process with infection rate β2, and massive tumor cell killing with necroptotic death rate β3. An uninfected tumor cell is infected when an oncolytic virus (OV) or a group of OVs enter this tumor cell and replicate upto 10,000-fold, initiating cell death program (either oncolysis and necroptosis) in response to OV therapy. Therefore, the indicator functions Ioncol and Inecrop reflect this infection process and the transition x→y. After infection and confirmation of the cell fate (either oncolysis or necroptosis), the infected tumor cells go through the subsequent, cell death process which is reflected in the y→n transition in the model. Here, Iapop,Ioncol,Inecrop are indicator functions for apoptotic, oncolytic, and necroptotic status, respectively, based on up- or down-regulation of NFκB, BAX, and RIP1:
Iapop={1if apoptosis 0otherwise ,Ioncol={1if oncolysis 0otherwise ,Inecrop={1if necroptosis 0otherwise . | (2.10) |
The exact definition of these status will be defined in the next section. Infected tumor cells (y) from the uninfected tumor population are cleared to dying tumor population (n) at a death rate δ and these necrotic population are cleared from the population system at a death rate μ. OVs are supplied at a rate uV and are able to replicate from lysis of infected cancer cells at a rate bδy with doubling number b in the absence of bortezomib (B). Injection of bortezomib in addition to OV therapy enhances viral replication at additional rate α1B, inducing synergistic anti-cancer effect. OVs are cleared from the system at a rate γ. Bortezomib is injected at a rate μB, internalized in uninfected and infected cancer cells at degradation rates μ1 and μ2, respectively, and cleared from the system at a decay rate μB.
A dimensional version of equations corresponding to Eqs (2.1)–(2.9) and the nondimensionalization provided in the Supplementary material (Supporting information S1 File). Matlab (Mathworks) software is used to simulate the transient dynamics of the model and to obtain the numerical results for the optimal control problems. Table 1 provides the full list of parameters, its meaning and corresponding values used in the model.
Par | Description | Values | Ref |
Intracellular dynamics | |||
λB | Bortezomib signaling scaling factor | 1.9 | [46] |
α | Inhibition strength of bortezomib by oHSV | 30 | [46] |
σ1 | Autocatalytic enhancement rate of IκB | 4 | [46] |
σ9 | Hill-type parameter | 1.0 | [46] |
σ4 | Inhibition strength of IκB by the NFκB-Bcl-2 complex | 2.2 | [46] |
σ7 | Signaling strength of the NFκB-Bcl-2 complex | 0.07 | [46] |
σ2 | Autocatalytic enhancement rate of the NFκB-Bcl-2 complex | 1.33 | [46] |
σ10 | Hill-type parameter | 1.0 | [46] |
σ5 | Inhibition strength of the NFκB-Bcl-2 complex by IκB | 1 | [46] |
ω1 | Decay rate of NFκB-Bcl-2 complex | 0.3 | [107,108] |
σ8 | Signaling strength of BAX | 0.0033 | [46] |
σ3 | Autocatalytic enhancement rate of BAX | 0.111 | [46] |
σ11 | Hill-type parameter | 1.0 | [46] |
σ6 | Inhibition strength of BAX by the NFκB-Bcl-2 complex | 1 | [46] |
ω2 | Decay rate of BAX | 0.02 | [107,108,109] |
σ12 | Signaling strength of RIP1 | 0.1 | [46] |
σ13 | Activation rate of RIP1 in the presence of OVs | 0.13 | [46] |
ω3 | Decay rate of RIP1 | 0.139 | [107,108,110] |
k | Hill type parameter of oHSV switching | 0.01 v∗ | [46] |
Cancer cells | |||
λ | proliferation rate of tumor cells | 4×10−3 | [46,57] |
x0 | Carrying capacity of uninfected tumor cells | 1 | [38,45,46,57] |
β1 | Bortezomib-induced apoptosis rate of tumor cells | 2×10−4 | [46] |
β2 | Oncolysis rate of tumor cells | 3×10−4 | [45,46,57] |
β3 | Necroptosis rate of tumor cells | 5×10−4 | [46] |
δ | infected cell lysis rate | 4.4×10−3 | [45,57] |
μ | Removal rate of dead cells | 2.3×10−3 | [45,46,57] |
Anticancer drugs | |||
b | Burst size of infected cells | 50 | [46,57] |
α1 | bortezomib-induced viral replication rate | 1 | [45,46] |
γ | clearance rate of viruses | 1.8×10−3 | [45,57] |
μ1 | consumption rate of bortezomib by uninfected tumor cells | 0.2075 | [45,46] |
μ2 | consumption rate of bortezomib by infected tumor cells | 0.2075 | [45,46] |
kB | Hill-type parameter | 1 | [45,46] |
μB | decay rate of bortezomib | 0.03 | [46,111,112] |
Reference values of main variables | |||
S∗ | Concentration of IκB | 0.05 μM | [113,114] |
F∗ | Concentration of the NFκB-Bcl2 complex | 0.5 μM | [113,114,115,116] |
A∗ | Concentration of BAX | 0.1 μM | [117] |
R∗ | Concentration of RIP1 | 5.0 μM | [118] |
x∗ | Uninfected cell density | 106cells/mm3 | [45,57,119] |
y∗ | Infected cell density | =x∗ | [45,57,119] |
n∗ | Dead cell density | =x∗ | [45,57,119] |
v∗ | Virus concentration | 2.2×108virus/mm3 | [45,57,119] |
B∗ | Bortezomib concentration | 1.0×10−11g/mm3 | [6,43,45] |
The optimal control theory was used to yield an optimal administration strategies of OVs and bortezomib that minimizes tumor cell populations (uninfected and infected tumor cells) and associated costs. Our goal is to obtain feasible therapeutic schedule of uV and uB to reduce the tumor size with minimal administrative costs. The controls uV and uB represent dosages of OVs and bortezomib, respectively. These controls are assumed to be bounded. That is, the control sets are defined as
uV(t)∈[0,umaxV]anduB(t)∈[0,umaxB]for allt∈[ts,tf], | (2.11) |
where umaxV and umaxB are the maximum dosage of OVs and bortezomib, respectively, which could be administered in a given treatment period [ts,tf] (where ts is the start time of treatment with optimal control applied, and tf is the end time).
Two distinct cost functionals are formulated to facilitate alternating administrations of OVs and bortezomib, enhancing their synergistic efficacy against cancer cells. In the first formulation, both integrands involve linear combination of uninfected x(t), infected y(t) and quadratic control. This quadratic form guarantees a strictly convex Hamiltonian and a unique minimizer, rendering the mathematical problem more tractable. Incorporating quadratic controls within the cost function effectively simulates the detrimental repercussions of excessive drug administration, a strategy observed in multiple studies [32,82,90,91,92,93]. However, the use of such quadratic functionals lacks robust biological justification, raising concerns about the validity of its simplifying assumptions [94]. The quadratic functional types may not be commonly motivated by biological systems making its simplifying model assumptions bit questionable and ambiguous [94]. However, its use also showed qualitatively significant outcomes in terms of incorporating the intrinsic nonlinearity of the given problem [90,95]. Thus, the quadratic controls in the cost function has been used in controlling the adverse effects of use of too much therapeutic drugs in various works [32,82,90,91,92,93]. In cancer immunotherapy, a type of isoperimetric optimal controls has been applied [79,80,81]. This leads to the formulation of an alternative cost functional, where linear controls are involved, supplemented by a penalty function serving as a guide for controlled implemented.
1) Quadratic controls This strategy proposes a sequential administration of OV and bortezomib, minimizing both their dosages and the levels of uninfected (x(t)) and infected cancer cells (y(t)). Employing quadratic controls lead to the following cost functional
J(uV(t),uB(t))=∫tfts(x(t)+y(t)+C1u2V(t)2)dt+∫t′ft′s(x(t)+y(t)+C2u2B(t)2)dt. | (2.12) |
2) Linear controls with constraint In this approach, linear controls are employed to emulate a biologically plausible regulatory mechanism. To achieve the desired effect of inducing necroptosis, a sigmoid function is integrated that ensures a sustained and optimal concentration of bortezomib. The cost functional yields
J(uV(t),uB(t))=∫tfts(x(t)+y(t)+C3uV(t))dt+∫t′ft′s(C4uB(t)+B51+ep(B−thB))dt. | (2.13) |
Our goal is to find optimal administration of both OV and bortezomib (u∗V(t) and u∗B(t)), identifying optimal dosages under the framework of optimal control theory. The principle technique for such an optimal control problem is to solve a set of "necessarycondition" that an optimal control and corresponding state must satisfy. The necessary conditions we derived were developed by Pontryagin. He introduced the concept of 'adjoint functions' to incorporate differential equations into the objective function. Adjoint functions serve a similar purpose to the Lagrange multiplier of multivariate calculus, which adds constraints to functions of several variables that are being maximized or minimized [62,96]. We perform numerical simulations by employing a fourth-order iterative Runge-Kutta method. Commencing with the initial conditions and an initial control estimate, we solve the state equations using the forward scheme. Furthermore, the adjoint equations are solved using the backward scheme, ensuring compliance with the transversality conditions. Control updates involve a convex combination of the previous controls and values derived from the characterizations. This methodology, recognized as the Forward-Backward Sweep Method (FBSM), has demonstrated convergence [94]. See supporting information S1 File for more details.
The fundamental structure of four phenotypic modes in response to various bortezomib levels in the absence and presence of OVs are represented in Figure 3. Specifically, Figure 3(A), (B) illustrate the steady states of three key intracellular variables: NFκB (represented by blue solid curve), BAX (depicted by the dashed red curve), and RIP1 (shown as the dotted yellow curve). These plots delineate the behavior of these variables under varying bortezomib (B) concentrations, when OVs are absent (v=0 in Figure 3(A)) and present (v=1 in Figure 3(B)).
In the absence of OV within the system, we observe elevated levels of NFκB, coupled with low levels of both BAX and RIP1 in response to lower levels of bortezomib. This dynamic pattern prevents the initiation of bortezomib-induced apoptosis, as the key apoptotic gene, BAX, remains suppressed. As the bortezomib concentration increases, a shift occurs: NFκB is down-regulated, while RIP1 maintains a low level; moreover, the BAX level transitions to an overexpression state, triggering programmed cell death, specifically apoptosis (Figure 3(A)). However, the intracellular dynamics of NFκB, BAX, and RIP1 undergo a significant transformation in the presence of OV when responding to bortezomib stimuli. Both NFκB and RIP1 consistently sustain high levels irrespective of bortezomib concentrations, while the BAX level remains consistently low (Figure 3(B)). This particular state signifies the activation of necroptotic signaling (depicted within the pink box in Figure 3(B)) under conditions of elevated bortezomib levels. Conversely, when bortezomib stimulus is limited, cell demise ensues through oncolysis (illustrated by the green box in Figure 3(B)), rather than necroptosis.
By defining thresholds for these intracellular variables, thF=1.7 for NFκB, thA=1.7 for BAX, thR=1.7 for RIP1, and thB=0.1 for bortezomib, we can delineatefour different phenotypic modes, namely the anti-apoptotic (Tt), apoptotic (Ta), necroptotic (Tn), and oncolysis (To) phases as follows:
Tt={(F,A,R)∈R3:F>thF,A<thA,R<thR}, | (3.1) |
Ta={(F,A,R)∈R3:F<thF,A>thA,R<thR}, | (3.2) |
Tn={(F,A,R,B)∈R4:F>thF,A<thA,R>thR,B>thB}, | (3.3) |
To={(F,A,R,B)∈R4:F>thF,A<thA,R>thR,B<thB}. | (3.4) |
Figure 3(C) illustrates the anti-apoptosis (yellow cube), apoptosis (blue cube), necroptosis (red cube), and oncolysis (green cube) modules in a NFκB-BAX-RIP1 state space. Figure 3(D), (E) show temporal dynamics of concentrations of NFκB and BAX, respectively, with various fixed bortezomib levels (0≤B≤1), illustrating convergent behaviors of two key intracellular variables. Sharp contrast in expression levels at final time (t=200) in both NFκB and BAX reflects different modes: anti-apoptosis (yellow cube in Figure 3(C)) for low bortezomib levels (0≤B≤0.46) and apoptosis (blue cube in Figure 3(C)) phases for high bortezomib levels (0.46≤B≤1). In Figure 3(F)–(I), we investigate the temporal dynamics of the intracellular module in response to a fluctuating bortezomib stimulus (B(t)=−0.5cos(π500t+0.5)). The concentration of IκB undergoes a sudden increase surpassing that of B(t), and reaches its peak coinciding with that of bortezomib's maximum concentration. Conversely, as B(t) decreases, the concentration of IκB takes a downward course, undergoing an abrupt reduction that surpasses the decline in bortezomib levels. This sequence culminates a cycle of bortezomib-triggered activation and deactivation. The inherent mutual inhibition between IκB and NFκB, as described in Eqs (2.1) and (2.2), gives rise to a pivotal dynamic. Specifically, the upward shift (from down-regulation to up-regulation) of IκB during the initial stimulus phase (0≤t≤500) induces a downward transition (from up-regulation to down-regulation) in NFκB and brings about an upward transition (from down-regulation to up-regulation) in BAX. This orchestrated sequence results in a significant phenotypic switch, steering the system from an anti-apoptosis state to an apoptosis mode (as depicted in Figure 3(G)).
Bortezomib has been demonstrated to induce cancer cell elimination through diverse mechanisms, including the suppression of the NFκB pathway. Furthermore, it effectively inhibits the NFκB-mediated tumor growth enhancement by suppressing the degradation of IκB, thereby manifesting its multifaceted antitumor effects [97,98,99,100,101]. Thus, bortezomib can successfully induce apoptotic death of cancer cells in the absence of OV as illustrated in Figure 3(F)–(H) and corroborated by experimental evidences [97,98,99,100,101]. As the value of B(t) is reduced during the second half of the period (500≤t≤1000), the backward switch (NFκB (low→high); BAX (high→low)) follows, leading to the reverse (Ta→Tt) transition (Figure 3(G)). This loop of forward and backward transitions with initial (red asterisk) and mid point (blue circle) on the trajectory curve is shown in Figure 3(H). This illustrates the dynamic movement of concentrations of key intracellular molecules in the bifurcation diagram, forming a loop of that characterizes a cell death program (Tt→Ta→Tt, Figure 3(H)), in response to a fluctuating bortezomib stimulus (Figure 3(I)).
In Figure 4, we investigate the effect of anti-NFκB antibodies on the dynamic progression of apoptotic cell death within cancer cells. These antibodies serve as valuable tools to study the activation and regulation of NFκB signaling pathways in various biological contexts, including cancer, inflammation, and immune responses. In therapeutic applications, targeting NFκB with specific antibodies may have potential applications in treating diseases characterized by aberrant NFκB activity, such as certain autoimmune disorders and cancers [102,103]. Figure 4(A)–(C) depict the steady states of intracellular variables (NFκB, BAX, and RIP1) for various bortezomib stimuli in the absence (Figure 4(A), D=0) and presence (D=0.45 in Figure 4(B) and D=1 in Figure 4(C)) of anti-NFκB drug in the system. The results demonstrate that the area of apoptosis expands as the dose of NFκB antibody increase from 0 to 1. This suggests that apoptosis can be induced even with a small amount of bortezomib, and its effect becomes more prominent as the dose of NFκB antibody increases. Figure 4(D), (E) show the time courses of concentrations of NFκB (Figure 4(D)) and BAX (Figure 4(E)) in response to various drug levels (0≤D≤1) when B=0.1 with nitial condition S(0)=0,F(0)=4.5,A(0)=0.5,R(0)=0.7. Figure 4(G), (H) illustrate the time courses of concentrations of NFκB (Figure 4(G)) and BAX (Figure 4(H)) in response to various drug levels (0≤D≤1) when B=0.1 with initial condition S(0)=5,F(0)=0.5,A(0)=4.5,R(0)=0.7. Figure 4(F) depicts the trajectories of solutions in the F−A plane when D=0.45,B=0.1. Initial conditions used for the red curve are S(0)=1.7,F(0)=0.82,A(0)=0.5,R(0)=0.7 while for the blue curve are S(0)=1.7,F(0)=0.83,A(0)=0.5,R(0)=0.7. The bistability region exist when the level of NFκB antibody is intermediate. Indeed, near the bistable point, small changes in the initial values can lead to convergence in opposite directions, resulting to different treatment outcomes. Therefore, when proposing a treatment strategy in the early stages, it is crucial to carefully consider the condition of intracellular signaling. The sensitivity to initial conditions near the bistable point highlights the need for precision and thorough analysis when designing therapeutic interventions to achieve the desired treatment outcomes. Figure 4(I) shows the characterization of apoptosis, bistability, anti-apoptosis regions in the B−D plane.
Within oncolytic virotherapy (OV therapy), the mere presence of active OVs proves adept at eradicating cancer cells through oncolysis, that avoids the initiation of necroptosis signaling pathways, even in the absence of bortezomib. Intriguingly, empirical evidence has illuminated the amalgamation of therapies, wherein OVs and bortezomib are combined, induces a synergistic impact, notably enhancing the efficacy of cancer cell eradication [6,43]. We investigate the temporal dynamics of intracellular variables in Eqs (2.1)–(2.4) in response to fluctuating OV supply in the absence and presence of bortezomib. In response to increasing OV input described by v(t)=−225cosπ100t+0.0008t+0.28 (Figure 5(A)), the system starting from the same apoptotic status (red asterisks in Figure 5(B), (C)) converges to oncolytic phenotype (blue asterisk in Figure 5(B)) and necroptotic mode (blue asterisk in Figure 5(C)) in the absence (B=0) and presence (B=1) of bortezomib, respectively. In other words, the typical apoptotic cell death program in response to bortezomib transits to more severe cell death program, i.e., necroptosis (red asterisk → blue asterisk in Figure 5(C)), when oncolytic virus is present in the tumor microenvironment. The phenotypic transition between apoptosis and necroptosis is bidirectional. When OV stimulus fluctuates according to v(t)=−12cosπ250t+0.5, the initial curve in v(t) leads to the first transition from the apoptotic mode (red asterisk in Figure 5(F)) to the necroptotic mode (green circle in Figure 5(F)) but it is followed by back-transition to the apoptotic mode (blue triangle in Figure 5(F)) in response to the declining OV level. The second wave of OV stimuli also induces a loop of transition series Ta→Tn→Ta by the forward and backward movements.
The fundamental structures of four phenotypic modes in response to various bortezomib levels in the absence and presence of OVs are represented in Figure 3. Here, we further scrutinized the fundamental structures of four phenotypic modes, discerning their responses to diverse levels of oncolytic viruses in both the absence and presence of bortezomib. Specifically, Figure 6 illustrates the steady states of three key intracellular variables (NFκB, BAX, and RIP1). These plots delineate the behavior of these variables under varying [oHSV] conditions in the absence (B=0 in Figure 6(A)) and presence (B=1 in Figure 6(B)) of bortezomib.
NFκB consistently sustains high levels irrespective of [oHSV], while the BAX level remains consistently low (Figure 6(A)). However, the level of RIP1 increases as the [oHSV] increases. This particular state signifies the activation of oncolysis signaling (depicted within the green box in Figure 6(A)) under conditions of elevated [oHSV] levels. Conversely, when [oHSV] stimulation is limited, the cell death program is not initiated, leading to the anti-apoptosis phenotype (illustrated by the yellow box in Figure 6(A)). However, the intracellular dynamics of NFκB, BAX, and RIP1 undergoes a significant transformation in the presence of bortezomib in response to [oHSV] stimuli. In the presence of bortezomib within the system, we observe elevated levels of BAX, coupled with low levels of both NFκB and RIP1 in response to lower levels of [oHSV] (Figure 6(B)). This dynamic pattern prevents the initiation of bortezomib-induced apoptosis, as the key apoptotic gene, BAX, remains suppressed. As the concentration of [oHSV] increases, we have a dramatic shift: NFκB and RIP1 are up-regulated, while the level of BAX remains low, leading anti-apoptosis (yellow box in Figure 6(B)). As the [oHSV] increases further, the RIP1 level transits to an overexpression state, triggering necroptosis (pink box in Figure 6(B)).
In the main text model, certain parameters lack experimental data and may influence simulation results. We considered all parameters in the intracellular signaling (B, v, λB, α, σ1, σ9, σ4, σ7, σ2, σ10, σ5, ω1, σ8, σ3, σ11, σ6, ω2, σ12, σ13, ω3) for sensitivity analysis. We examined the sensitivity of intracellular molecules' activities (IκB, NFκB, BAX, and RIP1) at t = 0.1, 1, 10,100 to these parameters. A range was selected for each parameter, divided into 10,000 intervals of uniform length. For each of the 20 parameters of interest, a partial rank correlation coefficient (PRCC) value was calculated. PRCC values range between -1 and 1, with the sign determining whether an increase in the parameter value will decrease (-) or increase (+) activities of the intracellular dynamics at a given time. The sensitivity analysis, as described below, followed the method outlined in [104]. In Figure 7, The BAX (activities of apoptosis inducer) level is positively correlated to ω1, σ8, σ3, σ11 but negatively correlated to ω2. In a similar fashion, the RIP1 (activities of necroptosis or oncolysis inducer) activity is positively correlated to σ7, σ2, σ10, σ12, σ13 but negatively correlated to ω1, ω3. These results are summarized in Table 2.
Parameter | B | v | λB | α | σ1 | σ9 |
PRCC(S, 0.1 h) | 0.28187 | −0.03828 | 0.26953 | −0.33367 | 0.94918 | 0.50929 |
PRCC(F, 0.1 h) | −0.02305 | 0.01961 | −0.02248 | 0.04351 | −0.27969 | −0.04106 |
PRCC(A, 0.1 h) | 0.00822 | 0.00416 | −0.01449 | −0.01347 | −0.01005 | −0.00355 |
PRCC(R, 0.1 h) | −0.00527 | 0.13183 | −0.00513 | 0.02775 | −0.06686 | −0.01338 |
PRCC(S, 1 h) | 0.26056 | −0.02550 | 0.25868 | −0.29694 | 0.82422 | 0.73676 |
PRCC(F, 1 h) | −0.07284 | 0.02242 | −0.08059 | 0.09171 | −0.48167 | −0.32707 |
PRCC(A, 1 h) | 0.02008 | 0.00165 | 0.01746 | −0.02042 | 0.11314 | 0.05378 |
PRCC(R, 1 h) | −0.03627 | 0.11745 | −0.04266 | 0.06355 | −0.27849 | −0.14751 |
PRCC(S, 10 h) | 0.37318 | −0.03499 | 0.35211 | −0.35990 | 0.55805 | 0.62897 |
PRCC(F, 10 h) | −0.09981 | 0.02151 | −0.10770 | 0.12735 | −0.32377 | −0.35500 |
PRCC(A, 10 h) | 0.04337 | 0.00249 | 0.04355 | −0.04101 | 0.16069 | 0.14905 |
PRCC(R, 10 h) | −0.06094 | 0.07032 | −0.07396 | 0.10494 | −0.23546 | −0.24755 |
PRCC(S, 100 h) | 0.38143 | −0.03454 | 0.36085 | −0.36390 | 0.51007 | 0.58888 |
PRCC(F, 100 h) | −0.09180 | 0.01882 | −0.10493 | 0.12160 | −0.27286 | −0.30714 |
PRCC(A, 100 h) | 0.04653 | −0.00022 | 0.02574 | −0.01787 | 0.11811 | 0.10076 |
PRCC(R, 100 h) | −0.04955 | 0.05732 | −0.06608 | 0.09572 | −0.17990 | −0.19896 |
Minimum | 0 | 0 | 0.19 | 3 | 0.4 | 0.1 |
Baseline | 0.5 | 0.01 | 1.9 | 30 | 4 | 1 |
Maximum | 5 | 0.1 | 5.7 | 60 | 8 | 5 |
Parameter | σ4 | σ7 | σ2 | σ10 | σ5 | ω1 |
PRCC(S, 0.1 h) | −0.17767 | −0.06597 | −0.24638 | −0.03404 | −0.00584 | 0.02214 |
PRCC(F, 0.1 h) | 0.00846 | 0.63726 | 0.96485 | 0.53649 | −0.19325 | −0.15541 |
PRCC(A, 0.1 h) | −0.01288 | −0.03460 | −0.19312 | −0.03172 | −0.00515 | 0.00624 |
PRCC(R, 0.1 h) | 0.02074 | 0.26024 | 0.79329 | 0.14504 | −0.05250 | −0.04657 |
PRCC(S, 1 h) | −0.40421 | −0.20533 | −0.50535 | −0.30275 | 0.12012 | 0.16696 |
PRCC(F, 1 h) | 0.12788 | 0.52405 | 0.85069 | 0.68899 | −0.35946 | −0.55150 |
PRCC(A, 1 h) | −0.02169 | −0.16545 | −0.45533 | −0.24162 | 0.08285 | 0.14085 |
PRCC(R, 1 h) | 0.06123 | 0.32027 | 0.72956 | 0.46364 | −0.19797 | −0.28509 |
PRCC(S, 10 h) | −0.36729 | −0.27416 | −0.40630 | −0.34399 | 0.15168 | 0.53178 |
PRCC(F, 10 h) | 0.18956 | 0.49155 | 0.67193 | 0.55311 | −0.26960 | −0.81199 |
PRCC(A, 10 h) | −0.06678 | −0.25292 | −0.36120 | −0.29522 | 0.11705 | 0.42062 |
PRCC(R, 10 h) | 0.13050 | 0.35427 | 0.54956 | 0.43078 | −0.19729 | −0.69605 |
PRCC(S, 100 h) | −0.33893 | −0.24777 | −0.38422 | −0.30025 | 0.12914 | 0.56675 |
PRCC(F, 100 h) | 0.17244 | 0.45223 | 0.63376 | 0.50030 | −0.23642 | −0.82047 |
PRCC(A, 100 h) | −0.06158 | −0.17745 | −0.25095 | −0.20439 | 0.06632 | 0.38276 |
PRCC(R, 100 h) | 0.11273 | 0.30119 | 0.48277 | 0.35914 | −0.15850 | −0.70620 |
Minimum | 0.22 | 0.007 | 0.133 | 0.1 | 0.1 | 0.03 |
Baseline | 2.2 | 0.07 | 1.33 | 1 | 1 | 0.3 |
Maximum | 6.6 | 1.4 | 6.65 | 5 | 5 | 1.5 |
Parameter | σ8 | σ3 | σ11 | σ6 | ω2 | σ12 |
PRCC(S, 0.1 h) | 0.00418 | 0.00133 | 0.00721 | −0.00730 | 0.00880 | −0.00095 |
PRCC(F, 0.1 h) | 0.00491 | −0.02376 | −0.00479 | −0.01447 | 0.00070 | 0.00029 |
PRCC(A, 0.1 h) | 0.94584 | 0.94967 | 0.43179 | −0.14146 | −0.02909 | 0.01632 |
PRCC(R, 0.1 h) | −0.00100 | −0.00278 | 0.00123 | 0.00560 | 0.00341 | 0.99028 |
PRCC(S, 1 h) | −0.00087 | 0.00042 | −0.00427 | 0.00687 | 0.00177 | 0.00779 |
PRCC(F, 1 h) | 0.00618 | −0.00739 | 0.00563 | −0.01924 | 0.00263 | −0.00287 |
PRCC(A, 1 h) | 0.91506 | 0.83297 | 0.64729 | −0.35019 | −0.17798 | 0.00315 |
PRCC(R, 1 h) | 0.00334 | −0.01925 | 0.00145 | 0.00094 | 0.00290 | 0.71206 |
PRCC(S, 10 h) | −0.00308 | −0.01061 | −0.00714 | 0.00995 | −0.01861 | 0.01484 |
PRCC(F, 10 h) | 0.01215 | −0.00269 | 0.00844 | −0.01508 | 0.00412 | −0.01259 |
PRCC(A, 10 h) | 0.88966 | 0.56291 | 0.46346 | −0.27714 | −0.70996 | 0.00142 |
PRCC(R, 10 h) | 0.01655 | −0.01123 | 0.00754 | −0.01035 | −0.00245 | 0.30818 |
PRCC(S, 100 h) | −0.00690 | −0.01292 | −0.00574 | 0.00764 | −0.01664 | 0.01243 |
PRCC(F, 100 h) | 0.01736 | 0.00037 | 0.00739 | −0.01191 | 0.00101 | −0.01150 |
PRCC(A, 100 h) | 0.81841 | 0.37472 | 0.30075 | −0.17214 | −0.87775 | −0.00371 |
PRCC(R, 100 h) | 0.02093 | −0.00869 | 0.00635 | −0.01199 | −0.00584 | 0.25393 |
Minimum | 0.00033 | 0.0111 | 0.1 | 0.1 | 0.002 | 0.01 |
Baseline | 0.0033 | 0.111 | 1 | 1 | 0.02 | 0.1 |
Maximum | 0.99 | 1.11 | 5 | 5 | 0.2 | 1 |
Parameter | σ13 | ω3 | ||||
PRCC(S, 0.1 h) | 0.01202 | −0.00260 | ||||
PRCC(F, 0.1 h) | 0.01100 | −0.00583 | ||||
PRCC(A, 0.1 h) | 0.00190 | 0.00014 | ||||
PRCC(R, 0.1 h) | 0.84920 | −0.25007 | ||||
PRCC(S, 1 h) | 0.00730 | 0.00780 | ||||
PRCC(F, 1 h) | 0.01072 | −0.01173 | ||||
PRCC(A, 1 h) | 0.01013 | 0.00761 | ||||
PRCC(R, 1 h) | 0.82148 | −0.49732 | ||||
PRCC(S, 10 h) | 0.00154 | 0.01567 | ||||
PRCC(F, 10 h) | 0.00974 | −0.01014 | ||||
PRCC(A, 10 h) | 0.00512 | 0.01810 | ||||
PRCC(R, 10 h) | 0.68681 | −0.75844 | ||||
PRCC(S, 100 h) | −0.00381 | 0.01276 | ||||
PRCC(F, 100 h) | 0.01531 | −0.00586 | ||||
PRCC(A, 100 h) | 0.00191 | −0.00058 | ||||
PRCC(R, 100 h) | 0.61909 | −0.74965 | ||||
Minimum | 0.013 | 0.0139 | ||||
Baseline | 0.13 | 0.139 | ||||
Maximum | 1.3 | 1.39 |
To investigate the effect of injection sequence, we consider two administration schedules: (ⅰ) Bortezomib + oncolytic-virus (bortezomib first) (Figure 8(A)–(D)) and (ⅱ) oncolytic-virus + Bortezomib (OV first) (Figure 8(E)–(H)). The approach involves partitioning the 30-day interval evenly, allocating 15 days for the first therapeutic administration followed by an additional 15 days for the subsequent treatment. The objective is to discern the therapeutic administration sequence that yields a synergistic effect in the eradication of cancer cells. In these cases, the cumulative amounts of OV and bortezomib are the same (∑t=30t=0uV=0.5 and ∑t=30t=0uB=0.75). Figure 8(A), (B) show the time courses of injection rate and concentration of OV (red curves) and bortezomib (blue curves) of the case where bortezomib is administered first. Figure 8(C) shows the time courses of intracellular signaling (NFκB (F), BAX (A), and RIP1 (R)) in response to bortezomib first administration. Simulations reveal a scenario where cancer cells might elude the apoptotic pathway over the 15-day course of bortezomib administration (as depicted in Figure 8(C)). Notably, the initiation of oncolysis becomes feasible only upon the subsequent implementation of oncolytic virus therapy (as illustrated in Figure 8(D)). This observation prompted to explore an alternative scenario, involving the administration of oncolytic virus (OV) prior to bortezomib. Figure 8(E), (F) display the time courses of injection rate and concentration of OV (red curves) and bortezomib (blue curves) of the OV first administration. Corresponding time course evolutions of intracellular signaling variables (NFκB (F), BAX (A), and RIP1 (R)) are depicted in Figure 8(G). Significantly, the induction of oncolysis ensues as a result of OV administration, subsequently leading to necroptosis following bortezomib administration. Despite the equal administration of OV and bortezomib, distinct profiles emerge within the cancer cell population (depicted in Figure 8(D), (H)) as well as intracellular signaling dynamics (illustrated in Figure 8(C), (G)). In the scenario where bortezomib is administered first, cancer cells can evade the apoptotic phenotype. Conversely, when OV takes precedence in administration before bortezomib, oncolysis followed by necroptosis can be promoted. Consequently, the synergistic effect of these therapeutic agents is achieved by administering OV before bortezomib. However, constant administration of bortezomib and OV, as illustrated in Figure 8, poses challenges within a clinical context, encompassing concerns related to drug toxicity and practical feasibility. Consequently, we intend to explore a scenario in which the administration durations for both OV and bortezomib are confined to a single day, with a sole administration of OV.
We examined a periodic bortezomib administration and a single OV administration within the time interval [0,30]. The 30-day duration is partitioned where OV is introduced for a single day, followed by a 14-day interval of rest. Subsequently, the remaining days are divided into five segments, during which bortezomib is administered five times. Given its inherent replication capacity, OV is administered solely on the initial phase within a 15-day interval. We have examined six distinct injection schedules, comprising a single administration of OV and five instances of bortezomib. The total amount of OV and bortezomib is same as in the constant sequential OV and bortezomib administration above. We denote OV and bortezomib as 'V' and 'B', respectively. For instance, 'BBBBBV' reflects five administration of bortezomib within the first 15 days and one time OV administration in the remaining 15 days. Figure 9(A)–(D) show the time courses of injection rates (uB,uV), concentration of OV and bortezomib, concentration of intracellular signaling variables (NFκB, BAX, and RIP1), and cancer cells (uninfected, infected, and total cancer cells) of the 'BBBBBV', respectively. Observe that in Figure 9(C), there are two phenotypes: apoptosis (blue region) and oncolysis (green region). Remarkably, despite employing an equivalent quantity of bortezomib as seen in the constant sequential administration of OV and bortezomib shown in Figure 8, distinct phenotypic outcomes can be elicited. Hence, administering a substantial quantity separately proves more effective than delivering a smaller dose over an extended period. Figure 9(E)–(H) showcase the outcomes for the 'VBBBBB' scenario, revealing a clear manifestation of the synergistic effect between OV and bortezomib. Upon each bortezomib injection (from day 15th), a rapid increase in the level of OV becomes evident (depicted in Figure 9(F)). This phenomenon arises from the combined action of OV and bortezomib treatment, which induces a necroptotic phenotype and consequently enhances viral infections. As a result, the replication term (second term in Eq (2.8)) contributes to the substantial increase in OV levels. In Figure 9(G), a distinct alternation between necroptosis and oncolysis emerges starting at the 15-day mark, coinciding with bortezomib injections. However, the discernible effect on the cancer cell population remains limited, primarily attributed to the supplementary role played by bortezomib, which acts as an enhancer, while OV alone displays only modest infectivity towards cancer cells (as demonstrated in Figures 8(H) and 9(H)). In the context of single administering OV and five bortezomib injections, an intriguing query arises: Should the OV injection be administered solely at the beginning or the end of bortezomib administration? To address this concern, we strategically place the OV injection between the bortezomib administrations. The temporal dynamics of phenotypic changes corresponding to diverse injection schedules are illustrated in Figure 9(I). Figure 9(I) illustrates the discrete phenotypic states of cancer cells resulting from various injection schedules. Figure 9(J) displays the normalized cancer cell (green: uninfected cancer cells; pink: infected cancer cells; red: total cancer cells (uninfected + infected) outcomes across all scenarios at the final time point. Evidently, the 'BBBBBV' scenario exhibits the least effectiveness in eradicating cancer cells. Conversely, the 'VBBBBB' scenario emerges as the most potent in targeting uninfected cancer cells for elimination. Generally, the strategy commencing with OV demonstrates a superior anti-cancer effect compared to the approach initiated with bortezomib. Nonetheless, an important query arises regarding the optimality of the strategy derived from these findings. Up to this point, the injection timings have been manually set, and the cumulative dosage of anti-cancer drugs has been predetermined. Furthermore, it's noteworthy that the bortezomib dosage surpasses our reference threshold of 1, consequently impacting patient toxicity levels.
As a result, the subsequent section will elucidate how the application of optimal control theory can yield favorable outcomes, further utilizing minimal quantities of OV and bortezomib.
Now, we investigate the optimal control problems for various objective functions in Eqs (2.12) and (2.13). As observed from the previous results, it is evident that administering OV first is significantly more effective compared to administering bortezomib first. The simulation outcomes and analysis support the idea that initiating the treatment with OV leads to superior therapeutic results in combating cancer progression and inducing necroptotic phenotype. This finding can have important implications for designing optimized treatment strategies in clinical settings, prioritizing OV administration for improved efficacy and better patient outcomes. Regarding the objective function, as described in the previous section, we will examine the following results using two different objective functions (Eqs (2.12) and (2.13)). Each objective function represents a distinct therapeutic goal, and by analyzing the outcomes based on these different criteria, we aim to gain comprehensive insights into the effectiveness of the treatment strategies in different scenarios. These objective functions allow us to assess the impact of the interventions from multiple perspectives, providing a more comprehensive evaluation of the treatment outcomes.
1) Quadratic controls
In this therapeutic strategy, we investigate the optimal control problem for objective function Eq (2.12). Figure 10(A), (B) shows the time courses of the injection rate and concentration of OV and bortezomib, respectively. We set that the state stays at necroptotic phenotype after injecting bortezomib (day 15th). That is, the level of bortezomib never falls under the threshold (thB=0.1). In Figure 9(G), necroptotic and oncolytic mode alternately appeared even though the total amount of bortezomib is 0.75. However, with the application of optimal control theory, necroptotic phenotype is maintained from the moment bortezomib is injected (Figure 10(C)). It is worth noting that with optimal control, total amount of bortezomib is just 0.6733. Because we set the upper bound of the control when applying the optimal control theory, the level of bortezomib does not exceed the reference value (Figure 10(B)). However, in this case, one-time OV is coupled with frequent bortezomib administrations, the inconvenience arises from the need for repetitive hospital visits solely for the purpose of bortezomib injections. Hence, we present an alternative approach, incorporating intervals of respite between bortezomib injections (depicted in Figure 10(E)–(H)). This strategy involves a one-time OV coupled with intermittent bortezomib administration. In this latter scenario, the injection intervals are designed, ensuring that the bortezomib dosage threshold will not be exceeded that could potentially lead to toxicity. The total amount of bortezomib, in this case, is calculated at 0.2823, significantly lower by over fifty percent compared to the previous scenario's value of 0.6733. Furthermore, there is a notable reduction in the total amount of OV, diminishing from 0.5 as seen in Figures 8 and 9, to 0.2938 as demonstrated in Figure 10. In the second scenario, the injection strategy of bortezomib shows four injections in 15 days. The possible actual administration of bortezomib could be bolus intravenous injection twice weekly for two weeks (days 1, 4, 8, and 11) followed by a ten-day rest period [105,106]. Hence, our mathematical model utilizing optimal control theory, aligns with the practical drug administration protocols and has the potential to provide valuable insights into optimizing drug utilization within real treatment scenarios.
2) Linear controls with constraint
In this therapeutic approach, we consider the impact of initially administering OV followed by bortezomib, employing a sigmoid function to effectively steer cells toward a necroptotic phenotype. We divided the entire time into two periods: the period of injecting only OV and the period of injecting only bortezomib. In the first integration term (in Eq (2.13)), a term for uninfected cancer cells (x(t)) is added because the main purpose is to infect cancer cells through OV. The objective function of the second integration term (in Eq (2.13)) considers a sigmoid term for bortezomib to keep the cells in the necroptotic phenotype. Figure 11 are the results for Eq (2.13). Figure 11(A), (B) shows the time courses of injection rate (uV and uB) and concentration of OV (v) and bortezomib (B), respectively. Figure 11(C), (D) illustrate the time courses of intracellular signaling (NFκB, BAX, RIP1) and cancer cell population (uninfected, infected, and total cancer cells), respectively. Bortezomib continues to maintain a level that satisfies the necroptotic phenotype (blue curves in Figure 11(B)). As a result, after 15 days, the necroptosis phenotype is maintained and more cancer cells can be infected (Figure 11(C), (D)). The mechanism of quadratic controls necessitates the dynamic adjustment of infusion rates over time, as both uV and uB exhibit continuous dynamics when the necessary conditions for optimal control are computed. In contrast, linear controls result to either 0 or the maximum injection rate which is due to the calculation of necessary conditions in the optimal control formulation. Consequently, linear controls provide practical administration protocol as it do not require temporal adjustments to the injection rate.
The optimal control framework herein is based on the mathematical models developed by [45,46] describing the intracellular dynamics and tumor suppression in response to a combination therapy involving bortezomib and OV in cancer patients. While bortezomib is quite effective in killing cancer cells in general, it can cause strong toxicity and serious side effects. One of our goals of this study is to analyze the cellular system behaviors in response to bortezomib-OV combination therapy via a delicate intracellular signaling pathway (NFkB-Bcl-2-BAX-RIP1). Another goal is to find optimal strategies that effectively kill cancer cells by OV and bortezomib through both apoptotic and necroptotic phenotype with minimal costs (side effects + administrative costs) using optimal control theory [46]. The associated optimal control problems are formulated such that the overall cancer cell population and the total cost of the two anticancer drugs are minimized.
Yoo et al. showed that OV replication, thus anti-tumor efficacy, can be improved by the unfolded protein response (UPR) after bortezomib treatment, inducing synergistic cancer cell killing [6]. Kim et al. [46] introduced the apoptosis and necroptosis cell death mechanism involved in this combination (OV + bortezomib) therapy. In this work, a detailed analysis of the model in this work suggests that injection of OVs followed by bortezomib can effectively increase anti-tumor efficacy due to the higher infection rate of cancer cells from the oncolytic mode and superior killing capacity of tumor cells from combined cell death (necroptosis + oncolysis) phase while the OV treatment at a latter time induces only oncolysis or a mild killing phase (apoptosis + oncolysis), leading to lower killing rates (Figures 8 and 9). Treatment of a tumor with OVs at an earlier time usually leads to better anti-tumor efficacy [6] and our study supports the idea of effectiveness of initial OV treatment even in the combination therapy by revealing the dynamics of detailed sub-cellular cell death program at an intracellular level. Relative order of OVs relative to bortezomib treatment in a combination therapy induces the cellular transition from the severe ON (oncolysis + necroptosis) mode to the intermediate AON phase (apoptosis + oncolysis + necroptosis), and to mild AO (apoptosis + oncolysis) phase (Figure 9). Therefore, delays in OV treatment decrease anti-tumor efficacy, suggesting the importance of OV injection time. Excessive use of OVs and bortezomib also has to be limited due to safety issues and drug-associated side effects, respectively, in addition to a need to minimize administrative costs. Anti-cancer strategies with optimal control in our work suggest that both the necroptosis-dominant phase from frequent injection of bortezomib and a mixed (necroptosis + oncolysis) mode from less frequent injection mode of bortezomib can effectively kill cancer cells in the 2nd cell death program a combination therapy with initial administration of OVs (Figures 10 and 11). Optimally controlled injection strategy of OVs and bortezomib in this work may lead to better understanding of fundamental mechanisms of the cell death program and effective anti-cancer efficacy in various cancers [105,106].
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
This work was supported by the National Research Foundation of Korea (NRF) grant funded by the Korea government (MSIT) (No. NRF-2021R1A2C1010891) (Y.K.).
The authors declare there is no conflict of interest.
[1] |
Maldonado-Soto AR, Oakley DH, Wichterle H, et al. (2014) Stem cells in the nervous system. Am J Phys Med Rehabil 93: S132-144. https://doi:10.3389/fcell.2020.00533 ![]() |
[2] |
Mira H, Morante J (2020) Neurogenesis from embryo to adult-lessons from flies and mice. Front Cell Dev Biol 8: 533. https://doi:10.1097/PHM.0000000000000111 ![]() |
[3] |
Kudla J, Becker D, Grill E, et al. (2018) Advances and current challenges in calcium signaling. New Phytol 218: 414-431. https://doi:10.1111/nph.14966 ![]() |
[4] |
Moreau M, Leclerc C (2004) The choice between epidermal and neural fate: a matter of calcium. Int J Dev Biol 48: 75-84. https://doi:10.1387/ijdb.15272372 ![]() |
[5] |
Somasundaram A, Shum AK, McBride HJ, et al. (2014) Store-operated CRAC channels regulate gene expression and proliferation in neural progenitor cells. J Neurosci 34: 9107-9123. https://doi:10.1523/JNEUROSCI.0263-14.2014 ![]() |
[6] |
Colella M, Gerbino A, Hofer AM, et al. (2016) Recent advances in understanding the extracellular calcium-sensing receptor. F1000Res 5. https://doi:10.12688/f1000research.8963.1 ![]() |
[7] |
Ermakov A, Daks A, Fedorova O, et al. (2018) Ca2+-depended signaling pathways regulate self-renewal and pluripotency of stem cells. Cell Biol Int 42: 1086-1096. https://doi:10.1002/cbin.10998 ![]() |
[8] |
Lee JH, Lee JE, Kahng JY, et al. (2018) Human glioblastoma arises from subventricular zone cells with low-level driver mutations. Nature 560: 243-247. https://doi:10.1038/s41586-018-0389-3 ![]() |
[9] |
Barami K (2007) Biology of the subventricular zone in relation to gliomagenesis. J Clin Neurosci 14: 1143-1149. https://doi:10.1016/j.jocn.2007.04.009 ![]() |
[10] |
Quinones-Hinojosa A, Chaichana K (2007) The human subventricular zone: a source of new cells and a potential source of brain tumors. Exp Neurol 205: 313-324. https://doi:10.1016/j.expneurol.2007.03.016 ![]() |
[11] |
Recht L, Jang T, Savarese T, et al. (2003) Neural stem cells and neuro-oncology: quo vadis?. J Cell Biochem 88: 11-19. https://doi:10.1002/jcb.10208 ![]() |
[12] |
Zarco N, Norton E, Quinones-Hinojosa A, et al. (2019) Overlapping migratory mechanisms between neural progenitor cells and brain tumor stem cells. Cell Mol Life Sci 76: 3553-3570. https://doi:10.1007/s00018-019-03149-7 ![]() |
[13] | Rodriguez SMB, Staicu GA, Sevastre AS, et al. (2022) Glioblastoma stem cells-useful tools in the battle against cancer. Int J Mol Sci 23. https://doi:10.3390/ijms23094602 |
[14] |
Samanta K, Parekh AB (2017) Spatial Ca2+ profiling: decrypting the universal cytosolic Ca2+ oscillation. J Physiol 595: 3053-3062. https://doi:10.1113/JP272860 ![]() |
[15] |
Pinto MC, Kihara AH, Goulart VA, et al. (2015) Calcium signaling and cell proliferation. Cell Signal 27: 2139-2149. https://doi:10.1016/j.cellsig.2015.08.006 ![]() |
[16] |
Berridge MJ, Bootman MD, Roderick HL (2003) Calcium signalling: dynamics, homeostasis and remodelling. Nat Rev Mol Cell Biol 4: 517-529. https://doi:10.1038/nrm1155 ![]() |
[17] | Marchant JS (2019) Ca2+ signaling and regeneration. CSH Perspect Biol 11: a035485. https://doi:10.1101/cshperspect.a035485 |
[18] |
Hao B, Webb SE, Miller AL, et al. (2016) The role of Ca2+ signaling on the self-renewal and neural differentiation of embryonic stem cells (ESCs). Cell Calcium 59: 67-74. https://doi:10.1016/j.ceca.2016.01.004 ![]() |
[19] |
Toth AB, Shum AK, Prakriya M (2016) Regulation of neurogenesis by calcium signaling. Cell Calcium 59: 124-134. https://doi:10.1016/j.ceca.2016.02.011 ![]() |
[20] |
Li H, Rao A, Hogan PG (2011) Interaction of calcineurin with substrates and targeting proteins. Trends Cell Biol 21: 91-103. https://doi:10.1016/j.tcb.2010.09.011 ![]() |
[21] |
Morgan PJ, Hubner R, Rolfs A, et al. (2013) Spontaneous calcium transients in human neural progenitor cells mediated by transient receptor potential channels. Stem Cells Dev 22: 2477-2486. https://doi:10.1089/scd.2013.0061 ![]() |
[22] | Rosenberg SS, Spitzer NC (2011) Calcium signaling in neuronal development. CSH Perspect Biol 3: a004259. https://doi:10.1101/cshperspect.a004259 |
[23] |
Parekh AB (2008) Ca2+ microdomains near plasma membrane Ca2+ channels: impact on cell function. J Physiol 586: 3043-3054. https://doi:10.1113/jphysiol.2008.153460 ![]() |
[24] |
Petersen OH, Courjaret R, Machaca K (2017) Ca2+ tunnelling through the ER lumen as a mechanism for delivering Ca2+ entering via store-operated Ca2+ channels to specific target sites. J Physiol 595: 2999-3014. https://doi:10.1113/JP272772 ![]() |
[25] | Bootman MD, Bultynck G (2020) Fundamentals of cellular calcium signaling: a primer. CSH Perspect Biol 12: a038802. https://doi:10.1101/cshperspect.a038802 |
[26] | Barak P, Parekh AB (2020) Signaling through Ca2+ microdomains from store-operated CRAC channels. CSH Perspect Biol 12: a035097. https://doi:10.1101/cshperspect.a035097 |
[27] |
Bagur R, Hajnoczky G (2017) Intracellular Ca2+ sensing: its role in Calcium homeostasis and signaling. Mol Cell 66: 780-788. https://doi:10.1016/j.molcel.2017.05.028 ![]() |
[28] |
Duan PF, Yuan XP, Gan S, et al. (2020) Time delay induces saw-tooth calcium wave and transition in intracellular calcium oscillation. Chinese J Phys 66: 150-156. https://doi:10.1016/j.cjph.2020.05.003 ![]() |
[29] |
Lin L, Duan WL (2018) Transport phenomena in intracellular calcium dynamics driven by non-Gaussian noises. Physica A 492: 431-437. https://doi:10.1016/j.physa.2017.10.037 ![]() |
[30] | Duan WL, Zeng CH (2017) Statistics for anti-synchronization of intracellular calcium dynamics. Appl Math Comput 293: 611-616. https://doi:10.1016/j.amc.2016.07.041 |
[31] |
Zhang Y, Wu Y, Zhang M, et al. (2023) Synergistic mechanism between the endoplasmic reticulum and mitochondria and their crosstalk with other organelles. Cell Death Discov 9: 51. https://doi:10.1038/s41420-023-01353-w ![]() |
[32] |
Bruce JIE (2018) Metabolic regulation of the PMCA: role in cell death and survival. Cell Calcium 69: 28-36. https://doi:10.1016/j.ceca.2017.06.001 ![]() |
[33] |
Giladi M, Tal I, Khananshvili D (2016) Structural features of ion ransport and allosteric regulation in sodium-calcium exchanger (NCX) proteins. Front Physiol 7: 30. https://doi:10.3389/fphys.2016.00030 ![]() |
[34] |
Mammucari C, Raffaello A, Vecellio Reane D, et al. (2018) Mitochondrial calcium uptake in organ physiology: from molecular mechanism to animal models. Pflug Arch Eur J Phy 470: 1165-1179. https://doi:10.1007/s00424-018-2123-2 ![]() |
[35] |
Morgan AJ (2016) Ca2+ dialogue between acidic vesicles and ER. Biochem Soc Trans 44: 546-553. https://doi:10.1042/BST20150290 ![]() |
[36] |
Prakriya M, Lewis RS (2015) Store-operated Calcium channels. Physiol Rev 95: 1383-1436. https://doi:10.1152/physrev.00020.2014 ![]() |
[37] |
Parekh AB, Putney JW (2005) Store-operated calcium channels. Physiol Rev 85: 757-810. https://doi:10.1152/physrev.00057.2003 ![]() |
[38] |
Lintschinger B, Balzer-Geldsetzer M, Baskaran T, et al. (2000) Coassembly of Trp1 and Trp3 proteins generates diacylglycerol- and Ca2+-sensitive cation channels. J Biol Chem 275: 27799-27805. https://doi:10.1074/jbc.M002705200 ![]() |
[39] |
Liu X, Bandyopadhyay BC, Singh BB, et al. (2005) Molecular analysis of a store-operated and 2-acetyl-sn-glycerol-sensitive non-selective cation channel. Heteromeric assembly of TRPC1-TRPC3. J Biol Chem 280: 21600-21606. https://doi:10.1074/jbc.C400492200 ![]() |
[40] |
Santulli G, Lewis D, des Georges A, et al. (2018) Ryanodine receptor structure and function in health and disease. Subcell Biochem 87: 329-352. https://doi:10.1007/978-981-10-7757-9_11 ![]() |
[41] |
Lu L (2021) Mobilizing ER IP3 receptors as a mechanism to enhance calcium signaling. Cell Mol Immunol 18: 2284-2285. https://doi:10.1038/s41423-021-00725-5 ![]() |
[42] |
Prakriya M (2009) The molecular physiology of CRAC channels. Immunol Rev 231: 88-98. https://doi:10.1111/j.1600-065X.2009.00820.x ![]() |
[43] |
Berridge MJ (2016) The Inositol Trisphosphate/Calcium signaling pathway in health and disease. Physiol Rev 96: 1261-1296. https://doi:10.1152/physrev.00006.2016 ![]() |
[44] |
Parekh AB (2010) Store-operated CRAC channels: function in health and disease. Nat Rev Drug Discov 9: 399-410. https://doi:10.1038/nrd3136 ![]() |
[45] |
Meissner G (2017) The structural basis of ryanodine receptor ion channel function. J Gen Physiol 149: 1065-1089. https://doi:10.1085/jgp.201711878 ![]() |
[46] |
Wu HH, Brennan C, Ashworth R (2011) Ryanodine receptors, a family of intracellular calcium ion channels, are expressed throughout early vertebrate development. BMC Res Notes 4: 541. https://doi:10.1186/1756-0500-4-541 ![]() |
[47] |
Woll KA, Van Petegem F (2022) Calcium-release channels: structure and function of IP(3) receptors and ryanodine receptors. Physiol Rev 102: 209-268. https://doi:10.1152/physrev.00033.2020 ![]() |
[48] | Pulli I, Asghar MY, Kemppainen K, et al. (2018) Sphingolipid-mediated calcium signaling and its pathological effects. BBA Mol Cell Res 1865: 1668-1677. https://doi:10.1016/j.bbamcr.2018.04.012 |
[49] |
Berridge MJ, Lipp P, Bootman MD (2000) The versatility and universality of calcium signalling. Nat Rev Mol Cell Biol 1: 11-21. https://doi:10.1038/35036035 ![]() |
[50] |
Rapizzi E, Donati C, Cencetti F, et al. (2007) Sphingosine 1-phosphate receptors modulate intracellular Ca2+ homeostasis. Biochem Biophys Res Commun 353: 268-274. https://doi:10.1016/j.bbrc.2006.12.010 ![]() |
[51] |
Uhlen P, Fritz N, Smedler E, et al. (2015) Calcium signaling in neocortical development. Dev Neurobiol 75: 360-368. https://doi:10.1002/dneu.22273 ![]() |
[52] | Lewis RS (2020) Store-operated calcium channels: from function to structure and back again. CSH Perspect Biol 12: a035055. https://doi:10.1101/cshperspect.a035055 |
[53] |
Dolmetsch RE, Pajvani U, Fife K, et al. (2001) Signaling to the nucleus by an L-type calcium channel-calmodulin complex through the MAP kinase pathway. Science 294: 333-339. https://doi:10.1126/science.1063395 ![]() |
[54] |
Yanez M, Gil-Longo J, Campos-Toimil M (2012) Calcium binding proteins. Adv Exp Med Biol 740: 461-482. https://doi:10.1007/978-94-007-2888-2_19 ![]() |
[55] | Schwaller B (2010) Cytosolic Ca2+ buffers. CSH Perspect Biol 2: a004051. https://doi:10.1101/cshperspect.a004051 |
[56] |
Gilabert JA (2020) Cytoplasmic calcium buffering: an integrative crosstalk. Adv Exp Med Biol 1131: 163-182. https://doi:10.1007/978-3-030-12457-1_7 ![]() |
[57] |
Park YJ, Yoo SA, Kim M, et al. (2020) The role of calcium-calcineurin-NFAT signaling pathway in health and autoimmune diseases. Front Immunol 11: 195. https://doi:10.3389/fimmu.2020.00195 ![]() |
[58] |
Hogan PG, Chen L, Nardone J, et al. (2003) Transcriptional regulation by calcium, calcineurin, and NFAT. Genes Dev 17: 2205-2232. https://doi:10.1101/gad.1102703 ![]() |
[59] |
Dolmetsch RE, Xu K, Lewis RS (1998) Calcium oscillations increase the efficiency and specificity of gene expression. Nature 392: 933-936. https://doi:10.1038/31960 ![]() |
[60] |
Lilienbaum A, Israel A (2003) From calcium to NF-kappa B signaling pathways in neurons. Mol Cell Biol 23: 2680-2698. https://doi:10.1128/MCB.23.8.2680-2698.2003 ![]() |
[61] |
Groth RD, Coicou LG, Mermelstein PG, et al. (2007) Neurotrophin activation of NFAT-dependent transcription contributes to the regulation of pro-nociceptive genes. J Neurochem 102: 1162-1174. https://doi:10.1111/j.1471-4159.2007.04632.x ![]() |
[62] |
Groth RD, Mermelstein PG (2003) Brain-derived neurotrophic factor activation of NFAT (nuclear factor of activated T-cells)-dependent transcription: a role for the transcription factor NFATc4 in neurotrophin-mediated gene expression. J Neurosci 23: 8125-8134. https://doi:10.1523/JNEUROSCI.23-22-08125.2003 ![]() |
[63] |
Vashishta A, Habas A, Pruunsild P, et al. (2009) Nuclear factor of activated T-cells isoform c4 (NFATc4/NFAT3) as a mediator of antiapoptotic transcription in NMDA receptor-stimulated cortical neurons. J Neurosci 29: 15331-15340. https://doi:10.1523/JNEUROSCI.4873-09.2009 ![]() |
[64] |
Oliveria SF, Dell'Acqua ML, Sather WA (2007) AKAP79/150 anchoring of calcineurin controls neuronal L-type Ca2+ channel activity and nuclear signaling. Neuron 55: 261-275. https://doi:10.1016/j.neuron.2007.06.032 ![]() |
[65] |
Kar P, Samanta K, Kramer H, et al. (2014) Dynamic assembly of a membrane signaling complex enables selective activation of NFAT by Orai1. Curr Biol 24: 1361-1368. https://doi:10.1016/j.cub.2014.04.046 ![]() |
[66] |
Lewis RS (2003) Calcium oscillations in T-cells: mechanisms and consequences for gene expression. Biochem Soc Trans 31: 925-929. https://doi:10.1042/bst0310925 ![]() |
[67] |
Jairaman A, Yamashita M, Schleimer RP, et al. (2015) Store-operated Ca2+ release-activated Ca2+ channels regulate PAR2-activated Ca2+ signaling and cytokine production in airway epithelial cells. J Immunol 195: 2122-2133. https://doi:10.4049/jimmunol.1500396 ![]() |
[68] |
Gorlach A, Bertram K, Hudecova S, et al. (2015) Calcium and ROS: a mutual interplay. Redox Biol 6: 260-271. https://doi:10.1016/j.redox.2015.08.010 ![]() |
[69] |
Trebak M, Kinet JP (2019) Calcium signalling in T cells. Nat Rev Immunol 19: 154-169. https://doi:10.1038/s41577-018-0110-7 ![]() |
[70] |
Carvalho EJ, Stathopulos PB, Madesh M (2020) Regulation of Ca2+ exchanges and signaling in mitochondria. Curr Opin Physiol 17: 197-206. https://doi:10.1016/j.cophys.2020.08.010 ![]() |
[71] |
Wacquier B, Combettes L, Van Nhieu GT, et al. (2016) Interplay between intracellular Ca2+ oscillations and Ca2+-stimulated mitochondrial metabolism. Sci Rep 6: 19316. https://doi:10.1038/srep19316 ![]() |
[72] |
Demarque M, Represa A, Becq H, et al. (2002) Paracrine intercellular communication by a Ca2+- and SNARE-independent release of GABA and glutamate prior to synapse formation. Neuron 36: 1051-1061. https://doi:10.1016/s0896-6273(02)01053-x ![]() |
[73] |
Manent JB, Demarque M, Jorquera I, et al. (2005) A noncanonical release of GABA and glutamate modulates neuronal migration. J Neurosci 25: 4755-4765. https://doi:10.1523/JNEUROSCI.0553-05.2005 ![]() |
[74] | Catterall WA (2011) Voltage-gated calcium channels. CSH Perspect Biol 3: a003947. https://doi:10.1101/cshperspect.a003947 |
[75] |
Prevarskaya N, Skryma R, Shuba Y (2018) Ion channels in cancer: Are cancer hallmarks oncochannelopathies?. Physiol Rev 98: 559-621. https://doi:10.1152/physrev.00044.2016 ![]() |
[76] |
Pitt GS, Matsui M, Cao C (2021) Voltage-Gated Calcium Channels in Nonexcitable Tissues. Annu Rev Physiol 83: 183-203. https://doi:10.1146/annurev-physiol-031620-091043 ![]() |
[77] |
Pedersen SF, Nilius B (2007) Transient receptor potential channels in mechanosensing and cell volume regulation. Method Enzymol 428: 183-207. https://doi:10.1016/S0076-6879(07)28010-3 ![]() |
[78] |
Liberati S, Morelli MB, Amantini C, et al. (2014) Advances in transient receptor potential vanilloid-2 channel expression and function in tumor growth and progression. Curr Protein Pept Sci 15: 732-737. https://doi:10.2174/1389203715666140704115913 ![]() |
[79] | Alexander SP, Peters JA, Kelly E, et al. (2017) The concise guide to pharmacology 2017/18: Ligand–gated ion channels. Brit J Pharmacol 174: S130-S159. https://doi:10.1111/bph.13879 |
[80] |
Flint AC, Dammerman RS, Kriegstein AR (1999) Endogenous activation of metabotropic glutamate receptors in neocortical development causes neuronal calcium oscillations. P Natl Acad Sci USA 96: 12144-12149. https://doi:10.1073/pnas.96.21.12144 ![]() |
[81] |
Pathak MM, Nourse JL, Tran T, et al. (2014) Stretch-activated ion channel Piezo1 directs lineage choice in human neural stem cells. P Natl Acad Sci USA 111: 16148-16153. https://doi:10.1073/pnas.1409802111 ![]() |
[82] |
Beckervordersandforth R, Tripathi P, Ninkovic J, et al. (2010) In vivo fate mapping and expression analysis reveals molecular hallmarks of prospectively isolated adult neural stem cells. Cell Stem Cell 7: 744-758. https://doi:10.1016/j.stem.2010.11.017 ![]() |
[83] |
Lacar B, Young SZ, Platel JC, et al. (2011) Gap junction-mediated calcium waves define communication networks among murine postnatal neural progenitor cells. Eur J Neurosci 34: 1895-1905. https://doi:10.1111/j.1460-9568.2011.07901.x ![]() |
[84] |
Kraft A, Jubal ER, von Laer R, et al. (2017) Astrocytic calcium waves signal brain injury to neural stem and progenitor cells. Stem Cell Rep 8: 701-714. https://doi:10.1016/j.stemcr.2017.01.009 ![]() |
[85] |
Domenichini F, Terrie E, Arnault P, et al. (2018) Store-operated calcium entries control neural stem cell self-renewal in the adult brain subventricular zone. Stem Cells 36: 761-774. https://doi:10.1002/stem.2786 ![]() |
[86] |
Machaca K (2010) Ca2+ signaling, genes and the cell cycle. Cell Calcium 48: 243-250. https://doi:10.1016/j.ceca.2010.10.003 ![]() |
[87] |
Llorente V, Velarde P, Desco M, et al. (2022) Current understanding of the neural stem cell niches. Cells 11: 3002. https://doi:10.3390/cells11193002 ![]() |
[88] |
MacDougall MS, Clarke R, Merrill BJ (2019) Intracellular Ca2+ homeostasis and nuclear export mediate exit from naive pluripotency. Cell Stem Cell 25: 210-224. https://doi:10.1016/j.stem.2019.04.015 ![]() |
[89] |
Broxmeyer HE (2019) Hematopoietic stem cell intracellular levels of Ca2+ to the rescue! What next?. Cell Stem Cell 25: 171-173. https://doi:10.1016/j.stem.2019.07.003 ![]() |
[90] |
Umemoto T, Hashimoto M, Matsumura T, et al. (2018) Ca2+-mitochondria axis drives cell division in hematopoietic stem cells. J Exp Med 215: 2097-2113. https://doi:10.1084/jem.20180421 ![]() |
[91] |
Shi J, Parada LF, Kernie SG (2005) Bax limits adult neural stem cell persistence through caspase and IP3 receptor activation. Cell Death Differ 12: 1601-1612. https://doi:10.1038/sj.cdd.4401676 ![]() |
[92] |
Mikoshiba K (2015) Role of IP3 receptor signaling in cell functions and diseases. Adv Biol Regul 57: 217-227. https://doi:10.1016/j.jbior.2014.10.001 ![]() |
[93] |
Maric D, Maric I, Barker JL (2000) Developmental changes in cell calcium homeostasis during neurogenesis of the embryonic rat cerebral cortex. Cereb Cortex 10: 561-573. https://doi:10.1093/cercor/10.6.561 ![]() |
[94] |
Serrano-Perez MC, Fernandez M, Neria F, et al. (2015) NFAT transcription factors regulate survival, proliferation, migration, and differentiation of neural precursor cells. Glia 63: 987-1004. https://doi:10.1002/glia.22797 ![]() |
[95] |
Zhao D, Najbauer J, Garcia E, et al. (2008) Neural stem cell tropism to glioma: critical role of tumor hypoxia. Mol Cancer Res 6: 1819-1829. https://doi:10.1158/1541-7786.MCR-08-0146 ![]() |
[96] |
Yuan JP, Zeng W, Huang GN, et al. (2007) STIM1 heteromultimerizes TRPC channels to determine their function as store-operated channels. Nat Cell Biol 9: 636-645. https://doi:10.1038/ncb1590 ![]() |
[97] |
Di Giorgi-Gerevini V, Melchiorri D, Battaglia G, et al. (2005) Endogenous activation of metabotropic glutamate receptors supports the proliferation and survival of neural progenitor cells. Cell Death Differ 12: 1124-1133. https://doi:10.1038/sj.cdd.4401639 ![]() |
[98] |
Chan CM, Aw JT, Webb SE, et al. (2016) SOCE proteins, STIM1 and Orai1, are localized to the cleavage furrow during cytokinesis of the first and second cell division cycles in zebrafish embryos. Zygote 24: 880-889. https://doi:10.1017/S0967199416000216 ![]() |
[99] |
Petrik D, Myoga MH, Grade S, et al. (2018) Epithelial sodium channel regulates adult neural stem cell proliferation in a flow-dependent manner. Cell Stem Cell 22: 865-878. https://doi:10.1016/j.stem.2018.04.016 ![]() |
[100] |
Harraz OF, Altier C (2014) STIM1-mediated bidirectional regulation of Ca2+ entry through voltage-gated calcium channels (VGCC) and calcium-release activated channels (CRAC). Front Cell Neurosci 8: 43. https://doi:10.3389/fncel.2014.00043 ![]() |
[101] |
Naffaa MM (2024) Significance of the anterior cingulate cortex in neurogenesis plasticity: Connections, functions, and disorders across postnatal and adult stages. Bioessays 46: e2300160. https://doi:10.1002/bies.202300160 ![]() |
[102] |
Naffaa MM, Khan RR, Kuo CT, et al. (2023) Cortical regulation of neurogenesis and cell proliferation in the ventral subventricular zone. Cell Rep 42: 112783. https://doi:10.1016/j.celrep.2023.112783 ![]() |
[103] |
Naffaa MM (2025) Neurogenesis dynamics in the olfactory bulb: deciphering circuitry organization, function, and adaptive plasticity. Neural Regen Res 20: 1565-1581. https://doi:10.4103/NRR.NRR-D-24-00312 ![]() |
[104] | Naffaa MM, Yin HH (2023) A cholinergic signaling pathway underlying cortical circuit regulation of lateral ventricle quiescent neural stem cells. bioRxiv . https://doi:10.1101/2023.09.02.556037 |
[105] |
Lecca D, Fumagalli M, Ceruti S, et al. (2016) Intertwining extracellular nucleotides and their receptors with Ca2+ in determining adult neural stem cell survival, proliferation and final fate. Philos Trans R Soc Lond B Biol Sci 371: 20150433. https://doi:10.1098/rstb.2015.0433 ![]() |
[106] |
Braun N, Sevigny J, Mishra SK, et al. (2003) Expression of the ecto-ATPase NTPDase2 in the germinal zones of the developing and adult rat brain. Eur J Neurosci 17: 1355-1364. https://doi:10.1046/j.1460-9568.2003.02567.x ![]() |
[107] |
Messemer N, Kunert C, Grohmann M, et al. (2013) P2X7 receptors at adult neural progenitor cells of the mouse subventricular zone. Neuropharmacology 73: 122-137. https://doi:10.1016/j.neuropharm.2013.05.017 ![]() |
[108] |
Stafford MR, Bartlett PF, Adams DJ (2007) Purinergic receptor activation inhibits mitogen-stimulated proliferation in primary neurospheres from the adult mouse subventricular zone. Mol Cell Neurosci 35: 535-548. https://doi:10.1016/j.mcn.2007.04.013 ![]() |
[109] |
Genzen JR, Platel JC, Rubio ME, et al. (2009) Ependymal cells along the lateral ventricle express functional P2X(7) receptors. Purinergic Signal 5: 299-307. https://doi:10.1007/s11302-009-9143-5 ![]() |
[110] |
Leeson HC, Kasherman MA, Chan-Ling T, et al. (2018) P2X7 receptors regulate phagocytosis and proliferation in adult hippocampal and SVZ neural progenitor cells: implications for inflammation in neurogenesis. Stem Cells 36: 1764-1777. https://doi:10.1002/stem.2894 ![]() |
[111] |
Yasuda T, Bartlett PF, Adams DJ (2008) K(ir) and K(v) channels regulate electrical properties and proliferation of adult neural precursor cells. Mol Cell Neurosci 37: 284-297. https://doi:10.1016/j.mcn.2007.10.003 ![]() |
[112] |
Fiorio Pla A, Maric D, Brazer SC, et al. (2005) Canonical transient receptor potential 1 plays a role in basic fibroblast growth factor (bFGF)/FGF receptor-1-induced Ca2+ entry and embryonic rat neural stem cell proliferation. J Neurosci 25: 2687-2701. https://doi:10.1523/JNEUROSCI.0951-04.2005 ![]() |
[113] |
Young SZ, Taylor MM, Wu S, et al. (2012) NKCC1 knockdown decreases neuron production through GABA(A)-regulated neural progenitor proliferation and delays dendrite development. J Neurosci 32: 13630-13638. https://doi:10.1523/JNEUROSCI.2864-12.2012 ![]() |
[114] |
Stock K, Garthe A, de Almeida Sassi F, et al. (2014) The capsaicin receptor TRPV1 as a novel modulator of neural precursor cell proliferation. Stem Cells 32: 3183-3195. https://doi:10.1002/stem.1805 ![]() |
[115] |
Roberts JC, Davis JB, Benham CD (2004) [3H]Resiniferatoxin autoradiography in the CNS of wild-type and TRPV1 null mice defines TRPV1 (VR-1) protein distribution. Brain Res 995: 176-183. https://doi:10.1016/j.brainres.2003.10.001 ![]() |
[116] |
Khatri P, Obernier K, Simeonova IK, et al. (2014) Proliferation and cilia dynamics in neural stem cells prospectively isolated from the SEZ. Sci Rep 4: 3803. https://doi:10.1038/srep03803 ![]() |
[117] |
Platel JC, Dave KA, Bordey A (2008) Control of neuroblast production and migration by converging GABA and glutamate signals in the postnatal forebrain. J Physiol 586: 3739-3743. https://doi:10.1113/jphysiol.2008.155325 ![]() |
[118] |
Song M, Yu SP, Mohamad O, et al. (2017) Optogenetic stimulation of glutamatergic neuronal activity in the striatum enhances neurogenesis in the subventricular zone of normal and stroke mice. Neurobiol Dis 98: 9-24. https://doi:10.1016/j.nbd.2016.11.005 ![]() |
[119] |
Nochi R, Kato T, Kaneko J, et al. (2012) Involvement of metabotropic glutamate receptor 5 signaling in activity-related proliferation of adult hippocampal neural stem cells. Eur J Neurosci 36: 2273-2283. https://doi:10.1111/j.1460-9568.2012.08128.x ![]() |
[120] |
Deisseroth K, Singla S, Toda H, et al. (2004) Excitation-neurogenesis coupling in adult neural stem/progenitor cells. Neuron 42: 535-552. https://doi:10.1016/s0896-6273(04)00266-1 ![]() |
[121] |
Platel JC, Dave KA, Gordon V, et al. (2010) NMDA receptors activated by subventricular zone astrocytic glutamate are critical for neuroblast survival prior to entering a synaptic network. Neuron 65: 859-872. https://doi:10.1016/j.neuron.2010.03.009 ![]() |
[122] |
Cavaliere F, Urra O, Alberdi E, et al. (2012) Oligodendrocyte differentiation from adult multipotent stem cells is modulated by glutamate. Cell Death Dis 3: e268. https://doi:10.1038/cddis.2011.144 ![]() |
[123] |
Etxeberria A, Mangin JM, Aguirre A, et al. (2010) Adult-born SVZ progenitors receive transient synapses during remyelination in corpus callosum. Nat Neurosci 13: 287-289. https://doi:10.1038/nn.2500 ![]() |
[124] |
Hughes EG, Orthmann-Murphy JL, Langseth AJ, et al. (2018) Myelin remodeling through experience-dependent oligodendrogenesis in the adult somatosensory cortex. Nat Neurosci 21: 696-706. https://doi:10.1038/s41593-018-0121-5 ![]() |
[125] |
Xiao L, Ohayon D, McKenzie IA, et al. (2016) Rapid production of new oligodendrocytes is required in the earliest stages of motor-skill learning. Nat Neurosci 19: 1210-1217. https://doi:10.1038/nn.4351 ![]() |
[126] |
Larson VA, Zhang Y, Bergles DE (2016) Electrophysiological properties of NG2+ cells: Matching physiological studies with gene expression profiles. Brain Res 1638: 138-160. https://doi:10.1016/j.brainres.2015.09.010 ![]() |
[127] |
Buchanan J, Elabbady L, Collman F, et al. (2022) Oligodendrocyte precursor cells ingest axons in the mouse neocortex. Proc Natl Acad Sci USA 119: e2202580119. https://doi:10.1073/pnas.2202580119 ![]() |
[128] |
Chen TJ, Kula B, Nagy B, et al. (2018) In vivo regulation of oligodendrocyte precursor cell proliferation and differentiation by the AMPA-receptor subunit GluA2. Cell Rep 25: 852-861 e857. https://doi:10.1016/j.celrep.2018.09.066 ![]() |
[129] |
De Biase LM, Nishiyama A, Bergles DE (2010) Excitability and synaptic communication within the oligodendrocyte lineage. J Neurosci 30: 3600-3611. https://doi:10.1523/JNEUROSCI.6000-09.2010 ![]() |
[130] |
Khawaja RR, Agarwal A, Fukaya M, et al. (2021) GluA2 overexpression in oligodendrocyte progenitors promotes postinjury oligodendrocyte regeneration. Cell Rep 35: 109147. https://doi:10.1016/j.celrep.2021.109147 ![]() |
[131] |
Kougioumtzidou E, Shimizu T, Hamilton NB, et al. (2017) Signalling through AMPA receptors on oligodendrocyte precursors promotes myelination by enhancing oligodendrocyte survival. Elife 6: e28080. https://doi:10.7554/eLife.28080 ![]() |
[132] |
Buchanan J, da Costa NM, Cheadle L (2023) Emerging roles of oligodendrocyte precursor cells in neural circuit development and remodeling. Trends Neurosci 46: 628-639. https://doi:10.1016/j.tins.2023.05.007 ![]() |
[133] |
Marisca R, Hoche T, Agirre E, et al. (2020) Functionally distinct subgroups of oligodendrocyte precursor cells integrate neural activity and execute myelin formation. Nat Neurosci 23: 363-374. https://doi:10.1038/s41593-019-0581-2 ![]() |
[134] |
Rungta RL, Chaigneau E, Osmanski BF, et al. (2018) Vascular compartmentalization of functional hyperemia from the synapse to the pia. Neuron 99: 362-375 e364. https://doi:10.1016/j.neuron.2018.06.012 ![]() |
[135] |
Lu TY, Hanumaihgari P, Hsu ET, et al. (2023) Norepinephrine modulates calcium dynamics in cortical oligodendrocyte precursor cells promoting proliferation during arousal in mice. Nat Neurosci 26: 1739-1750. https://doi:10.1038/s41593-023-01426-0 ![]() |
[136] |
McKenzie IA, Ohayon D, Li H, et al. (2014) Motor skill learning requires active central myelination. Science 346: 318-322. https://doi:10.1126/science.1254960 ![]() |
[137] |
Kong H, Fan Y, Xie J, et al. (2008) AQP4 knockout impairs proliferation, migration and neuronal differentiation of adult neural stem cells. J Cell Sci 121: 4029-4036. https://doi:10.1242/jcs.035758 ![]() |
[138] | Young SZ, Platel JC, Nielsen JV, et al. (2010) GABA(A) Increases calcium in subventricular zone astrocyte-like cells through L- and T-type voltage-gated calcium channels. Front Cell Neurosci 4: 8. https://doi:10.3389/fncel.2010.00008 |
[139] |
Liu X, Wang Q, Haydar TF, et al. (2005) Nonsynaptic GABA signaling in postnatal subventricular zone controls proliferation of GFAP-expressing progenitors. Nat Neurosci 8: 1179-1187. https://doi:10.1038/nn1522 ![]() |
[140] |
Daynac M, Chicheportiche A, Pineda JR, et al. (2013) Quiescent neural stem cells exit dormancy upon alteration of GABAAR signaling following radiation damage. Stem Cell Res 11: 516-528. https://doi:10.1016/j.scr.2013.02.008 ![]() |
[141] |
Resende RR, Adhikari A (2009) Cholinergic receptor pathways involved in apoptosis, cell proliferation and neuronal differentiation. Cell Commun Signal 7: 20. https://doi:10.1186/1478-811X-7-20 ![]() |
[142] |
Atluri P, Fleck MW, Shen Q, et al. (2001) Functional nicotinic acetylcholine receptor expression in stem and progenitor cells of the early embryonic mouse cerebral cortex. Dev Biol 240: 143-156. https://doi:10.1006/dbio.2001.0453 ![]() |
[143] |
Williams BP, Milligan CJ, Street M, et al. (2004) Transcription of the M1 muscarinic receptor gene in neurons and neuronal progenitors of the embryonic rat forebrain. J Neurochem 88: 70-77. https://doi:10.1111/j.1471-4159.2004.02117.x ![]() |
[144] |
Zhou C, Wen ZX, Shi DM, et al. (2004) Muscarinic acetylcholine receptors involved in the regulation of neural stem cell proliferation and differentiation in vitro. Cell Biol Int 28: 63-67. https://doi:10.1016/j.cellbi.2003.10.005 ![]() |
[145] |
Paez-Gonzalez P, Asrican B, Rodriguez E, et al. (2014) Identification of distinct ChAT+ neurons and activity-dependent control of postnatal SVZ neurogenesis. Nat Neurosci 17: 934-942. https://doi:10.1038/nn.3734 ![]() |
[146] |
Narla ST, Klejbor I, Birkaya B, et al. (2013) Activation of developmental nuclear fibroblast growth factor receptor 1 signaling and neurogenesis in adult brain by alpha7 nicotinic receptor agonist. Stem Cells Transl Med 2: 776-788. https://doi:10.5966/sctm.2012-0103 ![]() |
[147] |
Sharma G (2013) The dominant functional nicotinic receptor in progenitor cells in the rostral migratory stream is the alpha3beta4 subtype. J Neurophysiol 109: 867-872. https://doi:10.1152/jn.00886.2012 ![]() |
[148] |
Mudo G, Belluardo N, Mauro A, et al. (2007) Acute intermittent nicotine treatment induces fibroblast growth factor-2 in the subventricular zone of the adult rat brain and enhances neuronal precursor cell proliferation. Neuroscience 145: 470-483. https://doi:10.1016/j.neuroscience.2006.12.012 ![]() |
[149] |
Wang J, Lu Z, Fu X, et al. (2017) Alpha-7 nicotinic receptor signaling pathway participates in the neurogenesis induced by ChAT-Positive neurons in the subventricular zone. Transl Stroke Res 8: 484-493. https://doi:10.1007/s12975-017-0541-7 ![]() |
[150] |
Jimenez E, Montiel M (2005) Activation of MAP kinase by muscarinic cholinergic receptors induces cell proliferation and protein synthesis in human breast cancer cells. J Cell Physiol 204: 678-686. https://doi:10.1002/jcp.20326 ![]() |
[151] |
Zhao WQ, Alkon DL, Ma W (2003) c-Src protein tyrosine kinase activity is required for muscarinic receptor-mediated DNA synthesis and neurogenesis via ERK1/2 and c-AMP-responsive element-binding protein signaling in neural precursor cells. J Neurosci Res 72: 334-342. https://doi:10.1002/jnr.10591 ![]() |
[152] |
Rosenblum K, Futter M, Jones M, et al. (2000) ERKI/II regulation by the muscarinic acetylcholine receptors in neurons. J Neurosci 20: 977-985. https://doi:10.1523/JNEUROSCI.20-03-00977.2000 ![]() |
[153] |
Goffart N, Kroonen J, Rogister B (2013) Glioblastoma-initiating cells: relationship with neural stem cells and the micro-environment. Cancers (Basel) 5: 1049-1071. https://doi:10.3390/cancers5031049 ![]() |
[154] |
Yuan X, Curtin J, Xiong Y, et al. (2004) Isolation of cancer stem cells from adult glioblastoma multiforme. Oncogene 23: 9392-9400. https://doi:10.1038/sj.onc.1208311 ![]() |
[155] |
Leclerc C, Haeich J, Aulestia FJ, et al. (2016) Calcium signaling orchestrates glioblastoma development: Facts and conjunctures. Biochim Biophys Acta 1863: 1447-1459. https://doi:10.1016/j.bbamcr.2016.01.018 ![]() |
[156] |
dePadua M, Kulothungan P, Lath R, et al. (2022) Establishment and characterization of brain cancer primary cell cultures from patients to enable phenotypic screening for new drugs. Front Pharmacol 13: 778193. https://doi:10.3389/fphar.2022.778193 ![]() |
[157] |
Zeniou M, Feve M, Mameri S, et al. (2015) Chemical library screening and structure-function relationship studies identify bisacodyl as a potent and selective cytotoxic agent towards quiescent human glioblastoma tumor stem-like cells. PLoS One 10: e0134793. https://doi:10.1371/journal.pone.0134793 ![]() |
[158] |
Dong J, Aulestia FJ, Assad Kahn S, et al. (2017) Bisacodyl and its cytotoxic activity on human glioblastoma stem-like cells. Implication of inositol 1,4,5-triphosphate receptor dependent calcium signaling. Biochim Biophys Acta Mol Cell Res 1864: 1018-1027. https://doi:10.1016/j.bbamcr.2017.01.010 ![]() |
[159] |
Robil N, Petel F, Kilhoffer MC, et al. (2015) Glioblastoma and calcium signaling--analysis of calcium toolbox expression. Int J Dev Biol 59: 407-415. https://doi:10.1387/ijdb.150200jh ![]() |
[160] |
Aulestia FJ, Neant I, Dong J, et al. (2018) Quiescence status of glioblastoma stem-like cells involves remodelling of Ca2+ signalling and mitochondrial shape. Sci Rep 8: 9731. https://doi:10.1038/s41598-018-28157-8 ![]() |
[161] |
Terrie E, Deliot N, Benzidane Y, et al. (2021) Store-operated calcium channels control proliferation and self-renewal of cancer stem cells from glioblastoma. Cancers (Basel) 13: 3428. https://doi:10.3390/cancers13143428 ![]() |
[162] |
Powell AE, Shung CY, Saylor KW, et al. (2010) Lessons from development: a role for asymmetric stem cell division in cancer. Stem Cell Res 4: 3-9. https://doi:10.1016/j.scr.2009.09.005 ![]() |
[163] | Koguchi M, Nakahara Y, Ito H, et al. (2020) BMP4 induces asymmetric cell division in human glioma stem-like cells. Oncol Lett 19: 1247-1254. https://doi:10.3892/ol.2019.11231 |
[164] |
Wee S, Niklasson M, Marinescu VD, et al. (2014) Selective calcium sensitivity in immature glioma cancer stem cells. PLoS One 9: e115698. https://doi:10.1371/journal.pone.0115698 ![]() |
[165] |
Wang F, Wang AY, Chesnelong C, et al. (2018) ING5 activity in self-renewal of glioblastoma stem cells via calcium and follicle stimulating hormone pathways. Oncogene 37: 286-301. https://doi:10.1038/onc.2017.324 ![]() |
[166] |
Motiani RK, Hyzinski-Garcia MC, Zhang X, et al. (2013) STIM1 and Orai1 mediate CRAC channel activity and are essential for human glioblastoma invasion. Pflugers Arch 465: 1249-1260. https://doi:10.1007/s00424-013-1254-8 ![]() |
[167] |
Lee SH, Rigas NK, Lee CR, et al. (2016) Orai1 promotes tumor progression by enhancing cancer stemness via NFAT signaling in oral/oropharyngeal squamous cell carcinoma. Oncotarget 7: 43239-43255. https://doi:10.18632/oncotarget.9755 ![]() |
[168] |
Wang J, Zhao H, Zheng L, et al. (2021) FGF19/SOCE/NFATc2 signaling circuit facilitates the self-renewal of liver cancer stem cells. Theranostics 11: 5045-5060. https://doi:10.7150/thno.56369 ![]() |
[169] |
Gengatharan A, Malvaut S, Marymonchyk A, et al. (2021) Adult neural stem cell activation in mice is regulated by the day/night cycle and intracellular calcium dynamics. Cell 184: 709-722 e713. https://doi:10.1016/j.cell.2020.12.026 ![]() |
[170] |
Cabanas H, Harnois T, Magaud C, et al. (2018) Deregulation of calcium homeostasis in Bcr-Abl-dependent chronic myeloid leukemia. Oncotarget 9: 26309-26327. https://doi:10.18632/oncotarget.25241 ![]() |
[171] |
Li Y, Guo B, Xie Q, et al. (2015) STIM1 mediates hypoxia-driven hepatocarcinogenesis via interaction with HIF-1. Cell Rep 12: 388-395. https://doi:10.1016/j.celrep.2015.06.033 ![]() |
[172] |
Umemura M, Baljinnyam E, Feske S, et al. (2014) Store-operated Ca2+ entry (SOCE) regulates melanoma proliferation and cell migration. PLoS One 9: e89292. https://doi:10.1371/journal.pone.0089292 ![]() |
[173] |
Jiang Y, Song Y, Wang R, et al. (2019) NFAT1-mediated regulation of NDEL1 promotes growth and invasion of glioma stem-like cells. Cancer Res 79: 2593-2603. https://doi:10.1158/0008-5472.CAN-18-3297 ![]() |
[174] |
Shin HJ, Lee S, Jung HJ (2019) A curcumin derivative hydrazinobenzoylcurcumin suppresses stem-like features of glioblastoma cells by targeting Ca2+ /calmodulin-dependent protein kinase II. J Cell Biochem 120: 6741-6752. https://doi:10.1002/jcb.27972 ![]() |
[175] |
Song Y, Jiang Y, Tao D, et al. (2020) NFAT2-HDAC1 signaling contributes to the malignant phenotype of glioblastoma. Neuro Oncol 22: 46-57. https://doi:10.1093/neuonc/noz136 ![]() |
[176] |
O'Reilly D, Buchanan P (2019) Calcium channels and cancer stem cells. Cell Calcium 81: 21-28. https://doi:10.1016/j.ceca.2019.05.006 ![]() |
[177] |
Terrie E, Coronas V, Constantin B (2019) Role of the calcium toolkit in cancer stem cells. Cell Calcium 80: 141-151. https://doi:10.1016/j.ceca.2019.05.001 ![]() |
[178] |
Gross S, Mallu P, Joshi H, et al. (2020) Ca2+ as a therapeutic target in cancer. Adv Cancer Res 148: 233-317. https://doi:10.1016/bs.acr.2020.05.003 ![]() |
[179] |
Tajada S, Villalobos C (2020) Calcium permeable channels in cancer hallmarks. Front Pharmacol 11: 968. https://doi:10.3389/fphar.2020.00968 ![]() |
[180] |
Alptekin M, Eroglu S, Tutar E, et al. (2015) Gene expressions of TRP channels in glioblastoma multiforme and relation with survival. Tumour Biol 36: 9209-9213. https://doi:10.1007/s13277-015-3577-x ![]() |
[181] |
Scrideli CA, Carlotti CG, Okamoto OK, et al. (2008) Gene expression profile analysis of primary glioblastomas and non-neoplastic brain tissue: identification of potential target genes by oligonucleotide microarray and real-time quantitative PCR. J Neurooncol 88: 281-291. https://doi:10.1007/s11060-008-9579-4 ![]() |
[182] |
Bomben VC, Turner KL, Barclay TT, et al. (2011) Transient receptor potential canonical channels are essential for chemotactic migration of human malignant gliomas. J Cell Physiol 226: 1879-1888. https://doi:10.1002/jcp.22518 ![]() |
[183] |
Liu H, Hughes JD, Rollins S, et al. (2011) Calcium entry via ORAI1 regulates glioblastoma cell proliferation and apoptosis. Exp Mol Pathol 91: 753-760. https://doi:10.1016/j.yexmp.2011.09.005 ![]() |
[184] |
Liu Z, Wei Y, Zhang L, et al. (2019) Induction of store-operated calcium entry (SOCE) suppresses glioblastoma growth by inhibiting the Hippo pathway transcriptional coactivators YAP/TAZ. Oncogene 38: 120-139. https://doi:10.1038/s41388-018-0425-7 ![]() |
[185] |
Coronas V, Terrie E, Deliot N, et al. (2020) Calcium channels in adult brain neural stem cells and in glioblastoma stem cells. Front Cell Neurosci 14: 600018. https://doi:10.3389/fncel.2020.600018 ![]() |
[186] |
Lepannetier S, Zanou N, Yerna X, et al. (2016) Sphingosine-1-phosphate-activated TRPC1 channel controls chemotaxis of glioblastoma cells. Cell Calcium 60: 373-383. https://doi:10.1016/j.ceca.2016.09.002 ![]() |
[187] |
Marfia G, Campanella R, Navone SE, et al. (2014) Autocrine/paracrine sphingosine-1-phosphate fuels proliferative and stemness qualities of glioblastoma stem cells. Glia 62: 1968-1981. https://doi:10.1002/glia.22718 ![]() |
[188] |
De Bacco F, Casanova E, Medico E, et al. (2012) The MET oncogene is a functional marker of a glioblastoma stem cell subtype. Cancer Res 72: 4537-4550. https://doi:10.1158/0008-5472.CAN-11-3490 ![]() |
[189] | Joo KM, Jin J, Kim E, et al. (2012) MET signaling regulates glioblastoma stem cells. Cancer Res 72: 3828-3838. https://doi:10.1158/0008-5472.CAN-11-3760 |
[190] |
Jacques TS, Swales A, Brzozowski MJ, et al. (2010) Combinations of genetic mutations in the adult neural stem cell compartment determine brain tumour phenotypes. EMBO J 29: 222-235. https://doi:10.1038/emboj.2009.327 ![]() |
[191] |
Matarredona ER, Pastor AM (2019) Neural stem cells of the subventricular zone as the origin of human glioblastoma stem cells. Therapeutic implications. Front Oncol 9: 779. https://doi:10.3389/fonc.2019.00779 ![]() |
[192] |
Zhang Y, Cruickshanks N, Yuan F, et al. (2017) Targetable T-type calcium channels drive glioblastoma. Cancer Res 77: 3479-3490. https://doi:10.1158/0008-5472.CAN-16-2347 ![]() |
[193] |
Niklasson M, Maddalo G, Sramkova Z, et al. (2017) Membrane-depolarizing channel blockers induce selective glioma cell death by impairing nutrient transport and unfolded protein/amino acid responses. Cancer Res 77: 1741-1752. https://doi:10.1158/0008-5472.CAN-16-2274 ![]() |
[194] |
Oh MC, Kim JM, Safaee M, et al. (2012) Overexpression of calcium-permeable glutamate receptors in glioblastoma derived brain tumor initiating cells. PLoS One 7: e47846. https://doi:10.1371/journal.pone.0047846 ![]() |
[195] |
Venkataramani V, Tanev DI, Strahle C, et al. (2019) Glutamatergic synaptic input to glioma cells drives brain tumour progression. Nature 573: 532-538. https://doi:10.1038/s41586-019-1564-x ![]() |
[196] |
Venkatesh HS, Morishita W, Geraghty AC, et al. (2019) Electrical and synaptic integration of glioma into neural circuits. Nature 573: 539-545. https://doi:10.1038/s41586-019-1563-y ![]() |
[197] |
Spina R, Voss DM, Asnaghi L, et al. (2016) Atracurium besylate and other neuromuscular blocking agents promote astroglial differentiation and deplete glioblastoma stem cells. Oncotarget 7: 459-472. https://doi:10.18632/oncotarget.6314 ![]() |
[198] |
Torres A, Erices JI, Sanchez F, et al. (2019) Extracellular adenosine promotes cell migration/invasion of glioblastoma stem-like cells through A(3) adenosine receptor activation under hypoxia. Cancer Lett 446: 112-122. https://doi:10.1016/j.canlet.2019.01.004 ![]() |
[199] |
D'Alimonte I, Nargi E, Zuccarini M, et al. (2015) Potentiation of temozolomide antitumor effect by purine receptor ligands able to restrain the in vitro growth of human glioblastoma stem cells. Purinergic Signal 11: 331-346. https://doi:10.1007/s11302-015-9454-7 ![]() |
[200] |
Morelli MB, Nabissi M, Amantini C, et al. (2012) The transient receptor potential vanilloid-2 cation channel impairs glioblastoma stem-like cell proliferation and promotes differentiation. Int J Cancer 131: E1067-1077. https://doi:10.1002/ijc.27588 ![]() |
[201] |
Chow KH, Park HJ, George J, et al. (2017) S100A4 is a biomarker and regulator of glioma stem cells that is critical for mesenchymal transition in glioblastoma. Cancer Res 77: 5360-5373. https://doi:10.1158/0008-5472.CAN-17-1294 ![]() |
[202] |
Calinescu AA, Kauss MC, Sultan Z, et al. (2021) Stem cells for the treatment of glioblastoma: a 20-year perspective. CNS Oncol 10: CNS73. https://doi:10.2217/cns-2020-0026 ![]() |
[203] |
Ahmed AS, Sheng MH, Wasnik S, et al. (2017) Effect of aging on stem cells. World J Exp Med 7: 1-10. https://doi:10.5493/wjem.v7.i1.1 ![]() |
[204] |
Ullah M, Sun Z (2018) Stem cells and anti-aging genes: double-edged sword-do the same job of life extension. Stem Cell Res Ther 9: 3. https://doi:10.1186/s13287-017-0746-4 ![]() |
[205] |
Bigarella CL, Liang R, Ghaffari S (2014) Stem cells and the impact of ROS signaling. Development 141: 4206-4218. https://doi:10.1242/dev.107086 ![]() |
[206] |
Ryall JG, Cliff T, Dalton S, et al. (2015) Metabolic reprogramming of stem cell epigenetics. Cell Stem Cell 17: 651-662. https://doi:10.1016/j.stem.2015.11.012 ![]() |
[207] |
Tenchov R, Sasso JM, Wang X, et al. (2024) Aging hallmarks and progression and age-related diseases: a landscape view of research advancement. ACS Chem Neurosci 15: 1-30. https://doi:10.1021/acschemneuro.3c00531 ![]() |
[208] |
Behringer EJ, Segal SS (2017) Impact of aging on calcium signaling and membrane potential in endothelium of resistance arteries: a role for mitochondria. J Gerontol A Biol Sci Med Sci 72: 1627-1637. https://doi:10.1093/gerona/glx079 ![]() |
[209] |
Sukumaran P, Nascimento Da Conceicao V, Sun Y, et al. (2021) Calcium signaling regulates autophagy and apoptosis. Cells 10: 2125. https://doi:10.3390/cells10082125 ![]() |
[210] |
Sharpless NE, DePinho RA (2007) How stem cells age and why this makes us grow old. Nat Rev Mol Cell Biol 8: 703-713. https://doi:10.1038/nrm2241 ![]() |
[211] | Newton AC, Bootman MD, Scott JD (2016) Second messengers. CSH Perspect Biol 8: a005926. https://doi:10.1101/cshperspect.a005926 |
[212] |
Iino M (2010) Spatiotemporal dynamics of Ca2+ signaling and its physiological roles. Proc Jpn Acad Ser B Phys Biol Sci 86: 244-256. https://doi:10.2183/pjab.86.244 ![]() |
[213] |
Dupont G, Combettes L (2016) Fine tuning of cytosolic Ca2+ oscillations. F1000Res 5: 2036. https://doi:10.12688/f1000research.8438.1 ![]() |
[214] |
Smedler E, Uhlen P (2014) Frequency decoding of calcium oscillations. Biochim Biophys Acta 1840: 964-969. https://doi:10.1016/j.bbagen.2013.11.015 ![]() |
[215] | Ong HL, Ambudkar IS (2020) The endoplasmic reticulum-plasma membrane junction: a hub for agonist regulation of Ca2+ entry. CSH Perspect Biol 12: a035253. https://doi:10.1101/cshperspect.a035253 |
[216] |
Yoast RE, Emrich SM, Zhang X, et al. (2021) The Mitochondrial Ca2+ uniporter is a central regulator of interorganellar Ca2+ transfer and NFAT activation. J Biol Chem 297: 101174. https://doi:10.1016/j.jbc.2021.101174 ![]() |
[217] |
De Stefani D, Raffaello A, Teardo E, et al. (2011) A forty-kilodalton protein of the inner membrane is the mitochondrial calcium uniporter. Nature 476: 336-340. https://doi:10.1038/nature10230 ![]() |
[218] | Oliveira AG, Guimaraes ES, Andrade LM, et al. (2014) Decoding calcium signaling across the nucleus. Physiology (Bethesda) 29: 361-368. https://doi:10.1152/physiol.00056.2013 |
[219] |
Kar P, Parekh AB (2015) Distinct spatial Ca2+ signatures selectively activate different NFAT transcription factor isoforms. Mol Cell 58: 232-243. https://doi:10.1016/j.molcel.2015.02.027 ![]() |
[220] |
Bakhshinyan D, Savage N, Salim SK, et al. (2020) The strange case of Jekyll and Hyde: parallels between neural stem cells and glioblastoma-initiating cells. Front Oncol 10: 603738. https://doi:10.3389/fonc.2020.603738 ![]() |
[221] |
Karlstad J, Sun Y, Singh BB (2012) Ca2+ signaling: an outlook on the characterization of Ca2+ channels and their importance in cellular functions. Adv Exp Med Biol 740: 143-157. https://doi:10.1007/978-94-007-2888-2_6 ![]() |
[222] |
Ong HL, Subedi KP, Son GY, et al. (2019) Tuning store-operated calcium entry to modulate Ca2+-dependent physiological processes. Biochim Biophys Acta Mol Cell Res 1866: 1037-1045. https://doi:10.1016/j.bbamcr.2018.11.018 ![]() |
1. | Muhammad Khan, Xiuting Huang, Xiaoxin Ye, Donghui Zhang, Baiyao Wang, Anan Xu, Rong Li, Anbang Ren, Chengcong Chen, Jingjing Song, Rong Zheng, Yawei Yuan, Jie Lin, Necroptosis-based glioblastoma prognostic subtypes: implications for TME remodeling and therapy response, 2024, 56, 0785-3890, 10.1080/07853890.2024.2405079 | |
2. | Donggu Lee, Sunju Oh, Sean Lawler, Yangjin Kim, Bistable dynamics of TAN-NK cells in tumor growth and control of radiotherapy-induced neutropenia in lung cancer treatment, 2025, 22, 1551-0018, 744, 10.3934/mbe.2025028 |
Par | Description | Values | Ref |
Intracellular dynamics | |||
λB | Bortezomib signaling scaling factor | 1.9 | [46] |
α | Inhibition strength of bortezomib by oHSV | 30 | [46] |
σ1 | Autocatalytic enhancement rate of IκB | 4 | [46] |
σ9 | Hill-type parameter | 1.0 | [46] |
σ4 | Inhibition strength of IκB by the NFκB-Bcl-2 complex | 2.2 | [46] |
σ7 | Signaling strength of the NFκB-Bcl-2 complex | 0.07 | [46] |
σ2 | Autocatalytic enhancement rate of the NFκB-Bcl-2 complex | 1.33 | [46] |
σ10 | Hill-type parameter | 1.0 | [46] |
σ5 | Inhibition strength of the NFκB-Bcl-2 complex by IκB | 1 | [46] |
ω1 | Decay rate of NFκB-Bcl-2 complex | 0.3 | [107,108] |
σ8 | Signaling strength of BAX | 0.0033 | [46] |
σ3 | Autocatalytic enhancement rate of BAX | 0.111 | [46] |
σ11 | Hill-type parameter | 1.0 | [46] |
σ6 | Inhibition strength of BAX by the NFκB-Bcl-2 complex | 1 | [46] |
ω2 | Decay rate of BAX | 0.02 | [107,108,109] |
σ12 | Signaling strength of RIP1 | 0.1 | [46] |
σ13 | Activation rate of RIP1 in the presence of OVs | 0.13 | [46] |
ω3 | Decay rate of RIP1 | 0.139 | [107,108,110] |
k | Hill type parameter of oHSV switching | 0.01 v∗ | [46] |
Cancer cells | |||
λ | proliferation rate of tumor cells | 4×10−3 | [46,57] |
x0 | Carrying capacity of uninfected tumor cells | 1 | [38,45,46,57] |
β1 | Bortezomib-induced apoptosis rate of tumor cells | 2×10−4 | [46] |
β2 | Oncolysis rate of tumor cells | 3×10−4 | [45,46,57] |
β3 | Necroptosis rate of tumor cells | 5×10−4 | [46] |
δ | infected cell lysis rate | 4.4×10−3 | [45,57] |
μ | Removal rate of dead cells | 2.3×10−3 | [45,46,57] |
Anticancer drugs | |||
b | Burst size of infected cells | 50 | [46,57] |
α1 | bortezomib-induced viral replication rate | 1 | [45,46] |
γ | clearance rate of viruses | 1.8×10−3 | [45,57] |
μ1 | consumption rate of bortezomib by uninfected tumor cells | 0.2075 | [45,46] |
μ2 | consumption rate of bortezomib by infected tumor cells | 0.2075 | [45,46] |
kB | Hill-type parameter | 1 | [45,46] |
μB | decay rate of bortezomib | 0.03 | [46,111,112] |
Reference values of main variables | |||
S∗ | Concentration of IκB | 0.05 μM | [113,114] |
F∗ | Concentration of the NFκB-Bcl2 complex | 0.5 μM | [113,114,115,116] |
A∗ | Concentration of BAX | 0.1 μM | [117] |
R∗ | Concentration of RIP1 | 5.0 μM | [118] |
x∗ | Uninfected cell density | 106cells/mm3 | [45,57,119] |
y∗ | Infected cell density | =x∗ | [45,57,119] |
n∗ | Dead cell density | =x∗ | [45,57,119] |
v∗ | Virus concentration | 2.2×108virus/mm3 | [45,57,119] |
B∗ | Bortezomib concentration | 1.0×10−11g/mm3 | [6,43,45] |
Parameter | B | v | λB | α | σ1 | σ9 |
PRCC(S, 0.1 h) | 0.28187 | −0.03828 | 0.26953 | −0.33367 | 0.94918 | 0.50929 |
PRCC(F, 0.1 h) | −0.02305 | 0.01961 | −0.02248 | 0.04351 | −0.27969 | −0.04106 |
PRCC(A, 0.1 h) | 0.00822 | 0.00416 | −0.01449 | −0.01347 | −0.01005 | −0.00355 |
PRCC(R, 0.1 h) | −0.00527 | 0.13183 | −0.00513 | 0.02775 | −0.06686 | −0.01338 |
PRCC(S, 1 h) | 0.26056 | −0.02550 | 0.25868 | −0.29694 | 0.82422 | 0.73676 |
PRCC(F, 1 h) | −0.07284 | 0.02242 | −0.08059 | 0.09171 | −0.48167 | −0.32707 |
PRCC(A, 1 h) | 0.02008 | 0.00165 | 0.01746 | −0.02042 | 0.11314 | 0.05378 |
PRCC(R, 1 h) | −0.03627 | 0.11745 | −0.04266 | 0.06355 | −0.27849 | −0.14751 |
PRCC(S, 10 h) | 0.37318 | −0.03499 | 0.35211 | −0.35990 | 0.55805 | 0.62897 |
PRCC(F, 10 h) | −0.09981 | 0.02151 | −0.10770 | 0.12735 | −0.32377 | −0.35500 |
PRCC(A, 10 h) | 0.04337 | 0.00249 | 0.04355 | −0.04101 | 0.16069 | 0.14905 |
PRCC(R, 10 h) | −0.06094 | 0.07032 | −0.07396 | 0.10494 | −0.23546 | −0.24755 |
PRCC(S, 100 h) | 0.38143 | −0.03454 | 0.36085 | −0.36390 | 0.51007 | 0.58888 |
PRCC(F, 100 h) | −0.09180 | 0.01882 | −0.10493 | 0.12160 | −0.27286 | −0.30714 |
PRCC(A, 100 h) | 0.04653 | −0.00022 | 0.02574 | −0.01787 | 0.11811 | 0.10076 |
PRCC(R, 100 h) | −0.04955 | 0.05732 | −0.06608 | 0.09572 | −0.17990 | −0.19896 |
Minimum | 0 | 0 | 0.19 | 3 | 0.4 | 0.1 |
Baseline | 0.5 | 0.01 | 1.9 | 30 | 4 | 1 |
Maximum | 5 | 0.1 | 5.7 | 60 | 8 | 5 |
Parameter | σ4 | σ7 | σ2 | σ10 | σ5 | ω1 |
PRCC(S, 0.1 h) | −0.17767 | −0.06597 | −0.24638 | −0.03404 | −0.00584 | 0.02214 |
PRCC(F, 0.1 h) | 0.00846 | 0.63726 | 0.96485 | 0.53649 | −0.19325 | −0.15541 |
PRCC(A, 0.1 h) | −0.01288 | −0.03460 | −0.19312 | −0.03172 | −0.00515 | 0.00624 |
PRCC(R, 0.1 h) | 0.02074 | 0.26024 | 0.79329 | 0.14504 | −0.05250 | −0.04657 |
PRCC(S, 1 h) | −0.40421 | −0.20533 | −0.50535 | −0.30275 | 0.12012 | 0.16696 |
PRCC(F, 1 h) | 0.12788 | 0.52405 | 0.85069 | 0.68899 | −0.35946 | −0.55150 |
PRCC(A, 1 h) | −0.02169 | −0.16545 | −0.45533 | −0.24162 | 0.08285 | 0.14085 |
PRCC(R, 1 h) | 0.06123 | 0.32027 | 0.72956 | 0.46364 | −0.19797 | −0.28509 |
PRCC(S, 10 h) | −0.36729 | −0.27416 | −0.40630 | −0.34399 | 0.15168 | 0.53178 |
PRCC(F, 10 h) | 0.18956 | 0.49155 | 0.67193 | 0.55311 | −0.26960 | −0.81199 |
PRCC(A, 10 h) | −0.06678 | −0.25292 | −0.36120 | −0.29522 | 0.11705 | 0.42062 |
PRCC(R, 10 h) | 0.13050 | 0.35427 | 0.54956 | 0.43078 | −0.19729 | −0.69605 |
PRCC(S, 100 h) | −0.33893 | −0.24777 | −0.38422 | −0.30025 | 0.12914 | 0.56675 |
PRCC(F, 100 h) | 0.17244 | 0.45223 | 0.63376 | 0.50030 | −0.23642 | −0.82047 |
PRCC(A, 100 h) | −0.06158 | −0.17745 | −0.25095 | −0.20439 | 0.06632 | 0.38276 |
PRCC(R, 100 h) | 0.11273 | 0.30119 | 0.48277 | 0.35914 | −0.15850 | −0.70620 |
Minimum | 0.22 | 0.007 | 0.133 | 0.1 | 0.1 | 0.03 |
Baseline | 2.2 | 0.07 | 1.33 | 1 | 1 | 0.3 |
Maximum | 6.6 | 1.4 | 6.65 | 5 | 5 | 1.5 |
Parameter | σ8 | σ3 | σ11 | σ6 | ω2 | σ12 |
PRCC(S, 0.1 h) | 0.00418 | 0.00133 | 0.00721 | −0.00730 | 0.00880 | −0.00095 |
PRCC(F, 0.1 h) | 0.00491 | −0.02376 | −0.00479 | −0.01447 | 0.00070 | 0.00029 |
PRCC(A, 0.1 h) | 0.94584 | 0.94967 | 0.43179 | −0.14146 | −0.02909 | 0.01632 |
PRCC(R, 0.1 h) | −0.00100 | −0.00278 | 0.00123 | 0.00560 | 0.00341 | 0.99028 |
PRCC(S, 1 h) | −0.00087 | 0.00042 | −0.00427 | 0.00687 | 0.00177 | 0.00779 |
PRCC(F, 1 h) | 0.00618 | −0.00739 | 0.00563 | −0.01924 | 0.00263 | −0.00287 |
PRCC(A, 1 h) | 0.91506 | 0.83297 | 0.64729 | −0.35019 | −0.17798 | 0.00315 |
PRCC(R, 1 h) | 0.00334 | −0.01925 | 0.00145 | 0.00094 | 0.00290 | 0.71206 |
PRCC(S, 10 h) | −0.00308 | −0.01061 | −0.00714 | 0.00995 | −0.01861 | 0.01484 |
PRCC(F, 10 h) | 0.01215 | −0.00269 | 0.00844 | −0.01508 | 0.00412 | −0.01259 |
PRCC(A, 10 h) | 0.88966 | 0.56291 | 0.46346 | −0.27714 | −0.70996 | 0.00142 |
PRCC(R, 10 h) | 0.01655 | −0.01123 | 0.00754 | −0.01035 | −0.00245 | 0.30818 |
PRCC(S, 100 h) | −0.00690 | −0.01292 | −0.00574 | 0.00764 | −0.01664 | 0.01243 |
PRCC(F, 100 h) | 0.01736 | 0.00037 | 0.00739 | −0.01191 | 0.00101 | −0.01150 |
PRCC(A, 100 h) | 0.81841 | 0.37472 | 0.30075 | −0.17214 | −0.87775 | −0.00371 |
PRCC(R, 100 h) | 0.02093 | −0.00869 | 0.00635 | −0.01199 | −0.00584 | 0.25393 |
Minimum | 0.00033 | 0.0111 | 0.1 | 0.1 | 0.002 | 0.01 |
Baseline | 0.0033 | 0.111 | 1 | 1 | 0.02 | 0.1 |
Maximum | 0.99 | 1.11 | 5 | 5 | 0.2 | 1 |
Parameter | σ13 | ω3 | ||||
PRCC(S, 0.1 h) | 0.01202 | −0.00260 | ||||
PRCC(F, 0.1 h) | 0.01100 | −0.00583 | ||||
PRCC(A, 0.1 h) | 0.00190 | 0.00014 | ||||
PRCC(R, 0.1 h) | 0.84920 | −0.25007 | ||||
PRCC(S, 1 h) | 0.00730 | 0.00780 | ||||
PRCC(F, 1 h) | 0.01072 | −0.01173 | ||||
PRCC(A, 1 h) | 0.01013 | 0.00761 | ||||
PRCC(R, 1 h) | 0.82148 | −0.49732 | ||||
PRCC(S, 10 h) | 0.00154 | 0.01567 | ||||
PRCC(F, 10 h) | 0.00974 | −0.01014 | ||||
PRCC(A, 10 h) | 0.00512 | 0.01810 | ||||
PRCC(R, 10 h) | 0.68681 | −0.75844 | ||||
PRCC(S, 100 h) | −0.00381 | 0.01276 | ||||
PRCC(F, 100 h) | 0.01531 | −0.00586 | ||||
PRCC(A, 100 h) | 0.00191 | −0.00058 | ||||
PRCC(R, 100 h) | 0.61909 | −0.74965 | ||||
Minimum | 0.013 | 0.0139 | ||||
Baseline | 0.13 | 0.139 | ||||
Maximum | 1.3 | 1.39 |
Par | Description | Values | Ref |
Intracellular dynamics | |||
λB | Bortezomib signaling scaling factor | 1.9 | [46] |
α | Inhibition strength of bortezomib by oHSV | 30 | [46] |
σ1 | Autocatalytic enhancement rate of IκB | 4 | [46] |
σ9 | Hill-type parameter | 1.0 | [46] |
σ4 | Inhibition strength of IκB by the NFκB-Bcl-2 complex | 2.2 | [46] |
σ7 | Signaling strength of the NFκB-Bcl-2 complex | 0.07 | [46] |
σ2 | Autocatalytic enhancement rate of the NFκB-Bcl-2 complex | 1.33 | [46] |
σ10 | Hill-type parameter | 1.0 | [46] |
σ5 | Inhibition strength of the NFκB-Bcl-2 complex by IκB | 1 | [46] |
ω1 | Decay rate of NFκB-Bcl-2 complex | 0.3 | [107,108] |
σ8 | Signaling strength of BAX | 0.0033 | [46] |
σ3 | Autocatalytic enhancement rate of BAX | 0.111 | [46] |
σ11 | Hill-type parameter | 1.0 | [46] |
σ6 | Inhibition strength of BAX by the NFκB-Bcl-2 complex | 1 | [46] |
ω2 | Decay rate of BAX | 0.02 | [107,108,109] |
σ12 | Signaling strength of RIP1 | 0.1 | [46] |
σ13 | Activation rate of RIP1 in the presence of OVs | 0.13 | [46] |
ω3 | Decay rate of RIP1 | 0.139 | [107,108,110] |
k | Hill type parameter of oHSV switching | 0.01 v∗ | [46] |
Cancer cells | |||
λ | proliferation rate of tumor cells | 4×10−3 | [46,57] |
x0 | Carrying capacity of uninfected tumor cells | 1 | [38,45,46,57] |
β1 | Bortezomib-induced apoptosis rate of tumor cells | 2×10−4 | [46] |
β2 | Oncolysis rate of tumor cells | 3×10−4 | [45,46,57] |
β3 | Necroptosis rate of tumor cells | 5×10−4 | [46] |
δ | infected cell lysis rate | 4.4×10−3 | [45,57] |
μ | Removal rate of dead cells | 2.3×10−3 | [45,46,57] |
Anticancer drugs | |||
b | Burst size of infected cells | 50 | [46,57] |
α1 | bortezomib-induced viral replication rate | 1 | [45,46] |
γ | clearance rate of viruses | 1.8×10−3 | [45,57] |
μ1 | consumption rate of bortezomib by uninfected tumor cells | 0.2075 | [45,46] |
μ2 | consumption rate of bortezomib by infected tumor cells | 0.2075 | [45,46] |
kB | Hill-type parameter | 1 | [45,46] |
μB | decay rate of bortezomib | 0.03 | [46,111,112] |
Reference values of main variables | |||
S∗ | Concentration of IκB | 0.05 μM | [113,114] |
F∗ | Concentration of the NFκB-Bcl2 complex | 0.5 μM | [113,114,115,116] |
A∗ | Concentration of BAX | 0.1 μM | [117] |
R∗ | Concentration of RIP1 | 5.0 μM | [118] |
x∗ | Uninfected cell density | 106cells/mm3 | [45,57,119] |
y∗ | Infected cell density | =x∗ | [45,57,119] |
n∗ | Dead cell density | =x∗ | [45,57,119] |
v∗ | Virus concentration | 2.2×108virus/mm3 | [45,57,119] |
B∗ | Bortezomib concentration | 1.0×10−11g/mm3 | [6,43,45] |
Parameter | B | v | λB | α | σ1 | σ9 |
PRCC(S, 0.1 h) | 0.28187 | −0.03828 | 0.26953 | −0.33367 | 0.94918 | 0.50929 |
PRCC(F, 0.1 h) | −0.02305 | 0.01961 | −0.02248 | 0.04351 | −0.27969 | −0.04106 |
PRCC(A, 0.1 h) | 0.00822 | 0.00416 | −0.01449 | −0.01347 | −0.01005 | −0.00355 |
PRCC(R, 0.1 h) | −0.00527 | 0.13183 | −0.00513 | 0.02775 | −0.06686 | −0.01338 |
PRCC(S, 1 h) | 0.26056 | −0.02550 | 0.25868 | −0.29694 | 0.82422 | 0.73676 |
PRCC(F, 1 h) | −0.07284 | 0.02242 | −0.08059 | 0.09171 | −0.48167 | −0.32707 |
PRCC(A, 1 h) | 0.02008 | 0.00165 | 0.01746 | −0.02042 | 0.11314 | 0.05378 |
PRCC(R, 1 h) | −0.03627 | 0.11745 | −0.04266 | 0.06355 | −0.27849 | −0.14751 |
PRCC(S, 10 h) | 0.37318 | −0.03499 | 0.35211 | −0.35990 | 0.55805 | 0.62897 |
PRCC(F, 10 h) | −0.09981 | 0.02151 | −0.10770 | 0.12735 | −0.32377 | −0.35500 |
PRCC(A, 10 h) | 0.04337 | 0.00249 | 0.04355 | −0.04101 | 0.16069 | 0.14905 |
PRCC(R, 10 h) | −0.06094 | 0.07032 | −0.07396 | 0.10494 | −0.23546 | −0.24755 |
PRCC(S, 100 h) | 0.38143 | −0.03454 | 0.36085 | −0.36390 | 0.51007 | 0.58888 |
PRCC(F, 100 h) | −0.09180 | 0.01882 | −0.10493 | 0.12160 | −0.27286 | −0.30714 |
PRCC(A, 100 h) | 0.04653 | −0.00022 | 0.02574 | −0.01787 | 0.11811 | 0.10076 |
PRCC(R, 100 h) | −0.04955 | 0.05732 | −0.06608 | 0.09572 | −0.17990 | −0.19896 |
Minimum | 0 | 0 | 0.19 | 3 | 0.4 | 0.1 |
Baseline | 0.5 | 0.01 | 1.9 | 30 | 4 | 1 |
Maximum | 5 | 0.1 | 5.7 | 60 | 8 | 5 |
Parameter | σ4 | σ7 | σ2 | σ10 | σ5 | ω1 |
PRCC(S, 0.1 h) | −0.17767 | −0.06597 | −0.24638 | −0.03404 | −0.00584 | 0.02214 |
PRCC(F, 0.1 h) | 0.00846 | 0.63726 | 0.96485 | 0.53649 | −0.19325 | −0.15541 |
PRCC(A, 0.1 h) | −0.01288 | −0.03460 | −0.19312 | −0.03172 | −0.00515 | 0.00624 |
PRCC(R, 0.1 h) | 0.02074 | 0.26024 | 0.79329 | 0.14504 | −0.05250 | −0.04657 |
PRCC(S, 1 h) | −0.40421 | −0.20533 | −0.50535 | −0.30275 | 0.12012 | 0.16696 |
PRCC(F, 1 h) | 0.12788 | 0.52405 | 0.85069 | 0.68899 | −0.35946 | −0.55150 |
PRCC(A, 1 h) | −0.02169 | −0.16545 | −0.45533 | −0.24162 | 0.08285 | 0.14085 |
PRCC(R, 1 h) | 0.06123 | 0.32027 | 0.72956 | 0.46364 | −0.19797 | −0.28509 |
PRCC(S, 10 h) | −0.36729 | −0.27416 | −0.40630 | −0.34399 | 0.15168 | 0.53178 |
PRCC(F, 10 h) | 0.18956 | 0.49155 | 0.67193 | 0.55311 | −0.26960 | −0.81199 |
PRCC(A, 10 h) | −0.06678 | −0.25292 | −0.36120 | −0.29522 | 0.11705 | 0.42062 |
PRCC(R, 10 h) | 0.13050 | 0.35427 | 0.54956 | 0.43078 | −0.19729 | −0.69605 |
PRCC(S, 100 h) | −0.33893 | −0.24777 | −0.38422 | −0.30025 | 0.12914 | 0.56675 |
PRCC(F, 100 h) | 0.17244 | 0.45223 | 0.63376 | 0.50030 | −0.23642 | −0.82047 |
PRCC(A, 100 h) | −0.06158 | −0.17745 | −0.25095 | −0.20439 | 0.06632 | 0.38276 |
PRCC(R, 100 h) | 0.11273 | 0.30119 | 0.48277 | 0.35914 | −0.15850 | −0.70620 |
Minimum | 0.22 | 0.007 | 0.133 | 0.1 | 0.1 | 0.03 |
Baseline | 2.2 | 0.07 | 1.33 | 1 | 1 | 0.3 |
Maximum | 6.6 | 1.4 | 6.65 | 5 | 5 | 1.5 |
Parameter | σ8 | σ3 | σ11 | σ6 | ω2 | σ12 |
PRCC(S, 0.1 h) | 0.00418 | 0.00133 | 0.00721 | −0.00730 | 0.00880 | −0.00095 |
PRCC(F, 0.1 h) | 0.00491 | −0.02376 | −0.00479 | −0.01447 | 0.00070 | 0.00029 |
PRCC(A, 0.1 h) | 0.94584 | 0.94967 | 0.43179 | −0.14146 | −0.02909 | 0.01632 |
PRCC(R, 0.1 h) | −0.00100 | −0.00278 | 0.00123 | 0.00560 | 0.00341 | 0.99028 |
PRCC(S, 1 h) | −0.00087 | 0.00042 | −0.00427 | 0.00687 | 0.00177 | 0.00779 |
PRCC(F, 1 h) | 0.00618 | −0.00739 | 0.00563 | −0.01924 | 0.00263 | −0.00287 |
PRCC(A, 1 h) | 0.91506 | 0.83297 | 0.64729 | −0.35019 | −0.17798 | 0.00315 |
PRCC(R, 1 h) | 0.00334 | −0.01925 | 0.00145 | 0.00094 | 0.00290 | 0.71206 |
PRCC(S, 10 h) | −0.00308 | −0.01061 | −0.00714 | 0.00995 | −0.01861 | 0.01484 |
PRCC(F, 10 h) | 0.01215 | −0.00269 | 0.00844 | −0.01508 | 0.00412 | −0.01259 |
PRCC(A, 10 h) | 0.88966 | 0.56291 | 0.46346 | −0.27714 | −0.70996 | 0.00142 |
PRCC(R, 10 h) | 0.01655 | −0.01123 | 0.00754 | −0.01035 | −0.00245 | 0.30818 |
PRCC(S, 100 h) | −0.00690 | −0.01292 | −0.00574 | 0.00764 | −0.01664 | 0.01243 |
PRCC(F, 100 h) | 0.01736 | 0.00037 | 0.00739 | −0.01191 | 0.00101 | −0.01150 |
PRCC(A, 100 h) | 0.81841 | 0.37472 | 0.30075 | −0.17214 | −0.87775 | −0.00371 |
PRCC(R, 100 h) | 0.02093 | −0.00869 | 0.00635 | −0.01199 | −0.00584 | 0.25393 |
Minimum | 0.00033 | 0.0111 | 0.1 | 0.1 | 0.002 | 0.01 |
Baseline | 0.0033 | 0.111 | 1 | 1 | 0.02 | 0.1 |
Maximum | 0.99 | 1.11 | 5 | 5 | 0.2 | 1 |
Parameter | σ13 | ω3 | ||||
PRCC(S, 0.1 h) | 0.01202 | −0.00260 | ||||
PRCC(F, 0.1 h) | 0.01100 | −0.00583 | ||||
PRCC(A, 0.1 h) | 0.00190 | 0.00014 | ||||
PRCC(R, 0.1 h) | 0.84920 | −0.25007 | ||||
PRCC(S, 1 h) | 0.00730 | 0.00780 | ||||
PRCC(F, 1 h) | 0.01072 | −0.01173 | ||||
PRCC(A, 1 h) | 0.01013 | 0.00761 | ||||
PRCC(R, 1 h) | 0.82148 | −0.49732 | ||||
PRCC(S, 10 h) | 0.00154 | 0.01567 | ||||
PRCC(F, 10 h) | 0.00974 | −0.01014 | ||||
PRCC(A, 10 h) | 0.00512 | 0.01810 | ||||
PRCC(R, 10 h) | 0.68681 | −0.75844 | ||||
PRCC(S, 100 h) | −0.00381 | 0.01276 | ||||
PRCC(F, 100 h) | 0.01531 | −0.00586 | ||||
PRCC(A, 100 h) | 0.00191 | −0.00058 | ||||
PRCC(R, 100 h) | 0.61909 | −0.74965 | ||||
Minimum | 0.013 | 0.0139 | ||||
Baseline | 0.13 | 0.139 | ||||
Maximum | 1.3 | 1.39 |