1 Introduction

Cancer cells undergo profound metabolic alterations facilitating their proliferation, invasion, metastasis, and even drug resistance (Han et al. 2013; Rahman and Hasan 2015). Unlike normal cells, cancer cells derive a substantial amount of their energy from glycolysis, converting a majority of incoming glucose into lactate in the cytoplasm rather than metabolising it in the mitochondria through oxidative phosphorylation (OXPHOS) (Cairns et al. 2011; Simabuco et al. 2018). This metabolic adaptation, recognised as the Warburg effect or aerobic glycolysis, leads to decreased oxygen consumption required by mitochondrial respiration while generating an increased amount of lactate (Simabuco et al. 2018).

By favouring glycolysis over OXPHOS, cancer cells ensure the availability of essential building blocks for biomass synthesis and meet the energy demands necessary for their rapid growth (Hanahan and Weinberg 2011; Simabuco et al. 2018). While glycolysis can produce adenosine triphosphate (ATP), the major cellular energy unit, more rapidly than oxidative phosphorylation, it is significantly less efficient in terms of ATP generated per unit of glucose consumed (Cairns et al. 2011; Simabuco et al. 2018). Consequently, tumour cells increase their glucose uptake at an exceptionally high rate to adequately satisfy their elevated energy and biosynthesis needs (Cairns et al. 2011; Simabuco et al. 2018).

The glycolytic phenotype of cancer cells is influenced by various molecular mechanisms extending beyond hypoxic conditions. Disruptions in signalling pathways downstream of growth factor receptors have been observed to affect glucose metabolism in cancer cells (Zhong et al. 2000; Laughner et al. 2001). Specifically, the PI3K/AKT/mTOR pathway, which is activated in the vast majority of human cancers (Hennessy et al. 2005; Danielsen et al. 2015; Vara et al. 2004; Malinowsky et al. 2014; Wang et al. 2013), and seen to instigate the glycolytic activity of cancer cells by upregulating the hypoxia-inducible factor 1 (HIF1) and its downstream targets (Cairns et al. 2011; Valvona et al. 2016; Laughner et al. 2001; Zhong et al. 2000).

Another crucial event that can impact cancer metabolism and is commonly observed in cancer is the inactivation of the tumour suppressor gene p53. Depending on the cellular conditions, p53 suppresses tumorigenesis by multiple mechanisms, including cell cycle regulation, initiation of DNA repair, and induction of apoptosis (programmed cell death) (Wanka et al. 2012; Simabuco et al. 2018). Moreover, p53 has recently emerged as a significant metabolic regulator in cancer cells, whether by inhibiting the PI3K/AKT/mTOR pathway (Feng and Levine 2010), thereby disrupting the glycolytic phenotype or by supporting mitochondrial respiration activity (Vousden and Ryan 2009; Zhang et al. 2010; Lago et al. 2011; Wanka et al. 2012; Liang et al. 2013; Flöter et al. 2017; Simabuco et al. 2018; Liu et al. 2019). This idea was investigated by Matoba et al. in 2006, where they examined the impact of p53 alterations on the cellular metabolism of human colon cancer cells (Matoba et al. 2006). Experimental results revealed that p53-deficient cells produced nearly the same amount of ATP but with substantially higher levels of lactate and lower oxygen consumption, highlighting the influence of p53 mutations in changing the energy production mode to one favouring glycolysis (Matoba et al. 2006).

The metabolic response controlled by p53 is mediated through the AMP-activated protein kinase (AMPK), a sensor attuned to cellular metabolic stress conditions (Jones et al. 2005). When activated by AMPK, typically in response to metabolic adversity such as those experienced by cancer cells, p53 restrains cell growth and division, conserves energy, and shifts the cell towards oxidative phosphorylation for more efficient energy production (Feng and Levine 2010). This may explain why cancer cells with p53 mutations tend to rely more on glycolysis and have a higher ability to grow and survive even under stress conditions.

Previous mathematical modelling sheds light on different aspects of cancer metabolism -ranging from the effect of reactive oxygen species (ROS) on HIF1 stabilisation in ischemic conditions (Qutub and Popel 2008) to the identification of metabolic targets to hinder cancer migration (Yizhak et al. 2014). Despite these insights, the genetic complexities underpinning the Warburg effect remain elusive. Recently, Linglin et al. made a notable contribution by discussing the genetic regulation of the interplay between glycolysis and oxidative phosphorylation (Yu et al. 2017). Nonetheless, this work did not account for the crucial influence of p53 or demonstrate its impact on metabolic pathways as observed experimentally in cancer cells.

Undoubtedly, the tumour-suppressive role of p53 has been studied through several computational modelling approaches. Ma et al. modelled the oscillatory dynamics of p53, emphasising the correlation between the DNA damage severity and the average number of p53 pulses (Ma et al. 2005). Further studies by Zhang et al. have elucidated how the frequency of p53 pulses can influence cellular decisions between repair and apoptosis (Zhang et al. 2007) and described the transition from cell repair to apoptosis with an inverse relationship between the apoptosis time and damage strength (Zhang et al. 2009). Another study has introduced a two-phase dynamic of p53, distinguishing between partial and complete activation of p53 (Zhang et al. 2011). Our recent study expanded on these findings by demonstrating that p53 can switch among three dynamic modes in a DNA damage strength-dependent manner following chemotherapy (Abukwaik et al. 2023). Despite these advances, earlier works have mainly concentrated on p53 dynamic behaviour in response to DNA damage stimuli and the subsequent cell fate, leaving a substantial gap in understanding its mechanisms regarding cellular metabolism.

To address this gap, this study develops a mathematical model that unveils the intricate machinery behind the regulation of cancer glucose metabolism by p53. While numerous p53 targets involved in cellular metabolism have been identified, their complex molecular interactions across different scenarios remain largely unexplored. By transforming existing experimental data into a mathematical framework, we uncovered hard-to-detect mechanisms and quantitatively analysed the activities of glycolysis and OXPHOS pathways under different cellular states. This methodology provides valuable insights for developing targeted therapeutic approaches aimed at disrupting cancer metabolism and combating the aggressive behaviour of cancer. Although our primary focus is on colon cancer cells, the model’s applicability extends to many cancer types experiencing similar conditions.

2 Methods

We constructed a comprehensive theoretical framework aiming to delineate the role of p53 on cellular metabolism, particularly its involvement in the Warburg effect. While cancer cells engage in diverse metabolic pathways, the Warburg effect is closely associated with alterations in glucose metabolism. Consequently, our primary attention was devoted to glucose metabolism, investigating its main pathways: glycolysis and oxidative phosphorylation. Integrating information from literature, our model incorporated all well-established p53 targets that markedly manipulate these pathways alongside the signalling pathways commonly activated in cancer in response to growth factors or metabolic stress, influencing the decision-making between these pathways, see Fig. 1. The following section briefly discusses these signalling pathways and their involvement in glucose metabolism.

Fig. 1
figure 1

Schematic diagram depicting the signalling pathway of key molecules involved in glucose oxidation, spanning glycolysis, the tricarboxylic acid (TCA) cycle, and the electron transport chain (ETC). In our notation, cytoplasmic, nuclear, and mitochondrial molecular species are indicated by subscripts ‘c’, ‘n’, and ‘m’, respectively, while the ‘\(*\)’ superscript symbol is used to denote active species in the case of species that exist in two states (active and inactive)

2.1 Signalling Pathways in Glucose Metabolism

2.1.1 Growth Factor Signalling Pathway

In response to growth factors, extracellular molecules activate specific receptors on the cell membrane, initiating the intracellular activation of phosphoinositide 3-kinase (PI3K). PI3K then phosphorylates phosphatidylinositol 4,5-bisphosphate (PIP2) into phosphatidylinositol 3,4,5-trisphosphate (PIP3), facilitating the protein kinase B (AKT) activation (Danielsen et al. 2015; Vara et al. 2004; Carnero and Paramio 2014).

Upon AKT activation, a cascade of events occurs, ultimately activating the mechanistic target of rapamycin (mTOR) (Dan et al. 2014; Inoki et al. 2002), which subsequently promotes the synthesis of proteins required for cell growth and proliferation. One of these proteins is HIF1\(\alpha \) (Düvel et al. 2010; Laughner et al. 2001; Hudson et al. 2002; Treins et al. 2005), a transcription factor that drives the expression of genes essential for glycolytic metabolism. These genes include glucose transporter-1 and -3 (GLUT1) and (GLUT3), respectively (Ancey et al. 2018), lactate dehydrogenase (LDH) (Valvona et al. 2016), and pyruvate dehydrogenase kinase-1 and -3 (PDK1) and (PDK3), respectively (Anwar et al. 2021; Lu et al. 2011; Wang et al. 2021; Kim et al. 2006).

2.1.2 Metabolic Stress Signalling Pathway

Different forms of metabolic stress activate AMPK, a crucial enzyme that restores cellular energy balance by suppressing ATP-consuming processes and promoting ATP production (Hardie 2011; Hardie et al. 2012; Faubert et al. 2015; Li et al. 2015). Through this activation, AMPK restrains cellular growth and increases ATP production efficiency primarily by repressing mTOR activation and stimulating the transcriptional activity of p53 (Faubert et al. 2015; Li et al. 2015; Inoki et al. 2003).

By phosphorating p53, AMPK disrupts the interaction between p53 and its negative regulator, murine double minute 2 (MDM2), leading to p53 stabilisation (Jones et al. 2005; Imamura et al. 2001). Consequently, p53 accumulates and translocates to the nucleus, triggering the promoter activity of its target genes.

2.1.3 p53 Transcriptional Targets and Feedback Mechanisms

p53 exerts its effects through the transcriptional regulation of a wide array of target genes involved in various cellular processes. Some p53 targets form a negative feedback loop that dampens its stability, such as MDM2 and wild-type p53-induced phosphatase 1 (WIP1) (Barak et al. 1993; Batchelor et al. 2011). WIP1 dephosphorylates p53, increasing its susceptibility to MDM2-mediated degradation (Batchelor et al. 2011).

Conversely, p53 activates genes that promote its activation and simultaneously inhibit the glycolytic regulator HIF1. For instance, p53 induces the expression of proteins that activate AMPK, which in turn inhibits mTOR activity and its downstream target HIF1 (Budanov and Karin 2008; Sanli et al. 2012; Feng and Levine 2010). This process forms a positive feedback loop, as active AMPK phosphorylates p53, further enhancing its stabilization (Feng and Levine 2010). Another crucial target of p53 is phosphatase and tensin homolog (PTEN) (Stambolic et al. 2001), which attenuates the PI3K/AKT/mTOR pathway by converting PIP3 back to PIP2 (Mayo et al. 2002; Carnero and Paramio 2014; Feng and Levine 2010). The induction of PTEN also supports p53 by hindering the AKT-dependent translocation of MDM2 to the nucleus, thereby boosting the transcriptional activity of p53 (Mayo et al. 2002, 2005).

In addition, p53 regulates genes directly involved in glucose metabolic pathways. This includes diminishing the protein synthesis of GLUT1, GLUT3 (Ancey et al. 2018; Schwartzenberg-Bar-Yoseph et al. 2004; Kawauchi et al. 2008), and PDK2 (Anwar et al. 2021; Liang et al. 2020), while stimulating the expression of other molecules, such as the TP53-inducible glycolysis and apoptosis regulator (TIGAR) (Bensaad et al. 2006; Lee et al. 2015), and the synthesis of cytochrome c oxidase 2 (SCO2) (Matoba et al. 2006).

2.1.4 Glucose Metabolism Pathways

Glucose metabolism comprises three main stages: starting with glycolysis, advancing to the tricarboxylic acid (TCA) cycle, and finishing with the electron transport chain (ETC).

Glycolysis. Glucose is transported to the cells by specific glucose transporters located in the cell membrane (Schwartzenberg-Bar-Yoseph et al. 2004; Szablewski 2013; Ancey et al. 2018; Mamun et al. 2020), namely GLUT1 and GLUT3 in our model, which are negatively regulated by p53 and positively by HIF1, controlling glucose uptake rate (Ancey et al. 2018; Schwartzenberg-Bar-Yoseph et al. 2004; Kawauchi et al. 2008). Once inside the cell, glucose undergoes an irreversible conversion to glucose-6-phosphate (G6P), consuming one ATP molecule (Golias et al. 2019). G6P can then either proceed through glycolysis or enter the pentose phosphate pathway (PPP) (Jiang et al. 2014), a decision influenced by the p53 target TIGAR, which inhibits the enzyme catalysing the third step in the glycolysis pathway, diminishing glycolysis flux (Bensaad et al. 2006; Lee et al. 2015). Continuing with glycolysis, a molecule of G6P is converted into two molecules of pyruvate, yielding three net ATP and two reduced nicotinamide adenine dinucleotide (NADH) molecules (Valvona et al. 2016; Golias et al. 2019).

In the presence of oxygen, pyruvate derived from glycolysis typically enters the mitochondria for further energy production through oxidative phosphorylation. However, increased levels of LDH enzyme, driven by HIF1 activation, redirect pyruvate away from the mitochondria by catalysing its conversion to lactate and oxidising NADH back to NAD\(^{+}\) (Valvona et al. 2016; Golias et al. 2019). This shift allows cells to produce ATP less efficiently through glycolysis while consuming more glucose (Valvona et al. 2016).

TCA cycle. Within the mitochondria, the pyruvate dehydrogenase (PDH) complex facilitates the first step towards glucose respiration by catalysing the oxidative decarboxylation of pyruvate to acetyl-CoA, concurrently reducing one NAD\(^{+}\) into NADH (Wang et al. 2021; Woolbright et al. 2019; Rodrigues et al. 2015). Acetyl-CoA subsequently enters the TCA cycle, undergoing a series of chemical reactions that generate one energy molecule, three NADH, and one reduced flavin adenine dinucleotide (FADH2) molecule (Martínez-Reyes and Chandel 2020).

The function of the PDH enzyme is controlled by the PDK family, which phosphorylates the E1-\(\alpha \) subunit of PDH, blocking its decarboxylation activity (Wang et al. 2021; Woolbright et al. 2019; Rodrigues et al. 2015). HIF1 enhances the promoter activities of two PDK family members, namely PDK1 and PDK3 (Anwar et al. 2021; Lu et al. 2011; Wang et al. 2021; Kim et al. 2006), while p53 suppresses the expression of another PDK member called PDK2 (Anwar et al. 2021; Liang et al. 2020).

ETC. In the last stage, NADH and FADH2 are oxidised back into NAD\(^{+}\) and FAD via protein complexes in the inner mitochondrial membrane (Ahmad et al. 2018). This oxidation process involves electron transfer from NADH and FADH2 across these complexes, during which protons are pumped from the mitochondrial matrix to the intermembrane space. This action creates a proton gradient that drives protons back into the matrix, facilitating ATP production (Ahmad et al. 2018).

Among these complexes, Complex IV acts as the final electron acceptor, channelling the electrons to molecular oxygen by using 0.5 oxygen (O\(_{2}\)) molecules for each pair of electrons received from NADH or FADH2 (Ahmad et al. 2018). However, the activity of Complex IV is regulated by SCO2, a p53-regulated gene essential for its proper assembly and maturation. Thus, deficiency of SCO2 can impair Complex IV functionality and disrupt the electron flow within the ETC (Matoba et al. 2006; Wanka et al. 2012).

2.2 Mathematical Framework

Our model encompasses the entire glucose oxidation network, spanning three compartments-cytoplasm, nucleus, and mitochondria. The directional fluxes within these compartments drive the temporal changes in the concentrations of 33 molecules involved in glucose oxidation pathways, as illustrated in Fig. 1. Consequently, the system is governed by 33 differential equations where cytoplasmic, nuclear, and mitochondrial molecular species are represented by the subscripts ‘c’, ‘n’, and ‘m’, respectively, while the ‘\(*\)’ superscript symbol indicates active species for those existing in both active and inactive states. Each term in the model represented by ’\({{\varvec{v}}}\)’ notation corresponds to one of the reactions detailed in the model reactions section in the Appendix (A), where \({{\varvec{v}}}= v/V_{max}\) or v/k.

$$\begin{aligned} \frac{dP53_{c}}{dt}&= k_{1}-k_{2}Ampk_{c}^*[{{\varvec{v}}}_{1(P53_{c})}]-k_{3}Mdm2_{c}[{{\varvec{v}}}_{1(P53_{c})}]\nonumber \\&\quad +k_{4}Wip1_{n}[{{\varvec{v}}}_{1(P53_{n})}]-k_{5}P53_{c}, \end{aligned}$$
(1)
$$\begin{aligned} \frac{dP53_{n}}{dt}&=k_{2}Ampk_{c}^*[{{\varvec{v}}}_{1(P53_{c})}]-k_{6}Mdm2_{n}[{{\varvec{v}}}_{1(P53_{n})}]-k_{4}Wip1_{n}[{{\varvec{v}}}_{1(P53_{n})}], \end{aligned}$$
(2)
$$\begin{aligned} \frac{dMdm2_{c}}{dt}&= k_{7}+k_{8}[{{\varvec{v}}}_{2(P53_{n})}^{+}]-k_{9}Akt_{c}^*[{{\varvec{v}}}_{1(Mdm2_{c})}] +k_{10}[{{\varvec{v}}}_{1(Mdm2_{n})}]\nonumber \\&\quad -k_{11}Mdm2_{c}, \end{aligned}$$
(3)
$$\begin{aligned} \frac{dMdm2_{n}}{dt}&=k_{9}Akt_{c}^*[{{\varvec{v}}}_{1(Mdm2_{c})}]-k_{10}[{{\varvec{v}}}_{1(Mdm2_{n})}] -k_{11}Mdm2_{n}, \end{aligned}$$
(4)
$$\begin{aligned} \frac{dWip1_{n}}{dt}&= k_{12}+k_{13}[{{\varvec{v}}}_{2(P53_{n})}^{+}]-k_{14}Wip1_{n}, \end{aligned}$$
(5)
$$\begin{aligned} \frac{dPten_{c}}{dt}&= k_{15}+k_{16}[{{\varvec{v}}}_{2(P53_{n})}^{+}]-k_{17}Pten_{c}, \end{aligned}$$
(6)
$$\begin{aligned} \frac{dSco2_{m}}{dt}&= k_{18}+k_{19}[{{\varvec{v}}}_{2(P53_{n})}^{+}]-k_{20}Sco2_{m}, \end{aligned}$$
(7)
$$\begin{aligned} \frac{dTigar_{c}}{dt}&= k_{21}+k_{22}[{{\varvec{v}}}_{2(P53_{n})}^{+}]-k_{23}Tigar_{c}, \end{aligned}$$
(8)
$$\begin{aligned} \frac{dAmpk_{c}^*}{dt}&= k_{24}[{{\varvec{v}}}_{1(Ampk_{c})}]+k_{25}[{{\varvec{v}}}_{2(P53_{n})}^{+}][{{\varvec{v}}}_{1(Ampk_{c})}]-k_{26}[{{\varvec{v}}}_{1(Ampk_{c}^*)}], \end{aligned}$$
(9)
$$\begin{aligned} \frac{dPip3_{c}}{dt}&= k_{27}[{{\varvec{v}}}_{1(Pip2_{c})}]-k_{28}Pten_{c}[{{\varvec{v}}}_{1(Pip3_{c})}], \end{aligned}$$
(10)
$$\begin{aligned} \frac{dAkt_{c}^*}{dt}&= k_{29}Pip3_{c}[{{\varvec{v}}}_{1(Akt_{c})}]-k_{30}[{{\varvec{v}}}_{1(Akt_{c}^*)}], \end{aligned}$$
(11)
$$\begin{aligned} \frac{dMtor_{c}^*}{dt}&= k_{31}Akt_{c}^*[{{\varvec{v}}}_{1(Mtor_{c})}]-k_{32}Ampk_{c}^*[{{\varvec{v}}}_{1(Mtor_{c}^*)}]-k_{33}[{{\varvec{v}}}_{1(Mtor_{c}^*)}], \end{aligned}$$
(12)
$$\begin{aligned} \frac{dHif1\alpha _{c}}{dt}&=k_{34}+k_{35}Mtor_{c}^*-k_{36}Mtor_{c}^*Hif1\alpha _{c}-k_{37}Hif1\alpha _{c}, \end{aligned}$$
(13)
$$\begin{aligned} \frac{dHif1\alpha _{n}}{dt}&=k_{36}Mtor_{c}^*Hif1\alpha _{c}-k_{37}Hif1\alpha _{n}, \end{aligned}$$
(14)
$$\begin{aligned} \frac{dGlut1_{c}}{dt}&=k_{38}[{{\varvec{v}}}_{2(P53_{n})}^{-}]+k_{39}Hif1\alpha _{n}-k_{40}Glut1_{c}, \end{aligned}$$
(15)
$$\begin{aligned} \frac{dGlut3_{c}}{dt}&=k_{41}[{{\varvec{v}}}_{2(P53_{n})}^{-}]+k_{42}Hif1\alpha _{n}-k_{43}Glut3_{c}, \end{aligned}$$
(16)
$$\begin{aligned} \frac{dPdk13_{m}}{dt}&=k_{44}+k_{45}Hif1\alpha _{n}-k_{46}Pdk13_{m}, \end{aligned}$$
(17)
$$\begin{aligned} \frac{dPdk2_{m}}{dt}&=k_{47}[{{\varvec{v}}}_{2(P53_{n})}^{-}]-k_{46}Pdk2_{m}, \end{aligned}$$
(18)
$$\begin{aligned} \frac{dLdh_{c}}{dt}&=k_{48}+k_{49}Hif1\alpha _{n}-k_{50}Ldh_{c}, \end{aligned}$$
(19)
$$\begin{aligned} \frac{dPdh_{m}^*}{dt}&=k_{51}-k_{52}\big (Pdk13_{m}+Pdk2_{m}\big )[{{\varvec{v}}}_{1(Pdh_{m}^*)}]+k_{53}[{{\varvec{v}}}_{1(Pdh_{m})}]\nonumber \\ &-k_{54}Pdh_{m}^*,\end{aligned}$$
(20)
$$\begin{aligned} \frac{dPdh_{m}}{dt}&=k_{52}\big (Pdk13_{m}+Pdk2_{m}\big )[{{\varvec{v}}}_{1(Pdh_{m}^*)}]-k_{53}[{{\varvec{v}}}_{1(Pdh_{m})}]-k_{54}Pdh_{m}, \end{aligned}$$
(21)
$$\begin{aligned} \frac{dGlucose_{c}}{dt}&=k_{55}\big (Glut1_{c}+Glut3_{c}\big )[{{\varvec{v}}}_{3(Glucose)}]-k_{56}[{{\varvec{v}}}_{4(Glucose_{c},Atp)}], \end{aligned}$$
(22)
$$\begin{aligned} \frac{dG6p_{c}}{dt}&=k_{56}[{{\varvec{v}}}_{4(Glucose_{c},Atp)}]-k_{57}[{{\varvec{v}}}_{7(G6p_{c})}]-k_{58}G6p_{c}, \end{aligned}$$
(23)
$$\begin{aligned} \frac{dPyruvate_{c}}{dt}&=2k_{57}[{{\varvec{v}}}_{7(G6p_{c})}]-k_{59}[{{\varvec{v}}}_{2(Ldh_{c})}^{+}][{{\varvec{v}}}_{4(Pyruvate_{c},Nadh_{c})}]\nonumber \\&\quad -k_{60}[{{\varvec{v}}}_{6(Pyruvate)}], \end{aligned}$$
(24)
$$\begin{aligned} \frac{dPyruvate_{m}}{dt}&=k_{60}[{{\varvec{v}}}_{6(Pyruvate)}]-k_{61}Pdh_{m}^*[{{\varvec{v}}}_{5(Pyruvate_{m},Nad_{m})}], \end{aligned}$$
(25)
$$\begin{aligned} \frac{dAcetyl_{m}}{dt}&=k_{61}Pdh_{m}^*[{{\varvec{v}}}_{5(Pyruvate_{m},Nad_{m})}]-k_{62}[{{\varvec{v}}}_{7(Acetyl_{m})}], \end{aligned}$$
(26)
$$\begin{aligned} \frac{dNadh_{c}}{dt}&=2k_{57}[{{\varvec{v}}}_{7(G6p_{c})}]-k_{59}[{{\varvec{v}}}_{2(Ldh_{c})}^{+}][{{\varvec{v}}}_{4(Pyruvate_{c},Nadh_{c})}]\nonumber \\ &-k_{63}[{{\varvec{v}}}_{6(Nadh)}], \end{aligned}$$
(27)
$$\begin{aligned} \frac{dNadh_{m}}{dt}&=k_{61}Pdh_{m}^*[{{\varvec{v}}}_{5(Pyruvate_{m},Nadh_{m})}]+3k_{62}[{{\varvec{v}}}_{7(Acetyl_{m})}]+k_{63}[{{\varvec{v}}}_{6(Nadh)}]\nonumber \\&\quad -k_{64}Sco2_{m}[{{\varvec{v}}}_{7(Nadh_{m})}], \end{aligned}$$
(28)
$$\begin{aligned} \frac{dFadh_{m}}{dt}&=k_{62}[{{\varvec{v}}}_{7(Acetyl_{m})}]-k_{64}Sco2_{m}[{{\varvec{v}}}_{7(Fadh_{m})}], \end{aligned}$$
(29)
$$\begin{aligned} \frac{dLactate_{c}}{dt}&=k_{59}[{{\varvec{v}}}_{2(Ldh_{c})}^{+}][{{\varvec{v}}}_{4(Pyruvate_{c},Nadh_{c})}]-k_{65}[{{\varvec{v}}}_{3(Lactate)}], \end{aligned}$$
(30)
$$\begin{aligned} \frac{dLactate_{out}}{dt}&=k_{65}[{{\varvec{v}}}_{3(Lactate)}]-k_{66}Lactate_{out}, \end{aligned}$$
(31)
$$\begin{aligned} \frac{dAtp}{dt}&=-k_{56}[{{\varvec{v}}}_{4(Glucose_{c},Atp)}]+3k_{57}[{{\varvec{v}}}_{7(G6p_{c})}]+k_{62}[{{\varvec{v}}}_{7(Acetyl_{m})}]\nonumber \\&\quad +2.5k_{64}Sco2_{m}[{{\varvec{v}}}_{7(Nadh_{m})}]+1.5k_{64}Sco2_{m}[{{\varvec{v}}}_{7(Fadh_{m})}]-k_{67}Atp, \end{aligned}$$
(32)
$$\begin{aligned} \frac{dO2_{con}}{dt}&=0.5k_{64}Sco2_{m}[{{\varvec{v}}}_{7(Nadh_{m})}]+0.5k_{64}Sco2_{m}[{{\varvec{v}}}_{7(Fadh_{m})}]. \end{aligned}$$
(33)

This model applies to all cell types, both normal and cancerous, with certain restrictions. Under the assumption that normal cells exist in a healthy environment, metabolic stress and continuous activation of growth factor signals are exclusively attributed to cancer cells. Consequently, \(k_{24}\) and \(k_{27}\) are set to zero in normal cells. Additionally, to differentiate between cancer cell types, reducing \(k_{2}\) to zero in p53-mutated cancer cells can ensure the absence of p53 response in these cells.

For a comprehensive understanding, detailed descriptions of the model-including assumptions, reactions, parameter values, experimental justifications, and sensitivity analysis-are provided in the Appendix (A). Additionally, all numerical simulations in this study were performed using MATLAB’s ’ode’ routine and Gear’s method in the XPPAUT software.

3 Results

3.1 p53 Orchestrates the Metabolic Shift in Cancer: Enhancing Oxidative Phosphorylation, Suppressing Glucose Consumption and Lactate Production

To examine the normal and cancer cellular metabolism and investigate the influence of p53 deficiency on cancer metabolic pathways seen in many laboratory experiments, we simulated the experiment conducted by Wanka et al. (2012), using our mathematical model.

In-silico, different types of cells (normal, cancer p53\(^{+/+}\), and cancer p53\(^{-/-}\)) were exposed to limited glucose (2 mM) for an 8-hour duration. Throughout this timeframe, the glucose consumed, lactate produced, oxygen consumed, and ATP produced were systematically monitored and quantified to provide a comprehensive comparison of metabolic processes across these distinct cell types, as illustrated in Fig. 2.

Our simulations succeeded in clarifying the distinctions in glucose metabolic pathways between cancer and normal cells. Cancer cells exhibited heightened glucose consumption and elevated lactate secretion, signifying their commitment to the aerobic glycolysis phenotype. Conversely, in normal cells, glucose was mainly metabolised by oxidative phosphorylation, accounting for 92% of the total ATP produced. Moreover, cancer cells in our model displayed high sensitivity to glucose availability, experiencing a notable decline in metabolic activity as glucose levels decreased. However, normal cells maintained relatively stable metabolic levels that were minimally affected by glucose fluctuations.

Our finding further confirms the significant influences of losing p53 in cancer metabolism, which caused a high tendency towards the glycolytic pathway reducing the oxygen consumption required for glucose oxidation by 22.5%, compensating that by increasing glucose consumption and thus producing lactate at higher rates. A comparison of our simulation findings with experimental observations from various studies reveals a good match (Wanka et al. 2012; Matoba et al. 2006; Wu et al. 2016). Detailed insights are presented in Table 1, corroborating the consistency between simulated and experimentally observed data.

Fig. 2
figure 2

A comparison between normal cells and cancer cells (p53\(^{+/+}\), p53\(^{-/-}\)) regarding their metabolic pathways. It shows the time course of the glucose consumption, lactate production, oxygen consumption, and ATP production by each cell type exposed to 2 mM of glucose over 8 h

Table 1 Comparing our simulation results with experimental observations under normoxia (a) and hypoxia (b) after 8 h. By fitting the glucose uptake rate to that observed in p53-deficient cancer cells (HCT116 p53\(^{-/-}\)) under normoxic conditions, the detailed results were obtained

3.2 The Influence of Abundant Extracellular Glucose Level on Stimulating High-Energy Production in Cancer Cells

Considering the dynamic nature of cellular behaviour within the body, it is crucial to note that experiments may not comprehensively capture the full spectrum of cellular responses. In the experiment we reproduced, cells were subjected to a limited supply of glucose that depletes over time. However, this scenario contrasts with the relatively constant glucose level in the bloodstream that is readily accessible to cells within the body.

Accordingly, for a more realistic representation of cellular metabolism, we need to bridge the gap between the laboratory settings and the continuous physiological conditions experienced by cells within the body. In pursuit of this goal, we replicated previous simulations but this time assumed a consistent extracellular glucose level, maintaining it at the normal physiological glucose blood concentration (5 mM) (Grupe et al. 1995), regardless of the cellular consumption rate.

By adopting this methodology, our simulations demonstrated that the maintenance of stable glucose availability prompted both normal and cancer cells to exhibit a sustained rate of glucose consumption throughout the 8-hour duration, which revealed the distinctive ability of cancer cells to produce markedly higher levels of ATP compared to our previous simulations and even more than normal cells (Fig. 3, Left).

This insight suggests that lowering the glucose levels in the bloodstream by following a specific regime could substantially diminish the ATP production in cancer cells, limiting their ability to sustain and spread. In addition, a comparative analysis of the three cell types under both limited and constant glucose levels highlights that the more the cell relies on the glycolytic pathway, the more it is affected by reducing glucose availability.

Fig. 3
figure 3

Glucose metabolism under normoxic and hypoxic conditions. A comparison of key outputs of metabolic pathways is shown for normal and cancer (p53\(^{+/+}\), p53\(^{-/-}\)) cells, considering limited extracellular glucose level that depletes over time (top row) and constant extracellular glucose levels (bottom row). The glucose consumption, lactate production, oxygen consumption, and ATP production were measured by simulating each cell type for 8 h under normoxic/hypoxic conditions

3.3 Unravelling Hypoxia’s Metabolism: Adaptive Strategies, Energy Production, and mTOR Signalling Dynamics in Cancer Progression

In cancer progression, hypoxia emerges as a vital challenge faced by rapidly proliferating cancer cells due to the formation of regions within the tumour that are deprived of an adequate blood supply. In response, cancer cells exhibit remarkable adaptive strategies. They undergo complex molecular alterations, activating a cascade of signalling pathways that drive angiogenesis (the formation of new blood vessels) to restore oxygen balance and intensify the shift toward glycolytic energy production mode to offset the deficit in respiration (Xu et al. 2019). These dynamic responses are primarily governed by stabilising HIF1, a regulator suppressed under normal conditions in an oxygen availability-dependent manner (Valvona et al. 2016; Laughner et al. 2001; Xu et al. 2019).

To investigate the impact of hypoxic conditions on cellular metabolism, we mimicked the hypoxic environment by diminishing the oxygen-dependent degradation rate of HIF-1, factoring in the HIF-1 half-life observed under hypoxic conditions (Kubaichuk and Kietzmann 2023). In parallel, we attenuated the activity of the electron transport chain by an equivalent rate (50%), considering the inadequate availability of oxygen to facilitate the oxygen reduction process. Additionally, because hypoxia is often accompanied by a lack of blood supply to cancer cells, we reduced the glucose availability to cells within the body by the same percentage (from 5 mM to 2.5 mM).

By employing this approach, we anticipated and indeed observed a notable increase in lactate fermentation by both normal and cancer cells to maximise energy production as mitochondrial capacity diminishes, consequently escalating overall glucose consumption (Fig. 3, right). This adaptive strategy mirrors the metabolic response seen in normal cells during intense exercise, where lower oxygen availability prompts alternative energy pathways. Furthermore, our simulation closely aligned the observed glucose metabolism outcomes for colon cancer cells HCT116 (p53\(^{+/+}\) and p53\(^{-/-}\)) under hypoxic conditions (O\(_{2}\) 1%) (Wanka et al. 2012), providing a good estimation of the glucose consumption levels with a slight increase in lactate production, as detailed in Table 1, (b). The discrepancy in lactate production levels may be attributed to the potential conversion of some lactate back to pyruvate, especially in instances of extremely high lactate production not accounted for in our model.

On the other hand, our simulations revealed a prominent divergence in the response to hypoxia between normal and cancer cells regarding their energy production ability. While hypoxia led to a reduction in ATP production in normal cells, cancer cells displayed resilience, maintaining their energy productivity close to normal conditions (Fig. 3, second row). This intriguing observation prompted a thorough investigation into possible factors that may be missed in our signalling network influencing energy production under hypoxic conditions. Our investigation unveiled that hypoxia typically induces the expression of the hypoxia-responsive REDD1 gene (not incorporated in our model), which, in turn, disrupts mTOR activity as a major control point to inhibit energy-intensive processes like protein translation (Brugarolas et al. 2004; Connolly et al. 2006; DeYoung et al. 2008; Horak et al. 2010). This cascade leads to a decrease in HIF1 levels and a dampening of the glycolytic pathway (Brugarolas et al. 2004; Horak et al. 2010).

Motivated by these findings, we studied the impact of inhibiting mTOR activity on metabolic pathways and energy production levels under hypoxic conditions. We constructed diagrams showcasing the metabolic activity of glycolysis and OXPHOS and their contributions to energy production under different \(k_{35}\) rates (mTOR-dependent HIF1 synthesis rate), see Fig. 4. Analysing these diagrams confirms the mTOR involvement in producing high energy levels in hypoxic cancer cells, as impeding its activity drove the cell towards a similar energy level produced in our hypoxic normal cells.

Nevertheless, numerous studies have consistently reported resistance of transformed cells to mTOR inhibition under hypoxic conditions (Connolly et al. 2006). This phenomenon is seen to preserve the protein synthesis rates and promote cell proliferation and growth under hypoxia (Brugarolas et al. 2004; Connolly et al. 2006; DeYoung et al. 2008). The engagement in energy-demanding processes, such as protein synthesis and growth, underscores the cell’s proficiency in generating ample energy. This concurs with our findings regarding hypoxic cancer cells, where the maintenance of mTOR activity correlated with a remarkable ability to produce energy even in the face of oxygen deficiency.

Fig. 4
figure 4

The effect of mTOR on the cellular metabolism and energy production levels in hypoxic cancer cells (p53\(^{+/+}\), p53\(^{-/-}\)). The glucose consumption, lactate production, oxygen consumption, and ATP production were calculated under different mTOR-dependent HIF1 activation rates, \(k_{35}\)

3.4 Dual Stable Steady States in Cancer Cells, Contrasted by Singular Stability in Normal Cells

In previous sections, normal and cancerous cells, whether possessing wild-type p53 or mutated p53, manifest distinct metabolic profiles, signifying different stability states. To explore this further, we developed a phase space presenting the nullclines and potential steady states of key players in glucose metabolic pathways, p53, HIF1, and AMPK, across both normal and cancer cells (Fig. 5).

Considering the phase space diagrams, normal cells show a unique stability with no activation of p53 and HIF1, indicative of a healthy environment. Conversely, cancer cells display two stable steady states, with an unstable one in between. The first stable steady state lacks p53 activation but exhibits a high level of HIF1, representing the case when cancer cells have p53 mutations. In contrast, the other stable steady state shows high p53 activation with a lower level of HIF1, indicating the state of cancer cells with wild-type p53.

The transition between these two states in p53 wild-type cells is governed by the phosphorylation levels of AMPK, the protein responsible for instigating the p53-metabolic stress response. This dynamic is further elucidated by the bifurcation diagram, illustrating the levels of p53 and HIF1 under various AMPK phosphorylation rates denoted as \(k_{24}\) (Fig. 6).

Under low phosphorylation rates of AMPK, cells exhibit two stable steady states: high activation of p53 (Stable SS p53\(^{+/+}\)) and no activation of p53 (Stable SS p53\(^{-/-}\)). However, exceeding the bifurcation point by increasing the phosphorylation rate (\(k_{24}\)) induces p53 activation and shifts the cell into a unique stability regime, representing the p53-wild-type state.

Fig. 5
figure 5

Phase portrait of the system in normal and cancer cells. (Top row) Nullcline corresponding to nuclear p53 (p53\(_n\)) and active AMPK (AMPK\(_{c}^*\)). (Bottom row) Nullcline corresponding to nuclear HIF1 (HIF1\(_n\)) and active AMPK (AMPK\(_{c}^*\)). The green, yellow, and blue lines represent AMPK\(_{c}^*\), p53\(_n\), and HIF1\(_n\) nullclines, respectively. Solid and hollow magenta dots denote stable and unstable equilibria, respectively. The system exhibits a single stable equilibrium point in normal cells with no p53 and HIF1 activation, while in cancer cells, two stable and one unstable equilibria are observed. For cancer cells, the stable equilibrium point with low p53\(_n\)/high HIF1\(_n\) levels represents p53-mutated cancer cells. In contrast, the one with high p53\(_n\)/low HIF1\(_n\) concentrations indicates p53-wild-type cancer cells (Color figure online)

Fig. 6
figure 6

Bifurcation diagrams demonstrate nuclear p53 and HIF1 levels driven by AMPK phosphorylation rate, \(k_{24}\), in cancer cells. The diagrams reveal a bistability regime exhibiting low p53/high HIF1 and high p53/low HIF1 levels, which represent the wild-type p53 (p53\(^{+/+}\)) and mutated p53 (p53\(^{-/-}\)) states, respectively. However, with high AMPK activation surpassing the bifurcation point, unique stability emerges, transitioning wild-type cancer cells to high p53 activation levels

3.5 Restoring Normal Metabolism in Cancer Cells by Increasing the p53 Activation Levels

Beyond its traditional roles in DNA repair and apoptosis initiation, our study highlights the enhanced activation potential of p53 to counter the Warburg effect, restoring cancer cells to a more normal metabolic state. This transformative impact unfolds across three distinct phases, depicted in Fig. 7.

During the first phase (yellow area, Fig. 7), the elevation of nuclear p53 levels leads to a modest reduction in glucose consumption and lactate production. Nevertheless, energy production levels remain high due to improved glucose respiration, explaining the increase in oxygen consumption despite the lower amount of glucose consumed. Glycolysis maintains dominance in this phase, contributing to 47%-35% of the overall energy produced.

Advancing the p53 activation will shift the cells towards the next phase (magenta area, Fig. 7), further reducing glucose uptake and lactate formation. However, this time, the ATP production is negatively impacted as a balanced state between glycolysis and oxidative phosphorylation is achieved, with glycolysis responsible for 20–34% of ATP output.

In the third phase (blue area, Fig. 7), oxidative phosphorylation overcomes glycolysis, intensifying oxygen consumption while consistently diminishing glucose utilisation and lactate production. This transition guides the cell towards achieving the standards of normal cellular metabolism, represented by the dashed black line around \(k_{2}=0.9\). Along this line, glycolysis and OXPHOS are involved in producing energy with the same percentage seen in our normal cells, attaining standard rates of glucose consumption and lactate production. However, with high activation of TIGAR, glycolysis flux is lower than that of normal cells, resulting in reduced pyruvate production and overall ATP synthesis compared to the normal cellular state.

This finding sheds light on the crucial role of p53 activation levels in the cellular outcomes following chemotherapy that activates p53 to trigger apoptosis. Unlike cells with p53 mutation, those with intact p53 can manipulate cancer cells’ metabolism in response to chemotherapy, restraining the glycolytic pathway and decreasing the intracellular ATP levels, thereby boosting cells’ sensitivity to drugs.

Fig. 7
figure 7

The effect of p53 activation on cancer metabolism. These diagrams show the steady state levels of nuclear p53 and four key metabolic indicators: glucose consumption, lactate production, oxygen consumption, and ATP generation, under varying rates of p53 phosphorylation (\(k_{2}\)) in cancer cells. Each diagram is divided into three distinct regions: the yellow region, where glycolysis dominates, contributing to 47–35% of ATP production; the magenta region, indicating a balanced state between glycolysis and oxidative phosphorylation, with glycolysis contributing to 20–34% of total ATP; and the blue region, where OXPHOS becomes dominant, accounting for more than 80% of ATP production. A black dashed line within the diagrams marks the targeted normal cellular metabolism (Color figure online)

3.6 Targeting PI3K as an Alternative Player to p53 in Modulating the Metabolism of p53-Mutated Cancer Cells

In the context of addressing cancer metabolism in cells harbouring p53 mutations, our analysis suggests an alternative strategy by targeting the growth factors signalling pathway. This critical pathway plays a central role in instigating the HIF1 and its associated targets that mainly support the aerobic glycolysis of cancer (Lien et al. 2016). The initiation of this pathway involves the activation of PI3K, leading to the transformation of PIP2 into PIP3 (Danielsen et al. 2015; Vara et al. 2004; Lien et al. 2016). Thus, our investigation has focused on perturbing this pathway by simulating methodologies such as triggering PTEN or blocking PI3K activation with specific inhibitors, like idelalisib or copanlisib (Lannutti et al. 2011; Liu et al. 2013). The outcomes reveal a profound and systematic influence on cellular metabolism, manifesting across three phases (Fig. 8).

In the initial phase (yellow area, Fig. 8), the emphasis is placed on inhibiting the glycolysis pathway while leaving the oxidative phosphorylation unaffected, leading to a decrease in total energy production following a loss of more than 20% in PIP3 concentration. Despite the reduction in glycolysis activity, it is considered predominant in this phase, accounting for over 35% of cellular energy.

As PIP3 levels decrease, the glycolytic pathway continues to diminish, indirectly prompting the OXPHOS pathway to regain its functionality and bringing the two pathways into a balanced state (magenta area, Fig. 8). Restricting pyruvate flux to lactate is expected to elevate cytosolic pyruvate concentrations, redirecting them towards mitochondria and thus promoting pyruvate oxidation. This shift is reflected in the notable oxygen consumption boost and sustained ATP production levels despite lower glucose utilisation in this phase.

In the last phase (blue area, Fig. 8), glycolysis experiences a significant decline, allowing OXPHOS to overcome it, thus improving energy production efficiency. In this stage, glucose respiration becomes the preferred cellular pathway responsible for 80–90% of the total energy output. The black dashed line in this phase signifies the targeted normal cellular metabolism, with 11% of energy production attributed to glycolysis and 89% to OXPHOS.

Our simulations reveal the efficacy of targeting the growth factors signalling pathway and highlight the potency of PI3K inhibitors in disrupting the aerobic glycolysis in p53-mutated cancer cells, enhancing therapeutic outcomes.

Fig. 8
figure 8

The impact of disrupting the growth factors signalling pathway on p53-mutated cancer cells metabolism. These diagrams depict the steady state levels of PIP3 alongside key metabolic metrics-glucose consumption, lactate production, oxygen consumption, and ATP production-at different PIP2 phosphorylation rates (\(k_{27}\)) in cancer cells (p53\(^{-/-}\)). The diagrams are categorised into three zones: yellow for glycolytic predominance (accounting for over 35% of ATP production), magenta for a metabolic balance between glycolysis and oxidative phosphorylation (20–34% of ATP from glycolysis), and blue for oxidative phosphorylation supremacy (exceeding 80% of energy output). A black dashed line marks the standard for normal cellular metabolism, with energy contributions of 11% from glycolysis and 89% from OXPHOS (Color figure online)

3.7 SCO2: A Critical Component in Boosting the OXPHOS, Yet Alone Insufficient for Reversing the Warburg Effect

Numerous studies have emphasised the crucial role of SCO2, a p53 target, in the efficient functioning of the mitochondrial respiratory chain and cellular energy production (Matoba et al. 2006; Wanka et al. 2012). SCO2 is essential for the proper assembly and function of cytochrome c oxidase (Complex IV in ETC), which catalyses electron transfer to molecular oxygen in the inner mitochondrial membrane (Matoba et al. 2006; Wanka et al. 2012). Given its significance in cellular respiration, some studies have proposed targeting it as a potential strategy to rescue oxygen consumption in p53-deficient cells and modulate the Warburg effect (Matoba et al. 2006; Wanka et al. 2012).

Inspired by these insights, we delved into the impact of boosting SCO2 levels on the metabolic phenotypes of cancer cells, particularly those with p53 mutations. In-silico, we elevated the SCO2 expression levels of p53\(^{-/-}\) cells by increasing its basal production rate, \(k_{18}\) (Fig. 9).

Indeed, our simulations agreed with those studies’ observations (Matoba et al. 2006; Wanka et al. 2012), revealing a substantial activation of aerobic respiration in a SCO2 level-dependent manner. Additionally, we noticed that when SCO2 concentration achieves its level in wild-type p53 cells, the oxygen consumption activity of p53\(^{-/-}\) cells rises at a rate comparable to that in wild-type p53 cells. This is completely consistent with what was observed in Matoba’s study, which noted that the amount of SCO2 protein needed to rescue the deficit in mitochondrial respiration of the p53\(^{-/-}\) cells corresponded well to the physiological levels observed in the p53\(^{+/+}\) cells (Matoba et al. 2006).

On the other hand, our findings also indicate that increasing SCO2 alone is insufficient to eliminate or reverse the Warburg effect. Enhancing oxidative phosphorylation does not necessarily lead to efficient suppression of the glycolysis pathway, especially with continued incentives to consume large amounts of glucose and high activation of glycolysis enzymes. This clearly explains our results, which show a slight decline in glycolysis despite a striking increase in the oxidative phosphorylation pathway (Fig. 9). Consequently, solely targeting SCO2 may elevate energy production levels, as shown in our simulations, potentially promoting the proliferative capacity of cancer cells.

In brief, our results demonstrate that SCO2 may indeed play a robust role in transforming cancer cell metabolism, but in conjunction with targeting enzymes stimulating the glycolysis pathway.

Fig. 9
figure 9

The role of SCO2 in the metabolism of p53-deficient cancer cells. The diagrams represent steady state levels of SCO2 alongside key metabolic metrics-glucose consumption, lactate production, oxygen consumption, and ATP production-influenced by various SCO2 basal production rates (\(k_{18}\)) in p53\(^{-/-}\) cancer cells. Magenta triangles denote the baseline scenario of p53\(^{-/-}\) cancer cells, whereas green triangles signify the altered state after increasing SCO2 concentration to match levels observed in p53\(^{+/+}\) cells (Color figure online)

4 Discussion

The Warburg effect is a hallmark of cancer metabolism, granting cancer cells exceptional metabolic flexibility that enables their rapid adaptation and survival in hostile microenvironments. This phenomenon is pivotal in cancer research, with particular interest in the regulatory mechanisms that govern metabolic pathways. At the forefront of these is the tumour suppressor gene p53, whose role extends beyond cell cycle control and apoptosis to include metabolic processes. Our investigation delves into the critical role of p53 in modulating cancer cell metabolism, offering novel insights into its capacity to counteract the Warburg effect phenomenon.

In terms of existing research, Linglin et al. have made good strides in elucidating the influence of genetic regulation on glycolysis and oxidative phosphorylation, identifying a hybrid metabolic state unique to cancer cells (Yu et al. 2017). However, this study did not consider the vital influence of p53, nor did it conduct a quantitative analysis of how genetic factors impact metabolic outcomes or potential strategies to mitigate the Warburg effect. Our research bridges these gaps by constructing a comprehensive mathematical framework that dissects the mechanisms through which p53, alongside other genetic regulations, influences glycolysis and OXPHOS and quantitatively explores their impact on these pathways under various cellular conditions.

Our model analysis reveals distinct metabolic profiles characterised by different stability regimes, delineating clear metabolic distinctions between normal and cancer cells with or without p53 mutations. Importantly, our model successfully replicated experimental observations on glucose metabolism in both p53-mutated and wild-type colon cancer cells, underscoring its validity.

By exploring various scenarios, our study uncovers the mechanism of how diminished glucose availability massively curtails cancer cell proliferation and viability. We further identify adaptive tactics cancer cells employ under low-oxygen conditions to maintain energy production and growth, particularly emphasising the crucial role of mTOR activation. This adaptation starkly contrasts with the energy production downturn observed in normal cells under similar hypoxic conditions, highlighting the unique metabolic resilience of cancer cells.

Interestingly, we detect a novel aspect of chemotherapy resistance linked to insufficient p53 activation levels, suggesting that beyond apoptosis evasion, inadequate p53 activity also impedes the reversal of the Warburg effect, enhancing cellular resistance.

Moreover, this study discusses strategies to combat the Warburg effect in p53-mutated contexts, evaluating the efficacy of augmenting cellular respiration by increasing the SCO2 expression levels. While this approach indeed elevates mitochondrial respiration, it does so without a noticeable reduction in the glycolysis pathway, thereby boosting the overall ATP production and potentially supporting cancer cells even further. Alternatively, we suggest inhibiting the glycolysis pathway using a PI3K inhibitor, which has shown promising results in our simulations.

While our model has shown considerable success and offered valuable insights, it is important to acknowledge its limitations. Our model does not incorporate the competitive dynamics between p53 and HIF1 over transcriptional coactivators. Transcription factors like p53 and HIF1 depend on coactivators such as p300/CBP for gene regulation, which involves acetylating histones at specific gene promoters to facilitate the recruitment of the transcriptional machinery (Grossman 2001; Freedman et al. 2002). Given the finite availability of these coactivators, competition for access between p53 and HIF1 emerges, affecting their transcriptional activities (Schmid et al. 2004).

Furthermore, our current work concentrated exclusively on glucose metabolism, yet cells can utilise additional energy sources, such as glutamine and fatty acids. Integrating these energy sources and the p53 influence on their respective metabolic pathways might give a more comprehensive overview of the metabolism outcomes and analyse the p53 role much deeper. Future research aims to expand our signalling network to include these pathways, providing a more holistic view of p53 impact on cancer metabolism.

In conclusion, this study broadens our understanding of the Warburg effect through the lens of p53 regulatory mechanisms, introducing, for the first time, a mathematical model that captures the observed impact of p53 deficiency on cancer metabolism. This pioneering model unravels the metabolic underpinnings of cancer, thoroughly scrutinising glucose metabolic pathways across different scenarios. Additionally, model findings propose fresh perspectives to improve therapeutic approaches, significantly highlighting the importance of optimal p53 activation for reversing the Warburg effect and the efficacy of PI3K inhibitors in overcoming metabolic adaptations in p53-mutated cancer cells.