Log in

A mathematical model of angiogenesis and tumor growth: analysis and application in anti-angiogenesis therapy

  • Published:
Journal of Mathematical Biology Aims and scope Submit manuscript

Abstract

The purpose of this paper is to develop a new coupled mathematical model of angiogenesis (new blood vessel growth) and tumor growth to study cancer development and anti-angiogenesis therapy. The angiogenesis part assumes the capillary to be a viscoelastic continuum whose stress depends on cell proliferation or death, and the tumor part is a Darcy’s law model regarding the tumor mass as an incompressible fluid where the nutrient-dependent growth elicits volume change. For the coupled model, we provide both an inviscid analysis and a parameter sensitivity analysis of the angiogenesis model in response to a stationary hypoxic tumor, and a steady state analysis of the tumor growth in response to a fixed and long blood capillary. The analysis shows that the stable steady state tumor with an invading blood capillary exists if and only if the nutrient release rate divided by the decay rate is less than the tumor viable limit, and the full tumor encloses one part of the capillary in this steady state. Afterwards, we use the coupled model to simulate vascularized tumor growth and anti-angiogenesis therapy. The simulations show that the tumor tends to maximize the nutrient transfer by blood vessel co-option and the anti-angiogenesis treatment by using growth factor neutralizing antibodies would regress the neovasculature and shrink the tumor size. However, the shrunken tumor mass could survive by feeding on mature blood vessels that resist the treatment. This implies the limited efficacy of the anti-angiogenesis monotherapy and its effect on vessel normalization.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Subscribe and save

Springer+ Basic
EUR 32.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or Ebook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

Price includes VAT (Germany)

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12

Similar content being viewed by others

References

  • Altrock PM, Liu LL, Michor F (2015) The mathematics of cancer: integrating quantitative models. Nat Rev Cancer 15:730–745

    Article  Google Scholar 

  • Anderson ARA, Chaplain MAJ (1998) Continuous and discrete mathematical models of tumor-induced angiogenesis. Bull Math Biol 60:857–900

    Article  Google Scholar 

  • Billy F, Ribba B, Saut O et al (2009) A pharmacologically based multiscale mathematical model of angiogenesis and its use in investigating the efficacy of a new cancer treatment strategy. J Theor Biol 260:545–562

    Article  MathSciNet  Google Scholar 

  • Byrne H, Preziosi L (2003) Modelling solid tumour growth using the theory of mixtures. Math Med Biol 20(4):341–66

    Article  Google Scholar 

  • Cai Y, Xu S, Wu J, Long Q (2011) Coupled modelling of tumour angiogenesis, tumour growth and blood perfusion. J Theor Biol 279:90–101

    Article  Google Scholar 

  • Cai Y, Wu J, Li Z, Long Q (2016) Mathematical modelling of a brain tumour initiation and early development: a coupled model of glioblastoma growth, pre-existing vessel co-option. Angiogenesis and blood perfusion. PLoS ONE 11(3):e0150296

    Article  Google Scholar 

  • Carmelie P, Jain RK (2011) Molecular mechanisms and clinical applications of angiogenesis. Nature 473:298–307

    Article  Google Scholar 

  • Carmelie P, Jain RK (2011) Principles and mechanisms of vessel normalization for cancer and other angiogenic diseases. Nat Rev Drug Discov 10:417–427

    Article  Google Scholar 

  • Cristini V, Lowengrub JS, Nie Q (2003) Nonlinear simulation of tumor growth. J Math Biol 46:191–224

    Article  MathSciNet  Google Scholar 

  • Folkman J (1971) Tumor angiogenesis: therapeutic implications. N Engl J Med 285(21):1182–6

    Article  Google Scholar 

  • Folkman J, Kalluri R (2003) Tumor angiogenesis. In: Kufe DW, Pollock RE, Weichselbaum RR et al (eds) Holland-Frei cancer medicine, 6th edn. BC Decker, Hamilton Chapter 11

    Google Scholar 

  • Frieboes HB, Lowengrub JS, Wise SM et al (2007) Computer simulation of glioma growth and morphology. Neuroimage 37:S59–S70

    Article  Google Scholar 

  • Frieboes HB, ** F, Chuang YL et al (2010) Three-dimensional multispecies nonlinear tumor growth-II: tumor invasion and angiogenesis. J Theor Biol 264(4):1254–1278

    Article  MathSciNet  Google Scholar 

  • Friedman A, Hu B (2006) Asymptotic stability for a free boundary problem arising in a tumor model. J Differ Equ 227:598–639

    Article  MathSciNet  Google Scholar 

  • Gevertz JL, Torquato S (2006) Modeling the effects of vasculature evolution on early brain tumor growth. J Theor Biol 243:517–531

    Article  MathSciNet  Google Scholar 

  • Gilbarg D, Trudinger NS (1983) Elliptic partial differential equations of second order. S**er, New York

    Book  Google Scholar 

  • Greenspan HP (1976) On the growth and stability of cell cultures and solid tumors. J Theor Biol 56:229–242

    Article  MathSciNet  Google Scholar 

  • Hwang EI, Jakacki RI, Fisher MJ et al (2013) Long-term efficacy and toxicity of Bevacizumab-based therapy in children with recurrent low-grade gliomas. Pediatr Blood Cancer 60(5):776–82

    Article  Google Scholar 

  • Lieberman GM (1996) Second order parabolic differential equations. World Scientific, Singapore

    Book  Google Scholar 

  • Lowengrub JS, Frieboes HB, ** F et al (2010) Nonlinear modeling of cancer: bridging the gap between cells and tumors. Nonlinearity 23:R1–R91

    Article  Google Scholar 

  • Lyu J, Cao J, Zhang P, Liu Y, Cheng H (2016) Coupled hybrid continuum-discrete model of tumor angiogenesis and growth. PLoS ONE 11:10

    Google Scholar 

  • Macklin P, McDougall S, Anderson ARA et al (2009) Multiscale modelling and nonlinear simulation of vascular tumour growth. J Math Biol 58:765–798

    Article  MathSciNet  Google Scholar 

  • Morris KA, Golding JF, Axon PR et al (2016) Bevacizumab in neurofibromatosis type 2 (NF2) related vestibular schwannomas: a nationally coordinated approach to delivery and prospective evaluation. Neuro-Oncol Pract 3(4):281–289

    Article  Google Scholar 

  • Murray JD (2002) Mathematical biology I. An introduction, 3rd edn. Springer, New York

    MATH  Google Scholar 

  • Murray JD (2003) Mathematical biology II. Spatial model and biomedical applications, 3rd edn. Springer, New York

    MATH  Google Scholar 

  • Naumov GN, Bender E, Zurakowski D et al (2006) A model of human tumor dormancy: an angiogenic switch from the nonangiogenic phenotype. J Natl Cancer Inst 98(5):316–325

    Article  Google Scholar 

  • Persano L, Rampazzo E, Della Puppa A, Pistollato F, Basso G (2011) The three-layer concentric model of glioblastoma: cancer stem cells, microenvironmental regulation, and therapeutic implications. Sci World J 11:1829–1841

    Article  Google Scholar 

  • Preziosi L, Tosin A (2009) Multiphase modelling of tumour growth and extracellular matrix interaction: mathematical tools and applications. J Math Biol 58:625–56

    Article  MathSciNet  Google Scholar 

  • Sciannaa M, Bell CG, Preziosia L (2013) A review of mathematical models for the formation of vascular networks. J Theor Biol 333:174–209

    Article  MathSciNet  Google Scholar 

  • Sholley MM, Ferguson GP, Seibel HR, Montour JL, Wilson JD (1984) Mechanisms of neovascularization. Vascular sprouting can occur without proliferation of endothelial cells. Lab Investig 51:624–634

    Google Scholar 

  • Spill F, Guerrero P, Alarcon T, Maini PK, Byrne HM (2015) Mesoscopic and continuum modelling of angiogenesis. J Math Biol 70:485–532

    Article  MathSciNet  Google Scholar 

  • Tang L, van de Ven AL, Guo D et al (2014) Computational modeling of 3D tumor growth and angiogenesis for chemotherapy evaluation. PLoS ONE 9(1):e83962

    Article  Google Scholar 

  • Thomlinson RH, Gray LH (1955) The histological structure of some human lung cancers and the possible implications for radiotherapy. Br J Cancer 9(4):539–549

    Article  Google Scholar 

  • Xu J, Vilanova G, Gomez H (2016) A mathematical model coupling tumor growth and angiogenesis. PLoS One 11(2):e0149422

    Article  Google Scholar 

  • Ye W (2016) The complexity of translating anti-angiogenesis therapy from basic science to the clinic. Dev Cell 37(2):114–125

    Article  Google Scholar 

  • Yeh AC, Ramaswamy S (2015) Mechanisms of cancer cell dormancy—Another hallmark of cancer? Cancer Res 75(23):5014–5022

    Article  Google Scholar 

  • Zheng X, **e C (2014) A viscoelastic model of blood capillary extension and regression: derivation, analysis, and simulation. J Math Biol 68:57–80

    Article  MathSciNet  Google Scholar 

  • Zheng X, Wise S, Cristini V (2005) Nonlinear simulation of tumor necrosis, neo-vascularization and tissue invasion via an adaptive finite-element/level-set method. Bull Math Biol 67:211–259

    Article  MathSciNet  Google Scholar 

  • Zheng X, Koh GY, Jackson T (2013) A continuous model of angiogenesis: initiation, extension, and maturation of new blood vessels modulated by vascular endothelial growth factor, angiopoietins, platelet-derived growth factor-B, and pericytes. DCDS-B 18(4):1109–1154

    Article  MathSciNet  Google Scholar 

  • Zheng X, Sweidan M (2018) A numerical method for two-point boundary value problem with non-fitting mesh and second order truncation error. J Comput Appl Math (submitted)

Download references

Acknowledgements

Zheng thanks Chun**g **e and Institute of Natural Sciences of Shanghai Jiaotong University for the accommodation for his visit in April, May, and June, 2017. Part of this work was done during the visit.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to **aoming Zheng.

Appendices

Appendices

1.1 Appendix A. Proof of Theorem 1

Proof

It suffices to prove for \(\beta =1\). First, introduce \(w(X,t)=X+u(X,t)\). Second, perform the odd extension of w from \([0,L_V]\) to \([-L_V,L_V]\), that is, letting

$$\begin{aligned} {\tilde{w}}(X,t) = \left\{ \begin{array}{lc} w(X,t) &{}\quad X\in [0,L_V],\\ -w(-X,t) &{}\quad X\in [-L_V,0]. \end{array} \right. \end{aligned}$$
(64)

Similarly \(\psi \) is extended to \({\tilde{\psi }}\). Third, define \(\phi ={\tilde{w}}_X=1+\frac{\partial u}{\partial X}\) and denote \({\tilde{\Omega }}=[-L_V,L_V]\). Then the Eq. (28) leads to \({\tilde{w}}_t = \frac{{\tilde{w}}_{XX} }{{\tilde{w}}_{X}}\). Differentiate w.r.t. X and notice that \(\phi ={\tilde{w}}_X\). Then we get

$$\begin{aligned} \phi _t = \frac{\phi _{XX} \phi - \phi _{X}^2}{\phi ^2}, \end{aligned}$$
(65)

with the initial condition \(\phi (X,0)=1+\frac{\partial {\tilde{\psi }}}{\partial X}\) and the boundary condition \(\phi (\pm L_V)=e^{\beta ^*_{ec}t}+g\). Using the maximum principle [e.g., Lemma 2.3 in Lieberman (1996)], we obtain

$$\begin{aligned} \min _{\partial '{\tilde{\Omega }}_T} \phi \le \phi (X,t) \quad \text { for } (X,t)\in {\tilde{\Omega }}_T, \end{aligned}$$
(66)

where \({\tilde{\Omega }}_T=[-L_V, L_V]\times [0,T]\) and \(\partial '{\tilde{\Omega }}_T= \partial '{\tilde{\Omega }}_T \backslash \{(x,T): -L_V< X < L_V \}\). It is easy to check \(\frac{\partial {\tilde{\psi }}}{\partial X}\ge 0\). Thus, \(\phi (X,0)\ge 1\). Furthermore, \(\phi (\pm L_V,t)=e^{\beta ^*_{ec}t} +g>0\), therefore, \(\phi \ge 0\) in \({\tilde{\Omega }}_T\). The global existence of a unique solution has the same proof as in that of Theorem 1(a) of Zheng and **e (2014). \(\square \)

1.2 Appendix B. Details of parameter study of angiogenesis model

As the VEGF diffusion rate \(D_c\) is doubled, the VEGF concentration is higher (compare Figs. 2b, 13b), which promotes faster extension of the capillary. The capillary reaches \(x=1\) around \(t=4.2\). As \(D_c\) is halved, the VEGF concentration is only slightly greater than \(c_0=0.2\) near the root (Fig. 13d). Thus, the EC proliferation activity is mostly suppressed, and the capillary grows very slow. At \(t=7\), the tip only extends to \(x=0.44\) (Fig. 13c).

Fig. 13
figure 13

Modulation on VEGF diffusion rate \(D_c\). a, b \(D_c\) is doubled. From \(t=0,1,\ldots \)4, u in a goes from bottom to top, and c in b goes from top to bottom. c, d \(D_c\) is halved. In c, the curve of u at \(t=0,1,\ldots \)7 goes from bottom to top

As the VEGF decay rate \(\gamma _1\) is doubled, the VEGF concentration is lower (Fig. 14b). The capillary grows very slow and reaches only \(x=0.44\) at \(t=7\) (Fig. 14a). As \(\gamma _1\) is halved, the VEGF concentration is much higher (Fig. 14d). Thus, the EC proliferation activity is highly activated, and the capillary grows very fast. At \(t=4.3\), the tip extends to \(x=1\) ((Fig. 14c).

Fig. 14
figure 14

Modulation on VEGF decay rate \(\gamma _1\). a, b \(\gamma _1\) is doubled, u in a and c in b, \(t=0,1,\ldots ,7\). c, d \(\gamma _1\) is halved. From \(t=0,1,\ldots \)5, u in c goes from bottom to top and c in d goes from top to bottom

Fig. 15
figure 15

Modulation on EC proliferation rate \(\beta _{EC}\). a, b \(\beta _{EC}\) is doubled. From \(t=0,1,\ldots \)4, u in a and \(\rho \) in b both go from bottom to top. c, d \(\beta _{EC}\) is halved. From \(t=0,1,\ldots \)7, u in c and \(\rho \) in d both go from bottom to top

When the EC proliferation rate \(\beta _{EC}\) is doubled, the EC density \(\rho \) is much higher near the tip than the control case (compare Figs. 3a, 15b), and the capillary extends much faster and reaches the VEGF source at \(x=1\) around \(t=3.5\) (Fig. 15a). When \(\beta _{EC}\) is halved, the capillary grows very slow and reaches only \(x=0.54\) at \(t=7\).

When the capillary tip pulling force g is doubled, the capillary extends to \(x=1\) at \(t=4\) as shown in Fig. 16a. However, the Eulerian EC density is lower compared with the control (compare Figs. 3c, 16b). This is due to, on the one hand, the larger pulling force elongating the cells thus reducing density, and on the other hand, the faster capillary extension (\(t=4\) compared with \(t=7\) in control) decreasing the time duration for mass increase. When g is halved, the capillary extends to \(x=0.67\) at \(t=7\) (Fig. 16c), and the Eulerian EC density is higher (Fig. 16d).

Fig. 16
figure 16

Modulation on the pulling force g. a, b g is doubled, for \(t=0,1,\ldots ,4\). c, d g is halved, for \(t=0,1,\ldots ,7\)

Fig. 17
figure 17

Modulation on EC proliferation threshold \(c_0\). ac \(c_0\) is doubled, \(t=0,1,\ldots ,7\). df \(c_0\) is halved, \(t=0,1,\ldots ,5\)

If the VEGF proliferation threshold \(c_0\) is doubled, then VEGF concentration in the region \(x<0.4\) are merely slightly greater than \(c_0\) (Fig. 17b). Thus, the ECs do not proliferate much (Fig. 17c). The displacement of the capillary tip by \(t=7\) is only \(u=0.3\), approximately. This value is mainly contributed by the pulling force because the steady state of u for a constant EC density \(\rho =1\) at the tip is 0.3 when \(g=3\). On the other hand, when \(c_0\) is halved, the VEGF available for EC proliferation is increased and the EC density is larger than the control (compare at \(t=5\) Figs. 3a, 17f). The capillary reaches \(x=1\) at \(t=5\).

1.3 Appendix C. Steady state analysis of tumor growth without capillary or with a whole-space capillary

Consider the following equations,

$$\begin{aligned} 0= & {} D_n n_{xx} - \gamma _n n + \beta _n, \quad x\in (x_L, x_R), \end{aligned}$$
(67)
$$\begin{aligned} n(x_L)= & {} n_L, \quad n(x_R)=n_R, \end{aligned}$$
(68)
$$\begin{aligned} v= & {} - K p_x, \quad x\in (x_L, x_R), \end{aligned}$$
(69)
$$\begin{aligned} v_x= & {} \beta _{tc} (n-n_g), \quad x\in (x_L, x_R), \end{aligned}$$
(70)
$$\begin{aligned} p(x_L)= & {} p_L, \quad p(x_R)=p_R, \end{aligned}$$
(71)
$$\begin{aligned} \frac{d x_L(t)}{dt}= & {} v(x_L,t),\quad \frac{d x_R(t)}{dt}=v(x_R,t). \end{aligned}$$
(72)

When \(\beta _n=0\), the above equations represent the tumor growth without the effects of the blood capillary. When \(\beta _n>0\), it represents the blood capillary occupying the whole space \(\mathbb {R}\).

Let \({\hat{\gamma }}=\sqrt{\gamma _n/D_n}\), \({\hat{\beta }}=\beta _n/\gamma _n\). The nutrient solution of (67) and (68) is

$$\begin{aligned} n=\frac{(n_R-{\hat{\beta }}) \sinh ({{\hat{\gamma }}} (x-x_L)) - (n_L-{\hat{\beta }}) \sinh ({{\hat{\gamma }}}(x-x_R)) }{\sinh ({{\hat{\gamma }}} (x_R-x_L))} + {\hat{\beta }}. \end{aligned}$$
(73)

Inserting (69) into (70), we get

$$\begin{aligned} -p_{xx}= & {} \frac{\beta _{tc}}{K} (n-n_g) , \quad \text { in } \Omega _T. \end{aligned}$$
(74)

The above equation equipped with (71) is an elliptic problem. The pressure solution is

$$\begin{aligned} p(x)= & {} \frac{-\beta _{tc}}{K} \left( \frac{ (n_R-{\hat{\beta }}) \sinh ({{\hat{\gamma }}} (x-x_L)) - (n_L-{\hat{\beta }}) \sinh ({{\hat{\gamma }}}(x-x_R)) }{{\hat{\gamma }}^2 \sinh ({{\hat{\gamma }}}(x_R-x_L))}\right. \nonumber \\&\quad \left. - \frac{(n_g-{\hat{\beta }}) x^2}{2} + c_1 x + c_2 \right) , \end{aligned}$$
(75)

where

$$\begin{aligned} c_1= -\frac{K}{\beta _{tc} } \frac{p_R-p_L}{x_R-x_L} - \frac{n_R-n_L}{{\hat{\gamma }}^2 (x_R-x_L)} + \frac{(n_g-{\hat{\beta }}) (x_R+x_L)}{2}. \end{aligned}$$
(76)

The velocities at the tumor boundary are computed as \(v=-K p_x\), and they are equal to

$$\begin{aligned} v(x_L)= & {} \beta _{tc} \frac{(n_R-{\hat{\beta }}) - (n_L-{\hat{\beta }}) \cosh (\hat{V})) }{{\hat{\gamma }} \sinh (\hat{V})} -\frac{\beta _{tc}(n_R-n_L) }{{{\hat{\gamma }}} \hat{V}} \nonumber \\&+ \frac{\beta _{tc} (n_g-{\hat{\beta }}) \hat{V}}{2{{\hat{\gamma }}}} - K \frac{p_R-p_L}{x_R-x_L}, \end{aligned}$$
(77)
$$\begin{aligned} v(x_R)= & {} \beta _{tc} \frac{(n_R-{\hat{\beta }}) \cosh (\hat{V}) -(n_L-{\hat{\beta }}) }{{\hat{\gamma }} \sinh (\hat{V})} -\frac{\beta _{tc}(n_R-n_L) }{{{\hat{\gamma }}} \hat{V}}\nonumber \\&-\frac{\beta _{tc} (n_g-{\hat{\beta }}) \hat{V}}{2{{\hat{\gamma }}}} - K \frac{p_R-p_L}{x_R-x_L}, \end{aligned}$$
(78)

where \({\hat{V}}={\hat{\gamma }} (x_R-x_L)\) is the tumor size. Since \(\frac{d{\hat{V}}}{dt} = {\hat{\gamma }}\left( \frac{d x_R(t)}{dt} - \frac{d x_L(t)}{dt} \right) = {\hat{\gamma }}(v(x_R,t) - v(x_L,t))\), we get

$$\begin{aligned} \frac{d\hat{V}}{d\hat{t}} = \hat{V} \left( 2(1-\beta ^*) \frac{\cosh (\hat{V})-1}{\hat{V} \sinh (\hat{V})} - (n_g^*-\beta ^*) \right) , \end{aligned}$$
(79)

where \(\hat{t}=\beta _{tc} \frac{(n_L+n_R)}{2} t\),

$$\begin{aligned} \beta ^*=\frac{2{\hat{\beta }}}{n_L+n_R}=\frac{2\beta _n}{\gamma _n(n_L+n_R)}, \quad n_g^*=\frac{2n_g}{n_L+n_R}. \end{aligned}$$
(80)

If \(\beta ^*=1\), then \({\hat{V}}={\hat{V}}_0 e^{(1-n_g^*)t}\). In this case, there is a unique steady state \({\hat{V}}=0\), which is stable if \(1\le n_g^*\) and unstable if \(1>n_g^*\).

If \(\beta ^*\ne 1\), then the nonzero steady state \({\hat{V}}^*\) satisfies

$$\begin{aligned} g({\hat{V}}^*)=\frac{n_g^*-\beta ^*}{1-\beta ^*}, \end{aligned}$$
(81)

where

$$\begin{aligned} g(x)\triangleq 2\frac{ (\cosh (x)-1)}{x\sinh (x)}. \end{aligned}$$
(82)

Direct calculation shows that \(\lim \limits _{x\rightarrow 0}g (x)=1\) and \(\lim \limits _{x\rightarrow \infty }g(x)=0\) and \(g'(x)= \frac{2 (x-\sinh (x))(\cosh (x)-1)}{(x\sinh (x))^2}\). It is easy to check that \(g'(x)<0\) when \(x>0\). Thus in the domain \((0,\infty )\), g(x) is decreasing from 1 to 0 as x moves from 0 to infinity. We summarize these result in the following lemma.

Lemma 1

The function g(x) defined in (82) (the same as in (53)) is a monotonically decreasing function in the domain \((0,\infty )\) and \(\lim \limits _{x\rightarrow 0+}g (x)=1\) and \(\lim \limits _{x\rightarrow \infty }g(x)=0\).

The existence of the nonzero steady states requires \(0<\frac{n_g^*-\beta ^*}{1-\beta ^*}<1\). If \(\beta ^*<1\), then it is equivalent to \(\beta ^*<n_g^*<1\). In this case, the steady state \(\hat{V}^*=g^{-1}( \frac{n_g^*-\beta ^*}{1-\beta ^*})\) and is stable according to (79). If \(\beta ^*>1\), then \(0<\frac{n_g^*-\beta ^*}{1-\beta ^*}<1\) is equivalent to \(\beta ^*>n_g^*>1\). There also exists a unique nonzero steady state \({\hat{V}}^*\) satisfying (81), which is unstable according to (79).

In the nonzero steady state of \(\hat{V}\), we can get \(n_g^*-\beta ^* =2(1-\beta ^*) \frac{\cosh (\hat{V})-1}{\hat{V} \sinh (\hat{V})}\) from (79). Inserting this relation to the above velocities at the boundary leads to

$$\begin{aligned} v(x_L)=v(x_R)=\frac{\beta _{tc}}{{{\hat{\gamma }}}} \frac{n_R-n_L}{2} \left( \frac{\cosh (\hat{V}) + 1 }{\sinh (\hat{V}) } - \frac{2}{\hat{V}} \right) - K \frac{p_R-p_L}{x_R-x_L}. \end{aligned}$$
(83)

Remark 4

In the steady states of tumor size, the whole tumor mass moves with a constant speed. In (83), the first term on the right has the same sign as \(n_R-n_L\), and the second term has the opposite sign as \(p_R-p_L\). This shows the tumor always moves towards the region of higher nutrient concentration and escapes away from the higher pressure.

The analysis in this section can be summarized as follows.

Theorem 3

The following results hold for the tumor growth model (67)–(72).

  1. (1)

    If \(\beta ^*=1\), then there is a unique steady state \({\hat{V}}=0\), which is stable if \(1\le n_g^*\) and unstable if \(1>n_g^*\).

  2. (2)

    If \(\beta ^*<n_g^*<1\), then the tumor size has two steady states, which are \(\hat{V}=0\) (unstable) and \(\hat{V}^*=g^{-1}( \frac{n_g^*-\beta ^*}{1-\beta ^*})\) (stable).

  3. (3)

    If \(\beta ^*>n_g^*>1\), then the tumor size has two steady states, which are \(\hat{V}=0\) (stable) and \(\hat{V}^*=g^{-1}( \frac{n_g^*-\beta ^*}{1-\beta ^*})\) (unstable).

  4. (4)

    When the tumor size is in a nonzero steady state, the tumor moves with a constant speed given by (83).

  5. (5)

    In any other cases, there are no nonzero steady states of tumor size.

The tumor is truly stationary (both the tumor size and the location are invariant) only when the boundary velocity given by (83) is zero. This can occur if \(n_R=n_L\) and \(p_R=p_L\), or these boundary conditions of nutrient and pressure are delicately modulated to make the boundary velocity zero. However, even in the truly stationary state, the inner velocity of the tumor does not vanish. Indeed, by using \(v=-Kp_x\) and (75) we can obtain at any moment (including unsteady state)

$$\begin{aligned} v(x)= & {} \beta _{tc} \left( \frac{(n_R-{\hat{\beta }}) \cosh ({{\hat{\gamma }}}(x-x_L))- (n_L-{\hat{\beta }}) \cosh ({{\hat{\gamma }}}(x-x_R)) }{ {{\hat{\gamma }}} \sinh ({\hat{\gamma }} (x_R-x_L))} \right. \nonumber \\&\left. - \frac{ (n_R-n_L) }{{\hat{\gamma }}^2 (x_R-x_L)} + \frac{ (n_g-{\hat{\beta }}) (x_R+x_L-2x)}{2} \right) - K \frac{p_R-p_L}{x_R-x_L}. \end{aligned}$$
(84)

The typical graphs of nutrient, pressure, and velocity when the tumor is in the steady state are plotted in Fig. 18. Note the velocity on both the left and right sides of the tumor region is pointing towards the tumor center.

Fig. 18
figure 18

One steady state of tumor. \(K=1\), \(D_n=1\), \(\gamma _n=10\), \(\beta _n=0\), \(\beta _{tc}=1\), \(n_g=0.5\), \(n_R=n_L=1\), \(p_L=p_B=1\). Thus, \(\beta ^*=0\) and \(n_g^*=0.5\). The tumor region is where the nutrient \(n<1\), roughly [0.4, 1.6]. Outside the tumor, we let \(n=1\), \(p=0\), and \(u=0\)

From the velocity solution (84), we realize the cell motility constant K only shows up in Darcy’s term, \(K\frac{p_R-p_L}{x_R-x_L}\), which is in accordance with Darcy’s law of fluid velocity in porous medium (extracellular matrix in this case). This term provides a uniform background flow advecting the entire tumor volume. If the boundary pressures are equal, i.e., \(p_L=p_R\), then Darcy’s advection will vanish, which means the cell motility has no effect on tumor growth. But even if \(p_L\ne p_R\), Darcy’s term still has no effect on the rate of change of tumor size [see Eq. (79)]. The fundamental reason for this neglect is the constant density assumption. Indeed, due to the constant density assumption, the rate of change of the tumor volume \(\Omega _T\) is given by

$$\begin{aligned} \frac{d}{dt} \int _{\Omega _T} d\vec {x} = \int _{\Omega _T} \nabla \cdot \vec {v} d\vec {x} = \int _{\Omega _T} \beta _{tc} (n-n_g) d\vec {x}. \end{aligned}$$
(85)

Thus, the rate of change of tumor volume solely relies on proliferative parameters, and is irrelevant to cell motility or pressure. But there is one concern when the cell motility is small (e.g., cells are very viscous or permeability is small) but the divergence of velocity remains the same because the pressure magnitude would be very large according to Darcy’s law or solution (75). The high pressure would prevent cells from proliferation or even crush the cells. Therefore, the current model would be valid only for medium to large cell motility.

1.4 Appendix D. Numerical method for the coupled model of angiogenesis and tumor growth

The numerical algorithm is briefly stated below. First, set up the initial values of VEGF, nutrient, blood vessels, and tumor. At each time step \(t^k\), \(k=0,1,\ldots \), denote the tumor domain as \(\Omega _{T_k}=(x_L(t^k), x_R(t^k))\), VEGF as \(c(x,t^{k})\), EC density as \(\rho (X,t^{k})\), EC displacement as \(u(X,t^{k})\), where \(x\in (0, L_{tissue})\), \(X\in (0, L_V)\), then we do the following.

  1. (1)

    Solve the nutrient \(n(x,t^k)\) inside the tumor domain \(\Omega _T\) with Eqs. (20) and (25).

  2. (2)

    Solve the pressure \(p(x,t^k)\) from

    $$\begin{aligned} -K p_{xx} = \beta _{tc} (n-n_g), x\in (x_L, x_R) \end{aligned}$$
    (86)

    with the boundary value (24).

  3. (3)

    Solve the velocity \(v(x,t^k)\) on the tumor boundary \(x_L(t^k)\) and \(x_R(t^k)\) with Eq. (18).

  4. (4)

    Update the tumor boundary points \(x_R\) and \(x_L\) with (21).

  5. (5)

    Solve VEGF \(c(x,t^{k+1})\) with Eq. (15).

  6. (6)

    Solve EC density \(\rho (X,t^{k+1})\) in the capillary domain \((0,L_V)\) with Eq. (16).

  7. (7)

    Solve EC displacement \(u(X,t^{k+1})\) in the capillary domain \((0,L_V)\) with Eq. (17) and boundary conditions (23).

  8. (8)

    Let \(k=k+1\), then go to the next time step \(t^{k+1}=t^k+\Delta t\).

A second order numerical method is developed to solve the nutrient and pressure equations and the details are in Zheng and Sweidan (2018)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zheng, X., Sweidan, M. A mathematical model of angiogenesis and tumor growth: analysis and application in anti-angiogenesis therapy. J. Math. Biol. 77, 1589–1622 (2018). https://doi.org/10.1007/s00285-018-1264-4

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00285-018-1264-4

Keywords

Mathematics Subject Classification

Navigation