Citation: Ana M. Martín, Javier Domínguez, Julio Amador. Applying LIDAR datasets and GIS based model to evaluate solar potential over roofs: a review[J]. AIMS Energy, 2015, 3(3): 326-343. doi: 10.3934/energy.2015.3.326
[1] | Mahya Mohammadi, M. Soltani, Cyrus Aghanajafi, Mohammad Kohandel . Investigation of the evolution of tumor-induced microvascular network under the inhibitory effect of anti-angiogenic factor, angiostatin: A mathematical study. Mathematical Biosciences and Engineering, 2023, 20(3): 5448-5480. doi: 10.3934/mbe.2023252 |
[2] | Giulio Caravagna, Alex Graudenzi, Alberto d’Onofrio . Distributed delays in a hybrid model of tumor-Immune system interplay. Mathematical Biosciences and Engineering, 2013, 10(1): 37-57. doi: 10.3934/mbe.2013.10.37 |
[3] | Lifeng Han, Steffen Eikenberry, Changhan He, Lauren Johnson, Mark C. Preul, Eric J. Kostelich, Yang Kuang . Patient-specific parameter estimates of glioblastoma multiforme growth dynamics from a model with explicit birth and death rates. Mathematical Biosciences and Engineering, 2019, 16(5): 5307-5323. doi: 10.3934/mbe.2019265 |
[4] | Floriane Lignet, Vincent Calvez, Emmanuel Grenier, Benjamin Ribba . A structural model of the VEGF signalling pathway: Emergence of robustness and redundancy properties. Mathematical Biosciences and Engineering, 2013, 10(1): 167-184. doi: 10.3934/mbe.2013.10.167 |
[5] | Urszula Foryś, Yuri Kheifetz, Yuri Kogan . Critical-Point Analysis For Three-Variable Cancer Angiogenesis Models. Mathematical Biosciences and Engineering, 2005, 2(3): 511-525. doi: 10.3934/mbe.2005.2.511 |
[6] | Austin Baird, Laura Oelsner, Charles Fisher, Matt Witte, My Huynh . A multiscale computational model of angiogenesis after traumatic brain injury, investigating the role location plays in volumetric recovery. Mathematical Biosciences and Engineering, 2021, 18(4): 3227-3257. doi: 10.3934/mbe.2021161 |
[7] | Emad Attia, Marek Bodnar, Urszula Foryś . Angiogenesis model with Erlang distributed delays. Mathematical Biosciences and Engineering, 2017, 14(1): 1-15. doi: 10.3934/mbe.2017001 |
[8] | Hao Wang, Yang Kuang . Alternative models for cyclic lemming dynamics. Mathematical Biosciences and Engineering, 2007, 4(1): 85-99. doi: 10.3934/mbe.2007.4.85 |
[9] | John D. Nagy, Dieter Armbruster . Evolution of uncontrolled proliferation and the angiogenic switch in cancer. Mathematical Biosciences and Engineering, 2012, 9(4): 843-876. doi: 10.3934/mbe.2012.9.843 |
[10] | Marek Bodnar, Monika Joanna Piotrowska, Urszula Foryś, Ewa Nizińska . Model of tumour angiogenesis -- analysis of stability with respect to delays. Mathematical Biosciences and Engineering, 2013, 10(1): 19-35. doi: 10.3934/mbe.2013.10.19 |
In biology and medicine we may observe a wide spectrum of self-organization phenomena. This may happen at any scale; from the cellular scale of embryonic tissue formation, wound healing or tumor growth, and angiogenesis, to the much larger scale of animal grouping. Such phenomena are usually explained in terms of a collective behavior driven by "forces", either external and/or internal, acting upon individuals (cells or organisms). In most of these organization phenomena, randomness plays a major role; here we wish to address the issue of the relevance of randomness as a key feature for producing nontrivial organization of biological structures (see [8] for a general discussion on this topic). As a working example we offer a review of papers by the same authors in which tumor-driven angiogenesis has been analyzed [10], [4], [41]. In this case cells organize themselves as a capillary network of vessels, the organization being driven by a family of underlying fields, such as nutrients, growth factors and alike [22,25,13,19,21,15,14].
Actually an angiogenic system is extremely complex due to its intrinsic multiscale structure. When modelling such systems, we need to consider the strong coupling between the kinetic parameters of the relevant microscale branching and growth stochastic processes of the capillary network and the family of interacting macroscale underlying fields. Capturing the keys of the whole process is still an open problem while there are many models in the literature that address some partial features of the angiogenic process [2,36,30,31,27,37,38,35,10,39,34,20].
The importance of using an intrinsically stochastic model at the microscale to describe the generation of a realistic vessel network has been the subject of a series of papers by one of the present authors [8,11].
Typical models treat vessel cells on the extracellular matrix as discrete objects, and different cell processes like migration, proliferation, etc. occur with certain probabilities. The latter depend on the concentrations of certain chemical factors which satisfy systems of reaction-diffusion equations (RDEs) [2,24,42,34].
Viceversa, the RDEs for such underlying fields contain terms that depend on the spatial distribution of vascular cells. As a consequence, a full mathematical model of angiogenesis consists of the (stochastic) evolution of vessel cells, coupled with a system of RDEs containing terms that depend on the distribution of vessels. The latter is random and therefore the equations for the underlying fields are random RDEs, which are supposed to drive the kinetic parameters of the stochastic geometric processes of birth (branching), growth (vessel extension), and death (anastomosis).
This strong coupling leads to an highly complex mathematical problem from both analytical and computational points of view. A possibility to reduce complexity is offered by the so called hybrid models, which exploit the natural multiscale nature of the system.
The idea consists of approximating the random RDEs by deterministic ones, in which the microscale (random) terms depending on cell distributions are replaced by their (deterministic) mesoscale averages. In this way only a simple stochasticity of the geometric processes of branching -vessel extension -anastomosis is kept, driven now by deterministic kinetic parameters depending upon the above mentioned mean field approximation of the concentrations of the relevant fields [10,28].
Our analysis has shown that nontrivial problems arise when deriving deterministic equations for the mean spatial densities of the relevant stochastic entities modelling the vessels' evolution at the microscale.
In the literature there are examples of rigorous derivations of mean field equations of stochastic particle dynamics [29,40,17,6]. However, to the best of the authors' knowledge, the kind of models considered here have not yet been studied and require further investigation.
In angiogenesis, the leading mechanisms driven by the underlying fields consist of vessel branching, elongation and anastomosis. This last one has been the major and critical addition to the model proposed in [10]. In [4] anastomosis has been modelled as a death process of a tip that encounters an existing vessel and is therefore coupled with the density of the full vessel network. In order to cope with anastomosis, the vessel network has been modelled as a stochastic geometric process of Hausdorff dimension one, as opposed to the system of tips which form a usual stochastic particle system of Hausdorff dimension zero.
In [4] we have been able to derive (at least formally) a mean field equation for the spatial density of tips, as a function of spatial location and velocity. This equation is a parabolic integrodifferential equation of a Fokker-Planck type, having a source term and a noninvertible diffusion matrix; it is second order in the derivatives with respect to the velocities, and first order in the derivatives with respect to the position coordinates. Apparently, together with the mean field equations for the underlying fields, we have thus found an independent (deterministic) integrodifferential system whose solution can provide the required (deterministic) kinetic parameters, which drive the stochastic system for the tips, eventually leading to the stochastic vessel network, at the microscale.
Mean field equations that follow from the "propagation of chaos" assumption (law of large numbers) (see [29], [40], and references therein) are quite convenient as they hold for any given realization of the underlying self-averaging stochastic processes, but this requires a large number of tips at any location in the phase space, and at any time.
Unfortunately anastomosis is responsible for a significant decay of the number of tips, so to make inapplicable laws of large numbers on a single realization of the stochastic process (a replica of the system), according to the "propagation of chaos".
However by re-examining the derivation given in [4], in [41] we have noticed that the same deterministic description holds for vessel tip densities calculated by averaging over replicas. This is close to J.W. Gibbs' original ensemble average interpretation of equilibrium statistical mechanics [23], except that, of course, our system is always very far from equilibrium. In either cases, though with a different interpretation of the solution, the deterministic description consists of a reaction-diffusion equation for the tumor angiogenic factor (TAF) concentration coupled to a Fokker-Planck type equation for the vessel tip density. To conclude, randomness cannot be avoided; the deterministic description represents the evolution of an ideal "mean" angiogenesis system and the evolution of a particular replica may deviate significantly from it. These deviations are important and deserve further study, but this is outside the scope of the present paper (see [8]).
The paper is organized as follows. Section 2 describes how our stochastic model treats vessel branching, extension and anastomosis. In the subsequent three sections we present the mathematical ingredients to be used in the sequel. In Section6 we derive the mean field approximation based on the propagation of chaos assumption. In Section 7 a discussion is presented based on the outcomes of the numerical simulations of the stochastic model as opposed to the mean field model. In Section 8 the ensemble average approach is presented to derive an equation of Fokker-Planck type for the density of vessel tips and the TAF's RDE. In Section 9 numerical simulations of the deterministic averaged model are presented as opposed to the stochastic model. Finally Section 10 contains our conclusions.
Based on the above discussion, the main features of the process of formation of a tumor-driven vessel network that we have considered are (see [18,30,10])
ⅰ) vessel branching;
ⅱ) vessel extension;
ⅲ) chemotaxis in response to a generic tumor angiogenic factor (TAF), released by tumor cells;
ⅳ) haptotactic migration in response to fibronectin gradient, emerging from the extracellular matrix and through degradation and production by endothelial cells themselves;
ⅴ) anastomosis, the coalescence of a capillary tip with an existing vessel.
Let
$\mathbf{X}(t)=\displaystyle\bigcup\limits_{i=1}^{N(t)} \{\mathbf{X}^i(s), T^i\le s\le \min\{t, \Theta^i \} \}, \label{network} $ | (1) |
giving rise to a stochastic network. It is clear that
We may describe both random sets by means of random distributions á la Dirac-Schwarz [12];
For convenience of the reader, we remind here that, under sufficient regularity assumptions on a random set
$ \lambda_{\mathbf{\Xi}}(\mathbf{x})=\mathbb{E} [\delta_{\mathbf{\Xi}}](\mathbf{x}).\label{mean_density} $ | (2) |
TAF, fibronectin and matrix degrading enzymes activate the migration of endothelial cells.
TAF diffuses, and it decreases where endothelial cells are present; we will adopt the following model for the TAF evolution, according to which "consumption" (actually molecular binding) is due to the additional endothelial cells producing vessels' extension. Consumption is then taken proportional to the velocity
$
∂∂tC(t,x)=d0δA(x)+d2△C(t,x)−ηC(t,x)1NN(t)∑i=1|vi(t)|(KN∗δXi(t))(x).
$
|
(3) |
Parameters
Model (3) considers a dependence upon the (mollified) empirical distribution of the variation in length of the existing vessels, per unit time.
The parameter
Specific choices about the dependence of the kernel
Fibronectin is bound to the extracellular matrix and does not diffuse [3]. Degradation of fibronectin, characterized by a coefficient
$ \frac{\partial}{\partial t}f(t,\mathbf{x}) \,=\, -\zeta_1 m(t,\mathbf{x}) f(t,\mathbf{x}) + \zeta_2 \frac1N\sum\limits_{i=1}^{N(t)}({K_N \ast \delta_{ \mathbf{X}^i(t)}})(\mathbf{x}). \label{fibronectin_evolution} $ | (4) |
The MDE, once produced at a rate
$ \frac{\partial}{\partial t}m(t,\mathbf{x}) \,=\, \epsilon_1 \triangle m(t,\mathbf{x}) -\nu_2 m(t,\mathbf{x}) + \nu_1 \frac1N\sum\limits_{i=1}^{N(t)}({K_N \ast \delta_{\mathbf{X}^i(t)}})(\mathbf{x}). \label{mde_evolution} $ | (5) |
All these (random) partial differential equations are subject to suitable boundary and initial conditions.
For the sake of technical simplification, in the sequel we will not consider the last two underlying fields
Let us assume that, at any time
There are two fundamental random measures describing the system at any time
$ Q_N(t):=\frac{1}{N} \sum\limits_{k=1}^{N(t)}\epsilon_{(\mathbf{X}^k(t),\mathbf{v}^k(t))}. \label{def: _empirical measure_global} $ | (6) |
Here
Consequently, the random empirical measure of the process
$ T_N(t)=\frac{1}{N} \sum\limits_{k=1}^{N(t)}\epsilon_{\mathbf{X}^k(t)}= Q_N(t)(\cdot\times\mathbb R^d). \label{def: _empiricalmeasure_tips} $ | (7) |
Two kinds of branching have been identified; either from a tip or from a mature vessel; here for the sake of simplicity, we shall consider only tip branching.
The birth process of new tips can be described in terms of a marked point process (see e.g. [5]), by means of the random measure
$ \Phi((0,t]\times B): =\int_0^t\!\int_B\Phi(ds\times d{\mathbf{x}}\times d{\mathbf{v}}), \label{PHI} $ | (8) |
where
We assume that the probability that a tip branches from one of the
$ prob\!\left(N(t+dt)-N(t)=1 |\mathcal F_{t^-}\right) =\sum\limits_{i=1}^{N(t)}\alpha(C(t,\mathbf{X}^i(t)))\, dt, \label{prob_nuova_nascita_tip} $ | (9) |
where
$ \alpha(C)=\alpha_1\frac{C}{C_R+C}, $ | (10) |
where
Here
As a technical simplification, we will further assume that whenever a tip located in
We then claim that the compensator of the random measure
$\alpha(C(s,{\mathbf{x}}))\delta_{{\mathbf{v}}_0}(\mathbf{v})\, N Q_N(s)(d(\mathbf{{x}},\mathbf{{v}}))\, ds. \label{branching_compensator} $ | (11) |
We describe vessel extension by the Langevin system
$
dXk(t)=vk(t)dtdvk(t)=[−kvk(t)+F(C(t,Xk(t)))]dt+σdWk(t),
$
|
(12) |
for all those
$ \mathbf{F}(C)= \frac{d_1}{(1+\gamma_1C)^q}\nabla_x C. \label{Fvector} $ | (13) |
When a vessel tip meets an existing vessel it joins it at that point and time and it stops moving. This process is called tip-vessel anastomosis.
As in the case of the branching process, we have modelled this process via a marked counting process; anastomosis is modelled as a "death" process.
Let
$ \Psi((0,t] \times B) := \int_0^t \int_B \Psi(ds\times d{\mathbf{x}} \times d{\mathbf{v}}) \label{psiB} $ | (14) |
where
Thanks to the choice of a Langevin model for the vessels extension, we may assume that the trajectories are sufficiently regular and have an integer Hausdorff dimension
Hence [12] the random measure
$ A \in \mathcal B_{\mathbb R^d} \mapsto \mathcal H^1 ({\mathbf{X}}(t)\cap A) \in \mathbb R_+ $ | (15) |
admits a random generalized density
$ \mathcal H^1 ({\mathbf{X}}(t)\cap A)= \int_A \delta_{\mathbf{X}(t)}(\mathbf{x})d \mathbf{x} . $ | (16) |
Given the definition of
$
H1(X(t)∩A)=γ∫t0dsN(s)∑i=1IA(Xi(s))=γ∫t0dsN(s)∑i=1ϵXi(s)(A),
$
|
(17) |
where
The latter has the Dirac delta as its generalized derivative with respect to the usual Lebesgue measure on
$ \delta_{\mathbf{X}(t)}(\mathbf{x})= \gamma \displaystyle \int_0^t ds\sum\limits_{i=1}^{N(s)} \,\delta_{\mathbf{X}^i(s)}(\mathbf{x}). $ | (18) |
We assume that the probability per unit time that a typical tip
$ prob\!\left(N(t+dt)-N(t)=-1 |\mathcal F_{t^-}\right) = N^{-1} \sum\limits_{i=1}^{N(t)} ( K_N \ast \delta_{\mathbf{X}(t)} )(\mathbf{X}^i(t)) dt. \label{prob_morte_tip} $ | (19) |
We then claim that the compensator of the random measure
$ N^{-1} ( K_N \ast \delta_{\mathbf{X}(s)} )(\mathbf{x})N Q_N(s)(d({\mathbf{x}},{\mathbf{v}}))ds. $ | (20) |
Our stochastic model is thus described by a set of Ito stochastic differential equations (SDEs), a marked point process describing tip branching (a birth process), and a marked point process describing anastomosis (a death process). The latter process depends on the past history of a given realization of the overall stochastic process. In addition, the TAF concentration is itself a random process since it depends on the stochastic evolution of the tips as indicated by Equation (3). Hence
By Itô's formula (see e.g. [9], p. 270), we may obtain the evolution equation of the random measure
$
∫g(x,v)QN(t)(d(x,v))=∫g(x,v)QN(0)(d(x,v))+∫t0∫v⋅∇xg(x,v)QN(s)(d(x,v))ds+∫t0∫[F(C(s,x))−kv]⋅∇vg(x,v)QN(s)(d(x,v))ds+∫t0∫σ22Δvg(x,v)QN(s)(d(x,v))ds+∫t0∫α(C(s,x))δv0(v)g(x,v)QN(s)(d(x,v))ds−∫t0∫1N(KN∗δX(s))(x)g(x,v)QN(s)(d(x,v))ds+˜MN(t),
$
|
(21) |
where
$\tilde{M}_N(t)= \tilde{M}_{1,N}(t) + \tilde{M}_{2,N}(t) + \tilde{M}_{3,N}(t),$ |
with
$ \tilde{M}_{1,N}(t)= \!\int_0^t\!\frac{1}{N} \sum\limits_{k=1}^{N(s)}\!\nabla_v g(\mathbf{X}^k(s),\mathbf{v}^k(s))\cdot d\mathbf{W}^k(s),\label{mtilde} $ | (22) |
$
˜M2,N(t)=∫t0∫g(x,v)[ΦN(ds×dx×dv)−α(C(s,x))δv0(v)QN(s)(d(x,v))ds],
$
|
(23) |
$
˜M3,N(t)=∫t0∫g(x,v)[ΨN(ds×dx×dv)−1N(KN∗δX(s))(x)QN(s)(d(x,v))ds].
$
|
(24) |
All
By Doob's inequality
$\tilde M_N(t) \mathop{\longrightarrow}\limits ^{P}_{N \rightarrow \infty} 0.$ |
This is the substantial reason of the possible deterministic limiting behavior of the process, as
As a consequence, if we suppose that indeed, for a large value of the scale parameter
$Q_N(t)(d(\mathbf{x},\mathbf{v}))\rightarrow Q_{\infty}(t)(d(\mathbf{x},\mathbf{v}))=p(t,\mathbf{x},\mathbf{v})d\mathbf{x} \, d\mathbf{v}. \label{LLN_Q} $ | (25) |
Assuming that (25) holds true, the scaled stochastic distribution of vessels
$ \lambda(t,{\mathbf{x}})= \mathbb{E}[\delta_{\mathbf{X}(t)}] (\mathbf{x})= \gamma \displaystyle \int_0^t \tilde{p}(s,{\mathbf{x}})\, ds, $ | (26) |
where
$ \tilde{p}(t,\mathbf{x}) = \mathbb{E}[\delta_{\mathbf{X^i}(t)}](\mathbf{x}) =\int p(t,\mathbf{x},\mathbf{v}')\, d\mathbf{v}'\label{reduced_density} $ | (27) |
is the marginal density of
A formal justification of the above is the following one. According to a previous discussion
$ \frac{1}{N}\delta_{\mathbf{X}(t)}(\mathbf{x})= \gamma \int_0^t ds \displaystyle{\frac{1}{N} \sum\limits_{i=1}^{N(s)} \delta_{\mathbf{X}^i(s)}(\mathbf{x})} .$ |
In the mean field approximation we are assuming that a "law of large numbers" may be applied, so that
For objects of Hausdorff dimension
Finally the consumption term in (3) is an approximation of the flux density
$ \mathbf{j}(t,\mathbf{x})= \int \mathbf{v}' p(t,\mathbf{x},\mathbf{v}')\, d \mathbf{v'}, \label{flux_max} $ | (28) |
so that the evolution equation (3) may be approximated by its (deterministic) mean field version
$ \frac{\partial}{\partial t}\widetilde{C}(t,\mathbf{x}) = d_2 \Delta_x \widetilde{C}(t,\mathbf{x}) - \eta \widetilde{C}(t,\mathbf{x})|\mathbf{j}(t,\mathbf{x})| . $ | (29) |
TAF injection from the tumor is realized as a nonzero flux boundary condition for this equation [4], instead of including it explicitly as in (3).
Based on the above discussion, one might claim that the evolution equation for the density
$
∂∂tp(t,x,v)=α1˜C(t,x)CR+˜C(t,x)p(t,x,v)δv0(v)−γp(t,x,v)∫t0˜p(s,x)ds−v⋅∇xp(t,x,v)+k∇v⋅(vp(t,x,v))−d1∇v⋅[∇x˜C(t,x)[1+γ1˜C(t,x)]qp(t,x,v)]+σ22Δvp(t,x,v).
$
|
(30) |
The coupled system (29), (30) is subject to suitable boundary and initial conditions, to be discussed later.
This approach is called "hybrid", since we have substituted all stochastic underlying fields by their "mean" counterparts; most of the current literature could now be reinterpreted along these lines. Indeed, one should check that the hybrid system is fully compatible with a rigorous derivation of the evolution for the vessel densities. Nonlinearities in the full model are a big difficulty in this direction; a rigorous derivation of the convergence results requires a nontrivial mathematical analysis, which is under investigation; for this it is instrumental a proof of existence, uniqueness and sufficient regularity of the solution of the mean field equations (see [29], and [6]). We wish to stress that anyhow substituting mean geometric densities of tips or of full vessels to the corresponding stochastic quantities leads to an acceptable coefficient of variation (percentage error) only when a law of large numbers can be applied, i.e. whenever the relevant numbers per unit volume are sufficiently large; otherwise stochasticity cannot be avoided, and in addition to mean values, the mathematical analysis and/or simulations should provide confidence bands for all quantities of interest (see e. g. [7]). An interesting case in this direction is discussed in [11].
Figure 1 shows the outcome of a numerical simulation of the fully stochastic model. The values of all parameters are discussed in [4] and [41]; in particular all kernels have been taken as Gaussians centered at
However, there is another interpretation of the numerical simulations for which a deterministic description is correct. While the vessel network may look different for different replicas of the stochastic process, tip densities associated to averages over replicas are described by Equation (30) [41]. How is this possible? Let
$ Q_N^*(t,\mathbf{x},\mathbf{v},\omega) \,\,\!\!\!=\!\!\!\sum\limits_{i=1}^{N(t,\omega)}\delta_{\mathbf{X}^i(t,\omega)}(\mathbf{x})\,\delta_{\mathbf{v}^i(t,\omega)}(\mathbf{v}). \label{eq99_1} $ | (31) |
Here the number of tips at time
$ \tilde{Q}_N^*(t,\mathbf{x}, \omega)= \sum\limits_{i=1}^{N(t,\omega)}\delta_{\mathbf{X}^i(t,\omega)}(\mathbf{x}), \label{eq99_2} $ | (32) |
and
$ \delta_{\mathbf{X}(t,\omega)}(\mathbf{x}):= \gamma \int_0^t\sum\limits_{i=1}^{N(s,\omega)}\delta_{\mathbf{X}^i(s,\omega)}(\mathbf{x})ds $ | (33) |
represents the concentration of all vessels per unit volume in physical space, at time
By following a similar approach as in our previous paper [4], we may then obtain the weak formulation of the stochastic evolution of
$
∫g(x,v)Q∗N(t,x,v)dxdv=∫g(x,v)Q∗N(0,x,v)dxdv+∫t0∫v⋅∇xg(x,v)Q∗N(s,x,v)dxdvds+∫t0∫[F(C(s,x))−kv]⋅∇vg(x,v)Q∗N(s,x,v)dxdvds+∫t0∫σ22Δvg(x,v)Q∗N(s,x,v)dxdvds+∫t0∫α(C(s,x))δv0(v)g(x,v)Q∗N(s,x,v)dxdvds−∫t0∫δX(s)(x)g(x,v)Q∗N(s,x,v)dxdvds+¯MN(t),
$
|
(34) |
where we have omitted the variable
By mimicking the typical kernel density estimation approach in Statistics (see e.g. [33], page 489), we introduce the (random) empirical distribution of tips per unit volume in the
$
pN(t,x,v)=1NN∑ω=1(KN∗Q∗N(t,⋅,⋅,ω))(x,v).
$
|
(35) |
Here the mollifying kernel
$ \tilde{p}_{\mathcal N}(t,\mathbf{x})\,\,\!\!=\!\frac{1}{\mathcal{N}}\sum\limits_{\omega=1}^\mathcal{N}(K_{\mathcal N} \ast \widetilde{Q}_N^*(t, \cdot, \omega))(\mathbf{x}), \label{eq11} $ | (36) |
$ \mathbf{j}_{\mathcal N}(t,\mathbf{x})\,\,\!\!=\!\frac{1}{\mathcal{N}}\!\sum\limits_{\omega=1}^\mathcal{N}\!\sum\limits_{i=1}^{N(t,\omega)}\!\!\mathbf{v}^i(t,\omega) (K_{\mathcal N} \ast \delta_{\mathbf{X}^i(t,\omega)})(\mathbf{x}),\label{eq12} $ | (37) |
respectively. We now conjecture that the following limit exists
$
p(t,x,v)=limN→∞pN(t,x,v),
$
|
(38) |
and that
$
˜p(t,x)=limN→∞˜pN(t,x).
$
|
(39) |
Finally, we may obtain the deterministic version of the vessel tip flux as
$
j(t,x)=limN→∞jN(t,x).
$
|
(40) |
Figure 2 shows the marginal tip density
On the basis of the above convergence assumptions and considering
$ \lim\limits_{\mathcal{N}\to\infty} \frac{1}{\mathcal{N}}\!\sum\limits_{\omega=1}^\mathcal{N} \overline{M}_N(t, \omega)=0,$ |
we may expect that the ensemble average of the stochastic equation (34) tends in its strong form to the same equation of Fokker-Planck type as (30):
$
∂∂tp(t,x,v)=α1C(t,x)CR+C(t,x)p(t,x,v)δv0(v)−γp(t,x,v)∫t0˜p(s,x)ds−v⋅∇xp(t,x,v)+k∇v⋅[vp(t,x,v)]−d1∇v⋅[∇xC(t,x)[1+γ1C(t,x)]qp(t,x,v)]+σ22Δvp(t,x,v).
$
|
(41) |
Here the marginal vessel tip density,
$ \tilde{p}(t,\mathbf{x}) = \int p(t,\mathbf{x},\mathbf{v}')\, d\mathbf{v}',\label{eq19} $ | (42) |
is the marginal density of
$ \frac{\partial}{\partial t}C(t,\mathbf{x})=d_2 \Delta_x C(t,\mathbf{x})- \eta\, C(t,\mathbf{x})\!\left| \mathbf{j}(t,\mathbf{x})\right|\!,\label{eq20} $ | (43) |
where
$ \mathbf{j}(t,\mathbf{x})= \int \mathbf{v}' p(t,\mathbf{x},\mathbf{v}')\, d \mathbf{v'}. \label{eq21} $ | (44) |
A crucial point in the derivation of (41) as the ensemble average of Equation (34) is the approximation of nonlinear terms in the latter, e.g. the anastomosis term. Had the law of large numbers been applicable (on the single replica, according to the propagation of chaos assumption), the mean values of nonlinear terms could be factorized. With the ensemble average description, factorization of nonlinear terms requires further investigation, and this is out of the scope of the present paper. By assuming factorization occurs, we may expect that the ensemble average of Equation (34) tends in its strong form to (41).
By an abuse of notations, we are using in Equation (43) the same letter
We need to stress here that the
For appropriate initial and boundary data, it is possible to prove that (41) and (43) have a unique smooth solution [16].
We have solved the system of equations (41) and (43) in a two dimensional strip geometry using the initial and boundary conditions introduced in [4]. The strip is
$
∂∂xC(t,0,y)=0,∂∂xC(t,L,y)=c1(y)d2,
$
|
(45) |
and
$ C(0,x,y)=1.1\, C_R e^{-[(x-L)^2/c^2+y^2/b^2]},\label{eq24} $ | (46) |
for appropriate
As boundary conditions for the tip density we have taken
$
p+(t,0,y,v,w)=e−k|v−v0|2σ2∫∞0∫∞−∞v′e−k|v′−v0|2σ2dv′dw′×[j0(t,y)−∫0−∞∫∞−∞v′p−(t,0,y,v′,w′)dv′dw′],
$
|
(47) |
$
p−(t,L,y,v,w)=e−k|v−v0|2σ2∫0−∞∫∞−∞e−k|v′−v0|2σ2dv′dw′×[˜p(t,L,y)−∫∞0∫∞−∞p+(t,L,y,v′,w′)dv′dw′],
$
|
(48) |
$ p(t,\mathbf{x},\mathbf{v})\to 0 \mbox{ as } |\mathbf{v}|\to \infty, $ | (49) |
where
$ j_0(t,y)=\frac{v_0\, L}{\sqrt{v_0^2+w_0^2}}\alpha(C(t,0,y))\, p(t,0,y,v_0,w_0), \label{eq28} $ | (50) |
for the vector velocity
Finally as initial condition for the tip density we have taken
$
p(0,x,y,v,w)=2e−x2/l2xπ3/2lxσ2ve−|v−v0|2/σ2vN0∑i=11√πlye−|y−yi|2/l2y.
$
|
(51) |
As
$ p(0,x,y,\mathbf{v})\!=\! 2\delta^{\sigma_v}_{\mathbf{v}_0}(\mathbf{v})\delta_{0}(x)\sum\limits_{i=1}^{N_0}\delta_{y_i}(y),\label{eq30} $ | (52) |
where
We have used the parameter values that have been extracted from experiments as explained in [4]. The anastomosis coefficient
$
∂p∂t=AC1+Cpδv0(v)−Γp∫t0˜p(s,x)ds−v⋅∇xp−∇v⋅[(δ∇xC(1+Γ1C)q−βv)p]+β2Δvp,
$
|
(53) |
$
∂C∂t=κΔxC−χC|j|.
$
|
(54) |
The dimensionless parameters appearing in these equations are defined in Table 1 (see [4], [41]). In the computations for the generalized function
$ \delta_{v}(\mathbf{v})= \frac{1}{\pi\epsilon^2}e^{-|\mathbf{v}|^2/\epsilon^2}. \label{eq33} $ | (55) |
|
|
|
||||
|
|
|
|
|
||
1.5 | 5.88 | 22.42 | 0.145 | 1 | 0.0045 | 0.002 |
We have used
The nondimensional boundary conditions for
$ \frac{\partial C}{\partial x}(t,0,y)=0, \frac{\partial C}{\partial x}(t,1,y)=f(y), \, \lim\limits_{y\to\pm\infty}C= 0, \label{eq34} $ | (56) |
where
$ C(0,x,y)=1.1\, e^{-[(x-1)^2L^2/c^2+y^2L^2/b^2]},\label{eq35} $ | (57) |
with
$
p(0,x,y,v,w)=2Le−x2L2/l2xπ3/2lxϵ2e−|v−v0|2/σ2vN0∑i=1L√πlye−|y−yi|2L2/l2y,
$
|
(58) |
with
$
p+(t,0,y,v,w)=e−|v−v0|2∫∞0∫∞−∞v′e−|v′−v0|2dv′dw′×[j0(t,y)−∫0−∞∫∞−∞v′p−(t,0,y,v′,w′)dv′dw′]
$
|
(59) |
for
$
p−(t,1,y,v,w)=e−|v−v0|2∫0−∞∫∞−∞e−|v′−v0|2dv′dw′×[˜p(t,1,y)−∫∞0∫∞−∞p+(t,1,y,v′,w′)dv′dw′]
$
|
(60) |
for
$ j_0(t,y)=A\, v_0\frac{C}{1+C}\, p(t,0,y,v_0,w_0) \label{eq39} $ | (61) |
(
In [4], we have shown the consistency of the deterministic model by depicting TAF concentration, marginal tip density and overall network density at different times. Figure 3 shows that the marginal tip density obtained by numerically solving the deterministic equations (53)-(61) agrees quite well with the ensemble average over 50 replicas of the stochastic process depicted in Figure 2.
Figure 4 compares the deterministic and averaged stochastic descriptions of the marginal tip density at the
In our recent papers [4] and [41] we have explored the behavior of a stochastic angiogenesis model, and of its possible deterministic approximation. In this model, the tips undergo a stochastic process of tip branching, vessel extension and anastomosis whereas TAF is described by a reaction-diffusion equation with a sink term proportional to the tip flow.
In [4] the empirical measure describing the tip distribution had been assumed to satisfy a "law of large numbers" for any single replica of the process, i. e. the classical "propagation of chaos" assumption (see e.g. [40], and references therein), so that it admits a position-velocity density which is shown to satisfy a nonlinear integro-differential equation of a Fokker-Planck type, coupled with a reaction-diffusion equation for the TAF concentration, in which the stochastic tip flow has been replaced by its mean field approximation, deriving from the tip mean density.
On the other hand, in [41] we have solved numerically the stochastic model for many realizations (independent replicas of the system). Numerically calculated velocity fluctuations have revealed that they do not decay even as the number of initial vessel tips increases. This shows that the stochastic model is not self-averaging and therefore we cannot use the "propagation of chaos" assumption to derive a mean field deterministic approximation of the stochastic model. The main reason being that anastomosis eliminates many vessel tips, resulting in the fact that there never are enough tips for a law of large numbers apply on a single replica. The vessel network has shown to be quite different for different replicas of the stochastic process.
However by re-examining the derivation given in [4], we conclude that the same deterministic description holds for vessel tip densities calculated by averaging over replicas. The deterministic description consists of a reaction-diffusion equation for the TAF concentration coupled to a Fokker-Planck type equation for the vessel tip density. The latter contains a birth term corresponding to tip branching and a death integral term corresponding to anastomosis or tip merging. The coefficient of the latter term has been fitted by comparison with the stochastic description: optimal selection produces a good fit for the evolution of the total number of tips.
For the averaged deterministic reaction-diffusion system, boundary conditions have been proposed in [41], which describe the flux of vessel tips injected from a primary blood vessel in response to TAF emitted by the tumor and the tip density eventually arriving at the tumor. Numerical solution of the model in a simple geometry shows how tips are created at the primary blood vessel, propagate and proliferate towards the tumor and may or not reach it after a certain time depending on the parameter values. This is consistent with known biological facts.
Actually nontrivial additional investigation is required for a rigorous derivation of the deterministic approximation of the relevant empirical measures and their evolution equations. Anyhow we wish to convey a general message elicited by the proposed angiogenesis model: in stochastic models containing birth and death processes in addition to Brownian motion (Langevin equations), the death processes may preclude reaching the large number of individuals required to have self-averaging and a deterministic description based on the "propagation of chaos". Nevertheless, deterministic equations for macroscopic densities and fluxes may follow from using ensemble averages over a large number of replicas.
The significant consequence concerns the variance; though the mean behavior can be described by the same PDE, the case of self-averaging does not carry any variance; but the variance cannot be ignored in the case of ensemble averaging, which implies the use of confidence bands in predicting the evolution of a real vessel network! The authors are well aware of the limits of their own analysis, but they wish to stimulate more attention to the mathematical issues raised by this important biomedical problem. Only accurate models can support medical intervention for prevention and cure. Simulations are surely useful as a first insight, but therapies (optimal control) require accurate mathematical models, validated by comparison with real data (inverse problems -statistics of random geometric structures).
Apart from the specific application we have been dealing with, in this paper methodological contributions have been given for a sound mathematical modelling of stochastic vessel networks: a) the use of stochastic distributions, and their mean densities, describing the vessels -random objects of Hausdorff dimension
The authors acknowledge the contribution of the anonymous Referees for a significant improvement of the manuscript. This work has been supported by the Spanish Ministerio de Economí a y Competitividad grant MTM2014-56948-C2-2-P. VC has been supported by a Chair of Excellence UC3M-Santander at the Universidad Carlos Ⅲ de Madrid. It is a great pleasure to acknowledge fruitful discussions with Daniela Morale of the Department of Mathematics of the University of Milan.
[1] |
Amado M, Poggi F (2014) Solar Urban Planning: A Parametric Approach. Energy Procedia 48: 1539-1548. doi: 10.1016/j.egypro.2014.02.174
![]() |
[2] | Directive 2009/28/EC of the European Parliament and of the Council of 23 April 2009 on the promotion of the use of energy from renewable sources and amending and subsequently repealing Directives 2001/77/EC and 2003/30/EC. Official Journal of the European Union L140. pp. 16-62. |
[3] |
Bergamasco L, Asinari P (2011) Scalable methodology for the photovoltaic solar energy potential assessment based on available roof surface area: Application to Piedmont Region (Italy). Sol Energy 85: 1041-1055. doi: 10.1016/j.solener.2011.02.022
![]() |
[4] |
Nex F, Remondino F, Agugiaro G, et al. (2013) 3D Solarweb: A solar cadaster in the Italian Alpine landscape. Int Arch Photogramm Remote Sens Spatial Inform Sci XL-7/W2: 173-178. doi: 10.5194/isprsarchives-XL-7-W2-173-2013
![]() |
[5] | Kassner R, Koppe W, Schüttenberg T, et al. (2008) Analysis of the solar potential of roofs by using official LIDAR data. Int Arch Photogramm Remote Sens Spatial Inform Sci XXXVII-B4: 399-403. |
[6] | United Nations, Department of Economic and Social Affairs, Population Division (2014) World Urbanization Prospects: The 2014 Revision, Highlights. 32 p. |
[7] | Carneiro C, Morello E, Desthieux G (2009) Assessment of solar irradiance on the urban fabric for the production of renewable energy using LIDAR data and image processing techniques. In: Springer Berlin Heidelberg, editor. Adv GIScience: 83-112. |
[8] |
Jacques DA, Gooding J, Giesekam JJ, et al. (2014) Methodology for the assessment of PV capacity over a city region using low-resolution LIDAR data and application to the City of Leeds (UK). Appl Energy 124: 28-34. doi: 10.1016/j.apenergy.2014.02.076
![]() |
[9] | Directive 2010/31/EU of the European Parliament and of the Council of 19 May 2010 on the energy performance of buildings. Official Journal of the European Union L153. 13-35. |
[10] |
Amado M, Poggi F (2012) Towards solar urban planning: A new step for better energy performance. Energy Procedia 30: 1261-1273. doi: 10.1016/j.egypro.2012.11.139
![]() |
[11] | Charabi Y, Rhouma MBH, Gastli A (2010) GIS-based estimation of roof-PV capacity & energy production for the Seeb Region in Oman. IEEE Int Energ Conf: 41-44. |
[12] |
Nguyen HT, Pearce JM, Harrap R, et al. (2012) The application of LIDAR to assessment of rooftop solar photovoltaic deployment potential in a municipal district unit. Sensors 12: 4534-4558. doi: 10.3390/s120404534
![]() |
[13] |
Kucuksari S, Khaleghi AM, Hamidi M, et al. (2014) An Integrated GIS, optimization and simulation framework for optimal PV size and location in campus area environments. Appl Energy 113: 1601-1613. doi: 10.1016/j.apenergy.2013.09.002
![]() |
[14] | La Gennusa M, Lascari G, Rizzo G, et al. (2011) A model for predicting the potential diffusion of solar energy systems in complex urban environments. Energy Policy 39: 5335-5343. |
[15] |
Izquierdo S, Rodrigues M, Fueyo N (2008) A method for estimating the geographical distribution of the available roof surface area for large-scale photovoltaic energy-potential evaluations. Sol Energy 82: 929-939. doi: 10.1016/j.solener.2008.03.007
![]() |
[16] |
Africani P, Bitelli G, Lambertini A, et al. (2013) Integration of LIDAR data into a municipal GIS to study solar radiation. Int Arch Photogramm Remote Sens Spatial Inform Sci XL-1/W1: 1-6. doi: 10.5194/isprsarchives-XL-1-W1-1-2013
![]() |
[17] | Byrne J, Taminiau J, Kurdgelashvili L, et al. (2014) A review of the solar city concept and methods to assess rooftop solar electric potential, with an illustrative application to the city of Seoul. Renew Sust Energ Rev 41: 830-844. |
[18] |
Jo JH, Otanicar TP (2011) A hierarchical methodology for the mesoscale assessment of building integrated roof solar energy systems. Renew Energ 36: 2992-3000. doi: 10.1016/j.renene.2011.03.038
![]() |
[19] |
Miranda RFC, Szklo A, Schaeffer R (2015) Technical-economic potential of PV systems on Brazilian rooftops. Renew Energ 75: 694-713. doi: 10.1016/j.renene.2014.10.037
![]() |
[20] |
Sun Y, Hof A, Wang R, et al. (2013) GIS-based approach for potential analysis of solar PV generation at the regional scale: A case study of Fujian Province. Energy Policy 58: 248-259. doi: 10.1016/j.enpol.2013.03.002
![]() |
[21] |
Agugiaro G, Nex F, Remondino F, et al. (2012) Solar radiation estimation on building roofs and Web-Based solar cadastre. Int Arch Photogramm Remote Sens Spatial Inform Sci I-2: 177-182. doi: 10.5194/isprsannals-I-2-177-2012
![]() |
[22] |
Calvert K, Pearce JM, Mabee WE (2013) Toward renewable energy geo-information infrastructures: Applications of GIScience and remote sensing that build institutional capacity. Renew Sust Energ Rev 18: 416-429. doi: 10.1016/j.rser.2012.10.024
![]() |
[23] |
Hofierka J, Kaňuk J (2009) Assessment of photovoltaic potential in urban areas using open-source solar radiation tools. Renew Energ 34: 2206-2214. doi: 10.1016/j.renene.2009.02.021
![]() |
[24] | Jochem A, Höfle B, Hollaus M, et al. (2009) Object detection in airborne LIDAR data for improved solar radiation modeling in urban areas. Int Arch Photogramm Remote Sens Spatial Inform Sci XXXVIII-3/W8: 1-6. |
[25] |
Mainzer K, Fath K, McKenna R, et al. (2014) A high-resolution determination of the technical potential for residential-roof-mounted photovoltaic systems in Germany. Sol Energy 105: 715-731. doi: 10.1016/j.solener.2014.04.015
![]() |
[26] |
Kodysh JB, Omitaomu OA, Bhaduri BL, et al. (2013) Methodology for estimating solar potential on multiple building rooftops for photovoltaic systems. Sust Cities Soc 8: 31-41. doi: 10.1016/j.scs.2013.01.002
![]() |
[27] |
Gooding J, Edwards H, Giesekam J, et al. (2013) Solar City Indicator: A methodology to predict city level PV installed capacity by combining physical capacity and socio-economic factors. Sol Energy 95: 325-335. doi: 10.1016/j.solener.2013.06.027
![]() |
[28] |
Araya-Muñoz D, Carvajal D, Sáez-Carreño A, et al. (2014) Assessing the solar potential of roofs in Valparaíso (Chile). Energ Buildings 69: 62-73. doi: 10.1016/j.enbuild.2013.10.014
![]() |
[29] |
Bergamasco L, Asinari P (2011) Scalable methodology for the photovoltaic solar energy potential assessment based on available roof surface area: Further improvements by ortho-image analysis and application to Turin (Italy). Sol Energy 85: 2741-2756. doi: 10.1016/j.solener.2011.08.010
![]() |
[30] |
Jakubiec JA, Reinhart CF (2013) A method for predicting city-wide electricity gains from photovoltaic panels based on LIDAR and GIS data combined with hourly Daysim simulations. Sol Energy 93: 127-143. doi: 10.1016/j.solener.2013.03.022
![]() |
[31] |
Lukač N, Žlaus D, Seme S, et al. (2013) Rating of roofs' surfaces regarding their solar potential and suitability for PV systems, based on LIDAR data. Appl Energy 102: 803-812. doi: 10.1016/j.apenergy.2012.08.042
![]() |
[32] |
Lukač N, Seme S, Žlausa D, et al. (2014) Buildings roofs photovoltaic potential assessment based on LIDAR (Light Detection and Ranging) data. Energy 66: 598-609. doi: 10.1016/j.energy.2013.12.066
![]() |
[33] | Redweik PM, Catita C, Brito MC (2011) 3D Local scale radiation model based on urban LIDAR data. Int Arch Photogramm Remote Sens Spatial Inform Sci XXXVIII-4/W19: 265-269. |
[34] | European Commission, Photovoltaic Geographical Information System (PVGIS). Available from: http://re.jrc.ec.europa.eu/pvgis. [Accessed on 4 February 2015]. |
[35] | Borfecchia F, Caiaffa E, Pollino M, et al. (2014) Remote sensing and GIS in planning photovoltaic potential of urban areas. Eur J Remote Sens 47: 195-216. |
[36] | SoDa, Solar Energy Service for Professionals. Available from: http://www.soda-pro.com/es. [Accessed on 4 February 2015]. |
[37] |
Wiginton LK, Nguyen HT, Pearce JM (2010) Quantifying rooftop solar photovoltaic potential for regional renewable energy policy. Comput Environ Urban Syst 34: 345-357. doi: 10.1016/j.compenvurbsys.2010.01.001
![]() |
[38] | Gobierno de España, Ministerio de Fomento. Instituto Geográfico Nacional. Available from: http://www.ign.es/ign/main/index.do. [Accessed on 4 February 2015]. |
[39] | Carneiro C, Morello E, Ratti C, et al. (2009) Solar radiation over the urban texture: LIDAR data and image processing techniques for environmental analysis at city scale. In: Heidelberg SB, editor. 3D Geo-Information Sci: 319-340. |
[40] | Agugiaro G, Remondino F, Stevanato G, et al. (2011) Estimation of solar radiation on building roofs in mountainous areas. Int Arch Photogramm Remote Sens Spatial Inform Sci XXXVIII-3/W22: 155-160. |
[41] |
Xu Z, Coors V (2012) Combining system dynamics model, GIS and 3D visualization in sustainability assessment of urban residential development. Build Environ 47: 272-287. doi: 10.1016/j.buildenv.2011.07.012
![]() |
[42] |
Dorninger P, Pfeifer N (2008) A comprehensive automated 3D approach for building extraction, reconstruction, and regularization from airborne laser scanning point clouds. Sensors 8: 7323-7343. doi: 10.3390/s8117323
![]() |
[43] |
Haala N, Kada M (2010) An update on automatic 3D building reconstruction. ISPRS J Photogramm 65: 570-580. doi: 10.1016/j.isprsjprs.2010.09.006
![]() |
[44] | Xiao Y, Wang C, Xi XH et al. (2014) A comprehensive framework of building model reconstruction from airborne LIDAR data. IOP Conference Series: Earth and Environmental Science 17: 1-6. |
[45] |
Jochem A, Höfle B, Rutzinger M, et al. (2009) Automatic roof plane detection and analysis in airborne LIDAR point clouds for solar potential assessment. Sensors 9: 5241-5262. doi: 10.3390/s90705241
![]() |
[46] | Fan H, Yao W, Fu Q (2014) Segmentation of sloped roofs from airborne LIDAR point clouds using ridge-based hierarchical decomposition. Sensors 6: 3284-3301. |
[47] |
Awrangjeb M, Fraser CS (2014) Automatic segmentation of raw LIDAR data for extraction of building roofs. Remote Sens 6: 3716-3751. doi: 10.3390/rs6053716
![]() |
[48] | Ludwig D, Lanig S, Klärle M (2009) Location analysis for solar panels by LIDAR-Data with Geoprocessing-SUN-AREA. Environmental Informatics and Industrial Environmental Protection: Concepts, Methods and Tools; Shaker Verlag: 83-89. |
[49] | Voegtle T, Steinle E, Tovari D (2005) Airborne laser scanning data for determination of suitable areas for photovoltaic. Int Arch Photogramm Remote Sens Spatial Inform Sci XXXVI-3/W19: 215-220. |
[50] |
Sisaki J (2013) Knowledge-based modeling of buildings in dense urban areas by combining airborne LIDAR data and aerial images. Remote Sens 5: 5944-5968. doi: 10.3390/rs5115944
![]() |
[51] |
Wuming Zhang W, Wang H, Chen Y et al. (2014) 3D Building roof modeling by optimizing primitive’s parameters using constraints from LIDAR data and aerial imagery. Remote Sens 6: 8107-8133. doi: 10.3390/rs6098107
![]() |
[52] |
Tooke TR, Coops NC, Voogt JA, et al. (2011) Tree structure influences on rooftop-received solar radiation. Landsc Urban Plan 102: 73-81. doi: 10.1016/j.landurbplan.2011.03.011
![]() |
[53] |
Alam N, Coors V, Zlatanova S, et al. (2012) Shadow effect on photovoltaic potentiality analysis using 3D city models. Int Arch Photogramm Remote Sens Spatial Inform Sci XXXIX-B8: 209-214. doi: 10.5194/isprsarchives-XXXIX-B8-209-2012
![]() |
[54] | Minghetti A, Africani P, Paselli E (2011) Bologna Solar City, a web application for the analysis or potential energy: From estimating solar radiation to the realization of the application. 17th Building Services, Mechanical and Building Industry Days Urban Energy Conference Debrecen, Hungary: 25-30. |
[55] | Hofierka J, Šúri M. The solar radiation model for Open source GIS: implementation and applications; September 2002; Trento, Italy: 1-19. |
[56] |
Freitas S, Catita C, Redweik P, et al. (2015) Modelling solar potential in the urban environment: State-of-the-art review. Renew Sust Energy Rev 41: 915-931. doi: 10.1016/j.rser.2014.08.060
![]() |
[57] |
Šúri M, Huld TA, Dunlop ED (2005) PV-GIS: a web-based solar radiation database for the calculation of PV potential in Europe. InterJ Sust Energ 24: 55-67. doi: 10.1080/14786450512331329556
![]() |
[58] | An overview of the Solar Radiation tools, ArcGIS Help Library, ESRI. Available from: http://resources.arcgis.com/en/help/main/10.2/index.html#//009z000000t4000000. [Accessed on 10 February 2015]. |
[59] | Martín AM (2014) Modelo geográfico para la estimación del potencial fotovoltaico en tejados. Caso de estudio: Miraflores de la Sierra [MsC]. Madrid: Universidad Complutense. 82 p. Available from: http://eprints.ucm.es/25543/. [Accessed on 10 February 2015]. |
[60] |
Brito MC, Gomes N, Santos T, et al. (2012) Photovoltaic potential in a Lisbon suburb using LIDAR data. Sol Energy 86: 283-288. doi: 10.1016/j.solener.2011.09.031
![]() |
[61] |
Robinson D (2006) Urban morphology and indicators of radiation availability. Sol Energy 80: 1643-1648. doi: 10.1016/j.solener.2006.01.007
![]() |
[62] | Perez R, Seals R, Michalsky J (1993) All-weather model for sky luminance distribution - Preliminary configuration and validation. Sol Energy 50 (3): 235-245. |
[63] | DAYSIM, Advanced Daylight Simulation Software. Available from: http://daysim.ning.com. [Accessed on 10 February 2015]. |
[64] |
Melo EG, Almeida MP, Zilles R, et al. (2013) Using a shading matrix to estimate the shading factor and the irradiation in a three-dimensional model of a receiving surface in an urban environment. Sol Energy 92: 15-25. doi: 10.1016/j.solener.2013.02.015
![]() |
[65] |
Redweik P, Catita C, Brito M (2013) Solar energy potential on roofs and facades in an urban landscape. Sol Energy 97: 332-341. doi: 10.1016/j.solener.2013.08.036
![]() |
[66] | Erdélyi R, Wang Y, Guo W, et al. (2014) Three-dimensional SOlar RAdiation Model (SORAM) and its application to 3-D urban planning. Sol Energy 101: 63-73. |
[67] |
Verso A, Martín A, Amador J et al. (2015) GIS-based method to evaluate the photovoltaic potential in the urban environments: The particular case of Miraflores de la Sierra. Sol Energy 117: 236-245. doi: 10.1016/j.solener.2015.04.018
![]() |
[68] |
Tooke TR, Coops NC, Christen A et al. (2012) Integrated irradiance modelling in the urban environment based on remotely sensed data. Sol Energy 86: 2923-2934. doi: 10.1016/j.solener.2012.06.026
![]() |
[69] |
Karteris M, Slini T, Papadopoulos AM (2013) Urban solar energy potential in Greece: A statistical calculation model of suitable built roof areas for photovoltaics. Energ Buildings 62: 459-468. doi: 10.1016/j.enbuild.2013.03.033
![]() |
[70] | Chaves A, Bahill T (2010) Locating Sites for Photovoltaic Solar Panels: Pilot study uses DEM derived from LIDAR. ArcUser: The Magazine for Esri Sortware User 13: 24-27. |
[71] | Hillshade (3D Analyst), ArcGIS Help Library, ESRI. Available from: http://resources.arcgis.com/en/help/main/10.2/index.html#//00q900000036000000. [Accessed on 20 March 2015]. |
[72] | Nguyen HT, Pearce JM (2012) Incorporating shading losses in solar photovoltaic potential assessment at the municipal scale. Sol Energy 86: 1245-1260. |
[73] |
Ordóñez J, E. Jadraque E, Alegre J et al. (2010) Analysis of the photovoltaic solar energy capacity of residential rooftops in Andalusia (Spain). Renew Sust Energ Rev 14: 2122-2130. doi: 10.1016/j.rser.2010.01.001
![]() |
[74] |
Hong T, Koo C, Park J, et al. (2014) A GIS (Geographic Information System) - based optimization model for estimating the electricity generation of the rooftop PV (photovoltaic) system. Energy 65: 190-199. doi: 10.1016/j.energy.2013.11.082
![]() |
[75] |
Ko L, Wang JC, Chen CY et al. (2015) Evaluation of the development potential of rooftop solar photovoltaic in Taiwan. Renew Energ 76: 582-595. doi: 10.1016/j.renene.2014.11.077
![]() |
[76] |
Esclapés J, Ferreiro I, Piera J et al. (2014) A method to evaluate the adaptability of photovoltaic energy on urban façades. Sol Energy 105: 414-427. doi: 10.1016/j.solener.2014.03.012
![]() |
[77] | Government of Spain, Development Ministry (2009) Regulation Technical Building. Basic Document HE. |
[78] | Dean J, Kandt A, Burman K, et al. Analysis of Web Based Solar Photovoltaic Mapping Tools. ASME 3rd International Conference on Energy Sustainability. San Francisco, California 2009 July 19-23. |
[79] |
Kanters J, Wall M, Kjellsson E (2014) The Solar Map as a Knowledge Base for Solar Energy Use. Energy Procedia 48: 1597-1606. doi: 10.1016/j.egypro.2014.02.180
![]() |
[80] | Lisbon Solar Potential Map, LISBOA e-nova: Agéncia Municipal de Energia e ambiente. Available from: http://lisboaenova.org/cartasolarlisboa. [Accessed on 24 February 2015]. |
[81] | Bologna Solar City, SIT Comune di Bologna. Available from: http://sitmappe.comune.bologna.it/BolognaSolarCity/. [Accessed on 24 February 2015]. |
[82] | Cadastre Solaire de Paris. Available from: http://capgeo.sig.paris.fr/Apps/CadastreSolaire/. [Accessed on 4 February 2015]. |
[83] | Klärle M, Lanig S, Meik K (2011) Web-based solar roof cadastre goes international. GEOconnexion Int Mag 10: 30-33. |
[84] | Solardachpotenzial Osnabrück. Available from: http://geo.osnabrueck.de/solar/. [Accessed on 24 February 2015]. |
[85] | Dachflächeneignung für Solaranlagen, Stadt Gelsenkirchen. Available from: http://geo.gkd-el.de/website/solar/viewer.htm. [Accessed on 24 February 2015]. |
[86] | Braunschweig Solar. Available from: http://braunschweig.publicsolar.de/frames/. [Accessed on 24 February 2015]. |
[87] | Bristol Solar Map. Available from: http://www.bristol.gov.uk/page/environment/solar-energy. [Accessed on 24 February 2015]. |
[88] | Krüger A, Kolbe TH (2012) Building analysis for urban energy planning using key indicators on virtual 3D city model - The energy atlas of Berlin. Int Arch Photogramm Remote Sens Spatial Inform Sci XXXIX-B2: 145-150. |
[89] | Solaratlas Berlin, Berlin Business Location Center. Available from: http://www.businesslocationcenter.de/wab/maps/solaratlas/. [Accessed on 24 February 2015]. |
[90] | San Diego Solar Map. Available from: http://sd.solarmap.org/. [Accessed on 24 February 2015]. |
[91] | LA County Solar Map and Green Planning Tool. Available from: http://solarmap.lacounty.gov/. [Accessed on 24 February 2015]. |
[92] | New York City Solar Map. Available from: http://www.nycsolarmap.com/. [Accessed on 24 February 2015]. |
[93] | San Francisco energy Map. Available from: http://sfenergymap.org/. [Accessed on 24 February 2015]. |
[94] | Mapwell. Available from: http://www.mapdwell.com. [Accessed on 24 February 2015]. |
[95] | Zonnescan Provincie Noord-Brabant. Available from: http://www.zonnescanbrabant.nl. [Accessed on 24 February 2015]. |
[96] | SolarDachAtlas. Available from: http://www.solare-stadt.de/zgb/SolarDachAtlas. [Accessed on 24 February 2015]. |
[97] | Zonatlas. Available from: http://www.zonatlas.nl/home. [Accessed on 24 February 2015]. |
[98] | Denholm P, Margolis RM (2008) Land-use requirements and the per-capita solar footprint for photovoltaic generation in the United States. Energy Policy 36: 3531-3543. |
[99] | Choi Y, Rayl J, Tammineedi C, et al. (2011) PV Analyst: Coupling ArcGIS with TRNSYS to assess distributed photovoltaic potential in urban areas. Sol Energy 85: 2924-2939. |
[100] | Resch B, Sagl G, Törnros T, et al. (2014) GIS-Based planning and modeling for renewable energy: Challenges and future research avenues. ISPRS International Journal of Geo-Information 3: 662-692. |
[101] |
Pérez-Lombard L, Ortiz J, Pout C (2008) A review on buildings energy consumption information. Energ Buildings 40: 394-398. doi: 10.1016/j.enbuild.2007.03.007
![]() |
[102] |
Liang J, Gong J, Zhou J, et al. (2015) An open-source 3D solar radiation model integrated with a 3D Geographic Information System. Environ Modell Softw 64: 94-101. doi: 10.1016/j.envsoft.2014.11.019
![]() |
1. | Wen‐song Hong, Gang‐qing Zhang, Simulation analysis for tumor radiotherapy based on three‐component mathematical models, 2019, 20, 1526-9914, 22, 10.1002/acm2.12516 | |
2. | Luis L. Bonilla, Manuel Carretero, Filippo Terragni, Integrodifference master equation describing actively growing blood vessels in angiogenesis, 2020, 21, 2191-0294, 705, 10.1515/ijnsns-2019-0094 | |
3. | Vincenzo Capasso, Franco Flandoli, On stochastic distributions and currents, 2016, 4, 2325-3444, 373, 10.2140/memocs.2016.4.373 | |
4. | Luis L. Bonilla, Manuel Carretero, Filippo Terragni, 2019, Chapter 13, 978-3-030-15095-2, 413, 10.1007/978-3-030-15096-9_13 | |
5. | Federico Camerlenghi, Elena Villa, Large and moderate deviations for kernel–type estimators of the mean density of Boolean models, 2018, 12, 1935-7524, 10.1214/18-EJS1397 | |
6. | Marina V. Polovinkina, Amar Debbouche, Igor P. Polovinkin, Sergio A. David, Stability of stationary solutions for the glioma growth equations with radial or axial symmetries, 2021, 0170-4214, 10.1002/mma.7194 | |
7. | Meryem Alkama, Abdelilah Larrache, Mostafa Rachik, Ilias Elmouki, Optimal duration and dosage of BCG intravesical immunotherapy: A free final time optimal control approach, 2018, 41, 0170-4214, 2209, 10.1002/mma.4745 | |
8. | W. Duncan Martinson, Helen M. Byrne, Philip K. Maini, Evaluating snail-trail frameworks for leader-follower behavior with agent-based modeling, 2020, 102, 2470-0045, 10.1103/PhysRevE.102.062417 | |
9. | Luis L. Bonilla, M. Carretero, F. Terragni, 2018, Chapter 6, 978-3-319-76598-3, 97, 10.1007/978-3-319-76599-0_6 | |
10. | Haili Qian, Adriana Sujey Beltran, Mesoscience in cell biology and cancer research, 2022, 1, 2770-9183, 271, 10.1002/cai2.33 | |
11. | Franco Flandoli, Marta Leocata, Cristiano Ricci, The Mathematical modeling of Cancer growth and angiogenesis by an individual based interacting system, 2023, 562, 00225193, 111432, 10.1016/j.jtbi.2023.111432 |
|
|
|
||||
|
|
|
|
|
||
1.5 | 5.88 | 22.42 | 0.145 | 1 | 0.0045 | 0.002 |