=Paper= {{Paper |id=Vol-3065/paper6_147 |storemode=property |title=Intelligent Search for Personalized Cancer Therapy Synthesis: an Experimental Comparison |pdfUrl=https://ceur-ws.org/Vol-3065/paper6_147.pdf |volume=Vol-3065 |authors=Marco Esposito,Leonardo Picchiami |dblpUrl=https://dblp.org/rec/conf/aiia/EspositoP21 }} ==Intelligent Search for Personalized Cancer Therapy Synthesis: an Experimental Comparison== https://ceur-ws.org/Vol-3065/paper6_147.pdf
Intelligent Search for Personalized Cancer Therapy
Synthesis: an Experimental Comparison
Marco Esposito1 , Leonardo Picchiami1
1
    Computer Science Dept., Sapienza University of Rome, via Salaria 113, 00198, Italy


                                         Abstract
                                         CRC is one of the deadliest forms of tumor for adult humans, which often requires expensive and highly
                                         toxic treatments. In this work, we exploit a recent System Biology Markup Language (SBML) quanti-
                                         tative model of the tumor growth and its immune response to two immunotherapic drugs and define
                                         a non-linear, constrained optimization problem for the synthesis of personalized therapies. We com-
                                         pute a virtual population of physiologically admissible patients and define a parametric therapy model
                                         that enables the usage of standard parameter optimization algorithms. The whole approach is imple-
                                         mented with a novel methodology that requires a single tool, namely COPASI, for the therapy template
                                         modelling, the definition of the optimization problem and its solution. We present results of an exper-
                                         imental evaluation of our approach by comparing five search algorithms among those implemented in
                                         COPASI, in order to find out the most promising kind of algorithms for this problem. The results show
                                         that population-based search algorithms perform well and are able to compute personalized therapies
                                         within time bounds compatible with clinical practice.

                                         Keywords
                                         In Silico Clinical Trials, VPH models, AI search, Systems Biology, Simulation.




1. Introduction
In recent years, the scientific community has produced a large number of quantitative models
of human patho-physiology (see, e.g., the well-known BioModels [1] among the others).
   As these models become more and more accurate and gradually obtain certification from the
authorities (such as FDA, EMA, see, e.g., [2, 3]), new approaches to the design of pharmacological
treatments and the safety and efficacy assessment of therapies for high-impact diseases become
possible.
   Such computational methods, collectively referred to as In Silico Clinical Trials (ISCTs), exploit
Artificial Intelligence (AI) and Simulation-based Optimization techniques to design personalized
therapies based on the specific characteristics of each patient, thus enabling precision medicine.
Such methods, however, are still far from adoption in clinical contexts, due to both a lack of
specialized tools and low trust in modern computational techniques.
   The Virtual Phisiological Human (VPH) models used in ISCTs are generally designed to
consider inter-patient variability, i.e., difference in their features that cause, for instance, different

IPS-RCRA 2021: 9th Italian Workshop on Planning and Scheduling and 28th International Workshop on Experimental
Evaluation of Algorithms for Solving Problems with Combinatorial Explosion
" esposito@di.uniroma1.it (M. Esposito); piccchiami@di.uniroma1.it (L. Picchiami)
 0000-0003-4543-8818 (M. Esposito); 0000-0001-5477-6419 (L. Picchiami)
                                       © 2021 Copyright for this paper by its authors. Use permitted under Creative Commons License Attribution 4.0 International (CC BY 4.0).
    CEUR
    Workshop
    Proceedings
                  http://ceur-ws.org
                  ISSN 1613-0073
                                       CEUR Workshop Proceedings (CEUR-WS.org)
evolutions of the disease. The models also capture absorption of drugs and their interaction
with the patient’s organism, i.e., the Pharmaco-Kinetics/Dynamics (PKPD), which is crucial to
analyze the different responses to drugs.
   Quantitative VPH models can be seen as hybrid systems, where systems of Ordinary Differ-
ential Equations (ODEs) describe the dynamics of biological quantities, state and time events
capture discrete dynamics, and parameters encode the inter-patient variability.
   VPH models are often described and distributed in open-standard modelling languages, among
which System Biology Markup Language (SBML) [4] is the most widespread. There exist several
tools for the design and analysis of SBML models. One of the most complete is the open-source
software COPASI [5]. It supports most of the SBML standard for modelling biological processes,
several analysis tasks, such as time-course simulation, parameter optimization and estimation,
and sensitivity analysis. It also provides a graphical user interface as well as Application Program
Interfaces (APIs) for several programming languages.

1.1. Motivation
Pharmacological therapies, especially for high-impact and chronic diseases, are generally
considerably expensive and toxic. Thanks to the use of ISCTs, we can simulate the effects of
a large number of candidate therapies without the risks and costs of in vivo experiments on
human patients, with the goal of finding the best one for each individual.
   Given a VPH model and a particular virtual patient, the task of computing a personalized
pharmacological therapy that maximizes the efficacy of the treatment and its safety is a hard
one. It is often the case, in fact, that high dosing regimens have high chances of yielding
effective therapies, but provide low safety due to drug toxicity. Furthermore, the number of
possible therapies is generally infinite and, even if dosing values and time are discretized, grows
combinatorially in the number of drugs and the duration of the treatment.
   Naïve approaches to the modelling of the problem and the search for the best personalized
therapy fail to explore such space effectively and within time bounds compatible with clinical
practice. Given the complexity of the problem, it is unclear what search strategy is the most
promising, even if considering a particular VPH model or a single Virtual Patient (VP). Generally,
the search space is highly constrained, and the objectives are conflicting and strongly non-linear.

1.2. Contribution
In this paper, we consider a recent VPH model of the immune response to Colorectal Cancer
(CRC) and analyze the performance of different types of search algorithms to synthesize per-
sonalized therapies. We first generated a population of VPs, along the lines of [6] which makes
it possible to set up ISCTs. The definition of the search space, i.e., the set of candidate therapies,
is a crucial step, in that it severely influences the effectiveness of the synthesis. We designed
a parametric model of therapies and integrated it within the CRC SBML model. This enabled
a new methodology that, within a single tool, namely COPASI, supports the definition of the
therapy optimization problem and its solution with several different search algorithms, in order
to find the optimal therapy for any given VP.
1.3. Paper Outline
The paper is organized as follows. In Section 2 we recall the state of the art on ISCTs, optimization
and parameter estimation for VPH models, Section 3 introduces the CRC VPH model and in
Section 4 we describe how we modelled the therapies to enable the parameter optimization
approach. Then, in Section 5 we formalize the optimization problem and in Section 6 we show
the experimental results of different kinds of search algorithms to solve it. Finally, in Section 7
we draw our conclusions.


2. State of the art
A preliminary study for this work has been performed in [7]. In that work, the feasibility of
the approach has been studied using a simpler therapy template model and a smaller number
of algorithms. The problem of synthesizing personalized therapies using VPH models has
been approached with a variety of techniques, e.g., reinforcement learning [8] and genetic
and evolutionary algorithms [9, 10, 11, 12]. A simulation-based backtracking algorithm has
been proposed in [13, 14] for the problem of computing an optimal personalized therapy for
assisted reproduction (an area where therapeutic success depends on many factors hard to be
manually taken under control [15, 16, 17]) on a patient digital twin, defined as the subset of the
VPs of a complete population best matching (up to a used-defined error threshold) the human
patient clinical data. One of the most studied problems in systems biology is the estimation of
unknown model parameters to fit experimental data. [18, 19, 20, 21] present several specialized
optimization algorithms for this task. Real-world case studies include [22, 23, 24, 25]. Several
tools are explicitly designed to solve parameter optimisation problems in systems biology; the
most notable are AMICI [26], AMIGO2 [27], MEIGO [28], PESTO [29], Data2Dinamics [30],
dMod [31], pyABC [32] and SBML2Julia [33]. Most of these use SBML and allow the definition
of the optimisation problem via the PEtab [34] open standard. There exists a large number of
SBML simulators. Among the most commonly used, besides COPASI, we mention BioUML [35]
and LibRoadRunner [36]. SBML2Modelica [37] is a tool that focuses on the interoperability
between systems biology and (hybrid) Cyber-Physical Systems (CPSs) domains by automatically
translating SBML models to Modelica [38] and the FMI/FMU open standard. This enables the
exploitation of well-established techniques for CPS optimization and verification. For instance,
it is possible to use backtracking optimization algorithms [39] that exploit the efficient storing
and retrieval of intermediate simulator states [40], as well as verification algorithms for systems
also in presence of uncontrollable events (e.g., [41, 42, 43, 44, 45, 46]). The solution of large-
scale optimization problems via logic- [47], automata- [48] or constraint-based [49] formalisms
(e.g., [50, 51, 52, 53, 54] among others) has been largely studied in the literature, on several
industry-relevant application domains (e.g., [55, 56, 57, 58, 59, 60] among others). However,
when the problem model is a complex VPH model, such approaches cannot be applied, since the
model cannot be accurately defined an it is only available as a simulable black-box. The problem
of finding an optimal therapy can be stated as an optimal planning model in hybrid domains.
In fact, the VPH models we considered are hybrid systems whose inputs are discrete event
sequences [61, 62]. Existing approaches to solve this problem [63, 64, 65, 66] are not applicable
in our setting, as symbolic methods fail to capture the complexity of large ODE-based VPH
real-world models.


3. VPH model of the immune response to colorectal cancer
The VPH model proposed in [67] describes the dynamics of the growth of CRC and the immune
response of the patient’s immune system. The model also includes the PKPD of two drugs,
namely Atezolizumab and Cibisatamab, and the patient’s response to their administration. The
change of the tumor size over time, as well as the total quantity of drugs administered, can be
observed via simulation thanks to dedicated model variables.
   As it is common for models of this kind, the dynamics of the tumor growth and the immune
system’s response to the drugs vary significantly from patient to patient. The CRC model
presents 23 real-valued parameters, which jointly define the space of possible human patients.
An assignment to those parameters represents a possible patient, and the evolution of their
biological quantities can be computed by simulation.
   However, as discussed in [68, 69], not all assignments yield patients that are of interest and
plausible. In this model, it may be the case that a particular assignment leads to a patient that
does not develop a tumor or that shows dynamics that are incompatible with the laws of biology.
We used the approach from [6] to compute a population of physiologically admissible VPs. We
adopted the admissibility criteria from [67], i.e., admissible patients must develop a tumor of a
given size within 8000 days and have plausible dynamics. Figure 1 shows the relative growth of
the tumor in the 25 VPs we considered in our experiments, without any drug administration. It
can be observed that for all of them the condition gets progressively worse over time.
   We filtered out all VPs that showed no response to the two drugs among the whole virtual
population. In this model, in fact, 85% of all admissible patients lack the biological characteristic
to take advantage of immunotherapy. We excluded such patients because our ultimate goal is to
assess the ability of optimization algorithms to synthesize effective treatments and not to study
the efficacy of the drugs over the whole population (as done in [67]). Including non-responding
VPs in the population would significantly alter the results since no optimization technique could
ever find an effective therapy.
   The CRC model makes it possible to simulate any therapy over a given VP. However, safety
constraints exist over therapies (defined thanks to clinical trials on human patients), which deal
with the toxicity of the drugs. In particular, it is not safe to administer more than 24 mg/kg
of Atezolizumab and 160 mg of Cibisatamab in a single dose or cumulatively within a certain
period (four weeks for Atezolizumab and three weeks for Cibisatamab).
   In general, any VP is simulated for 400 days from the day in which the tumor has reached
the given threshold (one of the 23 parameters). Such a period is deemed long enough to verify
whether patients respond positively to therapies or not.
   Figure 1 shows the relative growth of the tumor diameter in the 25 VPs we used in our
experiments when no therapy is administered. For all of them, the condition get worse and
worse over time.
Figure 1: Relative growth of tumor size in the 25 patients without therapy.


4. Modelling the therapies
The set of all possible therapies involving any given number of drugs is infinite. It corresponds,
in fact, to the set of all functions mapping each time instant (inside a finite interval) to the
quantity of each drug administered at that time.
   In order to find a personalized therapy for a given VP, we chose to parametrize the space
of therapies. Defining a parametric template model for therapies enables, in fact, the use of
standard parameter optimization techniques to compute an optimized therapy. The choice of
one template model (among the many possible ones) is sensible since it has several repercussions
over the success of our approach and the usefulness of the result. For our purposes, a desirable
template model should meet a set of requirements.
Number of parameters. The number of parameters of the template must be as low as possible.
   The optimization problem generally gets harder and harder as the parameter space gets
   larger since there may be a combinatorial explosion in the number of possible therapies.

Expressiveness. The template model should be able to represent the largest possible portion
     of the therapies that are of interest. Since the goal is to find the optimal therapy, the set
     of expressible therapies should include it.

Constraints and symmetries. The number of constraints that define the space of admissible
     therapies should be as low as possible. Ideally, any assignment of the parameters should
     encode a therapy that is safe and of interest. Moreover, it would be desirable not to have
     symmetries (see, e.g., [70]), i.e., different parameter assignments that correspond to the
     same actual therapy.

Regularity and Explainability. The therapies that the template model can express must be
     as regular as possible. Very irregular therapies (e.g., doses changing drastically from
      week to week) may fail to obtain the physician’s trust, which is crucial as the ultimate
      responsibility for the treatment lies on them.

Integration. The therapy template model must be described using the SBML language and
     integrated within the CRC model.

   We designed a therapy template model that offers
                                                  (︀ 1 a good   trade-off between     all the require-
ments. We define 32 parameters: 30 of them, 𝑑𝑎 , . . . , 𝑑𝑎 , 𝑑𝑐 , . . . , 𝑑𝑐 , are real-valued and
                                                             15   1         15
                                                                               )︀

define the doses of each drug that can be administered during each of the 15 months, while the
remaining two, 𝑓𝑎 and 𝑓𝑐 , are integer numbers that define the frequency of administration of,
respectively, Atezolizumab and Cibisatamab. For each drug 𝑥, the administered dose at week
𝑘 ∈ [1, 60], denoted by adm𝑥 (𝜏, 𝑘), is equal to 𝑑𝑚𝑥 × a(𝑥, 𝑘), where 𝑚 = ⌈ 4 ⌉ and a(𝑥, 𝑘) is 1 if
                                                                                  𝑘

𝑘 mod 𝑓𝑥 = 1 and 0 otherwise.
   Within this therapy template, we model the safety requirements about the maximum dose
that can be administered in a single day as simple bound constraints (aka box constraints) over
the 30 dose parameters.
   The other two safety requirements, which limit the total quantity of drug administered within
a certain period, must be modelled as non-linear constraints. Although COPASI supports the
definition of non-linear constraints in the optimization problem, they are only evaluated after
the simulation and not symbolically. Hence, the most straightforward way to model them is as
functional (i.e., black-box) constraints as follows. Let 𝜏 be a therapy. We introduce two new
variables in the model, denoted by safety𝑎𝑡𝑒 (𝜏, 𝑡) and safety𝑐𝑖𝑏𝑖 (𝜏, 𝑡), which, at time 𝑡, will hold
0 if the safety constraint over, respectively Atezolizumab and Cibisatamab is satisfied and a
positive number otherwise.
   We introduce an additional linear constraint that makes sure that the feasible therapies
administer at least one of the two drugs within the first week as they would otherwise not be
realistic.
   The proposed model has a relatively low number of parameters and expresses very regular
and explainable therapies since the doses stay constant over each month. The fact that it
cannot represent some classes of therapies (e.g., with doses changing within one month) is not
a compelling limitation as the set of expressible therapies is still very large and diverse.
   We described this therapy template model using the SBML language and integrated it into
the model. Therefore, the resulting VPH model has 55 parameters that can be governed: 23 of
them define the VPs and the remaining 32 define the space of therapies.
   Finally, we note that our template model is able to represent all therapies for CRC (involv-
ing Atezolizumab and Cibisatamab) that have been tested in past clinical trials and that are
summarized in [67].


5. Optimization
Given a VP 𝜆 in the set of all possible VPs Λ ⊆ R23 , let 𝑡0 be the starting time of therapies for
𝜆, 𝐻 = 𝑡0 + 400 be the time horizon of the therapies for 𝜆, and T = R30 × N2 be the space
of the therapy template model parameters. We want to compute a therapy 𝜏 which is optimal
under two conflicting objectives: maximizing the efficacy and minimizing the total quantity of
administered drugs. We first define a measure of the inefficacy of a therapy 𝜏 as

                                                     tumorsize (𝜆, 𝜏, 𝐻)2
                                 ineff (𝜆, 𝜏, 𝐻) =                          ,
                                                      tumorsize (𝜆, 𝜏, 𝑡0 )

where tumorsize (𝜆, 𝜏, 𝑡) is a model variable that holds the size of the tumor at any time 𝑡 during
the simulation. We chose such a measure so that we can compare the inefficacy of therapies
over different patients. In fact, only considering the final tumor size as the inefficacy measure
would ignore the fact that the initial tumor size varies from patient to patient. Instead, our
measure also considers the tumor reduction, and thus, given two patients with the same final
tumor size, it gives a lower (better) inefficacy value to the one with the larger initial tumor.
   We now define a measure of the quantity of drugs that is administered with therapy 𝜏 . Thanks
to the safety constraints on each drug, we know the maximum total amounts of each drug that
can be administered while still keeping the therapy safe. Let Max𝑎𝑡𝑒 and Max𝑐𝑖𝑏𝑖 be such values.
We define
                                               total𝑎𝑡𝑒 (𝜏, 𝐻)          total𝑐𝑖𝑏𝑖 (𝜏, 𝐻)
                                      (︂                                                 )︂
           totdrugs (𝜏, 𝐻) = 100 × 0.5 ×                       + 0.5 ×
                                                   Max𝑎𝑡𝑒                   Max𝑐𝑖𝑏𝑖

where total𝑎𝑡𝑒 (𝜏, 𝑡) and total𝑐𝑖𝑏𝑖 (𝜏, 𝑡) are model variables that hold the total quantity of, re-
spectively Atezolizumab and Cibisatamab administered up to time 𝑡 with therapy 𝜏 .
  We are now able to define our objective function: 𝐽 : Λ × T → R such that

                      𝐽 (𝜆, 𝜏 ∈ T) = 𝛼 × ineff (𝜆, 𝜏, 𝐻) + 𝛽 × totdrugs (𝜏, 𝐻)

with 𝛼, 𝛽 ∈ R+ .
     Given the maximum single-day doses 𝜇𝑎𝑡𝑒 and 𝜇𝑐𝑖𝑏𝑖 for, respectively Atezolizumab and
(︀Cibisatamab,         solving the personalized        therapy synthesis problem means finding a therapy 𝜏 =
   𝑑1𝑎 , . . . , 𝑑15    1 , . . . , 𝑑15 , 𝑓 , 𝑓        that minimizes 𝐽 (𝜆, 𝜏 ) while satisfying all constraints.
                                                )︀
                  𝑎  , 𝑑𝑐            𝑐     𝑎 𝑐     ∈ T
In particular, 𝜏 must be such that

    • 0 ≤ 𝑑𝑖𝑥 ≤ 𝜇𝑥       ∀ 𝑖 ∈ [1, 15] ∀ 𝑥 ∈ {𝑎𝑡𝑒, 𝑐𝑖𝑏𝑖}

    • safety𝑥 (𝜏, 𝐻) ≤ 0 ∀ 𝑥 ∈ {𝑎𝑡𝑒, 𝑐𝑖𝑏𝑖}

    • 𝑑1𝑎 + 𝑑1𝑐 > 0.

The latter constraint ensures that at least one of the two drugs is administered in the first
week. In fact, regardless of the frequency of administration of drug 𝑥, if 𝑑1𝑥 is > 0, the therapy
administers a dose equal to 𝑑1𝑥 at day 1.

5.1. Optimization algorithms
Among the different parameter optimization algorithms included in COPASI, we chose five in
the attempt to cover different search techniques as much as possible.
  We now describe each algorithm and, in particular, the salient characteristic of their imple-
mentation in COPASI.
Random Search. We used this algorithm as a baseline for comparing all studied optimization
    methods. The algorithm samples a random assignment of the therapy parameters at each
    step and evaluates the objective function if all constraints are met.

Levenberg-Marquardt. This algorithm is an adaptive-step gradient descent method, a hybrid
     between Steepest Descent and Newton methods. At each step, it uses finite differences
     to estimate the first and second derivatives in the current point. Bound constraints are
     handled by forcing parameters to stay within their bounds at each step. On the other
     hand, functional constraints are converted to non-negative penalties on the objective
     function.

Nelder-Mead. The algorithm constructs a simplex, i.e., a polytope of 𝑁 + 1 vertices in 𝑁
     dimensions. At each step, the objective function is evaluated at each vertex, and the
     polytope is modified according to the results. Constraints are handled analogously to the
     Levenberg-Marquardt algorithm.

Genetic Algorithm. A population-based metaheuristic search inspired by evolution. The
    initial population is constructed by including randomly sampled individuals and one
    individual given by the user. A new generation is obtained by randomly pairing individuals
    and producing two new individuals from each pair. The number of crossover points is
    chosen randomly between zero and 50% of the number of parameters, while a random
    parameter is mutated. In the selection phase, each individual’s fitness (at this point, the
    number of individuals is double the initial number) is compared with that of a number of
    others equal to 20% of the population size. Then, all individuals are sorted by the number
    of individuals they outperformed, and finally, the population size is reduced to the initial
    number of individuals by removing those who performed worse. The algorithm always
    forces parameters inside their bounds, both when sampling random individuals and when
    applying mutations. If an individual does not meet all functional constraints, its fitness
    value is set to infinity, and thus it is effectively excluded. We set the population size to 64
    and the mutational variance to 0.1.

Particle Swarm Optimization. A population-based metaheuristic search method in which a
      set of particles (points in the parameter space) collectively explore the search space in
      order to find the global minimum. The algorithm terminates if the standard deviation of
      the objective values of the particles is smaller than a given value. This implementation
      handles both bound and functional constraints like the Genetic Algorithm implementation.
      We used a swarm size of 30 and a standard deviation for termination of 0.1.


6. Results
In this section we show the outcome of the execution of all search algorithms over the 25 VPs.
In each experiment, the initial point was set as the reference therapy described in [67], which
administers the maximum allowed doses of drugs. The coefficients 𝛼 and 𝛽 of the objective
function were set, respectively to 0.9 and 0.1. In order to mimic a real clinical setting, we
Figure 2: Average inefficacy and average total quantity of drug for each algorithm.


assigned to each experiment a time budget of 1 hour. All experiments were run on two identical
machines equipped with Intel (R) Xeon (R) 8-core CPU E5430 and 32 GB RAM.
   Figure 2 shows the average of the total quantity of drugs and the average inefficacy over
all patients of the optimal therapies synthesized by each algorithm. In the chart, 100% of
drugs quantity represents the maximum safe total quantity that can be administered, i.e., the
quantity of drugs administered by the reference therapy. For each patient, 100% of inefficacy
indicates the inefficacy value reached on that particular patient with the reference therapy
(which administers the maximum allowed quantity of drugs).
   We first note that the Levenberg-Marquardt algorithm never improves over the reference
therapy for all patients. Nelder-Mead, on the other hand, performs worse than Random Search,
with an average reduction of the total drug quantity of just over 20%, while Random Search is
able to reduce the total dose of around 72%. Genetic Algorithm and Particle Swarm Optimization
both manage to drastically reduce the average total drug quantity while significantly improve on
the inefficacy. The poor performance of the Levenberg-Marquardt and Nelder-Mead algorithms
may be explained by the naïve handling of the bound and functional constraints. In fact, both
algorithms were not designed to work in constrained space and the COPASI implementation
makes ineffective choices whenever constraints are found to be violated. We also note that
pure gradient-descent methods, like Levenberg-Marquardt, are naturally unfit for our problem,
since the objective function is highly non-linear and the space is constrained. Not being able to
escape from local minima and unfeasible regions is likely to make the problem unapproachable
for this whole family of algorithms.
   Figure 3 shows the Response Evaluation Criteria in Solid Tumors (RECIST) classification
of the VPs with the reference therapy and with the optimal therapies. RECIST [71] is a set of
classification criteria for the patients response to cancer therapies, based on the final diameter
of the tumor after the treatment. RECIST classifies patients in four different classes: Complete
Response (CR) if the final tumor diameter is < 10 mm, Partial Response (PR) if the tumor shrinks
by at least 30%, Stable Disease (SD) if the diameter decreases by less than 30% or increases
at most by 20% and Progressive Disease (PD) if the diameter grows by more than 20%. The
Figure 3: RECIST results with each algorithm.


RECIST classification of the 25 VPs with the optimal therapies are consistent with the previously
shown results. In particular, we note that, with Genetic Algorithm and Particle Swarm, 3 and 4
VPs respectively change from PD to CR when the optimal therapy is administered.
   Figure 4 shows the growth of the tumor of each VP with the reference therapy and with
the optimal therapies synthesized with each algorithm, excluding Levenberg-Marquardt. As
expected from the results shown in Figure 2, we observe that the tumor size tends to decrease
most with Random Search, Genetic Algorithm and Particle Swarm. However, the latter also
manages, for some patients, to reduce the tumor size more quickly, thus computing therapies
that are safer and more effective.
   Finally, Figure 5 shows the optimal therapies found for patient #3 with each algorithm. In
each plot, 100% of a drug indicates the maximum quantity of that drug that can be administered
in a single dose.


7. Conclusions
In this work, we exploited a SBML model of Colorectal Cancer to design an ISCT and compared
five different search algorithms in the computation of personalized pharmacological therapies.
The whole approach was implemented by using the tool COPASI. The approach required the
modelling of a non-trivial parametric model of therapies, which allowed to reduce the search
space to therapies that are actually of interest and enable the use of standard parameter opti-
mization methods. We conclude, in our experimental analysis, that gradient-based algorithms,
as well as algorithm that do not employ a smart handling of constraints, fail to solve the problem
effectively. On the other side, population-based algorithms such as Genetic Algorithms and
Particle Swarm Optimization show good and promising performance. Possible future research
directions include the comparison of other families of algorithms (also outside of COPASI) and
the modelling of more sophisticated therapy template models.
          (a) Reference.                       (b) RS.                         (c) NM.




                               (d) G.                         (e) PSO.
Figure 4: Relative tumor size over time with reference therapy and the optimal therapies synthesized
with each algorithm.




               (a) Random Search.                                 (b) Nelder-Mead.




              (c) Genetic algorithm.                      (d) Particle Swarm Optimization.
Figure 5: Optimal therapies found for patient #3.


Acknowledgments
This work was partially supported by: Italian Ministry of University and Research under grant
“Dipartimenti di eccellenza 2018–2022” of the Department of Computer Science of Sapienza Uni-
versity of Rome. INdAM “GNCS Project 2020”; Sapienza University projects RG11816436BD4F21,
RG11916B892E54DB, RP11916B8665242F; Lazio POR FESR projects E84G20000150006, F83G17000830007.
References
 [1] N. Le Novère, B. Bornstein, A. Broicher, M. Courtot, M. Donizelli, H. Dharuri, L. Li, H. Sauro,
     M. Schilstra, B. Shapiro, J. Snoep, M. Hucka, BioModels Database: A free, centralized
     database of curated, published, quantitative kinetic models of biochemical and cellular
     systems, Nucl. Ac. Res. 34 (2006). doi:10.1093/nar/gkj092.
 [2] U.S.A. Food and Drug Administration, Physiologically based pharmacokinetic analyses –
     format and content guidance for industry, 2018. FDA-2016-D-3969.
 [3] European Medicines Agency, Reporting of physiologically based pharmacoki-
     netic (PBPK) modelling and simulation, 2019. URL: https://www.ema.europa.eu/
     en/reporting-physiologically-based-pharmacokinetic-pbpk-modelling-simulation,
     eMA/CHMP/458101/2016.
 [4] M. Hucka, F. Bergmann, A. Dräger, S. Hoops, S. Keating, N. Le Novère, C. Myers, B. Olivier,
     S. Sahle, J. Schaff, L. Smith, D. Waltemath, D. Wilkinson, The Systems Biology Markup
     Language (SBML): Language specification for Level 3 Version 2 Core, JIB 15 (2018).
     doi:10.1515/jib-2017-0081.
 [5] C. Lee, L. Xu, M. Singhal, P. Mendes, S. Hoops, J. ù Pahle, N. Simus, R. Gauges, S. Sahle,
     U. Kummer, COPASI —- A COmplex PAthway SImulator, Bioinformatics 22 (2006).
     doi:10.1093/bioinformatics/btl485.
 [6] S. Sinisi, V. Alimguzhin, T. Mancini, E. Tronci, B. Leeners, Complete populations of
     virtual patients for in silico clinical trials, Bioinformatics 36 (2020). doi:10.1093/
     bioinformatics/btaa1026.
 [7] M. Esposito, L. Picchiami, Simulation-based synthesis of personalised therapies for col-
     orectal cancer, in: OVERLAY, 2021. To appear.
 [8] A. Jalalimanesh, H. S. Haghighi, A. Ahmadi, M. Soltani, Simulation-based optimization
     of radiotherapy: Agent-based modeling and reinforcement learning, Mathematics and
     Computers in Simulation 133 (2017) 235–248.
 [9] R. Schmucker, G. Farina, J. Faeder, F. Fröhlich, A. S. Saglam, T. Sandholm, Combination
     treatment optimization using a pan-cancer pathway model, bioRxiv (2020).
[10] N. Noman, P. Moscato, Designing optimal combination therapy for personalised glioma
     treatment, Memetic Computing 12 (2020) 317–329.
[11] T. Cassidy, M. Craig, Determinants of combination gm-csf immunotherapy and oncolytic
     virotherapy success identified through in silico treatment personalization, PLoS computa-
     tional biology 15 (2019) e1007495.
[12] A. L. Jenner, F. Frascoli, C.-O. Yun, P. S. Kim, Optimising hydrogel release profiles for viro-
     immunotherapy using oncolytic adenovirus expressing il-12 and gm-csf with immature
     dendritic cells, Applied Sciences 10 (2020) 2872.
[13] T. Mancini, F. Mari, A. Massini, I. Melatti, I. Salvo, S. Sinisi, E. Tronci, R. Ehrig, S. Röblitz,
     B. Leeners, Computing personalised treatments through in silico clinical trials. A case
     study on downregulation in assisted reproduction, in: RCRA 2018, volume 2271 of CEUR
     W.P., CEUR, 2018.
[14] S. Sinisi, V. Alimguzhin, T. Mancini, E. Tronci, F. Mari, B. Leeners, Optimal personalised
     treatment computation through in silico clinical trials on patient digital twins, Fundam.
     Inform. 174 (2020). doi:10.3233/FI-2020-1943.
[15] B. Leeners, T. Kruger, K. Geraedts, E. Tronci, T. Mancini, F. Ille, M. Egli, S. Roeblitz, L. Saleh,
     K. Spanaus, C. Schippert, Y. Zhang, M. Hengartner, Lack of associations between female
     hormone levels and visuospatial working memory, divided attention and cognitive bias
     across two consecutive menstrual cycles, Front. Behav. Neuro. 11 (2017). doi:10.3389/
     fnbeh.2017.00120.
[16] M. Hengartner, T. Kruger, K. Geraedts, E. Tronci, T. Mancini, F. Ille, M. Egli, S. Roeblitz,
     R. Ehrig, L. Saleh, K. Spanaus, C. Schippert, Y. Zhang, B. Leeners, Negative affect is
     unrelated to fluctuations in hormone levels across the menstrual cycle: Evidence from
     a multisite observational study across two successive cycles, J. Psycho. Res. 99 (2017).
     doi:10.1016/j.jpsychores.2017.05.018.
[17] B. Leeners, T. Krüger, K. Geraedts, E. Tronci, T. Mancini, M. Egli, S. Röblitz, L. Saleh,
     K. Spanaus, C. Schippert, Y. Zhang, F. Ille, Associations between natural physiological
     and supraphysiological estradiol levels and stress perception, Front. Psycol. 10 (2019).
     doi:10.3389/fpsyg.2019.01296.
[18] L. Schmiester, D. Weindl, J. Hasenauer, Efficient gradient-based parameter estimation for
     dynamic models using qualitative data, Bioinformatics (2021). URL: https://doi.org/10.
     1093/bioinformatics/btab512. doi:10.1093/bioinformatics/btab512.
[19] A. F. Villaverde, F. Fröhlich, D. Weindl, J. Hasenauer, J. R. Banga, Benchmarking optimiza-
     tion methods for parameter estimation in large kinetic models, Bioinformatics 35 (2019)
     830–838.
[20] Y. Schälte, P. Stapor, J. Hasenauer, Evaluation of derivative-free optimizers for parameter
     estimation in systems biology, IFAC-PapersOnLine 51 (2018) 98–101.
[21] A. Yazdani, L. Lu, M. Raissi, G. E. Karniadakis, Systems biology informed deep learning
     for inferring parameters and hidden dynamics, PLoS computational biology 16 (2020)
     e1007575.
[22] O. D. Sánchez, E. Ruiz-Velázquez, A. Y. Alanís, G. Quiroz, L. Torres-Treviño, Parameter
     estimation of a meal glucose–insulin model for tidm patients from therapy historical data,
     IET systems biology 13 (2019) 8–15.
[23] C. Audigier, T. Mansi, H. Delingette, S. Rapaka, V. Mihalef, D. Carnegie, E. Boctor, M. Choti,
     A. Kamen, D. Comaniciu, et al., Parameter estimation for personalization of liver tumor
     radiofrequency ablation, in: International MICCAI Workshop on Computational and
     Clinical Challenges in Abdominal Imaging, Springer, 2014, pp. 3–12.
[24] T. Chen, N. F. Kirkby, R. Jena, Optimal dosing of cancer chemotherapy using model
     predictive control and moving horizon state/parameter estimation, Computer methods
     and programs in biomedicine 108 (2012) 973–983.
[25] M. Raissi, N. Ramezani, P. Seshaiyer, On parameter estimation approaches for predicting
     disease transmission through optimization, deep learning and statistical inference methods,
     Letters in Biomathematics 6 (2019) 1–26.
[26] F. Fröhlich, D. Weindl, Y. Schälte, D. Pathirana, Ł. Paszkowski, G. T. Lines, P. Stapor,
     J. Hasenauer, Amici: High-performance sensitivity analysis for large ordinary differential
     equation models, arXiv preprint arXiv:2012.09122 (2020).
[27] E. Balsa-Canto, D. Henriques, A. Gábor, J. R. Banga, Amigo2, a toolbox for dynamic
     modeling, optimization and control in systems biology, Bioinformatics 32 (2016) 3357–
     3359.
[28] J. A. Egea, D. Henriques, T. Cokelaer, A. F. Villaverde, A. MacNamara, D.-P. Danciu, J. R.
     Banga, J. Saez-Rodriguez, Meigo: an open-source software suite based on metaheuristics
     for global optimization in systems biology and bioinformatics, BMC bioinformatics 15
     (2014) 1–9.
[29] P. Stapor, D. Weindl, B. Ballnus, S. Hug, C. Loos, A. Fiedler, S. Krause, S. Hroß, F. Fröhlich,
     J. Hasenauer, PESTO: Parameter EStimation TOolbox, Bioinf. 34 (2018) 705–707.
[30] A. Raue, B. Steiert, M. Schelker, C. Kreutz, T. Maiwald, H. Hass, J. Vanlier, C. Tönsing, L. Ad-
     lung, R. Engesser, et al., Data2dynamics: a modeling environment tailored to parameter
     estimation in dynamical systems, Bioinformatics 31 (2015) 3558–3560.
[31] D. Kaschek, W. Mader, M. Fehling-Kaschek, M. Rosenblatt, J. Timmer, Dynamic modeling,
     parameter estimation, and uncertainty analysis in r, Journal of Statistical Software 88
     (2019) 1–32.
[32] E. Klinger, D. Rickert, J. Hasenauer, pyabc: distributed, likelihood-free inference, Bioinfor-
     matics 34 (2018) 3591–3593.
[33] P. F. Lang, S. Shin, V. M. Zavala, Sbml2julia: interfacing sbml with efficient nonlinear julia
     modelling and solution tools for parameter optimization, arXiv preprint arXiv:2011.02597
     (2020).
[34] L. Schmiester, Y. Schälte, F. T. Bergmann, T. Camba, E. Dudkin, J. Egert, F. Fröhlich,
     L. Fuhrmann, A. L. Hauber, S. Kemmer, et al., Petab—interoperable specification of
     parameter estimation problems in systems biology, PLoS computational biology 17 (2021)
     e1008646.
[35] F. Kolpakov, I. Akberdin, T. Kashapov, l. Kiselev, S. Kolmykov, Y. Kondrakhin, E. Kutumova,
     N. Mandrik, S. Pintus, A. Ryabova, et al., Biouml: an integrated environment for systems
     biology and collaborative analysis of biomedical data, Nucleic acids research 47 (2019)
     W225–W233.
[36] E. Somogyi, J.-M. Bouteiller, J. Glazier, M. König, J. Medley, M. Swat, H. Sauro, libRoad-
     Runner: A high performance SBML simulation and analysis library, Bioinformatics 31
     (2015). doi:10.1093/bioinformatics/btv363.
[37] F. Maggioli, T. Mancini, E. Tronci, SBML2Modelica: Integrating biochemical models
     within open-standard simulation ecosystems, Bioinformatics 36 (2020). doi:10.1093/
     bioinformatics/btz860.
[38] P. Fritzson, V. Engelson, Modelica—a unified object-oriented language for system modeling
     and simulation, in: European Conference on Object-Oriented Programming, Springer,
     1998, pp. 67–90.
[39] T. Mancini, F. Mari, A. Massini, I. Melatti, F. Merli, E. Tronci, System level formal verification
     via model checking driven simulation, in: CAV 2013, volume 8044 of LNCS, Springer, 2013.
     doi:10.1007/978-3-642-39799-8_21.
[40] S. Sinisi, V. Alimguzhin, T. Mancini, E. Tronci, Reconciling interoperability with efficient
     verification and validation within open source simulation environments, Simul. Model.
     Pract. Theory 109 (2021). doi:10.1016/j.simpat.2021.102277.
[41] T. Mancini, F. Mari, A. Massini, I. Melatti, E. Tronci, SyLVaaS: System level formal
     verification as a service, in: PDP 2015, IEEE, 2015. doi:10.1109/PDP.2015.119.
[42] T. Mancini, F. Mari, A. Massini, I. Melatti, E. Tronci, SyLVaaS: System level formal
     verification as a service, Fundam. Inform. 149 (2016). doi:10.3233/FI-2016-1444.
[43] T. Mancini, F. Mari, A. Massini, I. Melatti, E. Tronci, Anytime system level verification
     via parallel random exhaustive hardware in the loop simulation, Microprocessors and
     Microsystems 41 (2016). doi:10.1016/j.micpro.2015.10.010.
[44] T. Mancini, F. Mari, A. Massini, I. Melatti, I. Salvo, E. Tronci, On minimising the maximum
     expected verification time, Inf. Proc. Lett. 122 (2017). doi:10.1016/j.ipl.2017.02.001.
[45] T. Mancini, F. Mari, I. Melatti, I. Salvo, E. Tronci, J. Gruber, B. Hayes, M. Prodanovic,
     L. Elmegaard, User flexibility aware price policy synthesis for smart grids, in: DSD 2015,
     IEEE, 2015. doi:10.1109/DSD.2015.35.
[46] T. Mancini, I. Melatti, E. Tronci, Any-horizon uniform random sampling and enumeration
     of constrained scenarios for simulation-based formal verification, IEEE TSE (2021). doi:10.
     1109/TSE.2021.3109842.
[47] C. Barrett, C. Tinelli, Satisfiability modulo theories, in: Handbook of Model Checking,
     Springer, 2018.
[48] E. Clarke, T. Henzinger, H. Veith, Handbook of Model Checking, Springer, 2016.
[49] F. Rossi, P. Van Beek, T. Walsh (Eds.), Handbook of Constraint Programming, Elsevier,
     2006.
[50] M. Cadoli, T. Mancini, Combining relational algebra, SQL, constraint modelling, and local
     search, TPLP 7 (2007). doi:10.1017/S1471068406002857.
[51] M. Cadoli, T. Mancini, F. Patrizi, SAT as an effective solving technology for constraint
     problems, in: ISMIS 2006, volume 4203 of LNCS, Springer, 2006.
[52] T. Mancini, M. Cadoli, D. Micaletto, F. Patrizi, Evaluating ASP and commercial solvers on
     the CSPLib, Constraints 13 (2008).
[53] T. Mancini, P. Flener, J. Pearson, Combinatorial problem solving over relational databases:
     View synthesis through constraint-based local search, in: SAC 2012, ACM, 2012. doi:10.
     1145/2245276.2245295.
[54] T. Mancini, F. Mari, I. Melatti, I. Salvo, E. Tronci, An efficient algorithm for network
     vulnerability analysis under malicious attacks, in: ISMIS 2018, Springer, 2018. doi:10.
     1007/978-3-030-01851-1_29.
[55] T. Mancini, E. Tronci, A. Scialanca, F. Lanciotti, A. Finzi, R. Guarneri, S. Di Pompeo,
     Optimal fault-tolerant placement of relay nodes in a mission critical wireless network, in:
     RCRA 2018, volume 2271 of CEUR W.P., CEUR, 2018.
[56] Q. Chen, A. Finzi, T. Mancini, I. Melatti, E. Tronci, MILP, pseudo-boolean, and OMT solvers
     for optimal fault-tolerant placements of relay nodes in mission critical wireless networks,
     Fundam. Inform. 174 (2020). doi:10.3233/FI-2020-1941.
[57] T. Mancini, F. Mari, I. Melatti, I. Salvo, E. Tronci, J. Gruber, B. Hayes, M. Prodanovic,
     L. Elmegaard, Demand-aware price policy synthesis and verification services for smart
     grids, in: SmartGridComm 2014, IEEE, 2014. doi:10.1109/SmartGridComm.2014.
     7007745.
[58] B. Hayes, I. Melatti, T. Mancini, M. Prodanovic, E. Tronci, Residential demand management
     using individualised demand aware price policies, IEEE Trans. Smart Grid 8 (2017). doi:10.
     1109/TSG.2016.2596790.
[59] T. Mancini, F. Mari, I. Melatti, I. Salvo, E. Tronci, J. Gruber, B. Hayes, L. Elmegaard, Parallel
     statistical model checking for safety verification in smart grids, in: SmartGridComm 2018,
     IEEE, 2018. doi:10.1109/SmartGridComm.2018.8587416.
[60] I. Melatti, F. Mari, T. Mancini, M. Prodanovic, E. Tronci, A two-layer near-optimal strategy
     for substation constraint management via home batteries, IEEE Trans. Ind. Elect. (2021).
     doi:10.1109/TIE.2021.3102431.
[61] T. Mancini, F. Mari, A. Massini, I. Melatti, E. Tronci, System level formal verification
     via distributed multi-core hardware in the loop simulation, in: PDP 2014, IEEE, 2014.
     doi:10.1109/PDP.2014.32.
[62] T. Mancini, F. Mari, A. Massini, I. Melatti, E. Tronci, Anytime system level verification
     via random exhaustive hardware in the loop simulation, in: DSD 2014, IEEE, 2014. doi:10.
     1109/DSD.2014.91.
[63] M. Fox, D. Long, Modelling mixed discrete-continuous domains for planning., JAIR 27
     (2006).
[64] M. Vallati, D. Magazzeni, B. De Schutter, L. Chrpa, T. McCluskey, Efficient macroscopic
     urban traffic models for reducing congestion: A PDDL+ planning approach, in: AAAI 2016,
     AAAI, 2016.
[65] S. Bogomolov, D. Magazzeni, A. Podelski, M. Wehrle, Planning as model checking in
     hybrid domains., in: AAAI 2014, AAAI, 2014.
[66] S. Bogomolov, D. Magazzeni, S. Minopoli, M. Wehrle, PDDL+ planning with hybrid
     automata: Foundations of translating must behavior., in: ICAPS 2015, AAAI, 2015.
[67] H. Ma, H. Wang, R. J. Sové, J. Wang, C. Giragossian, A. S. Popel, Combination therapy with
     t cell engager and pd-l1 blockade enhances the antitumor potency of t cells as predicted
     by a qsp model, Journal for immunotherapy of cancer 8 (2020).
[68] E. Tronci, T. Mancini, I. Salvo, S. Sinisi, F. Mari, I. Melatti, A. Massini, F. Davi’, T. Dierkes,
     R. Ehrig, S. Röblitz, B. Leeners, T. Krüger, M. Egli, F. Ille, Patient-specific models from
     inter-patient biological models and clinical records, in: FMCAD 2014, IEEE, 2014. doi:10.
     1109/FMCAD.2014.6987615.
[69] T. Mancini, E. Tronci, I. Salvo, F. Mari, A. Massini, I. Melatti, Computing biological model
     parameters by parallel statistical model checking, in: IWBBIO 2015, volume 9044 of LNCS,
     Springer, 2015. doi:10.1007/978-3-319-16480-9_52.
[70] T. Mancini, M. Cadoli, Detecting and breaking symmetries by reasoning on problem
     specifications, in: SARA 2005, volume 3607 of LNCS, Springer, 2005.
[71] L. H. Schwartz, S. Litière, E. De Vries, R. Ford, S. Gwyther, S. Mandrekar, L. Shankar,
     J. Bogaerts, A. Chen, J. Dancey, et al., Recist 1.1—update and clarification: From the recist
     committee, European journal of cancer 62 (2016) 132–137.