Vol.2, No.2, 137-146 (2010)
doi:10.4236/health.2010.22021
SciRes
Copyright © 2010 Openly accessible at http://www.scirp.org/journal/HEALTH/
Health
Cellular responding kinetics based on a model of gene
regulatory networks under radiotherapy
Jin-Peng Qi1,3*, Yong-Sheng Ding1,2,3*, Shi-Huang Shao1, Xian-Hui Zeng1, Kuo-Chen Chou1,2
1College of Information Sciences and Technology, Donghua University, Shanghai, China
2Gordon Life Science Institute, San Diego, USA; qipengkai@126.com, ysding@dhu.edu.cn
3Engineering Research Center of Digitized Textile & Fashion Technology, Ministry of Education, Donghua University, Shanghai,
China
Received 16 November 2009; revised 3 December 2009; accepted 8 December 2009.
ABSTRACT
Radiotherapy can cause DNA damage into cells,
triggering the cell cycle arrest and cell apop-
tosis through complicated interactions among
vital genes and their signal pathways. In order to
in-depth study the complicated cellular res-
ponses under such a circumstance, a novel mo-
del for P53 stress response networks is pro-
posed. It can be successfully used to simulate
the dynamic processes of DNA damage trans-
ferring, ATM and ARF activation, regulations of
P53-MDM2 feedback loop, as well as the toxins
degradation. Particularly, it has become feasible
to predict the outcomes of cellular response in
fighting against genome stresses. Consequently,
the new model has provided a reasonable
framework for analyzing the complicated regu-
lations of P53 stress response networks, as well
as investigating the mechanisms of the cellular
self-defense under radiotherapy.
Keywords: P53; MDM2; DNA Damage; IR;
Oscillations; Radiotherapy
1. INTRODUCTION
Like immunotherapy, chemotherapy, and surgery, radio-
therapy is one of the major tools in fighting against cancer.
As acute IR is applied, cell can trigger its self-defensive
mechanisms in response to genome stresses [1]. As one of
the pivotal anticancer genes within the cell, P53 can
control the transcription and translation of series genes,
and trigger cell cycle arrest and apoptosis through inter-
action with downstream genes and their complicated
signal pathways [2]. Under radiotherapy, the outcomes of
cellular response depend on the presence of functional
P53 proteins to induce tumor regression through apop-
totic pathways [3]. Conversely, the P53 tumor suppressor
is the most commonly known specific target of mutation
in tumorigenesis [4]. Abnormalities in the P53 have been
identified in over 60% of human cancers and the status of
P53 within tumor cells has been proposed to be one of the
determinant response to anticancer therapies [3,4]. Con-
trolled radiotherapy studies show the existence of a strong
biologic basis for considering P53 status as a radiation
predictor [3,5]. Therefore, the status of P53 in tumor cell
can be considered as a predictor for long-term bio-
chemical control during and after radiotherapy [6-8].
Recently, several models have been proposed to ex-
plain the damped oscillations of P53 in cell populations
[9-12]. However, the dynamic mechanism of the sin-
gle-cell responses is not completely clear yet, and the
complicated regulations among genes and their signal
pathways need to be further addressed, particularly under
the condition of acute IR.
Many studies have indicated that introducing novel
mathematical and computational approaches can stimu-
late in-depth investigation into various complicated bio-
logical systems (see, e.g., [13-23]). These methods have
provided useful tools for both basic research and drug
development [24-33], helping understanding many mar-
velous action mechanisms in various biomacromolecular
systems (see, e.g., [21, 34-39]).
Based on the existing models [9-12] and inspired by
the aforementioned mathematical and computational
approaches in studying biological systems, here a new
model is proposed for studying the P53 stress response
networks under radiotherapy at the cellular level, along
with the kinetics of DNA double-strand breaks (DSBs)
generation and repair, ATM and ARF activation, as well
as the regulating oscillations of P53-MDM2 feedback
loop (MDM2 is an important negative regulator of the
p53 tumor suppressor). Furthermore, the kinetics of the
oncogenes degradation, as well as the eliminations of the
mutation of P53 (mP53) and the toxins were presented.
Also, the plausible outcomes of cellular response were
J. P. Qi et al. / HEALTH 2 (2010) 137-146
SciRes Copyright © 2010 http://www.scirp.org/journal/HEALTH/
138
Openly accessible at
DSBs generation
and repair
ATM activation
IR
P53-MDM2
feedback loop
Oncogenes
over-expression ARF activation
Depression of abnormal
genes and cellular toxins
Figure 1. Illustration showing the integrated model of P53 stress response networks under ra-
diotherapy. It is composed of three modules, including DNA damage generation and repair, ATM
and ARF activation, as well as P53-MDM2 feedback loop. As acute IR is applied, ARF is acti-
vated by the over-expression of oncogenes, and ATM is activated with the cooperation of DSBCs
and ARF*. ATM* and ARF* corporately trigger the responding mechanism of P53-MDM2
feedback loop.
D
1
k
dc1
k
dc2
k
cf2
k
fw1
k
fw2
k
cd1
k
cd2
DSBs Fixed
D
2
C
1
C
2
F
r
F
w
DSBCs k
cf1
IR dose
DSBC transferring to
downstream
g
enes
Figure 2. Illustration showing the module of DNA repair process. It includes both a fast repair
pathway and a slow one. DSB can be in one of four states: intact DSB (DSB), DBSC, Fr and Fw.
Subscripts ‘1’ and ‘2’ refer to the fast kinetics and slow one.
analyzed under different IR dose domains.
It is instructive to mention that using differential
equations and graphic approaches to study various dy-
namical and kinetic processes of biological systems can
provide useful insights, as indicated by many previous
studies on a series of important biological topics, such as
enzyme-catalyzed reactions [18,40], low-frequency in-
ternal motions of biomacromolecules [41-46], protein
folding kinetics [47,48], analysis of codon usage [49,50],
base distribution in the anti-sense strands [51], hepatitis B
viral infections [52], HBV virus gene missense mutation
[53], GPCR type prediction [54], protein subcellular
location prediction [55], and visual analysis of SARS-
CoV [7,56].
2. METHODS
2.1. Model Review
Under the genome stresses, many efforts have been made
to enhance P53-mediated transcription through some
models [9-12,58,59]. However, the interactions in a real
system would make these models [60] extremely com-
plicated. Therefore, a new feasible model is needed in
order to incorporate more biochemical information. To
realize this, let us take the following criteria or assump-
tions for the new model: 1) only the vital components and
interactions are taken into account; 2) all the localization
issues are ignored; 3) the simple linear relations are used
to describe the interactions among the components con-
cerned; and 4) there are enough substances to keep the
system ‘‘workable’’ [58].
The new integrated model thus established for the P53
stress response networks under radiotherapy is illustrated
in Figure 1. Compared with the previous models [9-12],
the current model contains more vital components, such
as oncogenes, ARF and mP53, as well as their related
regulating pathways. In the DSBs generation and repair
module, the acute IR induces DSBs stochastically and
forms DSB-protein complexes (DSBCs) at each of the
J. P. Qi et al. / HEALTH 2 (2010) 137-146
SciRes Copyright © 2010 http://www.scirp.org/journal/HEALTH/Openly accessible at
139
damage sites after interacting with the DNA repair pro-
teins [2,3]. As a sensor of genome stress, ATM is acti-
vated by the DSBCs signal transferred from DSBs.
Meanwhile, the over-expression of oncogenes prompted
by acute IR can trigger the activation of ARF, further
prompting the ATM activation [2,7]. The cooperating
effects of active ATM (ATM*) and active ARF (ARF*)
switch on or off the P53-MDM2 feedback loop [2,7,9],
further regulating the downstream genes to control the
cell-cycle arrest and the cell apoptosis in response to
genome stresses [8]. Here, we use the superscript * to
represent the activate state as done in [61].
2.2. DSBs Generation and Repair
Under the continuous effect of acute IR dose, DSBs occur
and trigger two major repair mechanisms in eukaryotic
cells: homologous recombination (HR) and nonhomolo-
gous end joining (NHEJ) [62,63]. About 60-80% of DSBs
are rejoined quickly, whereas the remaining 20-40% of
DSBs are rejoined more slowly [64,65]. As shown in
Figure 2, the module of DSBs generation and repair
process contains both the fast and slow kinetics, with each
being composed of a reversible binding of repair proteins
and DSB lesions into DSBCs, and an irreversible process
from the DSBCs to the fixed DSBs [62,65]. DSBCs are
synthesized by binding the resulting DSBs with repair
proteins (RP), which is the main signal source to transfer
the DNA damage to P53-MDM2 feedback loop by ATM
activation [2].
Due to the misrepair part of DSBs (Fw) having the
profound consequences on the subsequent cellular vi-
ability and the cellular response in fighting against ge-
nome stresses [1,3], we obviously distinguish between
correct repair part of DSBs (Fr) and Fw [9,10,12].
Moreover, we further deal the total Fw in both repair
processes as a part of toxins within the cell [2,4,11],
which can be eliminated by the regulatory functions of
P53 during and after radiotherapy, and treated as an in-
dicator of outcomes in cellular response to genome
stresses [2].
Some experimental data suggest that the quantity of the
resulting DSBs within different IR dose domains obey a
Poisson distribution [11]. In accordance with the ex-
periments, we assume that the stochastic number of the
resulting DSBs per time scale is proportional to the
number generated by a Poisson random function during
the period of acute radiation [11]. The DSBs generation
process is formulated as follows:
[DT] Poissrnd( IR)
tir
dka
dt
 
(1)
where [DT] is the concentration of total resulting DSBs
induced by IR in both fast and slow repair processes. kt is
the parameter to set the number of DSBs per time scale,
and air is the parameter to set the number of DSBs per IR
dose.
Moreover, we assume that the limited repair proteins
are available around DSBs sites, and 70% of the initial
DSBs are fixed by the fast repair process. Each DSB can
be in one of the four states: intact DSB, DSBC, Fr and Fw
[9,10,12]. Thus, we have the following differential equa-
tions:
1
11
[D ][D ][C ]
tcd
dak
dt

1
dc1 1cross12
[RP]([D ]([D ][D]))kk

(2)
2
2t cd22
[D ][D ][C]
dak
dt

dc2 2cross12
[RP]([D ]([D][D])kk

(3)
1
dc11 cd11 cf11
[C ][D ][C ][C ]
dkkk
dt

(4)
2
dc2 2cd22cf22
[] [D ][C ][C ]
dC kkk
dt

(5)
pcd1 1cd22
[RP] [C ][C]
r
dSk k
dt
 
dc1 1dc22cross12
[RP]([D][D]([D ][D]))kk k
 
(6)
w
w1 1fw22
[F ][C ][C]
f
dkk
dt

(7)
where [D], [C], and [Fw] represent the concentrations of
DSBs, DSBCs, and Fw in the fast and the slow repair
kinetics respectively, kdc, kcd, kcf, and kfw are the tran-
sition rates among the above three states; kdc, and kcross
represent the first-order and second-order rate constants in
both the fast and the slow repair kinetics respectively [65].
Srp is the basal induction rate of repair mRNA, and sub-
scripts ‘1’ and ‘2’ refer to the fast and the slow kinetics.
2.3. ATM and ARF Activation
As a DNA damage detector, ATM exists as a dimer in
unstressed cells. After IR is applied, intermolecular
autophosphorylation occurs, causing the dimer to disso-
ciate rapidly into the active monomers. The active ATM
monomer (ATM*) can prompt the P53 expression further
[64]. Meanwhile, ARF, another tumor suppressor, is ac-
tivated by hyperproliferative signals emanating from
oncogenes, such as Ras, c-myc etc., further prompting the
ATM activation [2, 7, 10]. Based on the existing model of
ATM switch [11], we present an ATM and ARF activation
module under IR. Shown in Figure 3 is the module
scheme of ATM and ARF activation, which includes five
components: ATM dimer, inactive ATM monomer, ATM*,
ARF, and ARF*. Compared with the previous studies in
J. P. Qi et al. / HEALTH 2 (2010) 137-146
SciRes Copyright © 2010 http://www.scirp.org/journal/HEALTH/
140
Openly accessible at
k
af
AT M
D
AT M
AT
k
undim
C
DSBCs
ARF ARF
*
Oncogenes over-expression
k
onf
F
(
C, [ARF*], [ATM*]
)
k
ar
k
dim
Figure 3. Illustration showing the module scheme of ATM and ARF activation under
constant IR. ARF is activated by the over-expression of oncogenes induced by acute
IR, and ATM is activated from ATM monomers under the cooperating effects of
DSBCs, ARF*, and self-feedback of ATM*.
[9-12], ARF, oncogenes, and the related signal pathways
are involved in this module [2,7]. Here, let us assume that
DSBCs is the main signal transduction from DSBs to
P53-MDM2 feedback loop through ATM activation, and
the rate of ATM activation is a function of the amount of
DSBCs, ARF* and the self-feedback of ATM*. Fur-
thermore, the total concentration of ATM is a constant,
including ATM dimer, ATM monomer and ATM, as
treated in {Ma, 2005 #1194].
As a detector of DNA damage, ATM activation plays
an important role in triggering the regulatory mechanisms
of P53 stress response networks [2,65]. After the acute IR
is applied, phosphorylation of inactive ATM monomers is
promoted first by DSBCs and then rapidly by means of
the positive feedback from ATM*, accounting for the
intermolecular autophosphorylation [11]. Meanwhile,
under the circumstance of continuous IR dose, ARF, a
detector of over-expression of oncogenes is activated by
hyperproliferative signals emanating from oncogenes,
further prompting the ATM activation [2,7,10], as can be
formulated as follows:
2
d
dim mdim d
[ATM ]1[ATM][ATM ]
2un
dkk
dt

(8)
2
m
undimd dimm
[ATM ]2[ATM] [ATM
dkk
dt
]
afmar
[ATM ][ATM*]kf k
(9)
af m
[ATM*] [ATM ][ATM*]
ar
dkf k
dt

(10)
arf adonf
[ARF] [ARF] [Onco][ARF]
dSk k
dt
 
(11)
onf pad
[ARF*] [Onco][ARF] [ARF*]
dkk
dt

(12)
12 3
(, [ATM*])[ATM*][ATM*]fCaC aaC

*][
4ARFa
(13)
where [ATMd], [ATM] and [ATM*] represent the concen-
trations of ATM dimer, ATM monomer, and active ATM
monomer respectively; [Onco], [ARF] and [ARF*] repre-
sent the concentrations of oncogenes, ARF, and active
ARF respectively; kundim, kdim, kar, and kaf are the rates
of ATM undimerization, ATM dimerization, ATM
monomer inactivation, and ATM monomer activation,
respectively. Sarf, konf, kad and kpad are the rates of ARF
basal induction, ARF activation triggered by Oncogenes,
ARF degradation, and ARF* degradation, respetively. In
addition, f is the function of ATM activation, the term a1C
implies the fact that DSBs somehow activate ATM mole-
cules at a distance, a2 [ATM*] indicates the mechanism of
autophosphorylation of ATM, a3C [ATM*] represents the
interaction between the DSBCs and ATM* [9-12,66], and
a4 [ARF*] represents the regulating function of ARF* to
ATM activation [1,3,7].
2.4. Regulation of P53-MDM2 Feedback Loop
As shown in Figure 4, P53 and its principal antagonist,
MDM2 transactivated by P53, form a P53-MDM2 feed-
back loop, which is the core part in the integrated networks
[9-12]. ATM* elevates the transcriptional activity of P53
by prompting phosphorylation of P53 and degradation of
MDM2 protein [67]. Also, ARF* can indirectly prompt the
transcriptional activity of P53 by inhibiting the expression
of MDM2 and preventing P53 degradation [2,7,9]. With
the cooperating regulations of ATM* and ARF*, this
negative feedback loop can produce oscillations in re-
sponse to the sufficiently strong IR dose [11].
Especially, the mutation of P53 (mP53) triggered by
oncogenes is added in this module, and mP53 is further
dealt as another detector of outcomes in cellular response
J. P. Qi et al. / HEALTH 2 (2010) 137-146
SciRes Copyright © 2010 http://www.scirp.org/journal/HEALTH/
141
Openly accessible at
Onco
P53
D
P53
R
P53
P
P53
*
mP53
R
MDM2
D
MDM2
R
MDM2
P
Toxin
ATM
*
ARF
*
S
P53
S
MDM2
Figure 4. The directed graph of P53-MDM2 feedback loop under radiotherapy. P53 is translated from
P53mRNA and phosphorylated by ATM* and ARF*. MDM2 protein promotes a fast degradation of
P53 protein and a slow degradation of P53*. In addition, ATM*and ARF* stimulate the degradation of
MDM2, and then indirectly increase the regulatory activation of P53* further. Especially, oncogenes,
toxins and mP53 are decreased directly by the regulatory functions of P53*.
to acute IR. To account for a decreased binding affinity
between inactive P53 and P53*, we assume that
MDM2-induced degradation of inactive P53 is faster than
that of P53*, and only P53* can induce target genes to
depress the over-expression of oncogenes and further
eliminate the toxins within the cell [3,4,9-12]. The main
differential equations used in this module are as follows:
R
P53rp Rrp R
[P53 ][P53 ][P53 ]
dSdk
dt
 
(14)
P
rp Rp*ppp P
[P53 ][P53][P53*][P53 ]
dkk d
dt
 
P
app*mp P
Pp Pd
[P53 ][P53 ]
[ATM*][MDM2 ]
[P53 ]+[P53 ]+
kk
k

P
k
(15)
P
app*p*p pp*
Pp
[P53]
[P53*] [ATM*][P53*] [P53*]
[P53]+
dkk
dt k
d
p*ppp*mp* P
d*
[P53*]
[P53*][P53*][MDM2][P53*]+
kdk k
 
(16)
n
R
mdm2 p*mnn
[MDM2 ][P53*]
+[P53*] +
dSk
dt k
mrpR mrR
[MDM2 ][MDM2]kd
(17)
P
mrp Rmp P
[MDM2 ][MDM2][MDM2 ]
dkd
dt

mat marP
at ar
[ATM*] [ARF*]
(
[ATM*]+ [ARF*]+
kk
kk
)[MDM2]
(18)
onIR onp
[Onco] [Onco][IR] [Onco][P53*]
dkk
dt

(19)
tfw wpt
[Toxin ][F][P53*][Toxins]
ds
kk
dt

(20)
mp Rpmd P*
[mP53] [P53 ][Onco][P53][mP53]
dkk
dt

(21)
where [P53R], [P53P], [P53*], [MDM2R], and [MDM2P]
represent the concentrations of P53 mRNA, P53 protein,
active P53, MDM2 mRNA, and MDM2 protein, respec-
tively; [Onco], [Toxins], and [mP53] represent the con-
centrations of oncogenes, Fw and mP53, respectively.
SP53, and SMDM2 represent the basal induction rates of
P53 mRNA and MDM2 mRNA, respectively; k, and d
represent the regulation and degradation rates among
genes and proteins, respectively. The other parameters are
presented in Tables 1-3.
3. RESULTS AND DISCUSSIONS
3.1. Kinetics of DSBCs Synthesizing
During the simulation process, the continuous 2, 5, and
7Gy IR are applied into a cell respectively. As shown in
Figure 5(a), owing to the condition that many DSBs
occur and the limited RP are available around damage
sites, the concentration of RP begins to decrease as IR
dose overtakes 5Gy, and trends to zero versus radiation
time. Meanwhile, the kinetics of DSBCs synthe sizing is
shown in Figure 5(b). We can see that the rates of DSBCs
synthesis keep increasing under 2, and 5Gy IR, whereas,
it begins to decrease and trend to constant after about
120min under 7Gy IR dose.
3.2. Kinetics of ARF and ATM Activation
The ARF activation is used to describe the mechanisms in
cellular response to the over-expression of oncogenes
J. P. Qi et al. / HEALTH 2 (2010) 137-146
SciRes Copyright © 2010 http://www.scirp.org/journal/HEALTH/
142
Openly accessible at
020406080100120140160180200
0
2
4
6
8
10
12
14
16
18
the dynamic traces of RP avaliable around DSBs under 2,5,7Gy, respectively
(a) Radiation Time(Min)
Concentration
under IR=2Gy
under IR=5Gy
under IR=7Gy
(a)
050 100 150 200 250 300
0
2
4
6
8
10
12
14
16
18
the kinetics of DSBCs synthesizing under 2,5,7Gy IR, respectively
(b) Radiation Time(Min)
Concentration
under IR=2Gy
under IR=5Gy
under IR=7Gy
Under 7Gy IR, the rate of DSBCs
synthesizing begin to decrease
after this plausible time threshold
(b)
Figure 5. The kinetics of DSBs repairing and transferring under
continuous effect of 2, 5, 7Gy IR. (a) The dynamics of RP
available around the resulting DSBs under different IR dose
domains. (b) The kinetics of DSBCs synthesized by DSBs and
RP versus continuous radiation time under different IR dose
domains.
induced by acute IR [2,7]. The kinetics of ARF activation
is shown in Figure 6(a). Owing to the over-expression of
oncogenes without depressing functions of P53*, ARF is
activated fast and ARF* keeps increasing followed by
trending to dynamic equilibrium versus radiation time.
Meanwhile, the ATM activation module was estab-
lished to describe the switch-like dynamics of the ATM
activation in response to DSBCs increasing, and the
regulation mechanisms during the process of the ATM
transferring DNA damage signals to the P53-MDM2
feedback loop. Under the cooperative function of DSBCs,
ARF*, and the positive self-feedback of ATM*, the ATM
would reach the equilibrium state within minutes due to
the fast phosphorylation [2,11,67]. Kinetics of ATM ac-
tivation is shown in Figure 6(b). ATM is activated rap-
idly and switches to “on” state with respective rates, and
then trends to the saturation state. The step- like traces
suggest that the ATM module can produce an on-off
switching signal, and transfer the damage signal to the
P53-MDM2 feedback loop [3]. Furthermore, under the
cooperation effects of ATM* and ARF*, DNA damage
0100 200300 400 500 600 700800
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
the kinetics of ARF activation under different IR dose domains
(a) Radiation Time(Min)
Concentration
under IR=2Gy
under IR=5Gy
under IR=7Gy
(a)
0100 200300 400 500 600 700 800
0
0.2
0.4
0.6
0.8
1
Concent ration
(b) Radiation Time(Min)
the kinetics of ATM activation under different IR dose domains
under IR=2Gy
under IR=5Gy
under IR=7Gy
the rate of ATM activation begin to decrease
after this plausible time threshold under 7Gy IR
(b)
Figure 6. The kinetics of ARF and ATM activation under 2, 5,
7Gy IR. (a) The kinetics of ARF activation in response to
over-expression of oncogenes induced by different IR dose. (b)
The switch-like kinetics of ATM activation, ATM* reach satu-
ration and trend to constant state in response to continuous
radiation time of different IR dose domains.
signals can be further transferred to the downstream
genes and their signal pathways more efficiently [2,7].
3.3. Outcomes of Cellular Responding
Radiotherapy
The P53-MDM2 feedback loop is a vital part in control-
ling the downstream genes and regulation pathways to-
fight against the genome stresses [6,67,68]. In response to
the input signal of ATM* and ARF*, the P53-MDM2
module generates one or more oscillations. The response
traces of P53 and MDM2 protein under continuous appli-
cation of 2, 5, and 7Gy IR from time 0 are shown in Figure
7(a). Upon the activation by ATM*, ARF* and decreased
degradation by MDM2, the total amount of P53 proteins
increases quickly. Due to the P53-dependent induction of
MDM2 transcription, the increase of MDM2 proteins is
sufficiently large to lower the P53 level, which in turn
reduces the amount of the MDM2 proteins.
J. P. Qi et al. / HEALTH 2 (2010) 137-146
SciRes Copyright © 2010 http://www.scirp.org/journal/HEALTH/Openly accessible at
143
The oscillation pulses shown in Figure 7(a) have a pe-
riod of 400 min, and the phase difference between P53
and MDM2 is about 100 min. Moreover, the first pulse is
slightly higher than the second, quite consistent with the
experimental observations [2,7,11] as well as the previous
simulation results [9,10,12,69].
Also, by comparing these simulation results, we can
see that the strength and swing of these oscillations begin
to decrease as IR overtakes 7Gy, suggesting that the
ability of cellular responding genome stresses begin to
0100 200300 400 500 600 700800
0
0.5
1
1.5
2
2.5
3
3.5
4
4.5
5
the kinetics of Fw elimination with the regulating functions of P53* under different IR
(b) Radiation Time(Min)
Concentration
Fw under IR=2Gy
Fw under IR=5Gy
Fw under IR=7Gy
(a)
0100 200 300 400 500600 700 800
0
0. 1
0. 2
0. 3
0. 4
0. 5
0. 6
0. 7
0. 8
0. 9
the oscillating kinetics of P53* and MDM2 in response to different IR dose domains
(a) Radiation Time(Min)
Conce ntratio n
P53* under IR=2Gy
P53* under IR=5Gy
P53* under IR=7Gy
MDM2 under IR=2Gy
MDM2 under IR=5Gy
MDM2 under IR=7Gy
(b)
0100 200 300 400 500600 700 800
0.1 5
0.1 6
0.1 7
0.1 8
0.1 9
0.2
0.2 1
0.2 2
0.2 3
0.2 4
0.2 5
the kinetics of Oncogenes depressed by the functions of P53* in response to different IR
(c) Radiation Time(Min)
Concentration
Onco under IR=2Gy
Onco under IR=5Gy
Onco under IR=7Gy
(c)
0100 200 300 400 500 600 700 800
0.48
0.5
0.52
0.54
0.56
0.58
0.6
0.62
0.64
0.66
the kinetics of mP53 degradated by P53* under different IR dose domains
(d) Radiation Time(Min)
Concentration
mP53 under IR=2Gy
mP53 under IR=5Gy
mP53 under IR=7Gy
(d)
Figure 7. The outcomes of cellular responding 2, 5, 7Gy IR
under radiotherapy. (a) The oscillating kinetics of P53* and
MDM2 in response to the cooperative effect of ATM* and
ARF* under different IR dose domains; (b) The kinetics of
toxins elimination triggered by the functions of P53*; (c) The
depressing dynamics of oncogenes over-expression with the
regulations of P53*; (d) The kinetics of mP53 elimination
triggered by the effect of P53*.
decrease as IR dose exceeds a certain threshold.
Furthermore, because in the current model the toxins,
mP53 and oncogenes can be degraded directly by P53* in
this module, we can plot the predictable outcomes of
cellular response in fighting against genome stresses
under different IR dose domains. As shown in Figure
7(b), Fw remaining within the cell keeps decreasing with
respective rate, and trends to zero versus continuous
radiation time under 2 and 5Gy IR. Whereas, when IR
exceeds 7Gy, Fw begins to increase slightly with some
oscillations. Also, the kinetics of oncogenes degrading is
plotted in Figure 7(c). As we can see, owing to the nega-
tive regulations of P53*, the expression level of onco-
genes keeps decreasing after the first climate under 2 and
5Gy IR dose, and then begins to increase slowly under
7Gy IR dose. Meanwhile, as shown in Figure 7(d), quite
similar to the results in Figure 7(b) and Figure 7(c),
mP53 keeps decrease after reaching the first maximum
under 2 and 5Gy IR dose, and then begins to increase
slowly under 7Gy IR dose. All these results obtained by
the above simulations based on the new model indicate
that that P53* indeed acts an important role in regulating
downstream genes and their signal pathways, whereas its
capabilities in cellular responding DNA damage under
radiotherapy begin to decrease as the strength of IR ex-
ceeds a certain maximal threshold.
4. CONCLUSIONS
A new model was proposed to simulate the P53 stress
response network under radiotherapy. It is demonstrated
according to our model that ATM and ARF exhibits a
strong sensitivity and switch-like behavior in response to
J. P. Qi et al. / HEALTH 2 (2010) 137-146
SciRes Copyright © 2010 http://www.scirp.org/journal/HEALTH/
144
Openly accessible at
the number of DSBs, fully consistent with the experi-
mental observations. Interestingly, it is shown in this
study that after the DNA damage signals transferring,
P53-MDM2 feedback loop will produce oscillations, then
triggering the cellular self-defense mechanisms to de-
grade the toxins remaining within the cell, such as Fw,
oncogenes, and mP53. Particularly, under different IR
dose domains, the new model can reasonably predict
outcomes of cellular response in fighting against genome
stresses, and hence providing a framework for analyzing
the complicated regulations of P53 stress response net-
works, as well as the mechanisms of the cellular self-
defense under radiotherapy.
5. ACKNOWLEDGMENT
The current work was supported in part by Specialized Research Fund
for Natural Science Foundation of Shanghai (No.10ZR1401600), the
Doctoral Program of Higher Education from Ministry of Education of
China (No.20060255006), Project of the Shanghai Committee of Sci-
ence and Technology (No. 08JC1400100), and the Open Fund from the
Key Laboratory of MICCAI of Shanghai (06dz22103).
REFERENCES
[1] Li, L., Story, M. and Legerski, R.J. (2001) Cellular re-
sponses to ionizing radiation damage. Int J Radiat Oncol
Biol Phys, 49, 1157-1162.
[2] Kohn, K.W. and Pommier, Y. (2005) Molecular interac-
tion map of the p53 and Mdm2 logic elements, which
control the Off-On switch of p53 in response to DNA
damage. Biochem Biophys Res Commun, 331, 816-827.
[3] Perez, C.A. and Purdy, J.A. (1998) Treatment planning in
radiation oncology and impact on outcome of therapy.
Rays, 23, 385-426.
[4] Tjebbes, G.W., Kreijveld, P.A., Tilanus, M.G., Hordijk, G.
J. and Slootweg, P.J. (2002) P53 tumor suppressor gene
mutations in laryngeal cancer and in recurrent disease
following radiation therapy. Oral Oncol, 38, 296-300.
[5] Duchting, W., Ulmer, W. and Ginsberg, T. (1996) Cancer:
a challenge for control theory and computer modelling.
Eur J Cancer, 32A, 1283-1292.
[6] Ritter, M.A., Gilchrist, K.W., Voytovich, M., Chappell, R.
J. and Verhoven, B.M. (2002) The role of p53 in radiation
therapy outcomes for favorable-to-intermediate-risk
prostate cancer. Int J Radiat Oncol Biol Phys, 53,
574-580.
[7] Lindstrom, M.S. and Wiman, K.G. (2003) Myc and E2F1
induce p53 through p14ARF-independent mechanisms in
human fibroblasts. Oncogene, 22, 4993-5005.
[8] Pauklin, S., Kristjuhan, A., Maimets, T. and Jaks, V. (2005)
ARF and ATM/ATR cooperate in p53-mediated apoptosis
upon oncogenic stress. Biochem Biophys Res Commun,
334, 386-394.
[9] Qi, J.P., Shao, S.H., Li, D.D. and Zhou, G.P. (2007) A
dynamic model for the p53 stress response networks under
ion radiation. Amino Acids, 33, 75-83.
[10] Qi, J.P., Shao, S.H., Xie, J. and Zhu, Y. (2007) A
mathemtical model of P53 gene regulatory networks un-
der radiotherapy. Biosystems, 90, 698-706.
[11] Ma, L., Wagner, J., Rice, J.J., Hu, W., Levine, A.J. and
Stolovitzky, G.A. (2005) A plausible model for the digital
response of p53 to DNA damage. Proc Natl Acad Sci USA,
102, 14266-14271.
[12] Qi, J.P., Shao, S.H. and Shen, Y.Z. (2008) Cellular re-
sponding DNA damage: An improved modeling of P53
gene regulatory networks under ion radiation (IR). Ap-
plied Mathematics and Computation, 205, 73-83.
[13] Chou, K.C. and Zhou, G.P. (1982) Role of the protein
outside active site on the diffusion-controlled reaction of
enzyme. Journal of American Chemical Society, 104,
1409-1413.
[14] Chou, K.C. (1988) Review: Low-frequency collective
motion in biomacromolecules and its biological functions.
Biophysical Chemistry, 30, 3-48.
[15] Chou, K.C. (1989) Graphical rules in steady and non-
steady enzyme kinetics. J. Biol. Chem., 264, 12074-
12079.
[16] Chou, K.C. (1990) Review: Applications of graph theory
to enzyme kinetics and protein folding kinetics. Steady
and non-steady state systems. Biophysical Chemistry, 35,
1-24.
[17] Ginsberg, T., Ulmer, W. and Duchting, W. (1993) Com-
puter simulation of fractionated radiotherapy: further re-
sults and their relevance to percutaneous irradiation and
brachytherapy. Strahlenther Onkol, 169, 304-310.
[18] Chou, K.C., Kezdy, F.J. and Reusser, F. (1994) Review:
Steady-state inhibition kinetics of processive nucleic acid
polymerases and nucleases. Analytical Biochemistry, 221,
217-230.
[19] Chou, K.C. (2004) Review: Structural bioinformatics and
its impact to biomedical science. Current Medicinal
Chemistry, 11, 2105-2134.
[20] Chou, K.C. In Structural bioinformatics and its impact to
biomedical science and drug discovery, In: Frontiers in
Medicinal Chemistry. Bentham Science Publishers: The
Netherlands, 3, 2006.
[21] Huang, R.B., Du, Q.S., Wang, C.H. and Chou, K.C. (2008)
An in-depth analysis of the biological functional studies
based on the NMR M2 channel structure of influenza A
virus. Biochem. Biophys. Res. Comm., 377, 1243-1247.
[22] Chou, K.C. and Shen, H.B. (2008) Cell-PLoc: A package
of web-servers for predicting subcellular localization of
proteins in various organisms. Nature Protocols, 3,
153-162.
[23] Wang, J.F., Gong, K., Wei, D.Q., Li, Y.X. and Chou, K.C.
(2009) Molecular dynamics studies on the interactions of
PTP1B with inhibitors: From the first phosphate binding
site to the second one. Protein Engineering Design and
Selection, 22, 349-355.
[24] Althaus, I.W., Chou, J.J., Gonzales, A.J., Diebel, M.R.,
Chou, K.C., Kezdy, F.J., Romero, D.L., Aristoff, P.A.,
Tarpley, W.G. and Reusser, F. (1993) Steady-state kinetic
studies with the non-nucleoside HIV-1 reverse transcrip-
tase inhibitor U-87201E. J. Biol. Chem., 268, 6119-6124.
[25] Althaus, I.W., Gonzales, A.J., Chou, J.J., Diebel, M.R.,
Chou, K.C., Kezdy, F.J., Romero, D.L., Aristoff, P.A.,
Tarpley, W.G. and Reusser, F. (1993) The quinoline U-
78036 is a potent inhibitor of HIV-1 reverse transcriptase.
J. Biol. Chem., 268, 14875-14880.
[26] Althaus, I.W., Chou, J.J., Gonzales, A.J., Diebel, M.R.,
Chou, K.C., Kezdy, F.J., Romero, D.L., Aristoff, P.A.,
Tarpley, W.G. and Reusser, F. (1993) Kinetic studies with
the nonnucleoside HIV-1 reverse transcriptase inhibitor
J. P. Qi et al. / HEALTH 2 (2010) 137-146
SciRes Copyright © 2010 http://www.scirp.org/journal/HEALTH/Openly accessible at
145
U-88204E. Biochemistry, 32, 6548-6554.
[27] Chou, K.C., Wei, D.Q. and Zhong, W.Z. (2003) Binding
mechanism of coronavirus main proteinase with ligands
and its implication to drug design against SARS. (Erratum:
ibid., 2003, Vol.310, 675). Biochem Biophys Res Comm,
308, 148-151.
[28] Du, Q.S., Wang, S.Q., Jiang, Z.Q., Gao, W.N., Li, Y.D.,
Wei, D.Q. and Chou, K.C. (2005) Application of bioin-
formatics in search for cleavable peptides of SARS-CoV
Mpro and chemical modification of octapeptides. Me-
dicinal Chemistry, 1, 209-213.
[29] Zhang, R., Wei, D.Q., Du, Q.S. and Chou, K.C. (2006)
Molecular modeling studies of peptide drug candidates
against SARS. Medicinal Chemistry, 2, 309-314.
[30] Du, Q.S., Sun, H. and Chou, K.C. (2007) Inhibitor design
for SARS coronavirus main protease based on “distorted
key theory”. Medicinal Chemistry, 3, 1-6.
[31] Zheng, H., Wei, D.Q., Zhang, R., Wang, C., Wei, H. and
Chou, K.C. (2007) Screening for new agonists against
Alzheimer's disease. Medicinal Chemistry, 3, 488-493.
[32] Wei, H., Wang, C.H., Du, Q.S., Meng, J. and Chou, K.C.
(2009) Investigation into adamantane-based M2 inhibitors
with FB-QSAR. Medicinal Chemistry, 5, 305-317.
[33] Gong, K., Li, L., Wang, J.F., Cheng, F., Wei, D.Q. and
Chou, K.C. (2009) Binding mechanism of H5N1 influ-
enza virus neuraminidase with ligands and its implication
for drug design. Medicinal Chemistry, 5, 242-249.
[34] Chou, K.C. (1987) The biological functions of low- fre-
quency phonons: 6. A possible dynamic mechanism of
allosteric transition in antibody molecules. Biopolymers,
26, 285-295.
[35] Chou, K.C. (1989) Low-frequency resonance and coop-
erativity of hemoglobin. Trends in Biochemical Sciences,
14, 212.
[36] Schnell, J.R. and Chou, J.J. (2008) Structure and mecha-
nism of the M2 proton channel of influenza A virus. Na-
ture, 451, 591-595.
[37] Gordon, G. (2008) Extrinsic electromagnetic fields, low
frequency (phonon) vibrations, and control of cell func-
tion: a non-linear resonance system. Journal of Biomedi-
cal Science and Engineering (JBiSE), 1, 152-156 (open
accessible at http://www.srpublishing.org/journal/jbise/).
[38] Du, Q.S., Huang, R.B., Wang, C.H., Li, X.M. and Chou, K.
C. (2009) Energetic analysis of the two controversial drug
binding sites of the M2 proton channel in influenza. A
virus Journal of Theoretical Biology, 259, 159-164.
[39] Pielak, R.M., Jason, R., Schnell, J.R. and Chou, J.J. (2009)
Mechanism of drug inhibition and drug resistance of in-
fluenza A M2 channel. Proceedings of National Academy
of Science, USA, 106, 7379-7384.
[40] Chou, K.C. and Jiang, S.P. (1974) Studies on the rate of
diffusion-controlled reactions of enzymes. Scientia Sinica,
17, 664-680.
[41] Chou, K.C. (1984) The biological functions of low- fre-
quency phonons: 3. Helical structures and microenvi-
ronment. Biophysical Journal, 45, 881-890.
[42] Chou, K.C., Kiang, Y.S. (1985) The biological functions
of low-frequency phonons: 5. A phenomenological theory.
Biophysical Chemistry, 22, 219-235.
[43] Zhou, G.P. (1989) Biological functions of soliton and extra
electron motion in DNA structure. Physica Scripta, 40,
698-701.
[44] Chou, K.C., Maggiora, G.M., Mao, B. (1989) Quasi-
continuum models of twist-like and accordion-like low-
frequency motions in DNA. Biophysical Journal, 56,
295-305.
[45] Chou, K.C., Zhang, C.T., Maggiora, G.M. (1994) Solitary
wave dynamics as a mechanism for explaining the internal
motion during microtubule growth. Biopolymers, 34,
143-153.
[46] Sinkala, Z. (2006) Soliton/exciton transport in proteins. J
Theor Biol, 241, 919-927.
[47] Chou, K.C. and Shen, H.B. (2009) FoldRate: A web- server
for predicting protein folding rates from primary sequence.
The Open Bioinformatics Journal, 3, 31-50 (openly acces-
sible at http://www.bentham.org/open/tobioij/).
[48] Shen, H.B., Song, J.N. and Chou, K.C. (2009) Prediction
of protein folding rates from primary sequence by fusing
multiple sequential features. Journal of Biomedical Sci-
ence and Engineering (JBiSE), 2, 136-143 (openly ac-
cessible at http://www.srpublishing.org/journal/jbise/).
[49] Chou, K.C. and Zhang, C.T. (1992) Diagrammatization of
codon usage in 339 HIV proteins and its biological im-
plication AIDS. Research and Human Retroviruses, 8,
1967-1976.
[50] Zhang, C.T. and Chou, K.C. (1994) Analysis of codon
usage in 1562 E. Coli protein coding sequences. Journal
of Molecular Biology, 238, 1-8.
[51] Zhang, C.T. and Chou, K.C. (1996) An analysis of base
frequencies in the anti-sense strands corresponding to the
180 human protein coding sequences. Amino Acids, 10,
253-262.
[52] Xiao, X., Shao, S.H. and Chou, K.C. (2006) A probability
cellular automaton model for hepatitis B viral infections.
Biochem. Biophys. Res. Comm., 342, 605-610.
[53] Xiao, X., Shao, S., Ding, Y., Huang, Z., Chen, X. and
Chou, K.C. (2005) An Application of Gene Comparative
Image for Predicting the Effect on Replication Ratio by
HBV Virus Gene Missense Mutation. Journal of Theo-
retical Biology, 235, 555-565.
[54] Xiao, X., Wang, P. and Chou, K.C. (2009) GPCR-CA: A
cellular automaton image approach for predicting G-
protein-coupled receptor functional classes. Journal of
Computational Chemistry, 30, 1414-1423.
[55] Xiao, X., Shao, S.H., Ding, Y.S., Huang, Z.D. and Chou,
K.C. (2006) Using cellular automata images and pseudo
amino acid composition to predict protein subcellular lo-
cation. Amino Acids, 30, 49-54.
[56] Wang, M., Yao, J.S., Huang, Z.D., Xu, Z.J., Liu, G.P.,
Zhao, H.Y., Wang, X.Y., Yang, J., Zhu, Y.S. and Chou, K.
C. (2005) A new nucleotide-composition based finger-
print of SARS-CoV with visualization analysis. Medicinal
Chemistry, 1, 39-47.
[57] Gao, L., Ding, Y.S., Dai, H., Shao, S.H., Huang, Z.D. and
Chou, K.C. (2006) A novel fingerprint map for detecting
SARS-CoV. Journal of Pharmaceutical and Biomedical
Analysis, 41, 246-250.
[58] Tyson, J.J. (1999) Models of cell cycle control in eu-
karyotes. J Biotechnol, 71, 239-244.
[59] Tyson, J.J. and Novak, B. (2001) Regulation of the eu-
karyotic cell cycle: molecular antagonism, hysteresis, and
irreversible transitions. J Theor Biol, 210, 249-263.
[60] Magne, N., Toillon, R.A., Bottero, V., Didelot, C., Houtte,
P.V., Gerard, J.P. and Peyron, J.F. (2006) NF-kappaB
modulation and ionizing radiation: mechanisms and future
directions for cancer treatment. Cancer Lett, 231,
J. P. Qi et al. / HEALTH 2 (2010) 137-146
SciRes Copyright © 2010 http://www.scirp.org/journal/HEALTH/Openly accessible at
146
158-168.
[61] Chou, K.C., Watenpaugh, K.D. and Heinrikson, R.L.
(1999) A Model of the complex between cyclin- de-
pendent kinase 5 (Cdk5) and the activation domain of
neuronal Cdk5 activator. Biochemical & Biophysical Re-
search Communications, 259, 420-428.
[62] Rapp, A. and Greulich, K.O. (2004) After double-strand
break induction by UV-A, homologous recombination and
nonhomologous end joining cooperate at the same DSB if
both systems are available. J Cell Sci, 117, 4935-4945.
[63] Rothkamm, K., Kruger, I., Thompson, L.H. and Lobrich,
M. (2003) Pathways of DNA double-strand break repair
during the mammalian cell cycle. Mol Cell Biol, 23,
5706-5715.
[64] Budman, J. and Chu, G. (2005) Processing of DNA for
nonhomologous end-joining by cell-free extract. EMBO J,
24, 849-860.
[65] Daboussi, F., Dumay, A., Delacote, F. and Lopez, B.S.
(2002) DNA double-strand break repair signalling: the
case of RAD51 post-translational regulation. Cell Signal,
14, 969-975.
[66] Oren, M. (2003) Decision making by p53: Life, death and
cancer. Cell Death Differ, 10, 431-442.
[67] Lev Bar-Or, R., Maya, R., Segel, L.A., Alon, U., Levine,
A.J. and Oren, M. (2000) Generation of oscillations by the
p53-Mdm2 feedback loop: a theoretical and experimental
study. Proc Natl Acad Sci U S A, 97, 11250-11255.
[68] Weller, M. (199) Predicting response to cancer chemo-
therapy: the role of p53. Cell Tissue Res, 292, 435-445.
[69] Lahav, G., Rosenfeld, N., Sigal, A., Geva-Zatorsky, N.,
Levine, A.J., Elowitz, M.B. and Alon, U. (2004) Dynam-
ics of the p53-Mdm2 feedback loop in individual cells.
Nat Genet, 36, 147-150.