Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Physica A 391 (2012) 5997–6007

Contents lists available at SciVerse ScienceDirect

Physica A
journal homepage: www.elsevier.com/locate/physa

Irreversibility in biophysical and biochemical engineering


Umberto Lucia
Dipartimento Energia, Politecnico di Torino, Corso Duca degli Abruzzi 24, 10129 Torino, Italy

article info abstract


Article history: The thermodynamic analysis of open systems is fundamental in engineering. For the
Received 7 March 2012 open systems at their steady state, two apparently opposed principles for the rate of
Received in revised form 11 June 2012 entropy production have been used: the minimum entropy production rate derived by
Available online 10 July 2012
Prigogine, used in the description of various processes in physics, chemistry and biology,
and the maximum entropy production, used in many other cases and now considered more
Keywords:
general. Both principles involve an extreme value of the rate of entropy production in an
Constructal law
Dissipative systems
open system at the steady state under non-equilibrium conditions. In this paper, a link
Entropy between these two approaches is developed and their synthesis with the constructal law is
Irreversibility proposed. An application to ATP synthesis in anaerobic fermentation for biogas production
is presented.
© 2012 Elsevier B.V. All rights reserved.

1. Introduction

Biology has been defined as both the ‘physics of the twenty-first century’, because it presents open problems on the
possible link between the grand unified goals of physics to the comprehension of life and its mechanisms by a mathematical
and formal approach [1], and ‘the computer science of the twenty-first century’, because it seeks to understand the logical
structure of life by directed rules of assembly and interactions among fundamental constituents [1].
The biochemical reactions produce or consume external metabolites, accumulated inside the system, and they connect
internal metabolites, in constant concentrations in the cells at their steady states [2].
In relation to the scaling law a hierarchical hypothesis has been introduced; indeed, fractal-like branching networks
seems to evolve by natural selection in order to minimize power loss when delivering resources to the cells of the body [1].
Regarding the GRNs, cell differentiation was related to the regulation, elucidating the molecular mechanisms for individual
genes and providing models of ontogenetic patterning [3,4] and the link between the structure of GRNs and the patterns of
phenotypic evolution.
In biology the lowest levels have direct impact on the highest levels and vice versa [1], but, from the experimental results,
Krakauer et al. [1] have pointed out that regularities exist only at an aggregate level of description. Levin et al. [5] have
suggested that one of the central issues consists in understanding how detail at one scale makes its signature felt at other
scales, and how to relate phenomena across scales.
In living systems communication is fundamental in relation to autopoietic processes because it is a prerequisite in order
to obtain stationary states [6]; at cellular level, communication is established by chemical reactions.
In biology, the spatial and the temporal scales interact [1], so the evolution of the biological system is represented by the
endpoints on several paths along which basic physical symmetries have been broken. A cell with optimal performance is
expected as a result of an evolutionary process of selection in an environment [7–9]: the cells reach their optimality by means
of a redistribution of the flux pattern through the metabolic network using the pattern of catalytic and regulatory proteins;
indeed, mutations and genetic rearrangements are fundamental for an organism to adapt to environmental conditions [10].

E-mail address: umberto.lucia@polito.it.

0378-4371/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
doi:10.1016/j.physa.2012.07.018
5998 U. Lucia / Physica A 391 (2012) 5997–6007

The previous comments suggest that a global approach, an engineering and statistical thermodynamic, could be useful
in obtaining a theory which allows us to understand the fundamental topics of biology. The topics which would be analysed
are:
1. metabolic analysis, sensitivity analysis and bifurcation analysis: a system-level understanding of native biological
systems with not only their structures but also their dynamics [11], finding also a general principle to obtain the
trajectories of the evolution of the systems in phase space [1].
2. cell analysis: a system-level understanding of pathology and malfunction in order to control the state from the cell to the
whole body [12,13];
3. biotechnology: a system-level approach to design biological systems [14].
The approach to life is developed by analysing the mechanisms in cells and organisms. Many of the mechanisms of individual
sequences and proteins and of inheritance and of development are universal [15].
Systems biology is considered the most powerful theoretical approach to biology systems [16,17]; indeed, L. Von
Bertalanffy applied general systems theory to biology because problems of organization cannot be studied by an analysis of
their singular parts in isolation, N. Weiner introduced the system approach to develop the control and communication in
the animals, while A.M. Turing used the system theory in the chemical analysis of morphogenesis [13].
For an open system at a steady state, two apparently opposed principles for the rate of entropy production have been
proposed: the minimum entropy production rate derived by Prigogine [18], used to the description of various processes in
physics, chemistry and biology, and the maximum entropy production, used in many other cases and now considered more
general [19]. Both principles involve an extreme value of the rate of entropy production in an open system at the steady
state under non-equilibrium conditions.
In this paper, an explanation for the maximum and minimum principles is suggested in relation the evolution in time or
space of the system. An application to ATP synthesis in anaerobic fermentation for biogas production is presented.

2. Entropy in biochemical systems

Life is an integrated process which involves nested living systems (complex systems of basic materials, cells, organisms,
ecosystems) and their environments: there exists a continuous interaction between living systems and environment both
in time and in space [6]. Consequently, living systems have been defined as open systems with life-processes, that consist
in capturing exergy, in exergy storage [20], in entropy exchange, in autopoietic processes [6].
From the results obtained in the analysis of living systems [6], it follows that the living systems can only exist:
1. where there are possibilities to convert an exergy source to entropy;
2. in a state far from thermodynamic equilibrium;
3. as nested, consisting of subsystems;
and life is an organizational process, a result of system cooperation between components, with an interconnection between
sub-systems and super-systems, such that for survival the super-system must export as much or more entropy products
than its sub-systems produces, towards maximum conversion of available exergy sources to entropy products, with the
consequence to obtain a hierarchical structure [6,21]. Then, the living system must exchange exergy and matter through its
boundary, resulting a far-from-equilibrium dissipative thermodynamics system. Knowledge of the steady state metabolic
flux distribution can be obtained through Elementary Mode Analysis (EMA), a theoretical approach useful to decompose the
overall metabolism of a cell into a set of unique, indivisible pathway fluxes called elementary modes [10,22,23]. Prigogine
et al. [24] introduced the dissipative structures to explain biochemical oscillations, cellular rhythms and morphogenesis.
The metabolic flux can be described as a weighted average of all elementary modes as:
n

Rj = pi rij (1)
i=1

where Rj is the jth reaction rate of the metabolic network, rij are the reaction rates defined by the n elementary modes, and
pi is the probability, described by the Boltzmann distribution, of the ith elementary mode to the reaction j [2]. The individual
elementary modes are related to the cellular regulation which use the metabolic regulation of enzymes to support the
operation of the individual elementary modes at specific rates; consequently, the knowledge of all possible elementary
modes allows us to describe the cellular metabolism [2]. The metabolism can be described as a weighted average of all
possible elementary modes which generate the metabolic flux in the cell [25]. The overall growth is related to the overall
reaction stoichiometry that links the nutrients to the products. The pathways of the individual elementary modes have
different overall stoichiometries, related to the external metabolites [2].
Ruelle [26] considered a classical system with isokinetic time evolution described by the equation:

ξ − αp
   
d p dx
= ⇔ = Fξ (x) (2)
dt q p/m dt
U. Lucia / Physica A 391 (2012) 5997–6007 5999

with
ξ − αp
   
p
x= and Fξ (x) = (3)
q p/m

p ∈ R N and q ∈ R N momentum and position respectively, ξ a nongradient time independent force, m mass and (−α p) the
isokinetic thermostat mathematical expression with α defined as:

p · ξ(q)
α(x) = . (4)
p·p
So that [26]:
d p · p
= 0. (5)
dt 2m
The concept of isokinetic thermostat is useful for calculations, but it does not reproduce the Hamiltonian time evolution at
the equilibrium [26]. Under these conditions Ruelle [26] defined the entropy increment as
 ∞  t    
S (ξ + 1ξ) = dt dτ ρξ ∇x dx ◦ fξt −τ · δτ F(x) Φ (x) (6)
0
0 −∞
   
with δτ F a time-dependent small perturbation of F, ρξ ∇x dx ◦ fξt −τ · δτ F(x) probability distribution, fξt −τ the solution
0
of the Eq. (2) at the time t − τ corresponding to the initial conditions ξ 0 , Φ (x) = (N − 1)α .
Denbigh [27,28] expressed the fundamental processes of living systems, introducing an entropy approach:
dS = dSint + dSext (7)
where dS is the total entropy elementary variation, dS int is the entropy elementary production within the system due to
its metabolism of ingested exergy and dS ext is the entropy exchange with the environment. Entropy is a path independent
state function, and the overall reaction entropy 1SR can be evaluated by the macroscopic reaction stoichiometry between
external metabolites:
n
 n
 k
 n

1SR = pi 1Si = pi νl sli = c pi ln pi (8)
i=1 i=1 l =1 i =1

where 1Si = −c ln pi is the entropy of reaction, sli = (hli − gli )/T , with hli molar enthalpy and gli Gibbs molar energy, are
the molar entropies of the k reactants and products, νl are the stoichiometry coefficients, pi is the probability of the ith mode
and c is a constant related to the numbers of elementary modes and their reaction entropies. It represents the state of the
fully evolved metabolic network [2,29].
When living systems increase in organization, they increase their entropy and far from equilibrium they have a high
exergy content [30]; indeed, considering two systems with the same mass and the same chemical composition, the one,
that has a large amount of organization, has also higher exergy content. During their evolution, the living systems, and also
ecosystems, increase their structure in organization, which is working information useful for resilience and integrity, and
also their efficiency in converting exergy to entropy, in order to reduce the applied exergy gradient, while their internal
entropic state continues to decrease [6,21]. Then, while dS int is always positive defined (dS int ≥ 0), dSext can have any sign.
The inner entropy production σ is defined as the local first time derivative of the [31] internal component of the entropy:
dSint
σ = . (9)
dt
If the irreversible processes are sufficiently slow, the Gibbs equation can be applied to any subsystem [31]:

T dS = dU + pdv − µi dni (10)
i

and the entropy can be expressed in terms of fluxes Ji and conjugated generalized forces Xi [31]:

Tσ = Ji Xi . (11)
i

The non-equilibrium stationary states, which are the states whose variables are independent of time, play a fundamental
role in the irreversible processes. After a characteristic time, the system achieves the equilibrium if no restraints are imposed
on it, while if a number of constant restraints are imposed, a steady state is attained [31]. In a steady state all functions of
the state are independent of time, consequently:
dS dSint dSext dSext dSint
= + =0⇒ =− (12)
dt dt dt dt dt
6000 U. Lucia / Physica A 391 (2012) 5997–6007

but
dSint dSext
≥0⇒ ≤0 (13)
dt dt
and it is possible to argue that the entropy production in a stationary system must be compensated by the liberation of
entropy to the surroundings. This means also that non-equilibrium steady states cannot occur in isolated systems because
these last systems do not allow exchange of entropy between the systems and the surroundings [31]. Prigogine proved
that [32–34]:

d2 Sint
dσ ≤ 0 ⇒ ≤ 0. (14)
dt 2
On the use of the Prigogine’s results there is little doubt that a mature organism may reach a stationary state; indeed, the
homeostasis of all self-regulating systems is interpreted as tendency to return from a perturbed state to that of highest
stability compatible with biological constraints [31].

3. The entropy generation analysis

In order to obtain a thermodynamic approach to biology engineering, a description of the system considered must be
summarized. The system considered [6]:
1. is open, because it exchanges energy and mass flows through its boundaries;
2. is far from equilibrium, because it is a source of high exergy values and basic materials;
3. has a continuous communication, because it has information channels between its different components;
4. is on autopoietic pathways, because there exists continuous cycles for generation and autocatalytic feedbacks;
5. has exergy enhancement or maintenance, because it exports entropy products which exceed or equal the entropy
production of the ingested free energy source and it decreases its internal entropy;
6. presents material conservation and maintains its physical components, because it maintains its structural basis for
storing the acquired organizational exergy.
The stability of an open system can be measured in terms of a certain dimensionless parameter (i.e. the Reynolds number
for fluid flows) and when such a parameter exceeds a certain critical value, the system is no longer stable against small
perturbations [35]. If ξ denotes the dimensionless of observation and ξlim denotes a characteristic constant for the formation
of an unstable structure in the non-linear phenomena it follows that ξ > ξlim . Another condition of the stability of the steady
states is related to the time scale of observation. If τ denotes the time scale of observation and τlim denotes a characteristic
time constant for the formation of a unstable structure in the non-linear phenomena it follows that τ > τlim . When the
two conditions are satisfied, instability develops fully in the system [35]. In phenomena out of equilibrium, irreversibility
manifests itself because the fluctuations of the physical quantities, which bring the system apparently out of stability,
occur symmetrically about their average values; the statistical physical analysis allows us to obtain that the probability
of occurrence follows a universal law and the frequency of occurrence is controlled by a quantity that has been related to
the entropy generation. According to this natural approach, all processes must follow the law of energy dispersal [36–39].
These considerations must be taken into account in the formulation of a thermodynamic model of analysis of the
biological processes and must be expressed introducing a quantity useful to obtain the stability of the steady states.
Moreover, this physical quantity must be also general enough to be used both in linear and in non-linear phenomena;
indeed:
1. an open irreversible real non-linear system with non-linear response must be considered;
2. each process has a finite lifetime τ [40];
3. what happens in each instant in the range [0, τ ] cannot be known, but what has happened after the time τ (the result of
the process) can be well known [24,33,40] (at least it is sufficient to wait and observe);
4. for open system the entropy cannot be defined [41], but the entropy balance equations can be introduced [40] and the
Ruelle [26] definition for the entropy of the non-equilibrium steady states is considered in relation to the Boltzmann
entropy [42] for the open system and the Gibbs entropy [10] only for the reaction evaluations;
5. the entropy balance equations are a balance of fluxes of entropy and energy [43];
6. the Gouy [36,37] principle works for real systems and has been used by Stodola in designing real machines (Gouy–Stodola
principle).
In engineering thermodynamics, the Gouy–Stodola theorem [36–39] relates the work lost Wλ for irreversibility, during the
finite lifetime τ of a process which occurs in an open system, to the entropy generation Sg by means of the relation Wλ = Ta Sg ,
with:

Wλ = Ta Sg . (15)
U. Lucia / Physica A 391 (2012) 5997–6007 6001

Considering that the entropy production must be an extremum along the thermodynamic pathway, and consequently
the entropy generation is an extremum too, we can derived the following equivalent relations:

∂ Sg ∂ Sg
δ Sg = dt + ∇ζ Sg · dζ = 0 ⇒ = −∇ζ Sg · ζ̇
∂t ∂t (16)
∂σ ∂σ
δσ = dt + ∇ζ σ · dζ = 0 ⇒ = −∇ζ σ · ζ̇
∂t ∂t
where ζ is the considered spatial variables (x or V , etc.) as in constructal theory [44].
Moreover, following the constructal approach [45] it is possible to argue:
τ = min (17)
and [44]
∇ζ Wλ = min (18)
as a consequence:

σ ∂ Sg
τ= = min ⇒ Sg = max = ≥0
Sg ∂t
τ (19)
∇ζ Wλ = min ⇒ T0 ∇ζ Sg ≤ 0 ⇒ ∇ζ Sg ≤ 0
from which it is possible to argue:
1. if the open system does not grow in space, but evolves only in time, then ∇ζ Sg = 0 and its entropy generation variation
results dSg /dt = ∂ Sg /∂ t and the entropy generation is maximum;
2. if the open system evolves also in space the entropy generation is maximum or minimum in consequence of the prevailing
of the spatial or time term.

4. Applications

4.1. The ATP synthesis

In this section the ATP synthesis, in anaerobic glycolysis and in respiration, will be considered to show an application of
the maximum generation approach; the reaction is:

{glucose} + [2ADP + 2P+ ]/{2lactate} + [2ATP]


(20)
{6O2 + glucose} + [36ADP + 36P+ ]/{6CO2 + 6H2 O} + [36ATP].
The steady state is characterized by two coupled processes with generalized fluxes (J, J) and corresponding generalized
forces (X1 , X2 ); considering the near-equilibrium irreversible thermodynamics, entropy generation results [46]
S g = J1 X1 + J2 X2 . (21)
Following Santillán et al. [47] only internal irreversibility is considered because it has been shown that it is responsible for the
k=1 Lik Xk , with i = 1, 2 and Lik Onsager coefficients [46],
2
whole entropy increments of the universe. Considering that Ji =
such that Lik = Lik , from the principle of maximum entropy generation, it follows:

L12
X1 = − X2 . (22)
2L11
For this system the generalized fluxes and forces are [47]:

J1 = υ1
J2 = υ2
1g1
X1 = − (23)
T1
1g2
X2 = −
T2
where υ1 and υ2 are the reaction velocities, such that υ1 = υ2 = u, while 1g1 and 1g2 the molar Gibbs energies changes
of the corresponding reactions. For this reaction the linear phenomenological Onsager coefficients are [47]:

υmf b
L11 = L22 = L12 = L21 = (24)
R
6002 U. Lucia / Physica A 391 (2012) 5997–6007

where υmf is the maximum forward velocity, b a positive constant that takes into account the enzymatic processes and R
the ideal gas constant.
Introducing the definition of efficiency η = −(J1 X1 /J2 X2 ) and considering the above relations the following analytical
result can be obtained:
1 T2
η= . (25)
2 T1
Using the data obtained in Ref. [47] about the ATP synthesis, the following numerical value for the efficiency of the reaction
can be obtained:
η = 0.5
(26)
η = 0.7
while, experimentally they give [46]:

η = 0.53
(27)
η = 0.70.

4.2. Human metabolism in interaction with environment

Living systems have been defined as open systems with life-processes, that consist in capturing exergy, in exergy
storage [20], in entropy exchange, and in autopoietic processes [6]. Fanger [48] developed a steady state model based
on the hypothesis that the body is in thermal equilibrium with negligible heat storage [49]: this model does not consider
vasoregulation and shivering because the body core and the skin are considered as just one compartment. The power balance
of the thermal physiology of the human body, at the steady state, can be simplified as follows:

Ṁ − ẇ = q̇c + q̇c ,res + q̇r + q̇e + q̇e,res


   
(28)
where M is the metabolism (J/kg), w is the specific work (J/kg), q is the specific heat lost (J/kg), the subscript c means
convective, r radiative, e evaporative, res respirative and cr core body. The indoor comfort analysis must consider also the
transient condition and the time useful to reach a new comfort steady state. To do so it is necessary to introduce the heat
storage qcr in the core compartment and the heat storage in the skin Qsk , for which the following equations hold:

mcb dTcr

Q̇cr = mb Ṁ − ẇ − q̇c ,res − q̇e,res − Q̇cr −sk = (1 − α)
   
ADu dt (29)
mc dT
Q̇sk = Q̇cr −sk − mb (q̇c + q̇r + q̇e ) = α b sk

ADu dt
where Qcr −sk is the heat exchanged from the body core and the skin [49]:

Q̇cr −sk = (k + ṁbl cbl ) (Tcr − Tsk ) (30)


ṁbl is the blood flow, c is the specific heat (J kg−1 K−1 ), b means body and bl blood, k is the thermal conductivity (W m−1 K−1 ),
T is the temperature, ADu is the surface body area (m2 ) evaluate by the DuBois and DuBois formula [50]:

ADu = 0.202 · y0b.725 m0b.425 (31)


yb is the height (m), mb is the mass (kg) and α is the relative mass of the skin [51]:

0.745
α = 0.0418 + . (32)
3600 · ṁbl − 0.585
The body temperature can be evaluated as [52]:
Tb = α Tsk + (1 − α) Tcr . (33)
Last, during respiration the body exchanges [49]:

Q̇c ,res + Q̇e,res = 0.0014M (34 − Ta ) + 0.0173M (5.87 − pa ) . (34)


The evaporated heat loss from the skin is the result of the combined action of the sweat secreted in consequence of the
thermoregulatory control mechanisms and the spontaneous diffusion of water through the skin, whose body portion of
interest is evaluated as:

Q̇rsw
rrsw = (35)
Q̇e,max
U. Lucia / Physica A 391 (2012) 5997–6007 6003

where rsw means request sweating and the evaporative power loss [53] can be written as:

Q̇rsw = ṁrsw h (36)

and the maximum power transfer due to evaporation is [52]:

psk,s − pa
Q̇e,max = 1
(37)
Rcl + fcl he

where fcl is the correction factor for the increase in available surface area for heat exchange caused by clothing. Considering
that only (1 − wrsw ) of the skin is covered by sweat, the diffusive power loss results [52]:

Q̇diff = 0.06 (1 − wrsw ) Q̇e,max (38)

with 0.06 the value of the skin wetness due to diffusion for normal indoor conditions. The evaporative power loss Q̇e from
the skin is the result of Ref. [52]:
1. the evaporation of sweat output for thermoregulation, Q̇rsw
2. the natural diffusion of water through the skin, Q̇diff
consequently, it follows:

Q̇e = Q̇rsw + Q̇diff . (39)

Moreover, respiration determines the loss of power Q̇res for convection Q̇c ,res and evaporation Q̇e,res [52]:

Q̇res = Q̇c ,rsw + Q̇e,resf = 0.0014M (34 − Ta ) + 0.0173M (5.87 − pa ) . (40)

In order to live, the human body converts its metabolism into its energy needs and it is possible to develop a second law
analysis evaluating its exergy [49]:

Ex = H − H0 (41)

where H is the enthalpy and 0 is the reference state. The thermal exergy exchanged through the skin can be written as:

Exsk = Exr + Eth (42)

where Exr is the exergy for radiation:

T04
 
4T0
Exr = 1− + Qr (43)
3T 3T 4
and Eth is the exergy exchanged for convection:
 
T0
1− Qc
T constant
Exth =   for body temperature in the interaction with the environment. (44)
T0 T0 variable
1− ln Qc
T − T0 T
The chemical exergy load, Exch , is evaluated as the difference between the exergy of the inhaled air and the exhaled air, the
last being considered saturated [52]:

(1 + 1.608ϕ0s ) ϕ
 
Exch = 1.608Rw T0 [Ex0 − (ϕ0s − ϕ0 ) Exw − Ex0s ] ln (45)
(1 + 1.608ϕ) ϕ0c
where Rw is the gas constant for the vapour, p is the pressure, ϕ is the relative humidity ratio, while the subscript s means
saturated, 0 reference state and c convective. Consequently, the total exergy Exout , that flows out from the body, results:

Exout = Exsk + Exch = Exr + Exth + Exch . (46)

The input exergy Exin is the result of the metabolism, sensible heat and latent or wet exergy and it is the fundamental quantity
to be evaluated in order to study the effect of the variation of the environmental parameter on the human metabolism and,
consequently, on the sensation of difficulty in the case of bad thermal control in the art buildings:
 
Trm
Exin = 1− M + Rw Ta ṁe ln ϕ (47)
Tcr
where the subscript rm means room, cr body core, w water (vapour), a air, e evaporative.
6004 U. Lucia / Physica A 391 (2012) 5997–6007

Consequently, the exergy lost Exλ in the process results:

Exλ = Exin − Exout


T04
   
Trm 4T0
M + Rw Ta ṁe lnϕ − 4 4
 
= 1− 1− + 4
Rr Tsk − Trm
Tcr 3Tsk 3Tsk
(1 + 1.608ϕ0s ) ϕ
 
− 1.608Rw T0 [Ex0 − (ϕ0s − ϕ0 ) Exw − Ex0s ] ln
(1 + 1.608ϕ) ϕ0c
 
T0  
1− Rc Tsk − Ta
Tsk body constant
− T0 T0

  for body variable temperature related to the environment (48)
1− ln Rc Tsk − Ta
Tsk − T0 Tsk

with R thermal resistance and the subscripts r and c mean radiative and convective respectively.
Entropy generation can be related to the exergy loss as [36–39,54–56]:

Exλ
Sg = (49)
Ta
where Ta is the ambient temperature and can change. The principle of extremum entropy generation states that:

δ Sg = 0 (50)

which allows us to obtain the relation:

Exλ Exin − Exout


dExλ = dTa = dTa (51)
Ta Ta
which represents the response of the human body to the variation of the ambient temperature; introducing the following
hypothesis:

1. Trm = Ta ;
2. Tsk = constant, in the first time of interaction;
3. Tcr = constant, in the first time of interaction;
4. ϕ = constant, in the first time of interaction;
and integrating it follows:
 
Ta2 Ta2 − Ta1
1Exλ = ln − M + Rw (Ta2 − Ta1 ) ṁe ln ϕ
Ta1 Tcr
T04 4 4
   
4T0 Ta2 Ta2 − Ta1
4
− 1.608Rw T0 Ex0 − (ϕ0s − ϕ0 ) Exw

− 1− + 4
Rr Tsk ln −
3Tsk 3Tsk Ta1 4

(1 + 1.608ϕ0s ) ϕ 2 2
     
 Ta2 T0 Ta2 Ta2 − Ta1
− Ex0s ln ln − 1− Rc Tsk ln − (52)
(1 + 1.608ϕ) ϕ0c Ta1 Tsk Ta1 2

which represents the effect during the time between the ambient temperature variation and the human body answer.
Considering that the metabolic effect of the ambient temperature variation can be evaluated as:

1Exin = 1Exout + 1Exλ (53)

the input exergy variation due to the ambient temperature variation becomes:

T04
   
4T0 T0
1Exin = 4 4
   
1− + 4
Rr Ta2 − Ta1 − 1− Rc Ta2 − Ta1
3Tsk 3Tsk Tsk
T04
   
Ta2 Ta2 − Ta1 4T0
+ ln − M + Rw (Ta2 − Ta1 ) ṁe ln ϕ − 1− + 4
Ta1 Tcr 3Tsk 3Tsk
4 4
 
Ta2 Ta2 − Ta1
× Rr Tsk4 ln − − 1.608Rw T0 [Ex0 − (ϕ0s − ϕ0 ) Exw − Ex0s ]
Ta1 4
(1 + 1.608ϕ0s ) ϕ T 2 − Ta1
2
     
Ta2 T0 Ta2
× ln ln − 1− Rc Tsk ln − a2 . (54)
(1 + 1.608ϕ) ϕ0c Ta1 Tsk Ta1 2
U. Lucia / Physica A 391 (2012) 5997–6007 6005

Fig. 1. Input exergy percentage variation in relation (K−1 ) to the ambient temperature percentage variation (%).

Consequently, it follows that the term useful to evaluate the effect on the metabolism is the variation of Exin , which gives:

T04
   
Ta1 4T0
M + Rw Ta1 ṁe ln ϕ + 4 4
 
Exin2 = 1− 1− + 4
Rr Ta2 − Ta1
Tcr 3Tsk 3Tsk
   
T0 Ta2 Ta2 − Ta1
M + Rw (Ta2 − Ta1 ) ṁe ln ϕ
 
− 1− Rc Ta2 − Ta1 + ln −
Tsk Ta1 Tcr
T04 4 4
   
4T0 Ta2 Ta2 − Ta1
4
− 1.608Rw T0 Ex0 − (ϕ0s − ϕ0 ) Exw

− 1− + 4
Rr Tsk ln −
3Tsk 3Tsk Ta1 4

(1 + 1.608ϕ0s ) ϕ T 2 − Ta1
2
     
Ta2 T0 Ta2
− a2 .

− Ex0s ln ln − 1− Rc Tsk ln (55)
(1 + 1.608ϕ) ϕ0c Ta1 Tsk Ta1 2
The results can be evaluated for a human body characterized by the value reported in Ref. [49]: mass of 80 kg, air velocity
of 0.1 m s−1 , relative humidity of 50%, body energy production of 68 W m−2 K−1 , thermal resistance 0.14 m2 kW−1 , core
body temperature 36.8 °C, skin set temperature 33.7 °C. The ambient temperature is considered around 20 °C. The result is
represented in Fig. 1. It must be pointed out that for less than 1 °C of variation (about 0.3% ambient temperature variation)
the input exergy change is about 0.05%, which means that about the 0.05% of the metabolism must be modified. For greater
temperature variation the situation becomes more difficult for the body to support.

5. Conclusions

Carnot’s general conclusion [57] about heat engines is that there exists a certain limit for the conversion rate of the heat
energy into the kinetic energy and that this limit is inevitable for any natural systems. In 1889, Gouy proved that the lost
work (or lost exergy) in a process is proportional to the entropy generation [36–39] which gave a quantity useful to describe
the progress of dissipative and irreversible processes [54–56]; indeed, an open system develops towards the stationary state
following the thermodynamic path which maximizes its entropy generation under present constraints [54–56]. A system,
capable of assuming many conformations, will tend to assume the one, or frequently return to the one, that maximizes
the rate of dissipation of the powering energy gradients: consequently, the principle of entropy growth is not only about
increasing, but increasing as fast as possible. The exergy gradients are the motive forces of physical processes. Entropy
quantifies the system’s evolutionary course toward increasingly more probable states, while entropy generation describes
its irreversibility [40,54–56]. There are many principles of stability for open systems: the least entropy production, the
maximum entropy production, the minimum entropy generation and the maximum entropy generation.
The relationship between the minimum entropy production principle and maximum entropy generation is not simple;
indeed, these principles are absolutely different because they include different constraints and different variable parameters,
different approaches to the system; consequently, these principles are not mutually opposed because they are applicable to
different boundary conditions and they have also one differential order in time, so they must have different signs of their first
differentials. The Prigogine principle is a principle governing the evolution of non-equilibrium dissipative systems, while
the entropy generation principle can describe all real open system, using global measurable thermodynamic quantities,
representing a powerful, simple and general principle useful in technical physics, engineering thermodynamics, energy
6006 U. Lucia / Physica A 391 (2012) 5997–6007

engineering, biophysics, economy, ecology, and chemistry. Indeed, second law analysis takes the entropy generation into
consideration by including irreversibility. The entropy generation increases because of irreversibility, while it is not the
same for the entropy of the system. In relation to the universe (system and environment) the entropy increases, being
an isolated system. It represents a general principle of investigation for the stability of the open systems: this is a global
theoretical principle for the analysis of the stability of any open system. The entropy generation in irreversible processes
could represent the basis of a new approach to modern thermodynamics and statistical physics. The steady state is stationary
if the entropy generation is maximum. This principle is a maximum principle of entropy generation, but there exists the
Bejan’s least entropy generation theorem, too. They are different in approach and non in contrast one another. Indeed,
Bejan states that the condition of optimization to design a real system is to obtain the minimum entropy generation, which
means that the efficiency is maximum if the work lost because of dissipation is minimum. Moreover, Bejan and Lorente [44]
has developed the ‘‘constructal’’ theory, a powerful theory by which it is possible to predict some macroscopic shapes,
originated by the spatial organization, in Nature, both living and non-living. This law states that for a flow system to persist
in time, its configuration must change in time such that it provides easier access to its currents. This theory is based on
the thought that flow architecture is a consequence of a principle of maximization of flow access, in time, and in flow
configuration that are free to be realized [44]. Bejan’s minimum entropy generation and Bejan’s constructal theory teach how
engineers and Nature design a system (growth and evolution), while the entropy maximum principle describes the natural
thermodynamic behaviour of the systems (interaction). The first describes the internal needs of a system (path in the phase
space, minimum momentum), while the second describes the external interaction between the system and environment
(entropy generation). They are two different approaches to describe the same world: looking it from the internal point of
view (growing in space) and looking it from the external point of view (evolution in time). This can be mathematically
expressed by assigning the right sign to work lost (and, for the Gouy–Stodola theorem, to the entropy generation) in
accordance with the thermodynamic rule for the sign: positive if the work lost is evaluated from the environment, negative
if it is evaluated from the system. The result consists in evaluating a maximum from the environment and a minimum from
the system. The two approaches can be linked by a right interpretation of the mathematical formulation: so the result of
this paper consists in this link between these two approaches.

References

[1] D.C. Krakauer, J.P. Collins, D. Erwin, J.C. Flack, W. Fontana, M.D. Laubichler, J. Prohaska, G.B. West, P.F. Stadler, The challenges and scope of theoretical
biology, J. Theoret. Biol. 276 (2011) 269–276.
[2] F. Srienc, P. Unrean, A statistical thermodynamical interpretation of metabolism, Entropy 12 (2010) 1921–1935.
[3] R. Britten, E. Davidson, Gene regulation for higher cells: a theory, Science 165 (1969) 349–357.
[4] S.C. Materna, E.H. Davidson, Logic of gene regulatory networks, Curr. Opin. Biotechnol. 18 (2007) 351–354.
[5] S.A. Levin, B. Grenfell, A. Hastings, A.S. Perelson, Mathematical and computational challenges in population biology and ecosystems science, Science
275 (1997) 334–343.
[6] F. Günther, C. Folke, Characteristics of nested living systems, J. Biol. Syst. 1 (1993) 257–274.
[7] R.U. Ibarra, J.S. Edwards, B.O. Palsson, Escherichia coli K-12 undergoes adaptive evolution to achieve in silico predicted optimal growth, Nature 420
(2002) 186–189.
[8] S.S. Fong, J.Y. Marciniak, B.O. Palsson, Description and interpretation of adaptive evolution of Escherichia coli K-12 MG1655 using a genome-scale in
silico metabolic model, J. Bacteriol. 185 (2003) 6400–6408.
[9] B. Teusink, A. Wiersma, L. Jacobs, R.A. Notebaart, E.J. Smid, Understanding the adaptive growth strategy of lactobacillus plantarum by in silico
optimization, PLoS Comput. Biol. 5 (2009) e1000410.
[10] P. Unrean, F. Srienc, Metabolic networks evolve towards states of maximum entropy production, Metab. Eng. 13 (2011) 666–673.
[11] J.P. Kernevez, D. Thomas, Numerical analysis and control of some biochemical systems, Appl. Math. Optim. 1 (1975) 237–285.
[12] J.E. Bailey, Lessons from metabolic engineering for functional genomics and drug discovery, Nat. Biotechnol. 17 (1999) 616–618.
[13] A. Friboulet, D. Thomas, System biology—an interdisciplinary approach, Biosens. Bioelectron. 20 (2005) 2404–2407.
[14] J.E. Bailey, Towards a science of metabolic engineering, Science 252 (1991) 1668–1674.
[15] S.B. Carroll, Endless forms: the evolution of gene regulation and morphological diversity, Cell 101 (2000) 577–580.
[16] H. Kitano, Systems biology: a brief overview, Science 295 (2002) 1662–1664.
[17] H. Kitano, Computational systems biology, Nature 420 (2002) 206–210.
[18] I. Prigogine, Modération et transformations irreversible des systems ouverts, Bull. Acad. Roy. Belg. Cl. Sci. 31 (1945) 600–606.
[19] L.M. Martyushev, V.D. Seleznev, Maximum entropy production principle in physics, chemistry and biology, Phys. Rep. Rev. Sect. Phys. Lett. 426 (2006)
1–45.
[20] S.-E. Jörgensen, H. Meijer, A holistic approach to ecological modelling, Ecol. Modell. 7 (1979) 169–189.
[21] D.R. Brooks, J. Collier, B.A. Maurer, J.D.H. Smith, E.O. Wiley, Entropy and information in evolving biological systems, Biol. Philos. 4 (1989) 407–432.
[22] T. Pfeiffer, I. Sanchez-Valdenebro, J.C. Nuno, F. Montero, S. Schuster, METATOOL: for studying metabolic networks, Bioinformatics 15 (1999) 251–257.
[23] S. Schuster, D.A. Fell, T. Dandekar, A general definition of metabolic pathways useful for systematic organization and analysis of complex metabolic
networks, Nat. Biotechnol. 18 (2000) 326–332.
[24] I. Prigogine, R. Lefever, A. Goldbeter, M. Herschkowitz-Kaugman, Symmetry breaking instabilities in biological systems, Nature 223 (1969) 913–916.
[25] S. Schuster, C. Hilgetag, J.H. Woods, D.A. Fell, Reaction routes in biochemical reaction systems: algebraic properties, validated calculation procedure
and example from nucleotide metabolism, J. Math. Biol. 45 (2002) 153–181.
[26] D.P. Ruelle, Extending the definition of entropy to nonequilibrium steady states, PNAS 100 (2003) 3054–3058.
[27] K.G. Denbigh, Note on entropy, disorder and disorganization, British J. Philos. Sci. 40 (1989) 323–332.
[28] K.G. Denbigh, The many faces of irreversibility, British J. Philos. Sci. 40 (1989) 501–518.
[29] S.I. Sandler, H. Orbey, On the thermodynamics of microbial-growth processes, Biotechnol. Bioeng. 38 (1991) 697–718.
[30] R. Swenson, Emergent attractors and the law of maximum entropy production: foundations to a theory of general evolution, Syst. Res. 6 (1989)
187–197.
[31] A. Katchalsky, O. Kedem, Thermodynamics of flow processes in biological systems, Biophys. J. 2 (1962) 53–78.
[32] P. Glansdorff, I. Prigogine, Thermodynamic Theory of Structure, Stability and Fluctuations, John Wiley & Sons, New York, 1971.
[33] I. Prigogine, Etude Thermodynamique des Phénomènes Irrèversibles, Desoer, Liège, 1947.
[34] I. Prigogine, Introduction to Thermodynamics of Irreversible Processes, Interscience, New York, 1961.
U. Lucia / Physica A 391 (2012) 5997–6007 6007

[35] H. Ozawa, A. Ohmura, R.D. Lorenz, T. Pujol, The second law of thermodynamics and the global climate system: a review of the maximum entropy
production principle, Rev. Geophys. 41 (4) (2003) 1018–1041.
[36] G. Gouy, Sur les transformation et l’équilibre en thermodynamique [On the transformation and equilibrium in thermodynamics], C. R. Acad. Sci. Paris,
Sér I Math. 108 (10) (1889) 507–509 (in French).
[37] G. Gouy, Sur l’énergie utilizable [On useful energy], J. Physique 8 (1889) 501–518 (in French).
[38] P. Duhem, Sur les transformations et l’équilibre en thermodynamique. Note de M.P. Duhem [On the transformation and equilibrium in
thermodynamics. Comments of M.P. Duhem], C. R. Acad. Sci. Paris, Sér I Math. 108 (13) (1889) 666–667 (in French).
[39] G. Gouy, Sur l’énergie utilisable et le potentiel thermodynamique. Note de M. Gouy [On the useful energy and the thermodynamic potential. Comments
of M. Gouy], C. R. Acad. Sci. Paris, Sér I Math. 108 (10) (1889) 794.
[40] V. Kirillin, V. Syčev, A. Šejndlin, Technical Thermodynamics, MIR, Moscow, 1980.
[41] J. Serrin, Conceptual analysis of the classical second law of thermodynamics, Arch. Ration. Mech. Anal. 70 (1979) 355–371.
[42] J.L. Lebowitz, Boltzmann’s entropy and time arrow, Phys. Today 46 (1993) 32–38.
[43] A. Bejan, Advance Engineering Thermodynamics, John Wiley, NJ, 2006.
[44] A. Bejan, S. Lorente, The constructal law of design and evolution in nature, Phil. Trans. R. Soc. B. 365 (2010) 1335–1347.
[45] A. Bejan, The golden ratio predicted: vision, cognition and locomotion as a single design in nature, Int. J. Des. Nat. Ecodyn. 4 (2) (2009) 97–104.
[46] S.R. De Groot, P. Mazur, Non-Equilibrium Thermodynamics, North-Holland Publishing Company, Amsterdam, 1962.
[47] M. Santillán, L.A. Arias-Hernandez, F. Angulo-Brown, Some optimisation criteria for biological systems in linear irreversible thermodynamics, Il Nuovo
Cimento D 19 (1) (1997) 99.
[48] P.O. Fanger, Thermal Comfort, Danish Technical Press, Copenhagen, 1970.
[49] M. Prek, Thermodynamical analysis of human thermal comfort, Energy 31 (2006) 732–743.
[50] D. DuBois, E.F. DuBois, A formula to estimate the approximate surface area if height and weight be known, Arch. Int. Med. 17 (1916) 863–871.
[51] X. Wang, Thermal comfort and sensation under transient conditions, Ph.D. Thesis, The Royal Institute of Technology, Stockholm, 1994.
[52] M. Prek, M. Mazej, V. Butala, An approach to exergy analysis of human physiological response to indoor conditions and perceived thermal comfort,
in: 7th International Thermal Manikin and Modelling Meeting, University of Coimbra, September 2008.
[53] I. Dincer, A.Z. Sahin, A new model for thermodynamic analysis of drying process, Int. J. Heat Mass Transfer 47 (2004) 645–652.
[54] U. Lucia, Mathematical consequences of Gyarmati’s principle in rational thermodynamics, Il Nuovo Cimento B 110 (10) (1995) 1227–1235.
[55] U. Lucia, Probability, ergodicity, irreversibility and dynamical systems, Proc. R. Soc. A 464 (2008) 1089–1184.
[56] U. Lucia, Maximum or minimum entropy generation for open systems? Physica A 392 (2012) 3392–3398.
[57] S. Carnot, Rèflexion Sur la Puissance Motrice du feu sur le Machine a Dèvelopper Cette Puissance, Bachelier Libraire, Paris, 1824.

You might also like