You are on page 1of 20

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/318372826

Description of wet-to-dry transition in model ORC working fluids

Article  in  Applied Thermal Engineering · July 2017


DOI: 10.1016/j.applthermaleng.2017.07.074

CITATIONS READS

3 447

3 authors, including:

Axel Groniewsky Attila R. Imre


Budapest University of Technology and Economics Budapest University of Technology and Economics
9 PUBLICATIONS   24 CITATIONS    113 PUBLICATIONS   1,260 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

ESS studies View project

Condensation Induced Water Hammer View project

All content following this page was uploaded by Attila R. Imre on 11 October 2017.

The user has requested enhancement of the downloaded file.


Applied Thermal Engineering, 125 (2017) 963–971

Description of wet-to-dry transition in model ORC working fluids

Axel Groniewsky1, Gábor Györke1, Attila R. Imre1,2,*


1
Budapest University of Technology and Economics, Department of Energy Engineering
Muegyetem rkp. 3, H-1111 Budapest, Hungary
2
Thermohydraulics Department, MTA Centre for Energy Research, P.O. Box 49, Budapest
1525, Hungary
* corresponding author, imreattila@energia.bme.hu

Abstract:

Conventional steam power cycles have their limitations on recovering low grade waste heat,
therefore other alternatives are required in these cases. Organic Rankine Cycle (ORC) is
suitable for power generation based on various heat sources including solar, geothermal,
biomass or waste heat. ORC working fluids can be characterized as wet, dry or isentropic. The
aim of this paper is to give a method to find novel dry or isentropic working fluids based on
simple physical properties, like degree of freedom and isochoric heat capacity.

Highlights

- Conditions for working fluids to show wet, dry or isentropic character were described
for van der Waals fluid
- Wet-to-dry transition was found by increasing the degree of freedom of the molecules
- A limiting isochoric heat capacity (around 45 J/molK) was given for isentropic fluids

Keywords: van der Waals equation of State; Organic Rankine Cycle; isochoric heat capacity;
vibrational degree of freedom

1
Applied Thermal Engineering, 125 (2017) 963–971

Nomenclature

a first parameter in vdW equation of state, m6 kg-2 or m6 mol-2


b second parameter in vdW equation of state, m3 kg-1 or m3 mol-1
cv specific heat capacity at constant volume, J mol-1 K-1
e specific internal energy, kJ kg-1 or J mol-1
f degree of freedom
N number of atoms
p pressure, kPa
R specific gas constant, J kg-1 K-1
S entropy, kJ K-1
s specific entropy, kJ kg-1 K-1 or J mol-1 K-1
T absolute temperature, K
TH average temperature at which heat is supplied to the working fluid, K
V volume, m3
v specific volume, m3 kg-1 or m3 mol-1

Subscripts
r reduced property
cr critical point
0 initial state / reference state

Abbreviations

EoS Equation of state


HRSG Heat Recovery Steam Generator
ORC Organic Rankine Cycle
vdW van der Waals

Supersripts
' (apostrophe) variable property during differentiation or integration

2
Applied Thermal Engineering, 125 (2017) 963–971

1. Introduction

Since conventional steam power cycles have their limitations on recovering low grade waste
heat, other alternatives are required to serve such purpose. Organic Rankine Cycle (ORC) is a
technology, which is ideally suitable for power generation based on various sources, including
solar, geothermal, biomass or waste heat. Finding the optimal working fluid for a given heat
source – which is essential to achieve higher energy efficiency in ORC systems – has great
impact on the power plant, namely it does not just influence operational variables of the cycle
but also alters the layout and modifies the design of the equipment. Selection criteria of working
fluids generally include thermodynamic and physical properties [1-7], stability (in the desired
temperature range), compatibility, environmental impact [8], safety [9-10], availability and
cost-effectiveness [11-16], but most often all of these criteria are used together. Hence
isentropicity, wetness or dryness is only one parameter among the many others to find the most
appropriate working fluid. The selection is a multi-dimensional optimization problem [2,11].

Traditionally, the working fluid is selected by a trial and error procedure through experience of
chemically similar materials. A set of possible working fluid candidates are identified by using
heuristic knowledge of the process based on similar systems. These candidates are then assessed
in a process simulation scheme and optimized [17]. Such an approach runs the risk of excluding
the optimal working fluids from further consideration if the process objectives are not
considered in the initial selection step [18]. If not only the influence of thermodynamic
properties of specific working fluids – which is normally the foundation of trial and error
selection mechanisms – but also the parameters of state equations on cycle performance are
analysed first, potential candidates are less likely to drop out in the initial step.

Among the many other criteria, a physical-chemical one - based on the properties of the
expanding fluid - is the wetness/dryness/isentropicity of the working fluid [6,7,11,14]. The aim
of this paper is to give a recipe for finding novel dry or isentropic working fluids based on the
parameters of the applicable equations of state by pointing out what practical significance these
parameters have on cycle performance. Although our approach is generally applicable to avoid
mathematical artefacts – caused by complex state equations – a methodology is demonstrated
with one of the simplest equation of state, which is still suitable to describe an equilibrium two
phase system, namely the van der Waals equation of state.

2. Classification of ORC working fluids

ORC working fluids can be characterized as wet, dry or isentropic, depending on their
behaviour during the adiabatic expansion of dry vapour in the temperature-entropy diagram.
This type of classification is widely spread because it clearly shows the limitations of the
different categories of working fluids in power plant application. Although this classification is
frequently used, these types are not strictly defined, therefore first we have to define them.

There are two different definitions for the groups, both of them are based on the shape of the
phase diagram on specific entropy – temperature space (called T-s diagram). In these diagrams,

3
Applied Thermal Engineering, 125 (2017) 963–971

pure liquid, mixed liquid+vapour and pure vapour phases are separated by the saturation curve
(see Fig. 1/a). The point corresponding to the temperature maximum is the critical point,
dividing the curve into two parts. The part for smaller entropies are the liquid saturation curve,
separating pure liquid states from mixed liquid+vapour states, while the high-entropy part is
the vapour saturation curve, separating the mixed liquid+vapour states from the pure vapour
states. The liquid saturation part of the curve is always positive sloped, while the vapour part
can be negative, positive, infinite or alternating (Figs. 1/a-d). This is the basis of the first
distinction, i.e. the first definition is based on the slope of the saturation vapour curve in T-s
diagrams. For a wet fluid it is negative, for a dry fluid it is positive and for an isentropic fluid
it is infinite [1,10,19,20], see Fig. 1/a-c. This definition cannot be applied to fluids, which have
saturation vapour curve with alternating slope (Fig. 1/d). The second – and more technical
definition – is based on the final state of an ideal adiabatic expansion [11,19]. These expansions
are marked by arrows in Fig. 1/a-d, where 1 marks the initial, 2 marks the final point of the
expansion. The starting point is always in the dry vapour state; in ideal case it is located on the
saturation vapour curve. After a reversible adiabatic (isentropic) expansion step, the final state
might be inside of the two-phase region (wet vapour, i.e. vapour mixed with liquid droplets);
this is the case for wet fluids (Fig. 1/a). In this case, satisfying the first definition, the second
one will be satisfied too, i.e. with negative slope the expansion always ends in the wet vapour
region. For dry fluids, after the isentropic expansion, the final state will be dry vapour (Fig.
1/b). For this case, satisfying the criteria of dryness for the first definition (negative slope or
the saturation vapour curve), the dryness criteria for the second definition will be also satisfied.
In the third (isentropic) case, starting the expansion from a high-temperature point of the vapour
saturation line, the expansion runs together with the equilibrium line, and therefore ends on the
saturation line too, where only dry vapour can be found. The problem is that vapour saturation
curves with infinite slope do not exists or at least the slope is not infinite in a wide temperature
range. In real systems, the vapour side of the curve can be curved turning the slope from positive
to negative, while passing the infinite (ideal isentropic) part (Fig. 1/d). A real example can be
seen in Fig. 2, where the T-s diagram of butane is plotted (entropy data are taken from the NIST
database [21]) in the whole liquid range (from triple point to critical point).

In this case, although the starting point (1) is still on the dry vapour side of the saturation curve,
the end point – depending on the final temperature of the expansion – can be isentropic (point
2), dry (2*) or wet (2**). Since transition between 1-2* points can be handled as isentropic,
therefore we are going to call these type as real isentropic contrary to the one in Fig. 1/c, which
is the ideal isentropic. In this case the problem is, that fixing the initial or final point of the
expansion on the vapour saturation curve, the other point will be fixed too. Hence the
temperatures of live vapour and condensate will be connected, causing inefficient operation
either in the HRSG or in the condenser. Technically, the temperature of live vapour (represented
by point 1) should be only determined by the temperature of the heat source and pinch
temperature of the heat exchanger, while the temperatures of the condensate (represented by
point 2) should depend only on the temperature of the coolant and pinch point temperature of
the condenser.

4
Applied Thermal Engineering, 125 (2017) 963–971

Each type of working fluid need special technics upon application. To avoid erosion of the
turbine blades caused by excessive liquid formation during the expansion, wet working fluids
mostly require superheating. The use of a superheater decreases the wet steam energy loss in
the turbine as well as increases TH average temperature at which the heat is added to the system
and consequently increases the equivalent Carnot efficiency considerably. However, in turn,
this makes the working fluid less desirable for low grade waste heat recovering. Additionally,
in some cases, the temperature of the required superheating might be too high, requiring the use
of special alloys. A Rankine like cycle with a wet working fluid is shown in Fig. 3/a,b. One
should realize that expansion can be started at any state between the ones marked as 4 and 5,
but closer to 4, the vapour contains more droplets, causing higher erosion rate.

Due to the nature of their saturated vapour line, neither isentropic nor dry fluids require
superheater to avoid turbine erosion. However, if the expanded vapour leaves the turbine in a
highly superheated state rather than saturated one, it requires greater cooling capacity from the
condenser. To avoid higher cooling load, a recuperative heat exchanger is used, which cools
the superheated vapour before reaching the condenser and preheats the liquid before reaching
the boiler. This solution does not only decrease the cooling load of the condenser but increases
TH and consequently the equivalent Carnot efficiency. Since the rate of efficiency improvement
depends on the working fluid and temperature range in which the system is used, the application
of a recuperator is rather an economic than a technical issue. An ORC with isentropic and dry
fluid is shown in Fig. 3/c-f.

As in Fig. 3/c can be seen, technically for isentropic fluids a small superheating is used before
the expansion, i.e. the expansion line does not run on the saturation curve, rather on a parallel
trajectory, not too deep in the dry vapour zone (see points 4 and 5 and on the line between them
in Fig. 3/c). This superheating is a safety margin, required to avoid any unwanted condensation
before reaching the expander. It should be noted here that by using real isentropic fluids, this
would not be necessary, since – due to the curved shape of the saturation curve – the expansion
line is far enough from the mixed (wet+dry) vapour states to avoid condensation even in case
of a small accidental heat-loss. This is also true for real expanders, where the expansion lines
have negative slope (dS >0) instead of infinite one.

As it is demonstrated above, the character of the potential working fluid influences not only the
energy efficiency of an ORC but has a significant impact on the system design as well.
Therefore the shape of the saturated vapour curve has great importance in working fluid
selection. The following methodology describes a simple way to determine the equilibrium
phase diagram of a working fluid on the temperature-entropy plane.

3. Methodology

For pure substances phase equilibria requires identical pressure, temperature and chemical
potential among the coexisting phases. When instead of calculating the chemical potential –
which requires caloric information – Maxwell construction is applied [22,23], equilibrium

5
Applied Thermal Engineering, 125 (2017) 963–971

phase diagram of the pressure-volume plane can be derived entirely from the thermal equation
of state ( f ( p,V,T )=0). Fig. 4. shows a general equilibrium phase diagram for the van der Waals
thermal equation of state on the reduced pressure - volume plane. Reduced quantities are
calculated by dividing the real ones with the critical ones. Two isothermal lines are plotted
(Tr=0.87 and 0.95; for example for water, where critical temperature is 647.1 K (374 0C), these
values represent 563 K (289.8 0C) and 614.7 K (341.6 0C), respectively). Maxwell-construction
cuts the curves by an isobar (constant pressure line, dashed on the figure) obtaining two equal
grey area for the given isotherm. Connecting the intersects of the isotherms and isobars (and
omitting the middle ones, marking instable states), one can obtain the binodal curve (solid line
with circles in Fig. 4), representing equilibrium vapour (high volume) and liquid (low-volume)
states. From the pressure-temperature pairs obtained by the same process (for each isotherm,
one equilibrium pressure is determined), the well-known saturation- or vapour pressure curve
can be constructed in temperature-pressure diagram.

After pressure and volume data pairs of the phase diagram are determined, they are used to
transfer the phase diagram from the reduced pressure - volume plane to the reduced temperature
- entropy plane. Since s  s(v, T ) function does not only require thermal but also caloric
information beside the thermal state equation, another equation with embedded caloric
information such as cv  cv (v, T ) is required. Specific heat capacity at constant volume can be
expressed as:

 2 p(v, T )
v
cv (v, T )  cv (v0 , T )  T  

v v0
T 2
 dv . (1)

Since the second term (called residual term) of this expression is derived from the thermal state
equation, the caloric information is encoded in the first one (which is the ideal gas contribution).
If cv  cv (v, T ) is given, entropy of the points of the phase equilibrium curve might be calculated
using Eq. (2):

p(v, T ) cv (v, T )
v T
s(v, T )  s0  
v v0
T 
 dv  
T T0
T
 dT  . (2)
T T v v0

Accurate description for complex fluids (and here complex means everything, except a few
noble gases) requires quite difficult equation of state, see for example the multi-parameter
reference IAPWS (also called Wagner-Pruss) equation of state for water [24]. On the other hand
these equations are mathematically very complex, and therefore they might even produce
artificial results, like for example the virtual existence of a second stable liquid phase [25]. To
avoid mathematical artefacts – caused by complex state equations – we used van der Waals
equation of state for demonstration, even though in this way the obtained results will be only
qualitatively correct, rather than quantitatively. The general form of van der Waals equation of
state is [22,23]:

RT a
p  v, T    2. (3)
v b v

6
Applied Thermal Engineering, 125 (2017) 963–971

where p and T are the normal (not reduced) pressure and temperature, v is the specific volume
(usually given as the volume of one mol or one kg), R is the gas constant and finally a and b
are the material dependent van der Waals parameters. The caloric state equation of van der
Waals has a temperature dependent and a volume dependent term, therefore the change of
internal energy is calculated as

de  e v T  dv  e T v  dT . (4)

Considering the relation between the thermal and caloric state equations suggested by the
Second Law of Thermodynamics:

e v T  T p T v  p (5)

and substituting Eq. (3), it leads to the conclusion that the volume dependent term of internal
energy e(v)   a v . The only constraint on the temperature dependent term arises from the
physical behaviour of specific heat, requiring a positive temperature dependent term. Therefore,
caloric state equation was chosen to be

f a
e  v, T    R T  , (6)
2 v

where f is the internal degree of freedom. Here, for the mathematical solution we are assuming
that R, a and b are positive coefficients, which is physically a sound assumption. Since a and b
only influences the vertical and horizontal stretch of the phase diagram – by alteration of the
position of critical point – not the shape, solution might be extended by using reduced
properties, which are state variables normalized by state properties at the critical point. Critical
values can be found by determining the inflection points of Eq. (3). p  v, T  v  0 and

 2 p  v, T  v  0 gives vcr  3  b , Tcr  8  a 27  R  b and pcr  vcr , Tcr   a  27  b2  . To have


2

a more coherent solution, reduced internal energy is normalized by pcr  vcr . Resulting reduced
constitutive functions are:

p Tr  Tcr , vr  vcr  8  Tr  vr2  9  vr  3


pr   , (7)
pcr vr2   3  vr  1

e Tr  Tcr , vr  vcr  12  Tr  a  b  f  vr3  27  a  b  vr2  1


er   . (8)
pcr  vcr 9  a  b  vr3

To determine vr for every pr of the equilibrium phase diagram on the reduced pressure -
volume plane, an isotherm was calculated and intersected between its local minimum and
maximum with an isobar, then bisection method was used to relocate this isobar line until it
satisfied the Maxwell construction. After the calculation of pr , vr point pairs, they were
transferred and linked on the reduced temperature - entropy plane with help of Eq. (2) [26,27].

7
Applied Thermal Engineering, 125 (2017) 963–971

Fig. 5. shows an equilibrium phase diagram for the van der Waals equation of state with
cv  3 2  R on the reduced temperature - entropy plane. This heat capacity value represents a
monoatomic material, like argon [23].

4. Results and Discussion

Regardless of the internal degree of freedom all diagrams have the same entropy at the chosen
reference temperature (Tr=0,3); this would be true for any other arbitrarily chosen value (in
reduced temperature scale). The s  s(v, T ) entropy function depends on s0 initial state,
temperature and volume changes. Initial state ( s0 ) does not influence the slope of the entropy
functions, only shifts the curves horizontally. Since the initial and final states are on the same
isotherm, there is no entropy change due to temperature change. It means that all the entropy
change between initial and final states is caused by the change of volume. Since the volume
dependent entropy change only influenced by the van der Waals equation – as Eq.(2) indicates
– it will cause the same change of entropy independently of the caloric information.

Although the choice of Tr=0.3 is arbitrary, this initial temperature is low enough to demonstrate
the relation between the shape of the phase diagram and the heat capacity function.
Additionally, it is small enough to fit the total liquid range of fluids into the reduced 0.3-1 range
(for water, the triple point temperature is at Tr=0.42, i.e. water can be liquid in the 0.42-1
reduced temperature range).

Fig. 6/a. shows the phase diagrams with different caloric state equations; for better
comparability, the same curves can be seen in Fig. 6/b, where on x-axis, the difference from the
critical entropy can be seen (i.e. the curves are shifted along the x-axis). With increasing
internal degree of freedom the shape of the phase equilibrium curve changes and from wet ( f
=3 to f =7, marked as f-3 to f-7) it slowly becomes isentropic (f-9 to f-11) and finally dry (f-12
and above). Fig. 6/c displays the transition between wet and dry slope. Based on this result, one
might give a thumb-rule to find dry or isentropic working fluid by choosing a limiting degree
of freedom, and based on that value a limiting specific heat capacity. Since this calculation is
only qualitative (due to the simple van der Waals equation of state), it is not necessary to find
a very accurate value. One can roughly say that a fluid is dry or isentropic, when its degree of
freedom goes above f =10..12 and therefore the isochoric heat capacity in the temperature range
of interest goes above the range of (10/2)*R - (12/2)*R, i.e. 41.6-49.9 J/molK. For
demonstration, in Fig. 7/a one can see isochoric heat capacities of lower alkanes (from methane
to pentane) in their liquid range (from triple point to critical point) [21] compared to the dry-to-
wet approximate transition value established here ( f =10..12, cV =41.6-49.9 J/molK). Since van
der Waals equation is quantitatively not reliable, instead of a sharp line a diffuse f =10-12 range
are marked by the grey band. For methane, cV cannot reach the limiting value, therefore it is
expected to be wet. For ethane, at high temperature the cV is over the limiting value suggesting
transition to isentropic; in reality it is still wet, but the high-temperature part of the T-s diagram
on the vapour side is already very steep, close to the isentropic behaviour. For the other three
fluids, isentropic or even dry behaviour is expected. In reality, propane is approximately

8
Applied Thermal Engineering, 125 (2017) 963–971

isentropic, while butane and pentane are dry; their T-s diagrams can be seen in Fig. 7/b, data
are taken from the NIST Webbook [21]. This example shows that our simple thumb-rule might
be used with success to find novel dry and isentropic working fluids simply by searching among
fluids with high vapour isochoric heat capacity. Although the given cV value is obviously too
low (from Fig. 7/a, one would expect 70-80 J/molK as limiting value), this might be corrected
by using more realistic equation of state.

It would be useful to make similar comparison with other chemical homologues (like linear
alcohols), but reliable T-s and heat capacity data are not available for them. On the other hand,
the fact that the isochoric heat capacity of some well-known wet fluids are lower (taking room-
temperature dry vapour , cV , for water it is 25.9 J/molK, for carbon dioxide, it is 52.0 J/molK)
than for some well-known dry fluids (like benzene with 73.8 J/molK and toluene with 95.6
J/molK), suggesting a general trend for dryness among fluids with higher heat capacity.

To clarify the reason of wet-dry transition, one more look has to be taken at Eq. (2). It indicate,
that the more dominant the caloric information is compared to the thermal one, the dryer the
character of the working fluid becomes. One can conclude that using only the residual part of
cV, the residual entropy curve in T-s diagram would be always wet; for the isentropicity and
dryness only the ideal gas contribution is responsible; this is demonstrated in Fig. 8, by showing
the residual, ideal and total part of the entropy in case of f =3 and f =15. Although in most cases
only the total entropies are calculated, some programs (like the ThermoC [28-30]) are able to
calculate separate residual contributions too; calculations done by ThermoC shows similar
results to ours and even shows that the validity of this statement can be true for other equation
of states too, not only for van der Waals EoS.

Among the three equation of state parameters (a,b and f ), now we can give a criteria for f,
which can influence the wetness/dryness/isentropicity of the working fluid. The influence of a
b parameters in reduced property space are non-existent, but – through the temperature
dependence of vibrational degrees of freedom – in non-reduced space they also might influence
the characteristic of T-s curves. The inclusion of this dependence is one of the potential way to
improve our model and obtain better limiting cV; another way is the application of more realistic,
but still analytically solvable equation of states (like Peng-Robinson [29,31] or Redlich-Kwong
[29,32]).

In case of most vapours, vibrational degrees of freedom are not excited at “low” temperature.
Under these conditions the heat capacity depends only on whether the gas is composed of
isolated atoms, linear molecules, or nonlinear molecules, i.e. only translational and rotational
degrees of freedom can contribute for the isochoric specific heat capacity. The values for cV,
the heat capacity at constant volume are (3/2)*R, (5/2)*R and (6/2)*R, since the active degrees
of freedom at low temperature can be 3,5 or 6 (for monoatomic, multiatomic linear and
multiatomic nonlinear molecules, respectively) [33]. As the internal degree of freedom for most
gases at low temperature is low, the shape of the phase diagram at the bottom is mostly wet (see
the part of the vapour branch with negative slope in Fig.2 below 250 K). At “higher”
temperature, internal (vibrational) degrees of freedom can be activated, increasing the heat
capacity, producing higher cV. “Low” and “high” temperature are relative, internal freedom of
degrees can be activated on various levels at various temperatures depending on the
composition/complexity of the molecule including the type of chemical bonds; in some senses,

9
Applied Thermal Engineering, 125 (2017) 963–971

one might call the degree of freedom as a measure of the complexity of the molecules [26].
While direct information about the activation of various internal degrees of freedom can be
obtained directly for example by infrared spectroscopy [34], indirect information can be gained
by the observation of the isochoric heat capacity of the vapour state of the given material.
Obtaining cV higher than (3/2)*R, (5/2)*R and (6/2)*R indicates partial or full activation of one
or more internal degrees of freedom. The total degree of freedom (including translational,
rotational and vibrational) can be 3*N, where N is the number of atoms of the given molecule,
suggesting that mono-, di- and triatomic molecules (like argon, oxygen, nitrogen, carbon
dioxide, water) have to be always wet working fluids (being f <10..12, given as the isentropic
limit); even working fluids of four-atomic molecules (like ammonia) are probably wet rather
than dry.

As it has been already mentioned above, at room temperature, usually only translational and
rotational degrees of freedom are active, resulting a maximum of f=6. The transitional value
where dry-wet transition can be seen are in the f=10-12 range; definitely higher than for an
“average” molecule at room temperature. To reach this range, more complex molecules with
higher molar mass and easily excitable chemical bonds are required. It should be mentioned
here, that probably this condition is necessary, but not sufficient to have isentropic or dry
working fluid, but together with other parameters (like the existence of intermolecular hydrogen
bonds in fluids, frequent among wet fluids [35]) it could be a useful tool for working fluid
selection.

5. Summary

ORC working fluids can be wet, dry or isentropic, depending on their behaviour during the
adiabatic expansion of equilibrium (saturated) dry vapour. To avoid non-preferred droplet
formation, dry or isentropic working fluids should be used, whenever it is possible. To find
these kind of fluids is a challenge; usually it is done by trial and error methods starting from
already known preferable materials.

In this paper, a simple thumb-rule has been introduced to find new dry or isentropic fluid, based
on the internal degree of freedom and on the isochoric heat capacity of equilibrium (saturated)
vapour of the materials. By using van der Waals equation of state (which is not accurate
numerically, but qualitatively gives good results), it has been concluded, that molecules with
higher active degrees of freedom (f ) are probably dry or isentropic. Since cV depends on the
degree of freedom of the given molecules, we were even able to give a heat capacity limit for
wet-to-dry transition, which is in the range of 41.6-49.9 J/molK for the saturated vapour; this
range corresponds to f = 10-12. The applicability of the rule was demonstrated on linear alkanes,
showing the transition from wet to dry around propane suggesting dry characteristic for longer
alkanes.

Acknowledgements

10
Applied Thermal Engineering, 125 (2017) 963–971

This work was supported by the Hungarian NKFIH Grant under contract No. K116375 and by
the Hungarian FIEK Grant under contract No. 16-1-2016-0007. The authors are indebted to T.
Fülöp, T. Környey and U.K. Deiters for their helpful advices.

References

[1] I. H. Aljundi, Effect of dry hydrocarbons and critical point temperature on the
efficiencies of organic Rankine cycle, Renewable Energy 36 (2011) 1196-1202.

[2] V.A. Mazur, D. Nikitin, Sustainable Working Media Selection for Renewable Energy
Technologies in “World Renewable Energy Congress” (WREC 2011), Linköping, Sweden,
May 8–13, (2011) 856–866.

[3] C. He, C. Liu, H. Gao, H. Xie, Y. Li, S. Wu, J. Xu, The optimal evaporation
temperature and working fluids for subcritical organic Rankine cycle, Energy 38 (2012) 136-
143.

[4] J. Wang, J. Zhang, Z. Chen, Molecular entropy, thermal efficiency, and designing of
working fluids for Organic Rankine Cycles, International Journal of Thermophysics, 33
(2012) 970-985.

[5] Z.Q. Wang, N.J. Zhou, J. Guo, X.Y. Wang: Fluid selection and parametric
optimization of organic Rankine cycle using low temperature waste heat, Energy 40 (2012)
107-115.

[6] A.R. Imre, S.E. Quinones-Cisneros, U.K. Deiters, Adiabatic processes in the liquid–
vapor two-phase region - 1. Pure fluids, Industrial&Engineering Chemistry Research,
53(2014) 13529-13542.

[7] A.R. Imre, S.E. Quinones-Cisneros, U.K. Deiters, Adiabatic processes in the vapor–
liquid two-phase region - 2. Binary mixtures, Industrial&Engineering Chemistry Research, 54
(2015) 6559-6568.

[8] D. Luo, A. Mahmoud, F. Cogswell, Evaluation of Low-GWP fluids for power


generation with Organic Rankine Cycle, Energy 85 (2015) 481-488.

[9] E.H. Wang, H.G. Zhang, B.Y. Fan, M.G. Ouyang, Y. Zhao, Q.H. Mu: Study of working
fluid selection of organic Rankine cycle (ORC) for engine waste heat recovery, Energy, 36
(2011) 3406-3418.

[10] J. Bao, L. Zhao: A review of working fluid and expander selections for organic
Rankine cycle, Renewable and Sustainable Energy Reviews 24 (2013) 325–342.

[11] H. Chen, D.Y. Goswami, E.K. Stefanakos, A review of thermodynamic cycles and working
fluids for the conversion of low-grade heat, Renewable and Sustainable Energy Reviews 14
(2010) 3059–3067.

11
Applied Thermal Engineering, 125 (2017) 963–971

[12] Q. Zhu, Z. Sun, J. Zhou, Performance analysis of Organic Rankine Cycles using
different working fluids, Thermal Science, 19 (2015) 179-191.

[13] P. Garg, M.S. Orosz, P. Kumar, Thermo-economic evaluation of ORCs for various
working fluids, Applied Thermal Engineering, 109 (2016) 841-853.

[14] M. Roedder, M. Neef, C. Laux, K-P. Priebe, Systematic Fluid Selection for Organic
Rankine Cycles and Performance Analysis for a Combined High and Low Temperature
Cycle, Journal of Engineering for Gas Turbines and Power Transactions of the ASME, 138
(2016) 031701.

[15] K.A. Barse, M.D. Mann, Maximizing ORC performance with optimal match of
working fluid with system design, Applied Thermal Engineering, 100 (2016) 11-19.

[16] H. Yu, X. Feng, Y. Wang, Working Fluid Selection for Organic Rankine Cycle (ORC)
Considering the Characteristics of Waste Heat Sources, Industrial & Engineering Chemistry
Research 55 (2016) 1309−1321.

[17] M. Lampe, J. Gross, A. Bardow, Simultaneous process and working fluid optimization for
Organic Rankine Cycles (ORC) using PC-SAFT, Computer Aided Chemical Engineering 30
(2012) 572-576.

[18] – A. Hattiangadi, Working Fluid Design for Organic Rankine Cycle (ORC) Systems,
Master of Science Thesis, TU Delft, 2013.

[19] P.J. Mago, L.M. Chamra, K. Srinivasan, C. Somayaji: An examination of regenerative


organic Rankine cycles using dry fluids, Applied Thermal Engineering 28 (2008) 998–1007.

[20] S. Quoilin, M. Van Den Broek, S. Declaye, P. Dewallef, V. Lemort: Techno-economic


survey of Organic Rankine Cycle (ORC) systems, Renewable and Sustainable Energy Reviews
22 (2013) 168-186.

[21] NIST Chemistry WebBook, NIST Standard Reference Database Number 69


http://webbook.nist.gov/chemistry/

[22] D. C. Johnston, Advances in Thermodynamics of the van der Waals Fluid, Morgan &
Claypool Publishers, 2014.

[23] P.W. Atkins, Physical Chemistry, 5th Edition, W.H. Freeman and Company, New York,
1994.

[24] Wagner, W., Pruß, A., The IAPWS formulation 1995 for the thermodynamic
properties of ordinary water substance for general and scientific use, Journal of Physical and
Chemical Reference Data, 31 (2002) 387–535.

[25] A. R. Imre: Stability and negative pressure in bulk and confined liquids, in “NATO Science
Series, Alternative Water Resources in Arid Areas by retrieving Water from Secondary
Sources” (Eds.: L. Mercury, N. Tas and M. Zilberbrand), Springer, 2013, pp151-157.

12
Applied Thermal Engineering, 125 (2017) 963–971

[26] J. Harinck, A. Guardone, P. Colonna: The influence of molecular complexity on


expanding flows of ideal and dense gases, Physics of Fluids 21 (2009) 086101.

[27] J. M. Smith, H. C. Van Ness, M. M. Abbott, Introduction to Chemical Engineering


Thermodynamics, McGraw-Hill, Singapore, 2001.

[28] U. K. Deiters, A modular program for the calculation of thermodynamic properties of


fluids”, Chem. Eng. Technol. 23 (2000) 581–584.

[29] U. K. Deiters, T. Kraska, High-Pressure Fluid Phase Equilibria—Phenomenology and


Computation, Elsevier, Amsterdam 2012.

[30] ThermoC by U.K. Deiters, http://thermoc.uni-koeln.de/ (2015).

[31] D. Y. Peng, D. B. Robinson, A New Two-Constant Equation of State, Ind. Eng. Chem.
Fundam. 15 (1976) 59–64.

[32] O. Redlich, J.N.S. Kwong, On the thermodynamics of solutions. V: An equation of


state. Fugacities of gaseous solutions, Chem. Reviews 44 (1949) 233–244.

[33] R.H. Perry, D.W. Green, Don W. Perry's chemical engineers’ handbook 8th edition, Mc-
Graw-Hill, 2008.

[34] N.B. Colthup, L.H. Daly, S.E. Wiberley, Introduction to Infrared and Raman
Spectroscopy, 2nd Edition, Academic Press, New York, 1975.

[35] B. Liu, K. Chien, C. Wang, Effect of working fluids on organic Rankine cycle for
waste heat recovery. Energy, 29 (2004) 1207-1217.

13
Applied Thermal Engineering, 125 (2017) 963–971

Figures

Figure 1.: Schematic diagrams for various types of working fluids (wet, dry, ideal isentropic
and real isentropic). Arrows represent isentropic (ideal adiabatic) expansion from initial (1) to
final (2) states.

Figure 2: T-s diagram of butane covering the whole liquid range in temperature (from triple
point temperature to liquid-vapour critical temperature) showing the strongly curved vapour
branch with a local entropy-minimum (around 270 K) and maximum (around 400 K).

14
Applied Thermal Engineering, 125 (2017) 963–971

Figure 3/a-f : T-s diagrams (solid lines) for a wet, ideal isentropic and dry van der Waals fluid,
marking the steps of ORC (dashed lines) (a,c,e, respectively) and the schematic process flow
diagram of an ORC process, using wet, ideal isentropic or dry working fluid (b,d and e,
respectively).

15
Applied Thermal Engineering, 125 (2017) 963–971

Figure 4 Two sub-critical van der Waals isotherms (solid lines) in reduced volume-pressure
space, showing van der Waals loops (upper one corresponds to higher reduced temperature).
Equilibrium pressures are marked by dashed lines; the grey areas of the same loops are equal,
according to the Maxwell-construction. Intersects of isobar Maxwell-lines (dashed) and
isothermal curves (solid) marks the equilibrium states (circles). The dotted line connecting the
circles are the binodal line, representing liquid-vapour equilibrium states.

Figure 5. Construction of the T-s diagram (marked as binodal curve) by connecting equilibrium
liquid (low-entropy side) and vapour (high-entropy side) states. In this case, the fluid (a
monoatomic van der Waals fluid) is a wet one.

16
Applied Thermal Engineering, 125 (2017) 963–971

Figure 6/a: Reduced T-s diagrams of van der Waals fluids with different internal degree of
freedom, influencing their isochoric heat capacity. b: T-s diagram, with all critical point
shifted into one point. c: Magnified vapour branches for f=9, 10 and 11, showing a wet-
isentropic-dry transition.

17
Applied Thermal Engineering, 125 (2017) 963–971

Figure 7/a: Isochoric heat capacities of lower alkanes (from methane to pentane) in their liquid
range (from triple point to critical point), compared to the dry-to-wet approximate transition
value established here (f =10-12, cV =41.6-49.9 J/molK). Methane and ethane are wet working
fluids, although the high temperature part for the ethane T-s diagram is close to isentropic;
propane is approximately isentropic, while butane and pentane are dry. Their T-s diagrams
can be seen in Figure 7/b.

18
Applied Thermal Engineering, 125 (2017) 963–971

Figure 8: a; Residual (solid line with circles) and ideal gas (solid line) contribution of the
specific entropy for f=3 (monoatomic vdW fluid); the sum of these two parts gives the total
entropy curve (b), which has wet character. c: Residual (circles) and ideal gas (solid)
contribution of the specific entropy for f=15 (vdW fluid of an at least five-atomic molecule);
the sum of these two parts gives the total entropy curve (d), which has a real isentropic
character; at higher temperatures, it is dry.

19

View publication stats

You might also like