Sei sulla pagina 1di 9

Energy 36 (2011) 4616e4624

Contents lists available at ScienceDirect

Energy
journal homepage: www.elsevier.com/locate/energy

Exergy analysis of an isothermal heat pump dryer


Will Catton*, Gerry Carrington, Zhifa Sun
Department of Physics, University of Otago, 730 Cumberland Street, Dunedin 9016, New Zealand

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 12 November 2010
Received in revised form
9 March 2011
Accepted 12 March 2011
Available online 12 April 2011

A numerical simulation of a plate contact-type isothermal heat pump dryer (HPD) is used to examine the
energy efciency improvement obtainable from this system compared with a conventional HPD. While
we consider this system design to be entirely feasible, we are not aware of any existing practical
applications of the design. The simulation incorporates a detailed plate, product and air ow model,
solving the mass, momentum and energy balances within the drier, into a pre-existing model of the
remaining HPD components. The accuracy of an idealised drier-duct model used in a previous analysis is
assessed. Although the accuracy of the idealised model is found to be sensitive to local system
temperature variations, this is found not to lead to signicant error when it is integrated into the wholesystem HPD model. The energy efciency benet associated with the isothermal contact HPD is
conrmed to be a factor of between 2 and 3. An exergy analysis is used to determine the causes of this
perfomance gain. Contact heat transfer in isothermal HPD is found to reduce irreversibility within the
refrigerant cycle by roughly the same amount as that occurring in heat transfer from the condenser to the
product.
2011 Elsevier Ltd. All rights reserved.

Keywords:
Drying
Heat pump
Isothermal
Exergy
Energy efciency

1. Introduction
The present work centers on a simple idea for increasing the
energy efciency of heat pump dryers (HPDs) [1,2]. The idea arises
naturally from consideration of the Gouy-Stodola law [3]. In
a standard (adiabatic) heat pump drying system, recycled heat is
returned to the drying process by heating the dehumidied
airstream as it is recycled to the product. But from a second-law
viewpoint, this approach e using air for heat transfer e appears
wasteful. Heat transfer both to the air and from the air to the drying
process is responsible for a large part of the entropy creation in
such a system [4,5]; so are irreversible losses owing to the air ow
resistance of the system, and fan losses [6]. Jonassen et al. [7, 8]
have successfully reduced the effect of these irreversibilities by
constructing a two-stage non-adiabatic HPD e a full cycle of the
energy efciency work-ow discussed by Asprion [9]. The idea
behind the present work is that these irreversibilities could be
further reduced by providing heat directly to the drying process,
through a conductive plate connecting the refrigerant and the
product (condenser plates labelled CD2 in Fig. 1). This conguration, which we will refer to as the isothermal contact HPD (ICHPD)
mode, would be most applicable to the drying of products that can
be spread into thin layers, especially those that can be dried under

* Corresponding author. Tel.: 64 3 4797796; fax: 64 3 479 0964.


E-mail address: wcatton@physics.otago.ac.nz (W. Catton).
0360-5442/$ e see front matter 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.energy.2011.03.038

high-humidity conditions [10]. The system depicted in Fig. 1


becomes equivalent to a standard adiabatic HPD if the refrigerant
bypasses the condenser plates (dotted line 2e20 ).
The idea of ICHPD has itself been advanced in the literature
previously [11], but the modelling of ICHPD is not straightforward,
because of the tightness of the linkage between the refrigerant
circuit, the drying process, and the air circuit. Perhaps partly for this
reason, no detailed modelling has apparently been reported prior
to the work described here. Nor has there been any reported
attempt to build and test a prototype system. A previous preliminary assessment of ICHPD previously indicated that such a system
conguration offers the potential for energy performance benets
of 2e3 times compared with an adiabatic HPD [10]. However, that
assessment was based on a highly idealised model of the HPD
system. The heat pump heating COP was simply assumed to be
a constant percentage (50%) of the Carnot COP. Also the model
assumed that the temperatures of the product surface (Ts) and of
the bulk duct air (Tb) were everywhere equal to that at the drier
inlet, i.e. location D in Fig. 1, an assumption we call the idealised
isothermal case. Thus although the analysis was able to satisfactorily match measured data in the adiabatic case, substantial
uncertainty remained in its predictions for the isothermal mode.
In the present work we corroborate and extend the previous
analysis, by developing a detailed physical model of a stack of drier
ducts with embedded refrigerant condenser tubes, and incorporating this model into a detailed model of the remainder of an HPD
system, previously established by Carrington and Bannister [12].

W. Catton et al. / Energy 36 (2011) 4616e4624

4617

Fig. 1. Schematic of HPD system being modelled. Locations on the air cycle are denoted by letters, and on the refrigerant cycle are denoted by numbers.

We present and discuss results from the resulting comprehensive


whole-system model, including an exergy audit comparing the
adiabatic and isothermal modes, and clarifying the system-level
effects of contact heat transfer. The detailed model also enables us
to assess the impact on predicted system performance of deviations
from several of the idealisations that were used in the preliminary
assessment, in particular (1) deviations from the isothermal and
adiabatic duct idealisations, and (2) an additional pressure drop
that arises when refrigerant passes through the condenser plates
labelled CD2 in Fig. 1.
We now briey discuss the range of potential applications for
ICHPD technology. The modelling work presented in this paper
conrms that ICHPD promises an increase in drying energy efciency that could yield signicant operational cost reductions,
especially in HPD applications with large product throughput. On
the other hand, the relative capital cost of ICHPD is presently
unknown. An ICHPD of a given drying capacity would require
a signicantly smaller compressor than its adiabatic HPD counterpart (see below); however, the technology would employ a signicantly larger heat transfer area at the condenser than the adiabatic
HPD. Furthermore, the energy performance gain offered by ICHPD
is highly sensitive to the acceptable relative humidity at the
product, and to product thickness. These sensitivities restrict the
range of products to which ICHPD is likely to be applicable, and may
also impact on ICHPD capital costs. In Section 3, we present
a tentative economic analysis of ICHPD in a hypothetical lter-cake
sludge drying case study. The analysis shows, in particular, that the
viability of ICHPD as a substitute for adiabatic HPD may be sensitive
to electricity and waste disposal price trends.
The dimensions of the batch-operation tray drier conguration
considered throughout this paper are specied in Table 1. ICHPD
could of course be implemented under a range of different dryer
congurations. For instance an internal-drum contact dryer, operating continuously with product throughput maintained by rotating
wipers [13], could offer steady-state operation allowing the system
to be held near to optimal operating conditions [14], and also would
offer a mechanism for maintaining a thin product layer with good
thermal contact with the heat transfer surface. Such varying designs
would diverge in the specics of their performance, but our broad
conclusions are unlikely to require signicant modication.

2. Theory
Here we describe the detailed drier-duct model and, briey, the
whole-system model. In the detailed duct model, discretised
control-volume conservation equations are numerically solved in
a 1-D plug-ow model of a drier duct (between locations D and E in
Fig. 1), using the staggered-grid SIMPLER algorithm described by
Patankar [15], subject to boundary conditions given by the inlet air
ow velocity and psychrometric state, the outlet air pressure pE,
and the refrigerant inlet enthalpy hr,2, outlet pressure pr,20 , and mass
_ r . Employing Ficks and Fouriers laws, and the Newow rate m
tonian model of viscous stress, and adopting a 1-dimensional plugow model, the steady-state (species-k) mass, momentum and
energy balance equations can be expressed for the air-side control
volume depicted in Fig. 2, as follows [16e18]:

Z 

rk y  rDva
Sw



Z 
vuk
vu
_ m
rk y  rDva k dS  m
dS
k
vx
vx

(1)

Se

Table 1
Baseline values of key system parameters. Entries marked by asterisks are subject to
variation in the whole-system model.
Parameter (unit)

Baseline value

Condenser face area Aco (m2)


Evaporator face area Aev (m2)
Number of ducts ND
_ a;co (kg s1)
Product air ow m
Drier maximum temperature TD ( C)
Relative humidity at drier inlet fD (%)*
Tray drier length L (m)
Tray drier width w (m)
Air duct depth d (m)
Product thickness d (m)
Heating plate condenser tube diameter D (m)
Heating plate condenser tube midline depth xp (m)
Refrigerant circuits per plate nb (e)
Passes through plate per circuit p (e)
_ r (kg s1)*
Total refrigerant mass ow rate m
Refrigerant saturated condensing temperature Trsat,3 ( C)*

1.0
1.0
10
1.0
55.0
30.0
5.0
1.0
20.0  103
1.0  103
10.0  103
6.0  103
3
5
1.0
60.0

4618

W. Catton et al. / Energy 36 (2011) 4616e4624

at Tr  Ts nv Dhvap h$ Ts  Tb

Fig. 2. Close-up of a single duct. Temperature gradient from refrigerant to air is shown.
Surfaces (Sw, Se, Sn, Sm) of an air-side control volume are indicated.

Z 
Sw



Z 
vy
vy
ry2 p  2m
ry2 p  2m
dS
dS  Ff
vx
vx

Z  

r h y2 y k

Sw

Se

m
 Q_

(3)

In Eqs. (1)e(3), the following denitions have been used:

_ m
m
k



uk rys h$m rk;s  rk;b dS

(4)

Sm

Z
Ff 

1
Cf$ ry2 dS
2

(2)




Z  
vT
1
vT
r h y2 y k
dS
dS Q_
vx
2
vx

1
2

Eq. (8) states that, per unit area at the product surface, the rate of
heat delivery from the refrigerant, at(Tr  Ts), equals the heat
consumed by vaporization, nvDhvap, plus the heat lost at the product
surface to the air above, h$(Ts  Tb). The term at (1/a1 1/a2 1/
a3)1 is the overall plate-product heat transfer coefcient,
a1 arpD/l is an effective heat transfer coefcient from the refrigerant to the inside surface of the tubes, a2 2pkp =lOln
2l=pDsinh2pxp =l is the effective heat transfer coefcient
through the plate, and a3 kd/d is the effective heat transfer coefcient through the product [10]. Eq. (8) is also valid for the adiabatic
mode, which can be simulated by setting the plate thermal
conductivity, kp, to zero. Considering only the constant-activity
period of drying, the humidity ratio at the product surface, us, is the
saturated value at Ts. The mean refrigerant heat transfer coefcient
ar is evaluated using the method of Fischer and Rice [21, p. 51], where
the local refrigerant heat transfer coefcient a(hr,pr)is evaluated
using the method of Cavallini et al. [22].
The idealised drier model, which was used in the previous
analysis, plays an important role in the validation of the detailed
duct model. The idealised model is based on the following
expression for the drier outlet humidity ratio uE, using the
isothermal assumption Ts TD [10] or the adiabatic assumption
Ts Twb,D, i.e. in the adiabatic case, product surface temperature Ts
given by the wet-bulb temperature at D [23]:

uE usat Ts uD  usat Ts exp 

Se

(5)

(8)

hm ra ND w
L
_a
m


(9)

Eq. (9) can be derived (for the present geometry) using the
constant Ts assumptions described above, together with the
assumptions of constant mass transfer coefcient hm; constant dryair density ra; and the approximation uv  ua, which allows
m
to be neglected. The pyschroconvective enhancement of my
metric state at E is obtained using uE and the inlet wet-bulb
temperature Twb,D (in the adiabatic case), or the inlet temperature
TD in the isothermal case [10]. Varying L in Eq. (9) allows the
variation of air state within the duct to be evaluated.
The whole-system model used in the present work, which is the
model of Carrington and Bannister [12] modied to include the
detailed drier-duct model described above, incorporates the isentropic and volumetric efciencies of the ZR61K2-TFD scroll
compressor with R134a characterised by [24]. The NewtoneRaphson method is applied to the state vector x (x1, x2, x3),
where, in terms of the locations shown in Fig. 1:

Sm;n

Q_

h$ Ts  Tb dS

(6)

Sm;n
m

Q_

X
k a;y

_
hk m
k

(7)

In Eqs. (1)e(7), the bounding surface of the control volume is


divided into the entrance-surface Sw, the exit-surface Se, the mass
transfer (product) surface Sm, and the duct top surface Sn, as shown
in Fig. 2. In Eqs. (4)e(7), h$m , Cf$ and h$ are the local mass transfer
coefcient, friction factor and heat transfer coefcient, respectively,
all adjusted for high mass transfer rates by the method of Bird et al.
[18, p. 661]. These have been calculated using the DittuseBoelter
equation [19] and the ChiltoneColburn analogy [18]. The steadystate product surface temperature for each control volume is
evaluated by solving the energy balance at this surface [20, p. 222],
using

x1 Trsat;1

(10)

x2 Trsat;3

(11)

x3 uD

(12)

Using the compressor model and assuming isenthalpic throttling, the two saturated refrigerant states Trsat,1 and Trsat,3 are
sufcient to specify the refrigerant thermodynamic cycle and mass
_ r . The pressure drop over each refrigerant ow branch
ow rate m
within the heating plates (in CD2) is evaluated by the method of
Traviss et al. [25]. The remaining refrigerant pressure drops are
evaluated using the correlations of Carrington and Bannister [12].
Subcooling is assumed negligible, and the thermostatic expansion
valve is assumed set to a 5  C superheat at location 1. A constant fan
efciency of 50% is assumed. The air pressure at E is assumed equal
to ambient: pE 101,325 Pa. The air pressure drop within the drier
is evaluated using the SIMPLER algorithm; other air pressure drops
are estimated using correlations developed by Turaga et al. [26] and

W. Catton et al. / Energy 36 (2011) 4616e4624

4619

the dynamic loss coefcients (k-factors) determined by Carrington


et al. [6]. Venting is assumed to be controlled to maintain TD xed.
Thus the air state at D is specied by uD, pD. Air states around the
system are obtained using the drier model and energy and moisture
balances across the HP components.
The error vector D that is used in the NewtoneRaphson method
is:

D1 Q_ ev  fev yev ; Twb;E ; T4

(13a)



D2 Q_ co  fco yco ; TC ; Trsat;3

(13b)

D3 uD  uC

(13c)

where fev and fco are specied by Eqs. (4) and (7) of [12]. These are
empirical component heat transfer correlations based on the
incoming ows. Each iteration of the NewtoneRaphson method
occurs as follows. From the current estimate of the state vector, x,
a current estimate of the states throughout the system is formed,
and the error vector D is evaluated. An estimated value J of the
Jacobian matrix for the system is then used to update the statevector estimate using

x1 x0  r$J 1 D

(14)

where r represents the relaxation that is applied, and x1 is the new


estimate of the system state, as obtained from the previous estimate x0.
The convergence criterion for the whole-system model is that
the maximum error function, across all submodels of the wholesystem model, must be less than a threshold, which is typically set
to 105. Each updated variable is used to generate an error function
for that variable. In the duct model these are the control volume
source terms that arise from the SIMPLER algorithm. For the wholesystem model, these are the magnitudes of the error vector
components D1, D2, and D3. For the remaining system variables, the
error function is the relative change in the variable in the latest
iteration. Since fev and fco represent the heat transfer from the
condenser and evaporator that was measured as a function of the
incoming refrigerant and air ow conditions [12], the components
of the error vector D approach zero only if the heat transfer at these
heat exchangers is consistent with these empirical correlations. In
addition, the global convergence criterion requires that the mass,
momentum and energy balances within the drier ducts be satised.
Thus on convergence the model is guaranteed to satisfy mass and
energy balances across all system components.
3. Results and discussion
3.1. Drier-duct model e comparison with previous results
Before discussing the validation of the detailed drier-duct
model, we examine representative outputs showing key differences between the adiabatic and isothermal modes. Fig. 3 shows
a set of psychrometric paths obtained from the detailed model
under the air inlet condition TD 63  C, uD 0.069 (kg/kg-dry),
corresponding to a relative humidity of 44%. Two dry-air mass ow
_ a 2:25 kg s1 . The
_ a 0:75 kg s1 , and m
rates are considered: m
drier geometry is as specied in Table 1. Locations 0, 1, 2, 3, 4 and
5 m into the duct are indicated using circles for the low ow rate
scenarios and diamonds for the high ow rate scenarios. Variation
of the dashed lines from the idealised isothermal case (which
would yield vertical psychrometric paths) is evident. With contact
heat transfer, a low air ow rate enables the air to receive more heat
as it passes over the product, resulting in a greater air temperature

Fig. 3. Psychrometric paths predicted by the detailed duct model. Tin 63  C,


uin 0.069, m_ a;in 0:75 kg s1 (circles) and 2.25 kg s1 (diamonds).

at a given humidity. In contrast, in the adiabatic case the humidity


ratio and the temperature are functionally related by way of the
constant wet-bulb temperature of the air, and the psychrometric
path traced by the air is unaffected by the air ow rate, although the
duct location at which a given state is attained is affected by the air
ow rate. The isothermal moisture extraction rates (MERs) are
larger than those in the corresponding adiabatic cases.
The low-ow case described in the previous paragraph was
selected to match the steady-state inlet condition that occurs in the
rst ten hours of the timber-drying situation modelled by Sun et al.
[17], which itself was tested against measured data [27,28]. The
adiabatic timber kiln modelled had similar duct dimensions
(5.76 m  0.02 m as seen from the side) and an inlet air velocity of
4 m s1. With an inlet velocity of 4.0 m s1, the present detailed
duct model predicts a change in the vapour mass fraction of
0.0041 kg kg1, which is in fair agreement with the value of
0.0045 kg kg1 that can be read from Fig. 2 of [17]. The low-ow
case leads to a pressure drop in the isothermal mode of 33 Pa. In the
adiabatic mode with the same inlet condition, the pressure drop is
28 Pa. This latter value is also in fair agreement with the pressure
drop of 20e25 Pa found by Sun et al. [17]. The disagreement
between the predictions of the present model in the adiabatic
mode and Sun et al.s model can be attributed to the different
transfer correlations employed in the two models, and to the
absence of evaporation from the duct ceiling (Sn in Fig. 2).
3.2. Drier-duct model e comparison with analytical case
Since the idealised model represents a limiting case to which an
analytic solution exists, it can contribute to the validation of the
detailed duct model. Fig. 4 and Table 2 provide illustrative
comparison between outputs from the two models. The output of
the detailed air ow model, shown in Fig. 4(b), is seem to be in good
agreement with the idealised case, Fig. 4(a). Once again, the
modelled situation involves the drier geometry specied in Table 1.
Air and refrigerant inlet conditions are as specied in Table 1. The
detailed model incorporates several physical effects not accounted
for in the idealised model. The most important, which is visible in

4620

W. Catton et al. / Energy 36 (2011) 4616e4624

Fig. 4. Example proles (from D to E) (a) idealised model, (b) detailed model. 1: Tb (ISO); 2: Ts (ISO); 3: Tb (ADI); 4: u (ISO); 5: Ts (ADI); and 6: u (ADI).

Fig. 4, is that the product surface and bulk air temperatures may
vary substantially from their idealised values. The detailed model
shows a dip in the air temperature (line 1) due to a relatively low
product surface temperature at the inlet (line 2). This is due to
evaporation being most intense near the air inlet, where the
airstream is least humid. The product surface temperature
increases with position in the drier, and between 3 and 4 m into the
drier the surface temperature can be seen to exceed the air
temperature. Beyond this location the bulk air temperature
increases with position. Another effect visible in Fig. 4(b) is a attening of the humidity ratio curve (line 4) compared with its
counterpart in Fig. 4(a). Since the humidity ratio gradient is
proportional to the local drying rate, this implies that the drying
rate predicted by the detailed model is more uniform throughout
the drier than predicted by the simple model. This effect can be
understood as follows. The driving force for evaporation in the
isothermal mode is approximately proportional to the vertical
separation between lines 2 and 4 in Fig. 4, since line 2, which
represents the product surface temperature, also provides a rstapproximation measure of the surface vapour density. The positive
gradient of line 2 thus has the effect of reducing the variation in the
driving force for drying along the length of the drier, which is
reected in the reduced curvature of line 4.
The MERs and air outlet temperatures estimated by the detailed
and simple models are summarized in Table 2 for a range of inlet air
temperatures and relative humidities. In the isothermal mode the
outlet air temperatures do not deviate markedly from the inlet
temperatures, but the MERs do in some cases vary signicantly
from those predicted by the simple model. The key reason for these

variations again appears to be that the product surface temperature


may deviate signicantly from the air inlet temperature. This
hypothesis has been tested by forcing the product surface
temperatures to equal the air inlet temperature (in the adiabatic
case, the inlet wet-bulb temperature) in the detailed model, with
results shown in the third column of Table 2. These show close
agreement, within a few percent for all the scenarios tabulated
here, with the MER predicted by the idealised model. In addition, in
the isothermal case the curvature of the bulk relative humidity
matches the idealised model, which supports the discussion in the
previous paragraph.
The above analysis has shown that deviation of the product
surface temperature from its idealised isothermal value, as used in
the simple model, may have a signicant effect on the drying rate.
At moderate drying rates the idealised and detailed models are in
good agreement; however at very high and very low drying rates,
correspondingly depressed and elevated surface temperatures
(respectively) may have a substantial effect on evaporation within
the dryer, as is shown by Table 2. Taken all together, these results
indicate the following: (1) the idealised model provides a reasonable rst-order approximation; (2) the detailed model may be
required in order to obtain accurate predictions of HPD performance, because of signicant temperature effects that occur at low
and high drying rates; and (3) the other additional physical effects
that have been included in the detailed model can be regarded as
minor corrections, rather than as primary aspects of the situation
being modelled.

3.3. Whole-system model


Table 2
_ a 1.
Estimates of system behaviour. m

ISO fin 30%


Tin 55  C
ADI fin 30%
Tin 55  C
ISO fin 60%
Tin 40  C
ISO fin 60%
Tin 55  C
ISO fin 90%
Tin 55  C
ISO fin 60%
Tin 70  C

Idealised model

Detailed model

Detailed with Tideal

MER 199.3 kg h1


Tout 55.0  C
MER 21.2 kg h1
Tout 41.8  C
MER 50.4 kg h1
Tout 40.0  C
MER 122.6 kg h1
Tout 55.0  C
MER 33.1 kg h1
Tout 55.0  C
MER 314.7 kg h1
Tout 70.0  C

MER 179.4 kg h1


Tout 53.9  C
MER 19.3 kg h1
Tout 42.4  C
MER 73.8 kg h1
Tout 42.2  C
MER 131.9 kg h1
Tout 55.4  C
MER 80.2 kg h1
Tout 56.9  C
MER 206.6 kg h1
Tout 68.0  C

MER 200.3 kg h1


Tout 54.4  C
MER 20.0 kg h1
Tout 42.4  C
MER 50.4 kg h1
Tout 41.8  C
MER 120.6 kg h1
Tout 55.4  C
MER 30.7 kg h1
Tout 56.5  C
MER 315.4 kg h1
Tout 68.6  C

We now consider an illustrative output from the detailed wholesystem model produced by integrating the detailed dryer-duct
model into the HPD model established by Carrington and Bannister
[12], as described in the theory section above. In the case that we
consider, the two modes again have identical specications
(Table 1) aside from plate heat transfer and plate refrigerant pressure drop, both of which are zero in the adiabatic case. Fig. 5 shows
the refrigerant thermodynamic state-cycle, for the isothermal and
adiabatic modes. The corresponding psychrometric cycles are
shown in Fig. 6. The resulting system performance is summarized
in Table 4. In the isothermal mode the total refrigerant pressure
drop in the condenser (between the locations labelled 2 and 3 in
Fig. 5) is somewhat greater than that for the adiabatic mode, owing
to the pressure drop within the condenser plates. Despite this

W. Catton et al. / Energy 36 (2011) 4616e4624

4621

Fig. 5. Thermodynamic cycle of R134a in the isothermal and adiabatic modes.

additional pressure drop, the isothermal mode can be seen to


enable the system to operate over a signicantly narrower pressure
range, which contributes to an enhancement of the specic moisture extraction rate (SMER). Thus the model predicts that the
refrigerant-side tradeoff between heat- and momentum-transfer
irreversibilities does not signicantly impact on performance in the
isothermal mode. The nominal power rating of the compressor is
5 kW, but the compressor power varies with operating condition
(Table 4). Despite this fact, and despite some deviation from the
idealised temperatures, the psychrometric cycles shown in Fig. 6
_ P 5 kW psychrometric
are in good agreement with the W
cycles of the preliminary analysis [10].

Table 3
Specic exergy destruction by component; key performance indicators.
(103), kWh kg1

Adiabatic

Isothermal

Condenser and product


Compressor
Evaporator
Throttle
Fan friction
Venting and condensate

103.7
62.4
31.1
26.6
3.1
14.1

17.5
21.2
22.4
4.3
1.5
5.8

86.2
41.2
8.7
22.3
1.6
8.3

Table 3 shows the exergy destroyed per kg of moisture removed


from the product in key system processes and components, and the
difference in this quantity, for each component, between HPD
modes. In evaluating the exergies (and the effect of venting), the
environment has been taken to be moisture-saturated at 10  C.
About half of the irreversibility-avoidance achieved by the
isothermal mode is seen to be associated with the condenser and
the drying process. Since most of the exergy destruction in the
condenser and product is associated with the transfer of heat [4,5],
this portion of the avoided irreversibility can be attributed chiey
to the isothermal modes avoidance of heat transfer through air
thermal boundary layers. Most of the rest of the energy efciency
gain is at the compressor and throttle, and can be attributed to the
narrower temperature and pressure range of the refrigerant cycle
for the isothermal mode. Most of this temperature-range narrowing is due to the avoidance of air cooling in the drier ducts, and also
to the high humidity that prevails in the isothermal mode, as can be
seen by examining the cycles shown in Fig. 6. We can therefore
associate most of this latter improvement with the fact that the
isothermal mode avoids using air (with its small specic heat

Table 4
Performance of adiabatic and isothermal HPD.

Q_ co , kW
_ t , kW
W
Fig. 6. Psychrometric chart showing air property paths in baseline scenario for adiabatic and isothermal dryers.

MER, kg h1
SMER, kg kWh1

Adiabatic

Isothermal

27.8
5.1
21.2
4.2

43.5
4.0
55.4
13.7

4622

W. Catton et al. / Energy 36 (2011) 4616e4624

layers less than about 1e2 cm thick. However, Fig. 7(b) shows
a signicant potential benet of the ICHPD mode, where it is
applicable. As the gure indicates, ICHPD may enable energy
performance (SMER) and MER to be maximised simultaneously (by
using a thin product layer). This absence of a tradeoff between
SMER and throughput contrasts with adiabatic HPD systems, which
must be operated at relatively low drying rates to obtain good
energy performance [10].
3.4. Economic case study

Fig. 7. Relationship between d, SMER and MER.

capacity) as a heat carrier, and with the isothermal modes high


humidity.
In summary: (1) in the adiabatic mode, condenser and product
irreversibilities contribute much of the overall work requirement;
(2) the isothermal mode greatly reduces this irreversibility, by
about six-fold per kg moisture removed; (3) a signicant part of the
overall reduction of irreversibility nevertheless occurs at the
compressor and the throttle. This last result highlights the synergistic nature of a HPD, and also implies that the exergy destruction
in the condenser and product of an adiabatic HPD does not set an
upper bound on the SMER gain associated with ICHPD. (Indeed if it
did, then from the rst column of Table 3 the isothermal SMER
could not exceed 7.3 kg kWh1.)
We now consider an interesting feature of the impact of the
product thickness on ICHPD SMER. Fig. 7(a) shows the effect that
product thickness has on the energy performance of the isothermal
and adiabatic modes. Since we are considering the constantactivity drying period, and convective heat transfer to the product
is not affected by its thickness, the adiabatic performance is unaffected by product thickness. In contrast, a thick product layer
represents a signicant thermal resistance, which nullies the
benet of the isothermal mode. As has already been discussed, the
requirement that the product must be able to be spread thinly
(together with the typically very high humidities in the isothermal
mode) limits the products for which isothermal drying will be
appropriate; a glance at Fig. 7(a) suggests that the contact HPD
system being modelled would yield a signicant performance
advantage only in the drying of products that can be spread into

Finally, we use the system performance values presented above


to conduct a tentative analysis of the relative economics of ICHPD.
We consider an operation that produces 1000 kg of waste sludge
daily, with an initial moisture content of 0.65 kg/kg (dry-mass
basis). A waste sludge drying operation has been selected for the
following reasons. (1) Adiabatic HPD of lter-cake sludge is used
today. (2) The value that is added to the product by drying is typically not large compared with the energy cost involved. (3) Waste
sludges could be dried under ICHPD conditions that are set to
optimize energy performance. This sludge is to be dried to a nal
moisture content of 0.1 kg/kg, corresponding to a required drying
capacity of 25.625 kg/h, in order to reduce transport and landlling
costs [29]. Using the rule of thumb that the capital cost of adiabatic
HPD is approximately $1 per watt of heat provision at the condenser,
we estimate a capital cost of $33,250 for the adiabatic case. We
assume that this cost is nanced at an annualized rate of 7%.
Table 5 shows the value added per day (after electricity costs
have been met), the payback time, and the net present value (NPV)
of installation, for adiabatic and isothermal HPD systems, over
a range of scenarios. We do not consider any costs associated with
depreciation, servicing or labour, and in evaluating the NPV we
assume a long project lifetime. The parameters that are varied are
the cost per unit of electricity, the cost of sludge transport and
landlling, and the relative capital cost of an ICHPD system,
compared with adiabatic HPD. In each scenario each independent
parameter is either high or low. Scenarios 1e4 reect current
electricity prices of roughly 10c kWh1. Scenarios 5e8 correspond
to a signicantly higher electricity price of 30c kWh1, and could
represent a future scenario characterised by energy shortages and/
or a strong CO2-emission price signal [30,31]. Scenario 6 (in which
electricity is expensive, sludge disposal is cheap, and ICHPD is
costly to install) is the only scenario in which neither adiabatic HPD
nor ICHPD is economically feasible. ICHPD appears to be a sensible
investment in scenarios 1, 3, 5, 7, and 8. Both scenarios 3 and 4 have
high disposal costs. The economics favour adiabatic HPD somewhat
in 3, and isothermal HPD somewhat in 4. In all of the high-electricity cost scenarios 5e8, ICHPD is strongly favoured. In particular,

Table 5
Economics of adiabatic HPD and ICHPD.
Scenario
Electricity cost, $/kWh
Disposal costs, $/kg
Relative capital cost (ISO)

2
0.1
0.05
1

0.1
0.05
3

4
0.1
0.15
1

Value added, $/day


ADI
ISO

16.1
26.3

16.1
26.3

77.6
87.8

77.6
87.8

Payback time, years


ADI
ISO

7.2
4.0

7.2
18.1

1.2
1.1

1.2
3.5

53.4
108.1

53.4
41.6

384.5
439.2

384.5
372.7

Net present value, k$


ADI
ISO

0.1
0.15
3

0.3
0.05
1
13.2
17.3
e
6.6
104.2
59.8

0.3
0.05
3
13.2
17.3
e
e
104.2
-6.7

8
0.3
0.15
1

0.3
0.15
3

48.3
78.8

48.3
78.8

2.0
1.2

2.0
4.0

226.9
390.9

226.9
324.4

W. Catton et al. / Energy 36 (2011) 4616e4624

in scenario 5 ICHPD is a sensible investment while adiabatic HPD


yields a negative net value, and in scenario 8 ICPHD is preferable to
adiabatic HPD despite its much higher up-front cost. Electricity and
waste disposal costs vary with region, while the relative capital cost
of ICHPD is currently unknown. Our results indicate that at present
the economic viability of ICHPD strongly depends on its capital cost
being less than three times that of adiabatic HPD, but that this
dependency could be lessened or reversed by increases in the costs
of sludge disposal or of electricity. Viewed alternatively, Table 5
shows that the isothermal modes high energy efciency makes
its economics relatively less sensitive to the price of electricity,
a result which may be signicant in a time of energy-price
uncertainty.
4. Conclusions
This paper has described the development and testing of
a exible nite-volume drier-duct model, which solves the mass,
momentum and energy balances for the ow within a stack of drier
ducts. This detailed model has been used to examine in detail the
drying process in quasi-isothermal and adiabatic driers. The model
has shown that the idealisation of purely isothermal conditions
within a contact HPD is the most severe restriction applying to the
idealised drier-duct model that was used in a previous analysis of
isothermal contact heat pump dehumidiers (ICHPD). Other
assumptions have been shown to make a much smaller difference
to the accuracy of the idealised model predictions.
By comparing the isothermal drying behaviour as predicted by
the detailed model with the idealised model used in the previous
analysis, we have tested the air ow model against a limiting case
which admits of analytical solution techniques. By comparing
model outputs in the adiabatic mode with those of the model
developed by [17], which itself was tested directly against
measured data, we have indirectly tested the air ow model against
experimental observations. These tests justify our condence that
the model robustly represents drying within a duct-stack. Our
investigation has conrmed that under moderate drying conditions
the idealised model used previously provides a satisfactory rstapproximation model of the drying process. However it has also
provided insight into several ways in which a contact drier, operated as part of an ICHPD system, would deviate from the idealised
isothermal case.
A whole-system HPD model has been produced by combining
the detailed dryer-duct model with models of the remaining heat
pump components. The system model has been used to conrm the
previous nding that isothermal contact dehumidication drying
can increase drying energy efciency by 2e3 times compared with
conventional adiabatic dehumidication HPDs. Deviation from
purely isothermal behaviour has been found to inuence the MER
and SMER that is predicted by the whole-system model by only
a few percent, suggesting that the idealised duct model can legitimately be used in assessments of ICHPD performance.
An exergy analysis of the system has been performed for both
the isothermal and adiabatic modes. This has shown that the
isothermal mode derives its relative energy efciency partially, but
by no means entirely, from a reduction in the irreversibility associated with the transfer of heat to the drying process. Signicant
contributions to the SMER improvement also occur at the
compressor and throttle, and have been attributed mainly to the
high humidity throughout the ICHPD system and to the avoidance
of air as the primary path for heat transfer to the product.
We have examined the relationship between the ICHPD SMER,
the product thickness d, and the product throughput as indicated by
the MER. Our results indicate that ICHPD provides an opportunity
to avoid the adiabatic modes tradeoff between drying rate and

4623

energy efciency by using a thin product layer. Thus obtaining good


performance from an ICHPD system boils down to the materialshandling problem of maintaining a sufciently thin product layer
that has good thermal contact with the heat transfer surface.
We have presented a case study of the economics of ICHPD.
Since we are not yet in a position to assess the likely capital cost of
an effective ICHPD system, we are currently unable to further assess
its economic viability. However, we have shown that the viability of
ICHPD would become substantially less sensitive to capital costs if
either waste disposal or electricity prices increase signicantly, and
that isothermal HPD provides an opportunity to minimize risks due
to uncertain electricity prices.
References
[1] Carrington G, Scharpf E. Dehumidier drier for pastes, liquors and aggregate
materials. NZ Patent 526648; 2004.
[2] Catton W, Sun Z, Carrington G. Exergy analysis of an isothermal heat pump
dryer. Chem Eng Trans 2010;21:139e44.
[3] Bejan A. Advanced engineering thermodynamics. 2nd ed. John Wiley; 1997.
[4] Carrington G, Baines P. Second law limits in convective heat pump driers. Int J
Energy Res 1988;12:481e94.
[5] Vaughan G, Carrington C, Sun Z. Exergy analysis of a wood-stack during
dehumidier drying. Int J Exergy 2007;4(No. 2):151e67.
[6] Carrington G, Sun Z, Sun Q, Bannister P, Chen G. Optimizing efciency and
productivity of a dehumidier batch dryer. Part 1: capacity and airow. Int J
Energy Res 2000;24:187e204.
[7] Jonassen O, Kramer K, Strmmen I, Vagle E. Non-adiabatic two-stage countercurrent uidized bed drier with heat pump. In: Ninth international drying
symposium, Gold Coast; 1994, p. 511e517.
[8] Strmmen I, Eikevik T, Alves-Filho O. Operational modes for heat pump
drying e new technologies and production of a new generation of high quality
sh products. In: 21st International congress of refrigeration, Washington, DC,
vol. ICR0068; 2003. No. 3.
[9] Asprion N, Rumpf B, Gritsch A. Work ow in process development for energy
efcient processes. Chem Eng Trans 2009;21:1381e6.
[10] Catton W, Carrington G, Sun Z. Performance assessment of contact heat pump
drying. Int J Energy Res; 2010. doi:10.1002/ER.1704.
[11] Carrington G. Heat pump and dehumidication drying. In: Food drying
science and technology: microbiology, chemistry, applications. DEStech
Publications, Inc.; 2007.
[12] Carrington G, Bannister P. An empirical model for a heat pump dehumidier
drier. Int J Energy Res 1996;20:853e69.
[13] Chen G, Yue P, Mujumdar A. Sludge dewatering and drying. Dry Technol 2002;
20(No. 4):883e916.
[14] Strmmen B, Eikevik I, Neksa P, Pettersen J, Aarlien R. Heat pumping systems
for the next century. In: 20th International Congress of Refrigeration. Sydney:
IIR/IIF; 1999.
[15] Patankar S. Numerical heat transfer and uid ow. New York: McGraw-Hill
Book Company; 1981.
[16] Slattery J. Momentum, energy and mass transfer in continua. New York:
McGraw-Hill; 1972.
[17] Sun Z, Carrington G, Bannister P. Dynamic modelling of the wood stack in
a wood drying kiln. Trans IChemE 2000;78:107e17.
[18] Bird R, Stewart W, Lightfoot E. Transport phenomena. New York: Wiley; 1960.
[19] Rohsenow W, Hartnett J. Handbook of heat transfer. New York: McGraw-Hill;
1972.
[20] Keey R. Introduction to industrial drying operations. Pergamon Press; 1978.
[21] Fischer S, Rice C. The Oak Ridge heat pump models: I a steady-state computer
design model for air-to-air heat pumps. Technical report. USA: Oak Ridge
National Laboratory; 1983.
[22] Cavallini A, Del Col D, Doretti L, Matkovic M, Rossetto L, Zilio C. Condensation
in horizontal smooth tubes: a new heat transfer model for heat exchanger
design. Heat Transfer Eng 2006;27(No. 8):31e8.
[23] Menon A, Mujumdar A. Drying of solids: principles, classication, and selection of dryers. In: Mujumdar A, editor. Handbook of industrial drying. New
York: Marcel Dekker; 1987.
[24] Carrington G, Bannister P, Liu Q. Performance of a scroll compressor with
R134a at medium temperature heat pump conditions. Int J Energy Res 1996;
20:733e43.
[25] Traviss D, Rohsenow W, Baron A. Forced-convection condensation inside
tubes: a heat transfer equation for condenser design. Part 2A. ASHRAE Trans
1973;79(No. 1):157e65.
[26] Turaga M, Lin S, Fazio P. Correlation for heat transfer and pressure drop factors
for direct expansion air cooling and dehumidi-fying coils. ASHRAE Trans
1988;94(No. 2):616e29.
[27] Bannister P, Bansal B. Heat pump dehumidier timber drying. Final trial and
general summary of ndings: Atlas timber trials. Technical report EROL-RR03. Dunedin, New Zealand: Energy Research Otago Ltd., ISBN 1-877157-01-5;
1996.

4624

W. Catton et al. / Energy 36 (2011) 4616e4624

[28] Bannister P, Bansal B, Carrington G, Sun Z. Impact of kiln losses on a dehumidier drier. Int J Energy Res 1998;28(No. 22):515e22.
[29] Macolino P, Bianco B, Veglio F. Drying process of a biological industrial sludge:
experimental and process analysis. Chem Eng Trans 2009;17:699e704.
[30] Ghoshray A, Johnson B. Trends in world energy prices. Energy Econ 2010;32:
1147e56.
[31] Hauch J. Electricity trade and CO2 emission reductions in the Nordic countries.
Energy Econ 2003;25:509e26.

Symbols (units)
A: area (m2)
ADI, ISO: adiabatic, isothernal
Cf: friction factor (1/m2)
COP: coefcient of performance (e)
D: diffusivity (m2/s)
d: air duct depth (m)
D: heating plate refrigerant tube internal diameter (m)
Ff : friction force (N)
Dhvap: latent heat of vaporization (J/kg)
h: specic enthalpy of moist air (J/kg dry-air), product
surface heat transfer coefcient (W/m2 K)
hm: product surface mass transfer coefcient (m/s)
k: thermal conductivity (W/K m)
l: heating plate condenser tube spacing (m)
L: heating plate length (m)
_ mass ow rate (kg/s)
m:
MER, SMER: moisture extraction rate (kg/s),
specic moisture extraction rate (kg kWh)
n: mass ux (kg/m2$s), refrigerant circuits per plate
ND: number of ducts (e)
p: pressure (Pa), passes through plate per circuit
Q_ : heat ow rate (W)
S: surface

T: temperature (K)
y: speed (m/s)
w: heating plate width (m)
_ power input (W)
W:
x: distance through kiln (m)
xp: heating plate refrigerant tube centerline depth (m)
z: distance along refrigerant ow in CD2 (m)
a: heat exchange coefcient (W/m2 K)
d: product thickness (m)
f: relative humidity (e)
r: density (kg/m3)
m: dynamic viscosity of uid (N s/m2)
u: humidity ratio (kg vapour/kg dry-air)
D: change
Subscripts and superscripts
0: environment
1, 2, 20 , 3, 4: locations on refrigerant cycle
A, B, C, D, E, F: locations on air cycle
b: bulk
co, ev: condenser, evaporator
D, F, P: ducts, fan, compressor
in, out: inlet, outlet
k, a, v, w: species-k, dry-air, water-vapour, liquidewater
n, w, s, e: north, west, south, east (control volume boundaries)
m: mass exchange
p: heating plate
r: refrigerant
S, s: surface
sat: saturation condition
t: Total, effective
wb: wet-bulb
: modied for high mass transfer rates
d: products

Potrebbero piacerti anche