You are on page 1of 12

i n t e r n a t i o n a l j o u r n a l o f r e f r i g e r a t i o n 3 4 ( 2 0 1 1 ) 1 1 9 7 e1 2 0 8

available at www.sciencedirect.com

w w w . i i fi i r . o r g

journal homepage: www.elsevier.com/locate/ijrefrig

Ammoniaewater desorption heat and mass transfer


in microchannel devices

Matthew D. Determan, Srinivas Garimella*


Sustainable Thermal Systems Laboratory, George W. Woodruff School of Mechanical Engineering, Georgia Institute of Technology,
Atlanta, GA 30332, USA

article info abstract

Article history: This paper presents the results of an experimentally validated model for the prediction of
Received 27 September 2010 local heat and mass transfer rates in a microchannel ammoniaewater desorber. The
Received in revised form desorber is an extremely compact 178 mm  178 mm  0.508 m tall component capable of
19 January 2011 transferring the required heat load (w17.5 kW) for a residential heat pump system. The
Accepted 4 February 2011 model predicts temperature, concentration and mass flow rate profiles through the
Available online 11 February 2011 desorber, as well as the effective wetted area of the heat transfer surface. Previous
experimental and analytical research by the authors demonstrated the performance of this
Keywords: same microchannel geometry as an absorber. Together, these studies show that this
Desorption compact geometry is suitable for all components in an absorption heat pump, which would
Miniaturization enable the increased use of absorption technology in the small-capacity heat pump
Mass transfer market.
Microchannel 2011 Elsevier Ltd and IIR. All rights reserved.
Heat pump

Transfert de chaleur et de masse dans un desorbeur a`


ammoniac-eau a` microcanaux
Mots cles : Desorption ; Miniaturisation ; Transfert de masse ; Microcanal ; Pompe a` chaleur

1. Introduction complex thermodynamic cycles have achieved high theo-


retical COPs, this same increased complexity requires the
The residential and light-commercial space-conditioning use of numerous heat exchangers and control systems.
market has long been dominated by the vapor compression A lack of commercially available compact heat exchangers
refrigeration system. The increasing global concern about has impeded the adoption of these thermodynamically
the environmental impact of synthetic refrigerants used in attractive absorption cycles as residential and light-
vapor compression refrigeration systems, and the potential commercial space-conditioning systems. The design of
for low-grade thermal energy utilization has led to renewed compact absorption system components is challenging
interest in absorption systems. Although increasingly because it requires innovative high-flux concepts along with

* Corresponding author. Tel.: 1 404 894 7479.


E-mail address: sgarimella@gatech.edu (S. Garimella).
0140-7007/$ e see front matter 2011 Elsevier Ltd and IIR. All rights reserved.
doi:10.1016/j.ijrefrig.2011.02.004
1198 i n t e r n a t i o n a l j o u r n a l o f r e f r i g e r a t i o n 3 4 ( 2 0 1 1 ) 1 1 9 7 e1 2 0 8

Nomenclature G mass flowrate per length of tube (kg s1 m1)


d liquid film thickness (m)
a wetted area ratio
A area (m2) Subscripts
C concentration (mol m3) a ammonia
cp specific heat (kJ kg1 K1) bulk bulk
c~p molar specific heat (kJ mol1 K1) con concentrated solution
h enthalpy (kJ kg1) des desorbing
LMTD log-mean temperature difference (K) dil dilute solution
M molar mass (kg mol1) dry dry portion
m _ mass flowrate (kg s1) f liquid film
Nu Nusselt number glyc water/glycol solution
n_ 00 molar flux (mol m2 s1) in inlet
Q_ heat transfer rate (kW) int interface
R thermal resistance (K W1) l liquid
Re Reynolds number out outlet
T temperature ( C) p pass
U overall heat transfer coefficient (W m2 K1) s solution side
x liquid mass fraction seg segment
y vapor mass fraction t tube side
y~ vapor molar fraction T total
z~ evaporating flux molar fraction v vapor
w water
Greek symbols
wall tube wall
a heat transfer coefficient (W m2 K1)
wet wet portion
b mass transfer coefficient (m s1)

an understanding of the heat and mass transfer process at can be used for absorbers, desorbers, and other components
a local level. in the cycle. Some of these concepts have yielded high heat
A detailed survey of related literature on ammoniaewater and mass transfer rates in commercial applications, but while
heat and mass transfer was presented in prior papers on designs for use in residential systems must have favorable
miniaturization technology for absorption components heat and mass transfer rates, they must attain these rates
(Garimella, 1999, 2000; Meacham and Garimella, 2002a,b, 2003, with simple and compact geometries.
2004) by this group. A comprehensive review of the theoretical Although most of the literature on heat and mass transfer
and analytical models for heat and mass transfer in ammo- in absorption system components has focused on the
niaewater and LiBr/H2O systems was provided in Killion and absorption process, for potential residential and light-
Garimella (2001), while a similar review of experimental commercial use, compactness must be achieved in all
studies to understand these phenomena appears in Killion components of the system. This paper presents the results of
and Garimella (2003). The reader is referred to those works a combined experimental and analytical study to model
for the background that provides a context for the present ammoniaewater desorption in a prototype heat and mass
study. Thus, prior research efforts have primarily focused on exchanger that utilizes miniaturization technology to address
understanding heat and mass transfer in falling films of deficiencies in currently available designs. The concept was
various fluid pairs on horizontal and vertical tubes, for first reported by Garimella (1999, 2000, 2004) for potential
example (Beutler et al., 1996a; Beutler et al., 1996b; Hoffmann application as any of the components in an absorption cycle,
and Ziegler, 1996). Jeong and Garimella (2002) addressed the with an analytical illustration as an absorber. Experiments
lack of complete wetting of the tube surface by introducing conducted on the first prototype for this concept were repor-
a wetting ratio and studied LiBr/H2O absorption during the ted by Meacham and Garimella (2002a), while a detailed
falling film, droplet formation, and droplet fall regions. They analysis of the local variations in heat and mass transfer
later (Jeong and Garimella, 2005) used this multiple-flow- coefficients along the absorber was presented in Meacham
regime model to show that the use of small diameter tubes for and Garimella (2003). An improved design that addressed
commercial applications yields absorbers that are signifi- solution flow distribution problems and enabled high-speed
cantly smaller than conventional absorbers. Generalized flow visualization of the falling-film heat and mass transfer
models (Potnis et al., 1997) for GAX component and system process was developed and reported in Meacham and
simulation, useful for sizing equipment, estimating the cycle Garimella (2004). Determan et al. (2004) demonstrated
COP, and determining individual phase flow rates, have also through experiments the use of this same device as a compact
been developed. Among the innovations in compact compo- ammoniaewater desorber. The objective of the current work
nents for ammoniaewater systems is the corrugated and is to develop an experimentally validated heat and mass
perforated fin surface placed between rectangular coolant transfer model for ammoniaewater desorption in miniatur-
channels that was developed by Christensen et al. (1998) that ized geometries from their data.
i n t e r n a t i o n a l j o u r n a l o f r e f r i g e r a t i o n 3 4 ( 2 0 1 1 ) 1 1 9 7 e1 2 0 8 1199

capability of this concept to transfer heat duties representative


2. Description of novel features and prior of residential heat pumps in a small envelope, the performance
developments was somewhat lower than that predicted by the preliminary
model of Garimella (1999). In addition, modeling of the local
The miniaturized heat and mass transfer device consists of measured heat and mass transfer variations within the
short lengths of microchannel tubes placed in multiple square prototype absorber (Meacham and Garimella, 2003) indicated
arrays, with successive arrays oriented transversely perpen- that only 20e30% of the absorber surface area might have been
dicular to the adjacent arrays as shown in Fig. 1. The number effectively utilized. From these results, it was inferred that
of tube arrays is determined by design duty requirements. solution distribution problems at the top of the array might be
Hydronic fluid flows in parallel through individual tubes in the responsible for the potential solution mal-distribution. It was
arrays, from the bottom to the top of the device countercur- also clear that the absorber performance was solution-side
rent to the falling solution. Due to the small diameter of the limited, due to the relatively small influence of the coolant flow
tubes, this configuration allows for extremely high tube-side rate on absorber performance. An improved configuration of
heat transfer coefficients even though the flow is laminar. the absorber was developed based on these inferences,
Tube-side pressure drops are maintained within the required primarily with the goal of addressing solution-side flow
limits by flowing the fluid in parallel through multiple tube distribution and surface wetting issues. In addition, optical
arrays. In an absorber application, vapor introduced at the access to the absorption process was provided to enable
bottom of the absorber flows upward countercurrent to the confirmation of the improvement in flow distribution. Experi-
gravity driven falling solution. The vapor is absorbed into ments covering a wide range of solution and coolant flow rates
the dilute solution, yielding a concentrated ammoniaewater and vapor fractions in this revised absorber (Meacham and
stream at the bottom of the absorber. Similarly, in a desorber Garimella, 2004) were used to determine the overall and solu-
application, the concentrated solution is introduced at the top, tion-side heat and mass transfer coefficients. It was found that
with hot hydronic fluid flowing inside the tubes. Ammo- although the surface area of this improved absorber was only
niaewater vapor is desorbed in the process, and rises to the 0.456 m2, approximately 30% of the surface area of the original
top of the desorber due to buoyancy, while the dilute solution prototype absorber, it was able to transfer duties as high as
flows to the bottom due to gravity. This flow arrangement 15.1 kW, almost equaling the load of the original larger
leads to effective vapor-solution contact, minimizing heat and absorber. The solution-side heat transfer coefficient was
mass transfer resistances. The counterflow arrangement reported to be in the range 638e1648 W m2 K1. This signifi-
between the vapor and the solution also leads to higher outlet cant increase in performance is attributed to the substantially
ammonia vapor concentrations, thus reducing the rectifica- improved flow distribution. Visual documentation of the flow
tion load. The low solution-side heat and mass transfer in this absorber also confirmed the significantly improved flow
resistance, coupled with the high tube-side heat transfer distribution and higher participation of the surface in the heat
coefficient, allows the heat of absorption or desorption to be and mass transfer process.
transferred to or from the tube-side fluid effectively. Having demonstrated the absorber function, the original
In the initial implementation as an absorber, the heat and absorber first fabricated and tested by Meacham and
mass transfer model developed by Garimella (1999) was used to Garimella (2002a) was modified slightly (inlets, outlets and
estimate the size of the absorber required for a single-effect distribution device) to allow operation as a desorber. This
10.55 kW cooling load residential heat pump, i.e., an absorption desorber is shown in Fig. 2, with details of the geometry listed
load of about 19 kW, and a prototype absorber was fabricated in Table 1. The heat and mass transfer surface was not
based on these results. While the overall experimental results modified, and remains the same as was used by Meacham and
presented by Meacham and Garimella (2002a) indicated the Garimella (2002a). The microchannel tube array consists of 5

Fig. 1 e Microchannel Heat and Mass Exchanger Concept.


1200 i n t e r n a t i o n a l j o u r n a l o f r e f r i g e r a t i o n 3 4 ( 2 0 1 1 ) 1 1 9 7 e1 2 0 8

Fig. 2 e Prototype Desorber.

passes of 16 tube rows each with twenty-seven 1.575 mm the transverse orientation of the tube array. In the photo-
outside diameter  140 mm long tubes per row for a total graph, although two of the tubes appear warped due to the
surface area of 1.5 m2. Fig. 2(a) shows the desorber tube array brazing process, backlit photographs of the entire tube
before the header plates were brazed on, to reveal the details assembly indicated that light passed through almost every
of the tube spacing and pass arrangement. The coolant tubes tube in the array, confirming that only an extremely small
were brazed to the header assembly with a 4.76 mm tube pitch fraction of the tubes near the top were affected. Fig. 2(c) shows
in both the horizontal and vertical directions. Fig. 2(b) shows the desorber subassembly with the header plates brazed on.
The thermocouple ports for measurement of the hydronic
fluid temperature at each pass are visible as well as the fluid
distribution device above the top tube level.
Table 1 e Desorber geometry. Solution distribution and surface wetting problems
Geometric parameter Value encountered in the first tests of the absorber reported by
Meacham and Garimella (2002a) were attributed to the design
Tube outer diameter, mm 1.575
and fabrication of its drip tray. In that drip tray, minor warping
Tube inner diameter, mm 1.067
Tube wall material and thermal T304 stainless steel,
of the underside of the tray caused the solution to collect at
conductivity 15.4 W m1 K1 preferential locations before falling unevenly over the tube
Tube length, m 0.140 array. To address these solution distribution issues, and to
Number of tubes per row 27 provide passages for the desorbed vapor, a new distribution
Number of rows per pass 16 device (Fig. 2(d) e top view, (e) e bottom view) was developed
Number of passes 5
for the desorber in the present study. This drip tray has 648
Tube transverse pitch, mm 4.76
microchannel tubes (1.27 mm OD) of 15.9 mm length brazed
Row vertical pitch, mm 4.76
into the bottom. These tubes ensure that solution supply
i n t e r n a t i o n a l j o u r n a l o f r e f r i g e r a t i o n 3 4 ( 2 0 1 1 ) 1 1 9 7 e1 2 0 8 1201

droplets form directly above the tubes across the entire array recuperatively subcooled. This also ensures that the solution
and do not coalesce before detaching as large streams at does not flash as it flows through the mass flow meter.
preferential locations. In addition, the height of the tubes in The vapor and dilute solution streams rejoin upstream of the
the tray provides an even gravity head for the inlet solution absorber, which is a coiled tube-in-tube heat exchanger. The
pool before flowing uniformly through the dripper tubes to the heat of absorption is removed by a closed cooling water loop
tube array. The drip tray also has nine vapor bypass tubes which, in turn, is cooled by city water in the plate heat
(12.7 mm OD) that extend above the top edge of the tray exchanger. City and cooling water flow rate variations provide
(Fig. 2(d)) to ensure that the vapor generated within the tube the requisite control of the absorber heat sink. After exiting
array has an unimpeded exit path. During testing, the the absorber, the solution enters a shell-and-tube subcooler to
assembly is enclosed in a flanged stainless steel shell (Fig. 2(f)) ensure that the vapor has been completely absorbed into the
that is bolted to the base flange. The thermocouple probes as solution before it returns to the solution pump.
well as all of the fluid inlets and outlets, except for the vapor Temperature, pressure, and mass flow rate measurement
outlet, enter through the base flange. The vapor exits the locations are shown in Fig. 3. Micromotion Coriolis flow meters
device through a port in the top flange. (0.15% uncertainty) are used for mass flow measurements of
the concentrated solution, dilute solution and glycol/water
solution. The vapor flow rate is determined by the difference
3. Experimental approach and data analysis between the concentrated and dilute solution flow rates. The
absorber cooling water flow rate is measured with a Rose-
A schematic of the heat and mass transfer test facility used for mount magnetic flow meter (0.5% uncertainty). Absolute
the investigation of this desorber is shown in Fig. 3. A variable- pressure measurements of the solution and vapor streams are
speed pump delivers the desired concentrated solution flow made using Rosemount transducers with uncertainties of
rate to the solution heat exchanger where it is preheated 0.25% of span. The tube-side pressure drop in the desorber is
before it enters the desorber. The concentrated solution then measured with a Rosemount differential pressure transducer
flows to the desorber where it is distributed uniformly over the with an uncertainty of 0.075% of span. Type T thermocouples
tube array by the solution drip tray described above, and flows (0.5  C uncertainty) are used to measure the pass-wise
in the falling-film mode over the tubes. A steam-heated glycol/ hydronic fluid temperatures in the headers of the desorber, as
water solution flows upward through the tube array in coun- well as the solution temperatures in the solution drip tray and
terflow with the falling solution. Control of the steam pressure collection tray. The remaining solution temperatures, as well
allows for a range of desorption heat transfer rates. The as the glycol/water temperatures and the absorber cooling
glycol/water loop consists of a pump and a coiled tube-in-tube water temperatures are measured with RTDs (0.5  C to
heat exchanger where steam is condensed to provide the 0.8  C uncertainty). The city water temperatures are
desired heat input. As the ammonia vapor is desorbed, it flows measured with type T thermocouples. A TEMPSCAN data
upwards, countercurrent to the solution, due to buoyancy and acquisition system records all measured data for real-time
is removed at the top of the desorber. The resulting dilute data analysis. Plots of the temperatures and pressures versus
solution is collected in a tray at the bottom of the desorber and time are used to ensure that steady state conditions are
then flows through the solution heat exchanger where it is reached before data for any test condition are recorded.
Temperature and pressure measurements of the dilute
solution and vapor leaving the desorber were used to deter-
mine the concentration at these saturated states. Mass and
species balances were used to determine the concentration of
the entering solution. Once the concentrations were deter-
mined, temperatures and pressures at the desorber inlet and
outlet were used to determine the solution-side heat duty. The
temperatures and flow rate of the glycolewater solution were
also used to calculate the heat duty, to help establish an
energy balance between the two sides of the desorber. Once
the desorber heat duty was known, the log-mean temperature
difference (LMTD) between the solution and the heating fluid
was used to calculate the overall heat transfer coefficient (U ).
The tube-side heat transfer coefficient (at) was calculated
using the Churchill equation (Churchill, 1977a,b) for the Nus-
selt number. The solution-side desorption heat transfer
coefficient (as) was then calculated using a resistance network
approach with a wall resistance of 1.99  105 m2 K W1.
An error propagation technique (Taylor and Kuyatt, 1994)
was used to estimate the uncertainty in the solution-side heat
transfer coefficient. At a representative point with a concen-
trated solution flow rate of 0.0264 kg s1, a vapor flow rate of
6.80  103 kg s1 and a glycol/water flow rate of 0.153 kg s1, the
Fig. 3 e Test Facility Schematic. uncertainties in the measured temperature and pressure values
1202 i n t e r n a t i o n a l j o u r n a l o f r e f r i g e r a t i o n 3 4 ( 2 0 1 1 ) 1 1 9 7 e1 2 0 8

used to determine concentrations yield: xcon 0.4793  0.0021, In this analysis, the desorber is divided into five segments,
xdil 0.3049  0.0025, and y 0.9832  0.0006. These values, with each segment representing one tube-side pass. Also,
combined with the uncertainties in the flow rate, temperature allowance is made for the tube surface not being completely
and pressure measurements, yield a solution-side heat duty of wetted by the solution film. Thus, a is the fraction of the tube
13.09  0.347 kW. The corresponding glycol/water-side heat surface area that is wetted and Ap is the total tube surface area
duty is 12.93  0.355 kW, which results in an average desorber of one pass of the exchanger. Mass and species balances on
heat duty of 13.01  0.248 kW. This heat duty and an LMTD of the segment provide the outlet flow rates and concentrations
17.6  0.56  C yield an overall U of 493.9  20.11 W m2 K1. The as follows:
tube-side heat transfer coefficient, at, with an assumed uncer-
_ l;in m
m _ v;in m
_ l;out m
_ v;out (4)
tainty of 25%, is calculated to be 1419  354 W m2 K1. Finally,
the solution-side heat transfer coefficient is calculated to be
as 1037  293.2 W m2 K1. _ l;in xin m
m _ v;in yin m
_ l;out xout m
_ v;out yout (5)
~ the molar concentration
In the above equations, the value of z,
of the evaporating flux, is needed to calculate the desorption
4. Heat and mass transfer model rate. This is determined based on the heat transfer rate that
can be supported by the overall heat transfer coefficient. The
The experimental results obtained from the tests described solution-side desorption heat duty to be supplied by the
above were also analyzed using a model that accounted for heating fluid is determined by an enthalpy balance as follows:
the local variations in the heat and mass transfer resistances
through the length of the desorber. The basic approach is well Q_ seg m
_ l;out hl;out m
_ v;out hv;out  m
_ l;in hl;in  m
_ v;in hv;in (6)
documented by Price and Bell (1974), who in turn used the
This equation accounts for both the latent heat of desorption
technique developed by Colburn and Drew (1937) for the
and the sensible heating of the liquid stream. The relevant
condensation of a binary vapor with a miscible condensate.
mass flow rates are established by the mass transfer and mass
Other investigators (Kang and Christensen, 1994, 1995;
balance equations outlined above. The inlet and outlet liquid
Garrabrant and Christensen, 1997; Kang et al., 1997; Kang
enthalpies are determined from the corresponding equilib-
et al., 1998) have also adapted this technique to the design of
rium concentrations. The vapor inlet enthalpy is known from
ammoniaewater heat and mass transfer components.
the heat and mass transfer calculations for the previous
Meacham and Garimella (2003) used a similar approach for
segments. The heat transfer rate from the glycol in the
modeling the performance of the device under consideration
segment is calculated as follows:
here as an absorber. The geometry under consideration is
 
assumed to induce a well-mixed liquid film. Using this Q_ seg m
_ glyc cp;glyc Tglyc;in  Tglyc;out (7)
assumption, the desorption of vapor from the falling liquid
film is governed by the mass transfer resistance in the vapor After calculating an average tube wall temperature for the
phase and the concentration gradient between the vapor- segment using a resistance network approach, the heat
eliquid interface and the bulk vapor. Thus, the molar flux of transfer rate through the wetted portion of the tube is calcu-
the vapor being absorbed into the liquid film is given by: lated as follows:

" # 1
Rf (8)
00 z~  y~int af $a$Ap
n_ T CT bv ln (1)
z~  y~bulk
 
1  
where bv is the mass transfer coefficient, CT is the total molar Q_ wet $ LMTDwall;f (9)
Rf
concentration of the vapor, and z~ is the molar concentration
of the evaporating flux. The mass transfer coefficient is where the log-mean temperature difference (LMTDwall,f),
determined through a heat and mass transfer analogy. The between the wall and the liquid is the driving force. It should
corresponding vapor-phase heat transfer coefficient is calcu- be noted that the heat transfer coefficient af used here is based
lated from a correlation for crossflow over tubes by Churchill only on the wetted surface area, and is therefore different
and Bernstein (1977). The driving molar concentration from the as reported from the experimental results in the
difference in the above equation is a function of the molar previous section. Also, the treatment below is different from
concentration of the evaporating flux, z, ~ the equilibrium that used by Meacham and Garimella (2003) in the manner in
molar vapor concentration at the interface, y~int , and the bulk which the surface wetting is addressed. Meacham and Gar-
molar concentration of the vapor, y~bulk . The interfacial vapor imella applied a wetted area ratio to the entire exchanger, and
concentration is determined by assuming that the vapor at used one value of a to represent all the passes. Here, an area
the interface is in local equilibrium with the liquid. The ratio is computed separately for each pass to account for
individual ammonia and water desorption fractions are changes in the tube wetting characteristics along the length of
determined as follows: the desorber. In addition, the area ratio used by Meacham and
Garimella simply applied a uniform area reduction to the
_ des;a Ma n_ 00a aAp
m (2) overall physical surface area of the exchanger; thus, the
surface was either participating in the heat and mass transfer
process if it was wet, or not participating at all. In the present
_ des;w Mw n_ 00w aAp
m (3)
study, the direct heat transfer between the dry portion of the
i n t e r n a t i o n a l j o u r n a l o f r e f r i g e r a t i o n 3 4 ( 2 0 1 1 ) 1 1 9 7 e1 2 0 8 1203

tube wall and the bulk vapor is also accounted for in an to 17.5 kW as the solution flow rate was increased from
explicit manner, in addition to the heat and mass transfer in 0.0188 kg s1 to 0.0339 kg s1. Over the range of test conditions,
the wet portion. The heat transfer rate from the dry portion of the overall heat transfer coefficient (Fig. 4) increased roughly
the wall directly to the vapor is calculated as follows: linearly with the concentrated solution flow rate, ranging
from a minimum of 388 W m2 K1 at 0.0188 kg s1 to
1
Rv;dry (10) a maximum of 617 W m2 K1 at a concentrated solution flow
av $1  a$Ap
rate of 0.0351 kg s1. The solution mass flow rate has a large
 
1 impact on the performance of the desorber. Determan et al.
Q_ v;dry $LMTDwall;v (11)
Rv;dry (2004) also reported that the corresponding solution-side
heat transfer coefficient as (total surface area basis) varied
The LMTDwall,v between the wall and the vapor is used here.
from 659 to 1795 W m2 K1 over the range of conditions
The sensible heat transfer between the liquidevapor interface
tested. This range of as in the desorption mode is similar to the
on the wet portion of the tube and the bulk vapor is calculated
values reported by Meacham and Garimella (2004)
as follows:
(638 < as < 1648 W m2 K1) for absorption for a similar micro-
 
f channel configuration, although with somewhat different
Q_ v;wet av $a$Ap $LMTDv;int (12)
1  ef dimensional details. Although the tube-side heat transfer
coefficient is relatively high (1396e1466 W m2 K1) for all the
Here, av is the vapor-phase heat transfer coefficient, LMTDv,int
test conditions due to the use of microchannels, it still
is the log-mean temperature difference between the vapor
represents 40e65% of the overall heat transfer resistance.
and the interface, and the term in the brackets is the correc-
Thus, the desorber was found to be solution-side limited at
tion factor for the heat transfer coefficient due to mass
the low solution flow rates, while at the high end of the
transfer. The parameter f is given by:
ammoniaewater solution flow rate, the ammoniaewater heat
n_ a c~p;a n_ w c~p;w transfer coefficient is high enough that the tube-side resis-
f (13) tance is not negligible. This was illustrated by Determan
av
et al. (2004) with the increase in performance with an
The segment vapor outlet temperature is found by summing the
increase in solution flow rate, and a somewhat smaller
sensible heat transfer from the dry wall and the liquidevapor
dependence of the overall heat transfer coefficient on the
interface as follows:
coolant flow rate. The overall U and as were found to be
Q_ v Q_ v;dry Q_ v;wet (14) less sensitive to vapor fraction. This is because the vapor-
side velocities and more importantly, the vapor-phase
This total heat transfer to the vapor phase is used to calculate resistance, do not change appreciably in the range of
the segment vapor outlet temperature: conditions tested.
  These experimental results were analyzed in detail, on
Q_ v m
_ v cp;v Tv;in  Tv;out (15)
a pass-by-pass basis, using the model discussed above.
The correlation developed by Wilke (Wilke, 1962) is used to Measured solution and vapor state points at the desorber inlet
estimate the falling-film Nusselt number as follows: and outlet, and the progression of measured glycol tempera-
tures at the headers of each pass within the desorber were
Nul 0:029$Re0:533 $Pr0:344 (16) used as inputs for the analysis. The wetted area fraction, a,
The Reynolds number is defined as Re 4G/m, where G is the was then varied for each segment for data points representing
mass flow rate per tube, per length of wetted tube. The
segment heat transfer rate, accounting for the heat transfer
from the wet portion (wall-to-solution: Qwet) and the dry
portion (wall-to-vapor: Qdry) is then calculated as follows:

Q_ seg Q_ wet Q_ dry (17)

Closure to the model is obtained when the segment heat


transfer rates computed iteratively based on ammoniaewater
enthalpies (Equation (6)), glycol temperature drop (Equation
(7)), and heat and mass transfer resistances (Equation (17)) are
equal.

5. Results and discussion

The application of the heat and mass transfer model pre-


sented above is discussed here. The results from the experi-
ments first reported by Determan et al. (2004) are summarized
in Fig. 4, and serve as a starting point for the discussion. In the Fig. 4 e Measured Overall Heat Transfer Coefficient vs.
experiments, the desorber heat duties increased from 5.4 kW Concentrated Solution Flow Rate.
1204 i n t e r n a t i o n a l j o u r n a l o f r e f r i g e r a t i o n 3 4 ( 2 0 1 1 ) 1 1 9 7 e1 2 0 8

Fig. 5 shows the variation of the glycol heat duty and the
sensible heat duty of the vapor with desorber pass. This figure
shows that the sensible heat transfer between the vapor and
the liquid film is a small fraction of the heat load transferred
by the glycol. The glycol heat duty per pass is plotted based on
measured glycol temperatures at inlet/outlet headers of each
pass. It can be seen that the heat transfer rate is the highest at
the top of the desorber, decreases toward the middle, and
rises again at the bottom-most pass. These results can be
interpreted as follows. The measured temperature drops in
the glycol, which correspond to the pass heat duty, are 6.5, 6.0,
5.3, 5.1, 6.5  C from the top to the bottom, respectively.
The decreasing heat duty per pass in the first four passes
appears to be the result of several competing factors as illus-
trated by Fig. 6. As the liquid film flows down the tube array,
the total liquid flow rate is decreasing due to the vapor
generation. The liquid, however, could also be coalescing into
rivulets, partly due to surface tension, thus reducing the
wetted area of the tubes. Based on the measured pass-by-pass
Fig. 5 e Glycol and Vapor Heat Duties vs. Desorber Pass. glycol temperatures, the wetted area ratio (Fig. 7) decreases
from 0.82 in the top pass to 0.23 at the fourth pass. The
consolidation of the liquid film into rivulets, and therefore, the
increased mass flux for a given mass flow rate, tends to
a broad range of experimental conditions until the predicted increase the film Reynolds number. On the other hand, the
desorber outlet conditions matched the corresponding decrease in solution flow rate due to desorption tends to
measured values. decrease the film Re. The film Re is also simultaneously
This model validation approach is further illustrated using affected by fluid property changes due to temperature and
a data point near the design condition. This test case had concentration changes. Thus, the liquid viscosity changes
a dilute solution flow rate of 0.0223 kg s1 and a temperature from 4.88  104 kg m1 s1 in the top pass of the desorber to
and concentration of 83.7  C and 0.316. Glycol solution entered 3.53  104 kg m1 s1 in the bottom pass. These competing
the desorber at a mass flow rate of 0.1152 kg s1 and factors lead to an increase in the Reynolds number from
a temperature of 99.4  C. The vapor desorption rate was Re 71 in the top pass, to Re 277 in the next to last pass.
0.00529 kg s1, and it exited the desorber at a temperature of In the final pass, the increased glycol temperature drop
68.4  C, and an experimentally determined concentration of indicates an increase in the wetted area ratio to 0.43, and
0.9774. The desorber pressure was maintained at 664.7 kPa. a corresponding decrease in the Reynolds number to Re 158.
The predicted variations in the significant parameters The film heat transfer coefficient varies with Reynolds
within the desorber are now illustrated for this data point. number and the liquid film thickness. An increase in the

Fig. 6 e Illustration of Passwise Variation in Significant Parameters.


i n t e r n a t i o n a l j o u r n a l o f r e f r i g e r a t i o n 3 4 ( 2 0 1 1 ) 1 1 9 7 e1 2 0 8 1205

Fig. 8 e Average Wetted Areas for the Present Study.


Fig. 7 e Wetted Area Fraction by Pass.

that increasing the solution flow rate leads to larger wetted


areas. Fig. 9 shows the pass-wise variation in wetted area
Reynolds number would tend to increase the heat transfer
fractions for a representative data point at each of the 5
coefficient in wavy flows, but an increase in film thickness
nominal solution flow rates. All of these cases show the same
(due to coalescence) would decrease the heat transfer coeffi-
general behavior and the increase in wetted area fraction with
cient. The above discussion indicated an increase in the liquid
increasing solution flow rates is evident. For the cases with
Re as the solution flows down due to a decrease in the wetted
higher solution flow rates, the area fraction actually
area. However, this decrease in the wetted area would also
approaches 1 in the upper-most pass. (Another potential
increase the liquid film thickness, which would contribute to
explanation is that when the glycol solution entering at the
a decrease in the heat transfer coefficient. It appears that the
bottom makes the first bend to enter the next pass, some tube-
Reynolds number plays the more dominant role here, and the
side mal-distribution is introduced, which reduces local heat
film heat transfer coefficient increases from 2140 to
transfer and desorption rates, and is erroneously reflected as
3116 W m2 K1 from the top pass to the fourth pass, and then
smaller wetted area fractions. Additional local solution-side
decreases in the final pass to 2890 W m2 K1. It should be
and glycol-side measurements and flow visualization studies
pointed out, however, that even though the heat transfer
could confirm this interpretation.)
coefficient is the highest in the fourth pass, the overall heat
Fig. 10 shows the variation of the total and species
duty of this pass is the lowest among all passes. This is due to
desorption rates through the desorber. In the bottom section
the fact that the wetted area is the smallest in this pass and
therefore there is the least amount of heat transfer area
between the tubes and the liquid film. This illustrates that
increasing the liquid film heat transfer coefficient alone is an
insufficient design goal for falling-film heat and mass
exchangers. The wetted area ratio plays a major role in the
total heat and mass transfer rates and therefore the size of the
device. The average wetted area fraction over the entire
desorber for this case is 0.45. Over the range of conditions
investigated in this study, the average wetted area fraction
ranged from 0.25 to 0.69.
The increasing wetted area fraction at the bottom pass is
substantiated by the corresponding increase in pass heat duty
near the bottom of the desorber (Fig. 5). It should be noted that
pass-by-pass duties were measured using glycol tempera-
tures, thus these predictions are based on the actual
measured quantities. It is possible that the lower solution flow
rates and lower countercurrent vapor velocities at the bottom
of the desorber result in thinner solution films that also spread
over the tubes more uniformly. Fig. 8 shows the average
wetted area fraction for all 60 data points analyzed in this Fig. 9 e Wetted Area Fractions by Pass for Representative
study versus the concentrated solution flowrate. It can be seen Points.
1206 i n t e r n a t i o n a l j o u r n a l o f r e f r i g e r a t i o n 3 4 ( 2 0 1 1 ) 1 1 9 7 e1 2 0 8

Fig. 12 e Variation of Average Vapor and Liquid-Phase


Fig. 10 e Desorption Rates by Pass. Concentrations with Desorber Pass.

of the desorber, both the ammonia and water are being des-
orbed from the liquid solution. The water desorption rate equilibrium concentration and the temperature of the exiting
decreases from the bottom pass to the top pass, eventually solution. The temperature of the vapor initially increases from
even becoming negative in the top pass, indicating that 83.7  C to 85.0  C due to heat transfer from the exposed dry
rectification of the vapor is occurring simultaneously with portion of the wall, after which it steadily decreases through
desorption of ammonia. This increases the concentration of the length of the desorber due to heat and mass transfer with
the ammonia vapor leaving the desorber, and therefore the adjoining liquid film of an increasing concentration as it
reduces the rectification load. The effects of the varying moves upward. The predicted vapor outlet temperature at the
wetted area ratio are also evident in this graph, which corre- top of the desorber is 61.1  C, while the corresponding
sponds to the heat duties shown in Fig. 5. However, the measured value is 68.4  C, showing a reasonable agreement
balance between water desorption and reflux condensation between the model and the data. However, it should be noted
(or rectification) rates is a stronger function of local temper- that because of the relatively low mass flow rate and specific
ature differences (between the vapor and the solution or heat of the vapor, this temperature discrepancy represents an
between the vapor and the tube wall) than on the wetted area error in heat duty on the order of 90 W. Compared to the
fraction. This coupled with the lower water desorption rates overall heat duty of the desorber (11.1 kW), this is a relatively
that are primarily the consequence of local vaporeliquid small (0.8%) error. The concentrated solution entered the
equilibrium concentration, results in a fairly monotonous desorber at a subcooled inlet temperature of 53.9  C, whereas
trend in water desorption rates. the saturated temperature is 58.8  C. The difference of 2.3  C
The rectification, or cooling, of the vapor as it rises in the between the vapor outlet and the saturated liquid inlet shows
desorber is also clear in Fig. 11. The initial vapor desorbed at a very desirable, small departure from equilibrium between
the bottom of the desorber must necessarily be at the the vapor and the solution at the desorber outlet. (The glycol
temperatures shown in this figure are the actual measured
values.)
Fig. 12 shows the predicted average pass concentrations for
the bulk vapor, vapor interface, evaporating flux, z, and the
liquid solution. The liquid concentration steadily decreases
from 0.4358 to 0.3321 as it flows down through the desorber,
while the bulk vapor concentration increases from 0.9515 to
0.9839 as it flows upwards. The corresponding vapor concen-
trations at the interface are 0.9543 and 0.9868, respectively.
For this data point, the vapor exits the desorber at a concen-
tration of 0.9873. The evaporating flux concentration, z,
becomes greater than 1 in the top passes, indicating that
water is being absorbed into the liquid.

6. Conclusions

The results of a combined experimental and analytical


Fig. 11 e Temperature Profiles in the Desorber. study of ammoniaewater desorption in a microchannel
i n t e r n a t i o n a l j o u r n a l o f r e f r i g e r a t i o n 3 4 ( 2 0 1 1 ) 1 1 9 7 e1 2 0 8 1207

device were presented in this study. The model was able to Determan, M.D., Garimella, S., Lee, S., 2004, Experimental
predict the measured conditions at the component inlets demonstration of a microchannel desorber for ammonia-
and outlets, and also the incremental variations in mass water heat pumps. Seventeenth National Heat and Mass
Transfer Conference and Sixth ISHMT/ASME Heat and Mass
flow rates, concentrations, temperatures and heat duties
Transfer Conference, Kalpakkam, India, pp. 453e458.
along the length of the desorber. The technique utilized Garimella, S., 1999. Miniaturized heat and mass transfer
a pass-wise wetted area ratio for closure, and showed that technology for absorption heat pumps. Proceedings of the
the wetted area ratio of the tube array ranged from 25 to International Sorption Heat Pump Conference, Munich,
69% over the conditions investigated in this study. Germany, pp. 661e670.
Although the primary resistance to mass transfer consid- Garimella, S., 2000. Microchannel components for absorption
space-conditioning systems, 2000 ASHRAE Winter Meeting,
ered here is in the vapor phase, it is the wetted area ratio
Feb 5eFeb 9 2000, Dallas, TX, USA. Amer. Soc. Heating,
that ultimately limits the performance of the desorber. It
Ref. Air-Conditoning Eng. Inc., Atlanta, GA, USA, pp.
was also found that the vapor being generated is cooled 453e462.
through interactions with the adjacent liquid film as it Garimella, S., 2004. Method and Means for Miniaturization of
rises through the desorber, yielding a desirable high Binary-Fluid Heat and Mass Exchangers. 6,802,364, USA. Iowa
concentration of ammonia vapor at the desorber outlet. State University Research Foundation, Inc, Ames, IA.
Desorption loads of up to 17.5 kW were transferred, at Garrabrant, M.A., Christensen, R.N., 1997. Modeling and
experimental verification of a perforated plate-fin absorber for
overall heat transfer coefficients between 388 W m2 K1 and
aqua-ammonia absorption systems. Proceedings of the 1997
617 W m2 K1. This performance was achieved with low
ASME International Mechanical Engineering Congress and
coolant pressure drops (11e30 kPa) in spite of the use of Exposition, Nov. 16e21, 1997, Dallas, TX, USA, ASME, Fairfield,
extremely small diameter tubes, due to the parallel pass NJ, USA, pp. 337e347.
arrangement chosen. Additional performance improvements Hoffmann, L., Ziegler, F., 1996. Experimental investigation of
can be obtained by further increasing the wetted surface of heat and mass transfer with aqueous ammonia.
the tube array. The concept investigated here was initially International Ab-sorption Heat Pump Conference, Montreal,
Canada, pp. 383e392.
demonstrated as an absorber by the authors in previous
Jeong, S., Garimella, S., 2002. Falling-film and droplet mode heat
studies (Garimella, 1999, 2000; Meacham and Garimella, 2002a, and mass transfer in a horizontal tube Libr/water absorber.
2003; Garimella, 2004; Meacham and Garimella, 2004). Minor Int. J. Heat Mass. Tran. 45 (7), 1445e1458.
modifications sufficed to convert the absorber to the desorber Jeong, S., Garimella, S., 2005. Optimal design of compact
function, demonstrating the versatility of the concept. The horizontal tube Libr/water absorbers. HVAC&R. Res. 11 (1).
results of this study show that the subject microchannel heat Kang, Y.T., Chen, W., Christensen, R.N., 1997. A generalized
component design model by combined heat and mass
and mass transfer technology can be successfully used for
transfer analysis in NH3/H2O absorption heat pump
both desorption and absorption to package an absorption heat
systems. Proceedings of the 1997 ASHRAE Winter Meeting,
pump system for small-capacity applications in an extremely Jan 26e29 1997, Philadelphia, PA, USA, ASHRAE, Atlanta,
compact envelope. GA, USA, pp. 444e453.
Kang, Y.T., Christensen, R.N., 1994. Development of a counter-
current model for a vertical fluted tube gax absorber.
Proceedings of the International Absorption Heat Pump
references Conference, Jan 19e21, 1994, New Orleans, LA, USA, ASME,
New York, NY, USA, pp. 7e16.
Kang, Y.T., Christensen, R.N., 1995. Combined heat and mass
Beutler, A., Hoffmann, L., Ziegler, F., Alefeld, G., Gommed, K., transfer analysis for absorption in a fluted tube with a porous
Grossman, G., Shavit, A., 1996a, Experimental Investigation of medium in confined cross flow. Proceedings of the 1995 ASME/
Heat and Mass Transfer on Horizontal and Vertical Tubes. JSME Thermal Engineering Joint Conference. Part 1 (of 4), Mar
Proceedings of the International Ab-Sorption Heat Pump 19e24 1995, Maui, HI, USA, ASME, New York, NY, USA, pp.
Conference, Montreal, Canada, pp. 409e419. 251e260.
Beutler, A., Ziegler, F., Alefeld, G., 1996b, Falling Film Kang, Y.T., Kashiwagi, T., Christensen, R.N., 1998,
Absorption with Solutions of a Hydroxide Mixture. Ammoniaewater bubble absorber with a plate heat
International Ab-Sorption Heat Pump Conference, Montreal, exchanger. Proceedings of the 1998 ASHRAE Winter Meeting.
Canada, pp. 303e309. Part 2 (of 2), Jan 18e21, 1998, San Francisco, CA, USA, ASHRAE,
Christensen, R.N., Garimella, S., Kang, Y.T., Garrabrant, M.A., Atlanta, GA, USA, pp. 1565e1575.
1998, Perforated Fin Heat and Mass Transfer Device. 5,704,417, Killion, J.D., Garimella, S., 2001. A critical review of models of
USA, Gas Research Institute, Chicago, IL. coupled heat and mass transfer in falling-film absorption. Int.
Churchill, S.W., 1977a. Comprehensive correlating equations J. Refrigeration 24 (8), 755e797.
for heat, mass and momentum transfer in fully developed Killion, J.D., Garimella, S., 2003. A review of experimental
flow in smooth tubes. Ind. & Eng. Chem., Fundam. 16 (1), investigations of absorption of water vapor in liquid films
109e116. falling over horizontal tubes. HVAC&R. Res. 9 (2), 111e136.
Churchill, S.W., 1977b. Friction-factor equation spans all fluid- Meacham, J.M., Garimella, S., 2002a. Experimental demonstration
flow regimes. Chem. Eng. Prog. 84 (24), 91e92. of a prototype microchannel absorber for space-conditioning
Churchill, S.W., Bernstein, M., 1977. Correlating equation for systems. International Sorption Heat Pump Conference,
forced convection from gases and liquids to a circular cylinder Shanghai, China, pp. 270e276.
in crossflow. J. Heat Transf. Trans. ASME Series C 99 (2), Meacham, J.M., Garimella, S., 2002b. Miniaturized shell-and-
300e306. tube heat and mass exchangers for absorption heat pumps.
Colburn, A.P., Drew, T.B., 1937. The condensation of mixed 12th International Heat Transfer Conference, Grenoble,
vapours. AIChE Trans. 33, 197e212. France.
1208 i n t e r n a t i o n a l j o u r n a l o f r e f r i g e r a t i o n 3 4 ( 2 0 1 1 ) 1 1 9 7 e1 2 0 8

Meacham, J.M., Garimella, S., 2003. Modeling of local measured Price, B.C., Bell, K.J., 1974. Design of binary vapor condensers
heat and mass transfer variations in a microchannel using the colburn-drew equations. AIChE Symposium
ammoniaewater absorber. ASHRAE Trans. 109 (1), 412e422. Series e Heat Transfer e Research and Design 70 (138),
Meacham, J.M., Garimella, S., 2004. Ammoniaewater absorption 163e171.
heat and mass transfer in microchannel absorbers with visual Taylor, B.N., Kuyatt, C.E., 1994. Guidelines for evaluating and
confirmation. ASHRAE Trans. 110 (1), 525e532. expressing the uncertainty of nist measurement results.
Potnis, S.V., Anand, G., Gomezplata, A., Erickson, D.C., Papar, National Institute of Standards and Technology Technical
R.A., 1997. Gax component simulation and validation. Note 1297.
Proceedings of the 1997 ASHRAE Winter Meeting, Jan Wilke, W., 1962. Heat transfer to falling liquid films
26e29, 1997, Philadelphia, PA, USA, ASHRAE, Atlanta, GA, (Waermeuebergang an Rieselfilme). VDI e Forschungsheft
USA, pp. 454e459. 490, 36.

You might also like