Sei sulla pagina 1di 19

International Journal of Refrigeration 28 (2005) 635–653

www.elsevier.com/locate/ijrefrig
Review Article
Falling-film evaporation on horizontal tubes—a critical review
Gherhardt Ribatskia,*, Anthony M. Jacobib
a
Laboratory of Heat and Mass Transfer, Faculty of Engineering Science, Swiss Federal Institute of Technology, ME G1 465, Station 9,
CH Lausanne 1015, Switzerland
b
Department of Mechanical and Industrial Engineering, University of Illinois, 1206 West Green Street, Urbana, IL 61801, USA

Received 12 February 2004; received in revised form 1 December 2004; accepted 14 December 2004
Available online 1 February 2005

Abstract
A state-of-the-art review of horizontal-tube, falling film evaporation is presented; the review is critical, in an attempt to
uncover strengths and weaknesses in prior research, with the overall purpose of clearly identifying gaps in our understanding.
The review covers flow-pattern studies, and the experimental parameters that affect the heat transfer performance on plain
single tubes, enhanced surfaces and tube bundles. In addition, this paper presents a comprehensive review of the significant
efforts to develop mathematical models, and empirical correlations for the heat transfer coefficient. Emphasis is placed on
studies that are related to refrigeration applications.
q 2004 Elsevier Ltd and IIR. All rights reserved.

Keywords: Evaporation; Falling film; Horizontal tube; Finned tube; Bundle; Survey; Heat transfer; Correlation

Evaporation d’un film tombant sur des tubes horizontaux—passage


en revue critique
Mots clés: Évaporation; Film tombant; Tube horizontal; Tube aileté; Faisceau; Enquête; Transfert de chaleur; Corrélation

1. Introduction 1970’s. Since then, this technology has been studied by


many investigators; however, the focus during the 1970’s
Falling-film-type (or spray-film) horizontal-tube evap- was primarily on the use of falling-film evaporators for
orators have been utilized in the refrigeration, chemical, ocean thermal energy conversion, (OTEC) systems, with
petroleum refining, and desalinization industries. Although interest in the early 1980’s driven by the second oil crisis. As
the first patent of such an evaporator was registered in 1888 a result of the interest in OTEC applications, most of this
[1], only a few researchers worked in this area prior to the work in the 1970’s and 1980’s used water or ammonia as a
working fluid. Moreover, the heat fluxes and other operating
parameters were constrained to ranges relevant for OTEC
* Corresponding author. Tel.: C41 21 693 5984; fax: C41 21 693 systems.
5960. During the 1990’s, the CFC phase-out began to motivate
E-mail addresses: gherhardt.ribatski@epfl.ch (G. Ribatski), wider application of falling-film evaporators. In air-
a-jacobi@uiuc.edu (A.M. Jacobi). conditioning and refrigeration applications, the falling film
0140-7007/$35.00 q 2004 Elsevier Ltd and IIR. All rights reserved.
doi:10.1016/j.ijrefrig.2004.12.002
636 G. Ribatski, A.M. Jacobi / International Journal of Refrigeration 28 (2005) 635–653

Nomenclature
Ar Archimedes number based on tube diameter Greek letters
D3extg/n3l , (dimensionless) a thermal diffusivity, (m2 sK1)
Dext surface diameter, (m) f specific heat flux, (W mK2)
g gravitational acceleration, (m sK2) G liquid mass flow rate per unit length of tube
Ga modified Galileo number (or Kapitza number) (each side), (kg mK1 sK1)
rs3/m4g, (dimensionless) l instability wavelength given by Eq. (3), (m)
H liquid feeder height, (m) m dynamic viscosity, (kg mK1 sK1)
h heat transfer coefficient, (W mK2 KK1) n kinematics viscosity, (m2 sK1)
hlv latent heat of vaporization, (J kgK1) q angle along surface perimeter measured from
k thermal conductivity, (W mK1 KK1) the tube apex, (8)
K dimensionless defined by [52], Eq. (4) r density, (kg mK3)
n/A active nucleate site density, (mK2) s surface tension, (kg sK2)
Nu Nusselt number (h/kl)(n2l /g)1/3, (dimensionless) x capillary constant given by [s/(rlg)]1/2, (m)
p pressure, (N mK2)
Subscripts
Pr Prandtl number n/a, (dimensionless)
crit referred to the critical state
Re Reynolds number 4G/m, (dimensionless)
l liquid
s tube pitch, (m)
sat saturation
T temperature, (K)
v vapor
w tube wall

evaporator possesses the following clear advantages over arrangements and desalination systems, because our goal is
flooded tube bundles: to assess the state of the art as it is likely to affect air-
conditioning and refrigeration systems. Our goal is to
High heat transfer coefficient, (1) allowing the evapor- critically review the literature, for the purpose of clearly
ation temperature to increase, and thus improving the identifying opportunities for application and areas where
cycle efficiency; (2) minimizing the evaporator size, further research is needed.
resulting in reduced initial costs and space requirements. In the present work, by referring to the results of
Low refrigerant charge, (1) lowering the cost of the previous research, we begin with a discussion of the effects
refrigeration plant (including the refrigerant inventory); of heat flux, flow rate, film temperature, tube diameter and
(2) reducing risks associated with a leak, including the liquid feeder configuration on the falling-film evaporation
attendant maintenance costs. This reduced risk allows for single plain tubes. Then, the observed inter-tube flow
wider application of systems using toxic or flammable patterns are presented as well as proposed methods to
fluids, such as ammonia. predict inter-tube flow patterns. Following that, experimen-
tal studies concerning falling-film evaporation on enhanced
These advantages notwithstanding, falling-film evapor- surfaces and tube bundles are described, and their results
ators are not widely used in refrigeration and air- critically discussed. Some interesting prediction methods to
conditioning. In part, reticence to adopt falling-film estimate the heat transfer coefficient in subcooled and
technology in these systems is due to difficulties in liquid saturated conditions are reported in the last part of the
distribution and tube alignment, which affect flow uniform- article.
ity and dryout, especially in deep bundles. Furthermore,
decades of experience with flooded bundles in this industry
has allowed for their heuristic optimization, but heat transfer
surface, bundle geometry, and operating strategies have not 2. Heat transfer aspects on plain surfaces
been so refined for falling-film evaporators. At present,
these issues are motivating many researchers. Generally speaking, prior studies involving smooth tubes
A prior related review was provided by Thome [2], with have covered wide ranges of heat flux, flow rates, liquid
a focus on studies published from 1994 to 1999; to our feeder height, and tube spacing. The tube diameters have
knowledge, no other review has appeared in the open been similar to those used in heat exchangers. Tests were
literature. The current article provides a literature review of conducted for sub-cooling and saturated liquid, and under
horizontal-tube, falling-film-type evaporators, focusing on both boiling and non-boiling conditions. Except for a few
the technical difficulties identified above. We focus this studies, the experiments were conducted with water and
review on falling-film evaporators and exclude vertical tube ammonia, and most of the work focused on how the
G. Ribatski, A.M. Jacobi / International Journal of Refrigeration 28 (2005) 635–653 637

Table 1
Description of experimental studies involving falling film on plain single tubes

Fluid Surface Dext (mm) H (mm) G Tsat or Tl (8C) f (kW mK2)


material (!103 kg sK1 mK1)
[3] Water Cu–Ni 25.4, 50.8 – 0–314 49–127 0–63.1
[4] Sea water Cu–Ni 25.4, 50.8 – 0–228 49–127 0–63.1
[5] Water Stainless steel 25.4, 50.8 8–54 144–377 55–100 19–76
[6] Water Brass 25.4, 50.8 3.2 133–373 45–127 15.8–78. 8
[7] Ammonia Stainless steel 50.8 – 4–202 22.2 5–16
[8] Ammonia Low carbon 25.4 25.4 3.7–37.4 12.8–23.9 3.2–25.2
steel
[9] Water Copper 25.4 12.7, 25.4, 50.8 4–40 27, 50 0–83
[10] Ammonia Stainless steel 50.8 50.8 4–350 22.2 5.2–15.8
[11] Water Copper 25.4 3 37–110 99.4 2–100
[12] Water Brass 25.4, 50.8 – 2.78–7.62! 49–127 16–79
10K4m3 sK1 mK1
[13–15] Water Copper 25.4 3–63.5 21.3–156 99.4 2–208
[16] Water, ethyl Copper 25.4 12.7, 25.4, 50.8 2.5–50 57 0–80
alcohol
[17] Water, isopro- Copper 18 4.5–87 38–130 25, 21.5 18.4, 9.4
pyl alcohol
[18] R-113 – – – – 47 7–80
[19] Water – 38 – 40–400 40–100 15–75
[20] Water Aluminum 132 – – – –
[21] R-11 Copper 18.0 20 8–91 23.5 1–43
[22] Water Brass 25.4, 50.8 6.3 135–366 49–127 30–80
[23] R-113 Copper 22.0 – 32.1–96.3 47.2 7.3–72.7
[24,25] R-11 Copper 25.0 25.0 1–180 44.6 0.5–2.5
[26] R-134a Copper 12.7, 19.1 38–65!10K3 kg sK1 2.0 5–40
[27] R-134a Copper 19.1, 12.7 – 38–65!10K3 kg sK1 K14, 2 5–40
[28] Water, isopro- Copper 19.5 7.8–48.8 5–150 9–50 0–150
pyl alcohol
[29] Water, ethylene Brass 9.52, 12.7, 0–100 0–360 25–40 0–115
glicol, mixture 15.87, 19.
of water and 05, 22.22
ethylene glycol
and ethyl alco-
hol
[30] R-134a with a Copper 19.1 – 13 2 5–40
polyol-ester
32 cs/40 8C
uZ0, 1, 2 and
3.0%
[31] Water Aluminum 132 – – 17–50 71–158
[32] Ammonia Stainless steel 19.1 102, 51 7–39 K23.3–10 8–60
[33] Water, ethylene Brass 15.9, 5–50 0–360 25–40 0–115
glicol, mixture 19.01,
of water and 22.22
ethylene glycol
[34] Water Surface cov- 38 10–20 40–400 42–100 15–75
ered with a
constant foil
[35] Water Copper 19.5 7.8, 2.3, 48.8 10–150 20, 35, 50 0
[36] Wateri, R-11ii Copper 18 6 15–354 99.4i, 23.5ii 2–500
638 G. Ribatski, A.M. Jacobi / International Journal of Refrigeration 28 (2005) 635–653

experimental parameters affect the heat transfer. Table 1 1000 and 8000, Parken [6] and Parken et al. [22] also
describes the experimental data bank of studies concerning observed an increase in h with G. Chyu and Bergles [14], for
falling film evaporation on single plain tubes. high heat flux, and Moeykens and Pate [27], independent of
f, did not note flow rate effects on h. According to both
2.1. Heat flux effect studies, h is independent of G due to the predominance of
boiling effects.
For completely wetted surfaces in strictly convective Changes in the inter-tube flow mode (droplet, jet and
conditions, the heat flux, f, does not affect the heat transfer sheet), and partial surface dryout can affect the effects of
coefficient, h, [6,7,12,22,25,33]. On the other hand, under flow rate on the heat transfer coefficient. Hu and Jacobi [33]
boiling-dominated conditions, the heat transfer coefficient noted an increase in h with G for the droplet and jet modes.
increases with the heat flux [3,4,6,22,27,32]. Apparently, For the sheet mode, this behavior was noted solely for water.
this behavior is due to increased nucleation site density, n/A, It can be related to wave effects on the film surface [33],
and increased boiling area. For a low heat flux, bubbles which are high for water as the result of the larger Re
nucleate near the bottom of the tube. As f increases, achieved in the experiments with this fluid.
nucleation occurs nearer to the top of the tube [6,22]. With
an increase in heat flux, a temperature profile in the film
2.3. Temperature effect
sufficient for nucleation closer to the film impingement
region is possible. A strong dependence of Nusselt number,
For convective heat transfer, h increases with the liquid
Nu, on f, with high rates of increase in h with heat flux has
temperature. It seems that this temperature effect is closely
been noted at low temperatures [3]. This finding seems to
related to the decrease of viscosity and consequent decrease
agree with pool boiling experiments of Ribatski and Saiz
in the film thickness [5,6,12,22,28]. The increase in h occurs
Jabardo [37], according to whom the variation of h with f is
around the overall surface perimeter and can be correlated
comparatively high at low reduced pressure. Thus, one
by the following equation [6]:
expects parameters such as the contact angle, surface
 
material and roughness to affect the increase in h with hðTA Þ n ðT Þ n
heat flux for falling film evaporators. However, no work Z l A (1)
hðTB Þ nl ðTB Þ
concerning the possible effects of these parameters on
falling film has been reported. It should be noted that Ganić and Roppo [9] found that the
fluid temperature does not affect the heat transfer coefficient.
2.2. Flow rate effect For boiling conditions, temperature gradients near the
surface are increased as the film thickness decreases, and a
In convection-dominated conditions, as the flow rate, G, decreased film thickness can be caused by viscosity
increases the heat transfer coefficient decreases first, but reductions associated with bulk temperature increases. A
increases after a minimum value [7,14,25,38]. This steeper temperature profile can inhibit bubble growth (see
minimum might represent a transition from laminar to Cerza and Sernas [44]), and change the position along the
turbulent film flow [25,38]. However, according to Rogers et surface perimeter at which active nucleation sites appear.
al. [31], this assertion is a flawed, arbitrary assumption. On the other hand, in pool boiling the heat transfer
Even modeling that does not consider the flow regime coefficient increases with temperature, because of the
transition exhibits similar behavior [14,39,40]. According to increase in n/A. In the context of falling-film evaporators,
Brumfield and Theofanous [41], the film can be fully these two competing effects can either increase or decrease
turbulent only for Reynolds numbers, Re, higher than 6000. h. These opposing effects might be responsible for
They suggest that the film is controlled by wave structures disagreement in prior work, ostensibly conducted under
imposed on the base film flow. At low Reynolds numbers, similar conditions. Refs. [3,4,6,32] point out an increase in h
the waves and the base flow are laminar; at intermediate with film temperature for boiling conditions. Fletcher et al.
Reynolds numbers the base flow is laminar while the waves [3,4] noted this behavior for a tube diameter of 25.4 mm,
are turbulent, and at high Reynolds numbers, both regions while for an external diameter of 50.8 mm the effect of film
are turbulent. Carey [42] proposed a transitional Reynolds temperature seemed to be marginal. Parken et al. [22] noted
number of 1500. For falling film condensation, a similar an increase in h with the film temperature for a tube with
process, values of 1200, 1800 and 2000 also have been DextZ50.8 mm and a liquid temperature up to 100 8C. An
proposed for the Reynolds number at which the flow opposite behavior was observed for higher temperatures; i.e.
becomes turbulent (see Thome [43]). for DextZ25.4 mm, and Tl!100 8C, they observed a weak
The following behaviors are also noted under convec- effect of film temperature on h; while for temperatures
tion-dominated conditions: the heat transfer coefficient is higher than 100 8C, h increased with liquid temperature.
almost independent of the flow rate [5,8,11,32], and the heat These behaviors seem to be related to the effects of the
transfer coefficient increases with G [6,9,12,17,22,28,32, diameter and the liquid temperature in the relative weight of
34]. Under boiling conditions and when Re falls between convective and boiling mechanisms, as previously discussed.
G. Ribatski, A.M. Jacobi / International Journal of Refrigeration 28 (2005) 635–653 639

2.4. Tube diameter effect uniformity. For ammonia, at saturation temperatures greater
than K1.1 8C, and commercial low-pressure-drop nozzles,
Tube diameter affects the heat transfer coefficient Zeng et al. [32] observed higher heat transfer coefficients for
because it changes the thermal boundary layer development spray nozzles with a cone angle greater than 908.
length and the length of liquid impingement region, relative
to the overall flow length of pDext/2. Under non-boiling 2.6. The effect of the liquid feeder height
conditions, these regions are characterized by higher local h,
resulting in an increase in the heat transfer coefficient with a The liquid feeder height can affect the heat transfer coef-
decreasing tube diameter [6,12,22,33], because a larger frac- ficient by modifying the flow mode, or through an increased
tion of the surface area is subject to impingement and deve- impingement velocity. An increased feeder height, H, can also
lopment effects. Liu [5] did not note effects of diameter on h. allow better spray distribution and thereby mitigate misalign-
Conversely, under boiling conditions the relative area of ment effects.
active boiling increases with the surface diameter, because Under non-boiling conditions, the heat transfer coeffi-
development of the thermal profile allows bubble growth. cient increases with H [5,6,13,14,17,28]. This effect is less
Thus, the overall behavior of h with Dext depends on the pronounced for fluids with a high liquid viscosity [5].
local h in the boiling region relative to the values on other Parken [6] reported h increased with H solely in the liquid
surface regions. These effects can explain differing trends impact region, while according to other Refs. [5,17], and as
observed under similar conditions. In the presence of bubble shown in Fig. 1 the heat transfer coefficient increased with H
nucleation, Parken [6] and Parken et al. [22] observed an along the whole surface perimeter. For the jet flow mode Hu
increase in h with the tube diameter, but Fletcher et al. [3], and Jacobi [33] noted that the heat transfer coefficient
Moeykens [26], and Moeykens and Pate [27] noted an increased with H. In the case of droplet-and-film modes they
opposite behavior. observed a weak effect of H. Ganić and Roppo [9] carried
out experiments for droplet-and-jet modes. Although they
2.5. The effect of liquid feeder configuration noted the increase in h with H at some flow rates, they did
not relate this behavior to flow modes. In boiling conditions,
The device used for liquid feeding can affect the Zeng et al. [32] for film temperatures lower than K1.1 8C,
evaporator performance. In general, feeder misalignment Chyu [13], and Chyu and Bergles [14] pointed out a weak
and designs that fail to provide axial uniformity can increase effect of the feeder height on h.
refrigerant maldistribution in the bundle and thus influence
the heat transfer rate. For convective heat transfer, the liquid
2.7. Effects of a vapor flow
feeder configuration may increase the heat transfer coeffi-
cient by increasing convective effects related to the liquid
A vapor flow can affect the evaporator performance in
impingement on the tube surface. Parken [6] and Parken and
the following ways: it can change the flow mode and
Fletcher [12] for a feeder with wedge shape, and Fletcher et
promote the deflection of the liquid flow, droplet atomiz-
al. [3] using a perforated plate noted axial temperature
ation and droplet drag; it can affect the film velocity profile
variations along the tube apex due to the flow non-uniformity.
and promote waves on the film surface [45]. The fluid
This effect is more pronounced for smaller tubes [6,12]. It has
been reported that film imperfections affect only the liquid
impact region [6,24]; however, Liu [5] observed that a small
misalignment produced an asymmetrical temperature distri-
bution all the way around the surface perimeter.
Fujita and Tsutsui [24,25] found that a cylindrical feeder
with holes along the bottom produced a heat transfer
coefficient 20% lower than that produced by a porous
sintered tube, a tube with holes along the top, and a
perforated plate with one, two, or three dummy tubes.
Moeykens and Pate [27] reported results under boiling
conditions for wide-angle, low-pressure-drop type and
wide-angle, high-pressure-drop type commercial spray
nozzles. The high-pressure-drop nozzles had the best
performance, according to the authors mainly because of
the impingement effect. However, using this nozzle might
not lead to an increase in overall evaporator performance, Fig. 1. The effect of the liquid feeder height on the heat transfer
because the impingement effect is confined to the top tubes, coefficient distribution along the external perimeter of a horizontal
and in a deep bundle impingement on the top few tubes cylindrical surface submitted to falling film evaporation under non-
becomes less important than other factors, such as flow boiling conditions (based on experimental results of Liu [5]).
640 G. Ribatski, A.M. Jacobi / International Journal of Refrigeration 28 (2005) 635–653

maldistribution caused by the gas flow can lead to partial where


surface dryout. On the other hand, the enhancement of pffiffiffiffiffiffiffiffiffiffiffi
convective effects on the film surface can increase h. The lZx 4p2 n (3)
vapor flow direction (countercurrent, concurrent or cross-
current) can influence these effects. In Eqs.ffi (2) and (3), nZ2, x is the capillary constant given by
pffiffiffiffiffiffiffiffiffiffi
Under boiling conditions, Parken [6] noted a heat s=rl g, and dp is the diameter of primary drops, experi-
transfer enhancement and the suppression of bubble mentally determined for water and alcohol to be equal to 3x.
nucleation with an increase in vapor velocity from 9 to Experimental observations by Ganić and Roppo [9]
18 m sK1. These effects were insignificant for velocities up indicated that the transition from the droplet to the jet mode
to 9 m sK1. According to Parken, this behavior seems to be occurred over a relatively large range of Reynolds number
related to a reduction in the film thickness and an increase in around 180, and it was affected by the tube spacing. When
the liquid velocity. For concurrent vapor flow and based on studying film condensation, Kutateladze et al. [47]
the results from a numeric model, Liu [5] pointed out suggested that the modified Archimedes number based on
that the degree to which h increases with the vapor velocity the capillary constant controlled the transition of flow mode.
is higher in thin films. In addition, he concluded that its Mitrovic [17] pointed out that the transition of flow modes
effect is insignificant for typical gas velocities of evapor- was dependent on the flow rate, fluid thermophysical
ators. According to the same study, in countercurrent flow, properties and the tube spacing.
vapor effects can reduce the heat transfer coefficient due to Armbruster and Mitrovic [48] modeled the mode
an increase in the film thickness resulting from an adverse transitions among the droplet, jet and sheet modes,
velocity profile in the film. according to ReZAGa1/4, where A is an empirical constant.
For sub-cooled water with a countercurrent air flow, For sufficiently high stagnation pressure, they observed that
Armbruster and Mitrovic [28] noted an increase in h with air the radial flows from adjacent jets collide and the film
velocity. This effect was more pronounced at low partial surface is raised between the jets. The crests created at the
vapor pressure, and was insignificant at high relative top of the tube remain nearly unchanged around the tube,
humidity. The effects of air velocity and relative humidity and form the departure sites for jets leaving the tube. This
were small at high heat fluxes. According to them, the air flow mode is called the staggered jet mode. On the other
temperature seemed not to affect h. Experimental results of hand, for low pressure at the stagnation point, the liquid
Armbruster and Mitrovic [35] for water and an adiabatic stagnation region does not cause crests to form, and the in-
surface showed the effects of air velocity and relative line jet mode is observed. Like Dhir et al. [49], Armbruster
humidity on the temperature of the liquid to be small. These and Mitrovic noted mode transition hysteresis (mode
results clearly suggest that the main effects of air velocity transition at distinct G for increasing and reducing flow
and relative humidity on h are imparted during the liquid rate) solely for the transition from the jet to the sheet mode.
free fall. For concurrent air velocity up to 15 m sK1, Hu and In addition, Armbruster and Mitrovic found the transition
Jacobi [33] noted an increase in h with air velocity, but the solely dependent on the liquid flow rate.
effect was within the uncertainty in their experimental results. Fujita and Tsutsui [24,25] defined the two flow modes as
follows: a distinct droplet mode—liquid initially falls as jet
and then disintegrates into droplets as the falling velocity
increases; a disturbed jets mode—characterized by the
3. Flow patterns studies collapse of neighbor jets, resulting in a sheet, followed by its
rupture. They noted that the transition between the droplet
When a liquid film flows from one horizontal tube to and the proposed new droplet mode occurred at Reynolds
another below it, according to an increasing flow rate order, number around 100 independent of feeding method. In
the flow may take the form of droplets, circular jets, or a contrast, the other mode transitions were affected by the
continuous sheet, as shown in Fig. 2. The pattern is referred feeder configuration with a decrease in the effects with the
to as falling-film mode and may play an important role in the increase in the dummy tube number.
heat transfer process. Yung et al. [46] correlated the flow Based on extensive observations of flow mode tran-
rate for a transition from the droplet mode to the jet mode sitions, Hu and Jacobi [50] suggested the following flow
with the droplet production frequency set equal to the modes: the droplet mode, the droplet-jet mode, an unsteady-
capillary wave oscillation frequency. Yung and his jet mode—characterized by a steadiness in the location of
co-workers based their studies on results from a single the jet departure site—the inline jet mode, the staggered jet
tube with a single liquid detachment site in quiescent vapor. mode, the jet-sheet mode, and the sheet mode. The flow
The flow rate on one-side per unit length of the tube at the modes proposed by Hu and Jacobi are shown in Fig. 3. In
transition was given by contrast to earlier reports, Hu and Jacobi observed hysteresis
  for all transitions. The flow mode was found to be relatively
r pdp3 2ps 1=2 independent of geometric effects over the range of their
2G Z 0:81 l (2)
l 6 rl l3 experiments. The effects of gravitational, inertial, viscous
G. Ribatski, A.M. Jacobi / International Journal of Refrigeration 28 (2005) 635–653 641

Fig. 2. The idealized inter-tube falling-film modes (Mitrovic [17]): (a) the droplet mode-liquid leaving the tube intermittently; (b) the jet mode-
liquid leaving the tube as a continuous column; (c) the sheet mode-liquid fil forms an unbroken sheet between the tubes.

and surface tension forces are captured by correlating the flow on heat transfer enhanced surfaces have been
transitional Reynolds number to a modified Galileo number, reported. For R-113, normal propanol, and methanol on
using a power-law fit, ReZAGab, where A and b are an adiabatic finned surface (1064 fins mK1), Honda
empirical constants. Data obtained with a concurrent air et al. [52] related the mode transitions to the following
flow and velocities up to 15 m sK1 indicated that a flow in dimensionless parameter:
the surrounding vapor affects the transitions involving  1=4
droplet modes only. Recently, based on experimental results G g
KZ (4)
with a countercurrent air flow, Wei and Jacobi [51] pointed s3=4 rl
out that for low Galileo numbers an air flow destabilizes the
jet and sheet mode, and reduces transition hysteresis. From the droplet mode to the jet mode the transition K
Few studies concerning the investigation of the mode ranged from 0.06 for normal propanol to 0.13 for R-113.

Fig. 3. Intertube flow modes for ethylene glycol according to Hu and Jacobi [50]: (a) droplet; (b) droplet-jet; (c) in-line jet; (d) staggered jet; (e)
jet-sheet; (f ) sheet mode.
642 G. Ribatski, A.M. Jacobi / International Journal of Refrigeration 28 (2005) 635–653

For the three test fluids, the jet-sheet mode began at KZ0.32, literature related to falling films on horizontal, enhanced
and a complete sheet was observed when K reached from 0.37 surfaces and a summary of the experimental conditions.
to 0.47. In a similar report for R-113 on finned tubes, Honda Commercial surfaces developed for flooded evaporators
et al. [53] found a significant increase in Re for transition from (Turbo-B, Turbo-BII, High Flux and Thermoexcel-E),
the jet mode to the sheet mode with a downward-flowing condensers (Turbo-CII, GEWA-SE, GEWA-SC, Turbo-
vapor. However, for three-dimensional fins this effect was Chil and Thermoexcel-C) and finned tubes are listed in
dramatically reduced due to the promotion of the liquid Table 2. According to this table, no research has been
distribution by the surface structure. carried out for new refrigerants (zeotropic blends) or
More recently, Roques et al. [54] investigated adiabatic hydrocarbons. Although not listed in this table, the liquid
flow modes on plain and enhanced surfaces. They proposed feeder devices used in the experiments are similar to the
correlations similar to those of Hu and Jacobi [50]. Flow ones used on plain surfaces research.
mode hysteresis effects were not observed. Plain and Turbo- In general, convection-enhancing surfaces present higher
BII surfaces presented a lower Re at the transition from the heat transfer enhancements at low heat fluxes. According to
jet to the jet-sheet mode. Lower Reynolds numbers at Chyu and Bergles [15], the heat transfer enhancement
transitions concerning droplet and jet modes were noted for provided by GEWA-T surface is a consequence of the
the surface Thermoexcel-C. This behavior was related to the increase in the surface area by the fins. Their conclusion was
effects of small fins that enabled the jet mode to remain based on observations according to which h enhancement is
stable at low Re. For the Thermoexcel-C and Turbo-Chil the proportional to the increase in the overall surface area. For
complete sheet mode began at a relatively high Re when an increase of 100% in the surface area provided by helical
compared to the others structures. grooves, Conti [7] noted an increase of 3.5 times in h
Fig. 4 shows the mode transitions as given by different compared to a plain tube. For finned surfaces under
investigators. The figure reflects a significant scatter in tran- convectively dominated conditions, Liu and Yi [36] found
sition data; however, such results are reasonable given the an increase in h due to surface tension effects and an
subjective nature of interpreting two-phase flow regimes. increase in the surface area. For a surface with conical
cavities and water as refrigerant, they noted a similar
behavior, in this case due to the increase in the surface area
4. Heat transfer on enhanced surfaces [36]. For surfaces with longitudinal grooves, Putilin et al.
[34] and Rifert et al. [19] noted an increase between 30 and
The use of enhanced (structured) surfaces can provide 90% in h compared to plain tubes. Both studies report higher
heat transfer coefficients approximately 10 times higher than heat transfer coefficients in the region close to the end of the
those obtained on plain surfaces. Furthermore, they can also grooves, due to boundary layer disruption [19,34]. On the
improve the liquid refrigerant distribution. The mechanisms contrary, according to Sabin and Poppendiex [8], longitudi-
causing the enhanced behavior can be divided into the nal grooves have no effects on the heat transfer coefficient.
following categories: enhanced boiling, achieved by surface This apparent contradiction is related to the groove dimen-
structures that promote bubble nucleation; enhanced con- sions, because on the surfaces of Refs. [19,34] grooves
vection, due to the effect of surface structures on surface present a depth about 10 times higher than those of [8].
area, film velocity profile and turbulence, and interfacial Shallow grooves behave as a simple surface roughness, and
temperature variation and the attendant Marangoni effects. will not affect h for laminar flow. For the experimental
Summarized in Table 2 are several studies from the results of Sabin and Poppendiex [8], a diamond-knurled
surface gives the best performance.
At high heat flux, boiling effects dominate and surface
structures that promote bubble nucleation tend to provide
higher heat transfer coefficients. Therefore, at high heat
fluxes Thermoexcel-E and High Flux surfaces give higher
performance than does GEWA-T [13,15,56]. At reduced f,
although bubble nucleation was not observed, Sabin and
Poppendiex [8] used the High Flux surface and noted a heat
transfer coefficient three times that of plain tubes, but Chyu
et al. [11] reported a similar h for these surfaces. According
to Liu and Yi [36], at high heat flux, the surface with conical
cavities presents a better performance than the finned
surface and the Gewa-T surfaces studied by Chyu and
Bergles [15]. Wang et al. [18] optimized the dimensions of
JK-1 surfaces under boiling-dominated conditions. Accord-
Fig. 4. Comparison of correlations for flow mode transitions on ing to Tan et al. [23], the surfaces type JK-2, an evolution of
plain tubes; it neglects transition hysteresis. JK-1 obtained by machining grooves in the surface tunnels,
G. Ribatski, A.M. Jacobi / International Journal of Refrigeration 28 (2005) 635–653 643

Table 2
Description of experimental studies involving falling film on enhanced surfaces

Fluid Surface characteristics Dext G (x103 kg Tsat or Tl f (kWmK2)


(mm) sK1 mK1) (8C)
[4] Sea water Knurled 50.8 0–228 49–127 0–63.1
[7] Ammonia High flux and helically grooved tubes 50 3.7–204 22.2 5–16
with 1100 and 315 groovesmK1
[8] Ammonia High flux and the following surface 25.4 3.7–37.1 12.8–23.9 3.2–25.2
preparations: turned and stripe-
burnished, diamond knurled, straight
knurled and covered with rough nickel
[11] Water GEWA-T19C, GEWA-T26B, high flux 25.4 37–110 99.4 2–208
[13] Water High flux, GEWA-T19C, GEWA-T26B 25.4 21–210 99.4 0.6–208
and thermoexcel-E
[15] Water GEWA-T19C, GEWA-T26B, 25.4 28–212 99.4 1–130
thermoexcel-E, high flux
[16] Water and ethyl alcohol High flux 25.4 2.5–50 45, 57 0–80
[18] R-113 15 Distinct configurations of JK-1 tube – 32–86 47 7–80
that presents a porous structure
[19] Water Longitudinally-profiled tubes 38.0 40–400 40–100 15–75
[21] R-11 Similar to JK-1 with addition of a 18 8–91 23.5 1–43
protuberance. Tubes with distinct
protuberance configurations.
[23] R-113 JK-1 and JK-2 (a second generation of 22.0 32–96 47.2 7.3–72.7
JK-1 tubes)
[26,30] Mixtures of R-134a and the GEWA-SE, GEWA-SC, Turbo-B, 19 6.5 2 5–40
polyol-ester oils: 32 cs/40 8C Turbo-CII, finned tubes 1024 fins mK1,
and 68 cs/40 8C u up to 5% 1575 fins mK1
[34] Water 20 Distinct configurations of 30, 38 40–200 42–100 15–75
longitudinally profiled tubes
[36] Water and R-11 Surface with conical cavities, and a 18 15–354 23.5, 99.4 2–500
helically finned tube 1429 fins mK1
[54] Water, glycol, and Turbo-BIIi, Turbo-Chili, thermoexcel-Cii 12.7ii, 0–210 15–35 0
water glycol mixtures 19.1i
[55] Ammonia Finned 1575 fins mK1 and corrugated 19 6.95–39.4 K23.3–10 10–80
tubes

achieves a heat transfer coefficient that is 50–190% higher surfaces, and Zeng et al. [55] using finned and corrugated
than that of JK-1, as can be noted in Fig. 5. The experimental surfaces, observed the boiling regime only. On the other
points shown in this figure were digitized from the hand, Kuwahara et al. [21] pointed out a marginal effect of f
publication. This improvement results from greater rough- on a boiling-enhanced surface despite the occurrence of
ness in the tunnels of the JK-2 than in the JK-1. The bubble nucleation. Moeykens et al. [30] noted that the heat
roughness increases the number of nucleation sites. In transfer performance increased with f, reached a maximum,
addition, at high heat flux the internal grooves in JK-2 and then declined with a further increase in heat flux. The
improve the liquid distribution on the surface through decrease in h is most likely due to the partial dryout. A
capillary effects [23]. Moeykens et al. [30] observed that distinct behavior is observed for GEWA-SC whose
enhanced boiling surfaces gave a higher performance than performance is weakly dependent on f. These behaviors
finned tubes and a lower performance than the surfaces are also displayed in Fig. 6. The surfaces Thermoexcel-E,
developed for condensers. Fig. 6 summarizes these results JK-1 and JK-2 use superficial structures characterized by
and also shows the experimental data of Webb and Pais [57]. minute parallel tunnels with tiny holes (porous) to connect
According to Fig. 6, the pool boiling on the Turbo-B surface to the outside surface. The heat transfer enhancement
gives the highest performance. achieved by these surfaces is associated with the heat flux
Liu and Yi [36] defined the convective and boiling level and is related to liquid suction and bubble ejection
regimes for falling films. In the convective regime, h is through the pores in a manner similar to the mechanism
constant, while in the boiling regime h increases with f. proposed by Nakayama et al. [58] for pool boiling [15,18].
They observed both regimes independent of the surface On the basis of this understanding, Wang et al. [18]
configuration. Wang et al. [18] using boiling-enhanced suggested increasing the pore diameter with f.
644 G. Ribatski, A.M. Jacobi / International Journal of Refrigeration 28 (2005) 635–653

enhanced surfaces. For finned surfaces and standard nozzles,


Zeng et al. [55] observed that at high Tsat the effects of G are
related to the heat flux level. At high f, h increases with G as
a result of the decrease in the dry area on the tube. An
opposite trend is observed at low f due to the increase in the
film thickness on fin tips in the absence of partial dryout.
Using finned surfaces and R-11, Liu and Yi [36] noted a
maximum h value for a Reynolds number of 250 and a
constant heat transfer coefficient at ReO1000. For a surface
with longitudinal grooves, Rifert et al. [19] observed an
increase in h with G, independent of f the magnitude of this
effect was related to the groove dimensions. For a similar
surface structure, Putilin et al. [34] noted that the degree of
heat transfer enhancement in relation to the plain surface
Fig. 5. Falling film heat transfer results of the JK-1 and JK-2 increases with G.
enhanced surfaces from digitized data of Tan et al. [23]. R-113, At high saturation temperatures, Zeng et al. [55] noted an
TsatZ47.2 8C, GZ0.0963 kg mK1 sK1. increase in h with an increase in the nozzle height and a
reduction in the nozzle angle. According to Chyu and
In general, at low heat fluxes, the falling film provides a Bergles [15,56], the liquid feeder height does not affect the
higher heat transfer rate than does pool boiling. This differ- heat transfer coefficient for boiling-enhanced surfaces, due
ence increases for convective-enhanced surfaces [15,18, to boiling dominance effects, or for convective-enhancing
21,36,56], and it is related to increased convective effects for surfaces, due to the reduction in the effect of the impact of
falling film relative to free convection, the main heat transfer the liquid by the fins. In the case of a High Flux surface at a
mechanism in incipient pool boiling. At high heat fluxes, this low heat flux they noted that the nozzle height affected h in a
effect diminishes due to the dominance of boiling effects. The similar manner as for plain surfaces.
heat flux at which these mechanisms compete depends on the According to Moeykens et al. [30], small concentrations
surface structure [15,21,36,56]. Hysteresis effects (distinct f of a lubricant can considerably increase the heat transfer
vs. (TwKTsat) curves depending upon whether the surface has performance. The oil concentration at which the optimal
been going through a heating or cooling process) were noted performance was observed decreased with an increase in f.
for boiling-enhancing surfaces in the following Refs. [11,13, This behavior was related to the foaming effect of the oil and
15,23,56]. In contrast, Kurahara et al. [21] noted small was noted for the Turbo-CII and a finned surface
hysteresis effects, though they obtained experimental results (1575 fins mK1).
for boiling-enhanced surfaces.
Without partial dryout on the surface, the flow rate
affects h solely through its impact on convection [15,21,36,56]. 5. Falling film on bundles of horizontal tubes
Wang et al. [18] noted an increase in h with G for boiling-
In contrast to the studies of a single tube, for which the
surface is generally heated by an electric current, the heating
effect was usually provided by the flow of a heated fluid in
the tube bundle studies. Table 3 describes the experimental
data bank of studies directed at falling-film tube bundles.
The principal conclusions of these studies are discussed in
the following paragraphs.
In the ideal case, the falling film would be uniformly
distributed on the evaporator surface; however, maldistribu-
tion effects and even partial dryout occur in applications
where a uniform distribution is not achieved. When dryout
occurs, the heat transfer coefficients are higher for the upper
tubes due to partial dryout on the lower surfaces [38,58].
When all surfaces are wet, the upper tubes still exhibit a
higher h, due to the impact of the liquid from the feeder [70],
and the local increase in G from row to row because of the
Fig. 6. A comparison between the results of R-134a from Moeykens irrigation caused by interstitial drizzles [65]. In a square-
[26] for falling film evaporation on single plain, and enhanced pitch arrangement, Moeykens et al. [67,68] using R-134a
surfaces at TsatZ2 8C, and the data from Webb and Pais [57] for noted a higher performance in the first row, while for R-123
pool boiling on enhanced surfaces at 4.4 8C. they found an increase in h from row to row. A similar
G. Ribatski, A.M. Jacobi / International Journal of Refrigeration 28 (2005) 635–653 645

Table 3
Description of experimental studies involving falling film on tube bundles

Fluid Liquid feeder characteristics Tube bundle layout/tube characteris-


ticsa/tube diameter/tube material
[23] R-123, water, glycol, water–glycol A reservoir with the low region pre- 5 Horizontal plain tubes vertically
mixtures, alcohol, and hydraulic oil senting a form of a semi-cylinder with aligned 1%H%80 mm/1%Dext%80 -
regularly spaced holes in its lower mm/stainless steel
generatrix.
[26] R-134a, R-22, R-123 and mixtures of Wide-angle low-pressure-drop com- 20 Tubes in 4 vertical rows; triangular-
these refrigerants with oil lubricant. mercial nozzles; solid pattern, circular and square-pitch/finned tube
and square plum type; orifices diameter (1575 fins mK1), GEWA-SC, Turbo-B,
of 3.97, 4.76 and 5.56 mm nozzle height Turbo-CII and plain tube/Dextz19 mm/
of 41.3 and 66.7 mm copper
[36] Water and R-11 Feeder with wedge shape and nozzle 3 Tubes in a vertical row/plain tube,
width of 1 mm surface with conical cavities, and a
helically finned tube 1429 fins mK1/
DextZ18 mm/copper
[38] R-11 Porous sintered tube; tube with holes 5 Horizontal plain tubes vertically
along its bottom; and a plate with a row aligned/DextZ25 mm/copper.
of holes in its center line
[59] R-22, R-12 and R-113 A perforated tube inside another tube 1 and 3 horizontal rows with 5 vertical
with drilled holes in its upper generatrix rows, 22 horizontal rows vertically
aligned. Distinct triangular layouts
including 60 tubes/18 mm/stainless
steel
[60] Sea water – /16, 24 and 32 mm/copper
[61] Water A tray with 3 rows of 3 mm holes, 3 Horizontal rows with 5 vertical rows/
drilled 20 mm apart elliptical and circular/19 mm/aluminum
[62] R-12, R-22 Similar to [59] 5 Horizontal tubes vertically aligned/9
types of porous tubes with coating by
deposition, spraying, and sintering; and
4 types of jacketing tubes/20 mm/
stainless steel and copper
[63] Ammonia Spray nozzles 100 Columns each with 30 plain tubes;
308 triangular layouts/25.4 mm/tita-
nium
[64] Water Distance between the nozzles and tubes 3 Tubes in a vertical row; 3–2–3
of 76 mm triangular-pitch; and 3 by 3 square
pitch/plain tube, finned tube (1575 fins/
m) and longitudinally groove surface;/
19 mm/stainless steel
[65] Ammonia Standard- and wide-angle commercial 3!3 square-pitch and a 3–2–3 triangu-
round full-cone nozzles; distance lar-pitch/DextZ19 mm/stainless steel
between nozzles and tubes of 150 mm.
[66] R-134a Wide-angle commercial nozzles that 20 Tubes in 4 vertical rows; triangular-
promote circular and square plumes pitch/finned tubes 1575 fins mK1/
(orifices DZ4 mm DZ4.76 mm and DextZ19 mm/copper
DZ5.56 mm) distances between noz-
zles and tubes of 41.3 and 66.7 mm
[67] R-134a Wide-angle commercial round full- 20 Tubes in 4 vertical rows; triangular-
cone nozzles (orifice DZ5.56 mm), and square-pitch/finned tube
distance between nozzles and tubes of (1575 fins mK1), GEWA-SC, Turbo-B,
66.7 mm Turbo-CII and plain tube/DextZ19 mm/
copper.
[68] Mixtures of R-123 and a 305 SUS Wide-angle commercial round full- 20 Tubes in 4 vertical rows triangular-
naphthenic mineral oil; u up to 2.5% cone nozzles (orifice DZ4.76 mm and and square-pitch/plain; Turbo-B and
DZ5.56 mm), distance between noz- Turbo-CII/DextZ19 mm/copper
zles and tubes of 44.3 mm
646 G. Ribatski, A.M. Jacobi / International Journal of Refrigeration 28 (2005) 635–653

Table 3 (continued)

Fluid Liquid feeder characteristics Tube bundle layout/tube characteris-


ticsa/tube diameter/tube material
[69] Mixtures of R-134a and oil 340-SUS Similar to [67] 20 Tubes in 4 vertical rows; triangular-
polyol-ester ; R-22 and oil 300-SUS pitch/plain tube, GEWA-SC, Turbo-B,
alkyl-benzene; u up to 2.5% finned tube (1575 fins/m)/DextZ
19 mm/copper
[70] Ammonia Standard- and wide-angle commercial 3 by 3 square pitch/DextZ19.1 mm/
round full-cone nozzles; distances stainless steel
between nozzles and tubes of 50.8 and
102 mm
[71] R-141b Full cone circular spray nozzles with an 5 Tubes in a triangular-pitch arrange-
orifice diameter of 2 mm ment with 2 horizontal rows/plain tubes
with and without a peripheral liquid
collector/DextZ19 mm/copper
[72] R-11 A tube with holes at its bottom above 36 Tubes (1024 fins/m) in 5 horizontal
each column, and two tubes with holes rows; triangular-pitch arrangement/
on their surfaces positioned in the third DextZ19.1 mm/R-11 vapor in counter-
horizontal row current flow
[73,74] Ammonia Standard- (orifice DZ4.76 mm/spray 3–2–3 triangular-pitch/DextZ19 mm/
angle 908) and wide-angle (orifice DZ stainless steel
3.99 mm/spray angle 1108) commercial
round full-cone nozzles; distance
between nozzles and tubes up to
150 mm
[75] Water Similar to [36] 3 Horizontal tubes vertically aligned/
DextZ13, 20, and 30 mm/copper
a
When not specified, consider the experiments as conducted solely for plain tubes.

behavior was measured by Fujita and Tsutsui [38], Liu and Turbo-B surface reveal that the triangular-pitch bundle
Yi [36] and Liu et al. [75]. Danilova et al. [59] and Bukin gives higher bundle-averaged heat transfer coefficients than
et al. [62] did not note substantial variations along the tube does the square-pitch bundle at high heat fluxes. An opposite
bundle depth. behavior occurs at low heat fluxes. In the square-pitch
In a triangular-pitch arrangement, the flow rate distri- bundle, the effects of f and G are less prominent. On the
bution tends to be less uniform than in a square-pitch other hand, for R-123 and the same surface structure,
arrangement and thus row-to-row variation is larger [26]. Moeykens et al. [68] found that a square-pitch bundle gave a
Plain surfaces in a triangular-pitch bundle exhibit a decrease higher performance. A less uniform flow distribution on
of h from row to row [63,69]. This behavior is noted for low triangular-pitch bundles [64] seems related to higher f
flow rates [63] and high heat fluxes [69]. Finned tubes such effects for this arrangement. Zeng et al. [73,74] pointed out
as GEWA-SC can exhibit maldistribution effects, because that square-pitch bundle tends to provide a higher
the surface structure inhibits the longitudinal movement of performance than a triangular-pitch bundle at low saturation
liquid [64,67]. For the surfaces Turbo-CII and Turbo-B, temperatures, and the triangular-pitch bundle is more likely
Moeykens and co-workers found that the relative perform- to provide a better performance at high saturation tempera-
ance among rows varied with the refrigerant, heat flux, and tures, as shown in Fig. 7. The experimental points shown in
film flow rate [67–69]. For R-123 and low heat fluxes, the this figure were digitized from the publication. Fig. 7 also
surface Turbo-B gives the highest performance at the lowest reveals an increase in the heat transfer performance with the
row [68]. Some methods have been proposed to avoid partial saturation temperature. In this figure, contrary to Moeykens
dryout on the lower rows [71,72]. By adding spray tubes at [26], by comparing heat transfer performance of the distinct
the center of the bundle, Tatara and Payvar [72] found an tubes in the bundle, a less uniform flow rate distribution can
increase in the thermal performance, relative to a configur- be noted for triangular-pitch arrangement. Two distinct
ation that used only liquid drip tubes above the tube bundle; arrangements of the triangular-pitch bundle were compared
they also found, a reduction in the overfeed rate for a similar by Danilova et al. [59]. A comparison shows that the per-
global h. Chang and Chiou [71] added liquid collectors to formance increases with a reduction in the horizontal
the bottom of the tubes and thus reduced the difference in the distance between vertical rows.
local h along the bundle depth. Moeykens et al. [26], using R-134a on tube bundles,
The results of Moeykens et al. [67] for R-134a and reported the following order of performance as shown
G. Ribatski, A.M. Jacobi / International Journal of Refrigeration 28 (2005) 635–653 647

Fig. 7. Falling film heat transfer results of individual tubes on square (filled symbols) and triangular-pitch (blank symbols) plain-tube bundles
from Zeng et al. [73]. Ammonia at (a) TsatZK23.3 8C and (b) TsatZ10 8C, DextZ19.1 mm, Gz0.075 kg mK1 sK1, nozzle height of 50.8 mm,
pitch ratio of 1.25, and standard-angle nozzles.

in Fig. 8: Turbo-CII, Turbo-B, GEWA-SC, finned


(1575 fins mK1) and plain tubes. Contrary to what is
shown for single surfaces in Fig. 6, in which the
enhanced-condensation tubes give similar performances,
bundles using Turbo-CII provided a performance up to
100% higher than those using GEWA-SC. This behavior can
be related to the restriction of the longitudinal liquid
movement on GEWA-SC by the fins, resulting in a large
variation in the local flow rate that increases in the lower
rows, and promotes the formation of dry patches [63,66].
For R-123, the Turbo-B tube provides a better performance
than does the Turbo-CII [68]. Thome [2] suggested that the
reason why the Turbo-CII outperformed the Turbo-B for R-
134a but the opposite occurred for R-123 could be the large
surface tension difference between these refrigerants and the
distinct mechanisms of the liquid retention provided by
these surfaces.
How much the bundle-averaged heat transfer coefficient
is affected by the flow rate depends on the surface structure
and the bundle layout [67,68]. When the surfaces are wet, a Fig. 8. Falling film heat transfer results on tube bundles from
flow rate increase seems to reduce the global h due to an Moeykens [26]. R-134a, TsatZ2 8C, triangular-pitch, a nozzle
increased film thickness [67]. On the other hand, when dry height of 66.7 mm, and wide-angle nozzles.
648 G. Ribatski, A.M. Jacobi / International Journal of Refrigeration 28 (2005) 635–653

patches are present, the global h increases with G due to the the spray flow. The results of Moeykens and Pate [66] can be
decrease in the dry areas [9,28,38,62,71]. These behaviors related to both of these statements, since they noted that the
are noted in Fig. 8 for Turbo-CII. At low f, the performance square nozzle presented a higher global h than the round
at the lowest flow rate is high; at high f, the performance at nozzle, and an increase in the nozzle spacing reduced
the lowest flow rate is low. For the square-pitch bundle and differences in performance.
at Tsat higher than K23,3 8C, Zeng et al. [70] observed a Moeykens and his co-workers [68,69] also investigated
slight increase in global h with G, while for the triangular- the performance of the bundles with refrigerant/oil mixtures.
pitch bundle Zeng et al. [74] found an insignificant effect of For refrigerants R-134a and R-22, the addition of oil pro-
G, independent of Tsat. No effect of G on global h was noted moted foaming, which is suggested to enhance the heat
in the following Refs. [36,62,75]. Danilova et al. [59] found transfer because it helps maintain wetting of the bundle [68].
that the behaviors for global h with G on tube bundles were On the other hand, for R-123, the heat transfer performance
similar to those of a single tube, due to the dominance of decreased with the addition of oil on the Turbo-B bundle. In
boiling effects, and as pointed out by Fujita and Tsutsui [38], the case of Turbo-CII and plain tube bundles, the oil addi-
to the flow regime on the surfaces (laminar, turbulent). tion to R-123 improves the bundle performance at low f.
Danilova et al. [59] also suggested that the flow rate be For this refrigerant, foaming is observed solely with plain
as high as possible in order to obtain the maximum tubes [69]. Apparently, foaming heat transfer enhancement
evaporation-zone heat-transfer. with the addition of oil cannot be due to foaming alone.
The following has been noted for the global h with
increasing f: h decreases due to increased dry areas within
the bundle [38,68,71]; h increases when boiling effects are 6. Methods to predict the heat transfer rate
dominant [59,62,68,70,74]; h is almost constant under
strictly convective conditions [36,38,59]. An increase in the 6.1. Empirically based correlations
vapor flow with f, promoted by the higher evaporation rate,
can also increase the global h, due to the enhancement of the In general, most efforts that involve the development of
convective effects (caused by vapor shear), principally on heat transfer correlations for falling films on horizontal
the upper rows. The behavior of global h with f depends cylindrical surfaces are intended for direct application;
upon, among other factors, the heat flux itself, G (as shown therefore, such correlations provide reasonable results in
in Fig. 8), the refrigerant, the surface structure and the layout their range of applicability and can be implemented without
of the bundle [66–69]. much difficulty. However, because their development is
Vapor flow in the bundle can promote either an increase usually based on a restricted set of data, care must be taken
in the film thickness [72] or a reduction in it due to the drag to apply them within the parameter space of their
of liquid [45]. It can also promote either the formation of dry development—extrapolation of such correlations may not
patches or better liquid distribution [46]. Thus, the effects yield good predictions. Although existing correlations are
of a vapor flow can either increase or decrease the bundle limited in applicability, they are usually cast in terms of
average heat transfer coefficient. A majority of the studies dimensionless parameters such as the Nusselt, Prandtl,
listed in Table 3 were carried out on limited rows and do not Reynolds, and Archimedes numbers, along with dimension-
include vapor-shear effects. Therefore, when vapor-shear less geometric variables, with the hope that such an
effects are relevant to the bundle performance, care must approach will elucidate the physical mechanisms or
be exercised in trying to extrapolate these results to full relationships more clearly than a dimensional approach.
evaporator performance. Results from simulations using numerical models have also
Zeng et al. [70,74] observed a decrease in the global h been used to develop correlations [76,77].
with an increase in the nozzle height, which was enhanced at Proposed heat transfer correlations are summarized in
high temperatures. This behavior seems to be related to an Table 4, where the experimental data bank used to fit these
increase in maldistribution effects. For strictly convective correlations is also partially described. The selection of a
heat transfer, Liu and Yi [36] pointed out that the global h correlation should include an assessment of the flow regime
was not affected by the tube spacing. For square-pitch (laminar or turbulent) [10], and the falling-film flow mode
bundles and high temperatures, Zeng et al. [70,74] observed [33]. Furthermore, the prevailing heat transfer mechanisms
that a wide-angle nozzle provided a higher performance should be considered, e.g. for boiling conditions, different
than did a standard-angle nozzle; however, for a triangular- correlations are proposed [10,22,59]. Correlations deve-
pitch bundle the bundle performance was not affected by the loped from experimental results with water are widely
nozzle angle. Chyu et al. [65] developed an analytical available [17,22,28,31,33,75], as are correlations from
method to predict the liquid flow distribution on the first row experiments with ammonia [10,55,70,74]. Danilova and
of a bundle. In the same paper, they indicated that, in order co-workers [59] proposed correlations based on experi-
to cover the same area, more square nozzles than round mental results with the refrigerants of R-22, R-12 and R-113
nozzles are needed. However, they also suggested that the in a tube bank. The correlations proposed by Fujita and
square nozzles provided an advantage in the uniformity of Tsutsui [25,38] are based on experimental data using R-11.
G. Ribatski, A.M. Jacobi / International Journal of Refrigeration 28 (2005) 635–653 649

Table 4
Heat transfer correlations for for falling film evaporation on horizontal cylindrical surfaces

Equ- Correlation Data bank Comments


ation
[10] (5) NuZ2.2(H/Dext)0.1ReK1/3 Ammonia, single plain tubes Laminar flow Re!1.680PrK3/2
(6) NuZ 0:185ðH=Dext Þ0:1 Prl Turbulent flow ReR1.680PrK3/2
[17] (7) NuZ 0:0137ðReÞ0:349 Prl ððs=Dext Þ0:158 = Water, single plain tubes ReO320
ð1C expðK0:0032Re1:32 ÞÞ
[22] (8) NuZ 0:042Re0:15 Prl Water, single plain tubes Strictily convective DextZ25.4 mm
(9) NuZ 0:038Re0:15 Prl Strictily convective DextZ50.8 mm
(10) NuZ 0:00082Re0:10 Prl f0:4 Boiling conditions DextZ25.4 mm
(11) NuZ 0:00094Re0:10 Prl f0:4 Boiling conditions DextZ50.8 mm
[31] (12) NuZ 0:2071Re0:24 Prl ArK0:111 Water, single plain tubes diameter Properties evaluated at film average
effect according to [78] temperature
[33] (13) NuZ 0:113Re0:85 Prl ArK0:27 Water, ethylene glicol, mixture of water Droplet mode
ð1C s=Dext Þ0:04 and ethylene glycol, single plain tubes
(14) NuZ 1:378Re0:42 Prl ArK0:23 Jet mode
ð1C s=Dext Þ0:08
(15) NuZ 2:194Re0:28 Prl ArK0:20 Sheet mode
ð1C s=Dext Þ0:07
[38] (16) NuZ ½ReK2=3 C aRe0:3 Prl 1=2 R-11, vertical row of horizontal tubes Top tube aZ0.008 other tubes, aZ0.
010
[55] (17) NuZ 0:0568ReK0:0058 Prl ðpsat =pcrit Þ0:323 Ammonia, single 1575 fins mK1 tube
ðfDext =ðTcrit K Tsat Þkl Þ1:034
[59] (18) NuZ 0:03Re0:22 ðf=hlv rv nl ðnl =gÞ1=3 Þ0:04 R-22, R-12 and R-113 on tubes Strictily convective
Prl ðs=Dext Þ0:48 vertically aligned
(19) h=kl ðs=gðrl K rv ÞÞ1=2 Z 1:32 ! Boiling conditions
10K3 ðf=hlv rv nl ðs=gðrl K rv ÞÞ1=2 Þ0:63 !
ðpsat =sðs=gðrl K rv ÞÞ1=2 Þ0:72 Prl
[70] (20) NuZ 0:0495ReK0:00399 Prl ðpsat =pcrit Þ0:261 Ammonia, 3 by 3 square pitch
ðfDext =ðTcrit K Tsat Þkl Þ0:722
[74] (21) NuZ 0:0678Re0:049 Prl p0:456 r ðfDext = Ammonia, 3–2–3 triangular-pitch
ðTcrit K Tsat Þkl Þ0:704
[78] (22) NuZ 0:041Re0:30 Prl ArK0:04 Water on a vertical row of horizontal
tubes

Recently, Zeng and his co-workers [70,74] developed the flow regime, either laminar [6,8,13,14,40,75,79] or
correlations based on triangular- and square-pitch bundle turbulent [5,75,76,80]. A range of turbulence models has
spray evaporation with ammonia. been considered by those developing falling-film models.
Assuming the effects of waviness at the liquid–vapor
6.2. Analytically based models interface are similar to those for vertical falling films,
Kocamustafaogullari and Chen [40] and Rogers [79] applied
In addition to the empirical correlations cited above, a corrections for such effects to the empirical correlations that
number of models have been proposed. Typically, these were proposed by Zazuli [40] and Kutateladze and Gogonin
models classify the flow and heat transfer as shown in Fig. 9: a [81], respectively. It is interesting to note that Marangoni
free fall, jet impingement, thermal developing, and fully effect is not typically included, but some cases it might be
developed region. Local Nusselt numbers are obtained through important (such as for a subcooled film [79]). In most
the application of continuity, momentum and the energy models, heat transfer is considered to result in sensible
equations for each of these regions. Such modeling can either heating or thin-film evaporation; however, Lorenz and
result in simple correlations involving dimensionless para- Young [39] suggested a model based on the assumption that
meters, or a solution obtained through numerical methods. evaporation on the heating surface was due to the bubble
In general, differences between the models are due to nucleation. These effects are modeled through a correlation
how the flow is classified (i.e., the above mentioned for pool boiling [82] considering a uniform nucleation along
regions), the model used for the jet impingement region, the heating surface perimeter. Their approach does not agree
initial and boundary conditions, the falling film modes, and with experimental behaviors described earlier, according to
650 G. Ribatski, A.M. Jacobi / International Journal of Refrigeration 28 (2005) 635–653

Fig. 9. Illustration of falling film regions adopted in the models, and the heat transfer coefficient distribution along the surface perimeter
according to experimental results, and models from distinct authors.

which high intensity in bubble nucleation was noted on the turbulence effects increase with the flow rate. The observed
tube bottom. Parken [6] and Saiz Jabardo [83] suggested a trend at high Re for the model of Chyu and Bergles is related
study of the nucleation criteria for thin films, related to the to the changes in area associated to the development region
one proposed by Bergles and Rohsenow [84]. Finally, only that occur with changes in flow rate. It is interesting to note
the model developed by Liu [5] includes the effects of a that in the case of a sufficiently high flow rate, the fully
concurrent vapor flow. developed region is not present. Therefore, according to this
Fig. 9 also shows the local h along the surface perimeter model, the minimum in Nu with Re is not related to a
according to the experimental results of Liu [5] and Parken transition in the film flow regime from laminar to turbulent,
[6], and the models by Chyu and Bergles [14], Sabin and but rather it is due to development effects.
Popendiex [8], and Fujita and Tsutsui [80]. Generally
speaking, the models by Chyu and Bergles [14] and Fujita
and Tsutsui [80] deviate significantly from the experimental 7. Conclusions
results of Liu [5]. Distinct differences can be noted near the
tube bottom, due to the fact that the effects of heat transfer From this review, the following conclusions can be drawn:
enhancement in this region are not modeled. The model of
Chyu and Bergles does not predict the decrease in h near the The parameter space relevant to falling-film evaporator
tube bottom, related to the fact that the fully developed performance is large and complex; despite numerous
region is not present. In the case of the Sabin and studies, even some of the basic mechanisms responsible
Popendiex’s model, lower heat transfer coefficients can be
noted near top of the tube, because they considered only a
thermal developed region. Fig. 10 shows the flow rate
effects on the overall Nusselt, as predicted by the models of
Chyu and Bergles [14], Sabin and Popendiex [8], and Fujita
and Tsutsui [80]. The models of Sabin and Popendiex [8]
and Fujita and Tsutsui [80] for laminar film flow exhibit a
decrease in Nu with Re, independent of the Reynolds
number range. This behavior is simply due to an increase in
the film thickness with liquid flow rate. For Reynolds
numbers lower than 100, all the model results are similar.
Distinct trends can be noted at higher Reynolds numbers,
where the models proposed by Chyu and Bergles [14] and
Fujita and Tsutsui [80] for the turbulent flow predict the
Nusselt number to increase with Re. In the model of Fujita Fig. 10. Effects of film Reynolds number on the falling film Nusselt
and Tsutsui, the trend is caused by an increase in the film number of a plain tube according to distinct models. R-22, TsatZ
temperature gradients near the heating surface, because 0 8C, a feeder height of 0.003 and a tube diameter of 25.4 mm.
G. Ribatski, A.M. Jacobi / International Journal of Refrigeration 28 (2005) 635–653 651

for heat transfer behavior remain unclear. In particular, CNPq, Brazil. Both authors acknowledge support through
the conditions for incipient nucleate boiling need to be the Air Conditioning and Refrigeration Center at the
investigated, because the occurrence of nucleate boiling University of Illinois. We are also deeply grateful to Ms.
and its impact is obfuscated by the effects of flow rate, Xuli Tang who provided significant editorial assistance in
fluid properties, temperature and heat flux. the preparation of this manuscript.
In general, enhanced surfaces provide higher heat
transfer performance than do plain tubes; however,
References
confidence in the predicted enhancement is undermined
by complexity in the heat transfer dependence on
[1] O. Lyle, The efficient use of steam, H.M. Stationery Office,
geometry, tube layout, and operating conditions. Special London, 1947.
attention must be directed not only to the optimization of [2] J.R. Thome, Falling film evaporation: state-of-the-art review
enhanced tube geometry but also to the definition of the of recent work, J. Enhanced Heat Transfer 6 (1999) 263–277.
conditions in which enhancements can be clearly [3] L.S. Fletcher, V. Sernas, L. Galowin, Evaporation from thin
identified and quantified. water films on horizontal tubes, Ind Eng Chem, Process Des
Bundle depth effects related to liquid maldistribution and Develop 13 (1974) 265–269.
partial dryout remain unclear. Liquid distribution has a [4] L.S. Fletcher, V. Sernas, W.H. Parken, Evaporation heat
dramatic impact on evaporator performance. Because transfer coefficients for thin seawater films on horizontal tubes,
they are important and not well understood, additional Ind Eng Chem, Process Des Develop 14 (1975) 411–416.
[5] Liu PJP. The evaporating falling film on horizontal tubes. PhD
experiments on bundle-depth effects should be under-
Thesis, University of Wisconsin-Madison; 1975.
taken. The results of such studies can be used in the [6] W.H. Parken, Heat transfer to thin films on horizontal tubes.
development of methods to avoid dry regions and to PhD Thesis, Rutgers University; 1975.
quantify flow non-uniformity effects on performance. [7] R.J. Conti, Experimental investigation of horizontal tube
Apparently, the presence of oil can enhance the heat ammonia film evaporators with small temperature differen-
transfer coefficient by 100% under certain conditions, tials, Proceedings of the fifth ocean thermal energy conversion
but it can have a deleterious effect under other conditions. (OTEC), Miami Beach, vol. 3 1978, p. VI-161-80.
Systematic experiments with different refrigerant-oil [8] S.M. Sabin, H.F. Poppendiek, Film evaporation of ammonia
mixtures, surface structures, and geometric parameters over horizontal round tubes, Proceedings of the fifth OTEC
are necessary to understand this behavior and develop conference, Miami Beach, vol. 3 1978, p. VI-237-60.
[9] E.N. Ganić, M.N. Roppo, An experimental study of falling
reliable design tools.
liquid film breakdown on a horizontal cylinder during heat
Several models focusing on the prediction of h have been transfer, J Heat Transfer 102 (1980) 342–346.
proposed. However, in general, they do not include [10] W.L. Owens, Correlation of thin film evaporation heat transfer
Marangoni effect, vapor-shear effects, interfacial wavi- coefficients for horizontal tubes, Proceedings of the fifth
ness, or nucleate boiling effects. Empirical correlations OTEC conference, Miami Beach, vol. 3 1978, p. VI-71-89.
are strongly dependent on specific operating conditions [11] M.-C. Chyu, A.E. Bergles, F. Mayinger, Enhancement of
under which they were developed, and great care must be horizontal tube spray film evaporators, Proceedings of the
exercised in trying to generalize such relations. Further seventh international heat transfer conference, vol. 6 1982,
experimental work is needed before generalized corre- p. 275–80.
lations can be developed; carefully considered experi- [12] W.H. Parken, L.S. Fletcher, Heat transfer in thin liquid films
flowing over horizontal tubes, Proceedings of the seventh
ments should be undertaken to broaden the current
international heat transfer conference, Munich, vol. 4 1982,
parametric space for which data are available and to p. 415–20.
resolve contradictions in the extant data. [13] M.-C. Chyu, Falling film evaporation on horizontal tubes
The falling-film evaporator meets many of the needs of with smooth and structured surfaces, PhD Thesis, Iowa State
the air-conditioning and refrigeration industry. The University; 1984.
thermal performance of falling-film heat exchangers is [14] M.-C. Chyu, A.E. Bergles, An analytical and experimental
excellent—it is thermally superior to flooded evaporators study of falling-film evaporation on a horizontal tube, J Heat
and competitive with plate heat exchangers. Falling-film Transfer 109 (1987) 983–990.
heat exchangers are tolerant of contaminating gases and [15] M.-C. Chyu, A.E. Bergles, Horizontal-tube falling-film evapo-
can probably operate with a lower refrigerant charge ration with structured surfaces, J Heat Transfer 111 (1989)
518–524.
than plate heat exchangers.
[16] E.N. Ganić, D. Getachew, Effects of surface condition and
working fluid on liquid film breakdown during heat transfer,
Proceedings of the eighth international heat transfer con-
Acknowledgements ference, San Francisco, vol. 4 1986, p. 1931–36.
[17] J. Mitrovic, Influence of tube spacing and flow rate on heat
The first author gratefully acknowledges support through transfer from a horizontal tube to a falling liquid film,
a post-doctoral assistantship given by the Conselho Proceedings of the eighth international heat transfer con-
Nacional de Desenvolvimento Cientı́fico e Tecnológico, ference, San Francisco, vol. 4 1986, p. 1949–56.
652 G. Ribatski, A.M. Jacobi / International Journal of Refrigeration 28 (2005) 635–653

[18] G. Wang, Y. Tan, S. Wang, N. Cui, A study of spray falling water film on horizontal tubes, Exp Thermal Fluid Science 18
film boiling on horizontal mechanically made porous surface (1998) 183–194.
tubes, Proceedings of the international symposium of heat and [36] Z.H. Liu, J. Yi, Enhanced evaporation heat transfer of water
mass transfer enhancement and energy conservation (ISHTEEC), and R-11 falling film with the roll-worked enhanced tube
Guangzhou, 1988, p. 425–32. bundle, Exp Thermal Fluid Science 25 (2001) 447–455.
[19] V.G. Rifert, V.I. Podberezny, J.V. Putilin, J.G. Nikitin, [37] G. Ribatski, J.M. Saiz Jabardo, Experimental study of nucleate
P.A. Barabash, Heat Transfer in thin film-type evaporator boiling of halocarbon refrigerants on cylindrical surfaces, Int J
with profiled tubes, Desalination 74 (1989) 363–372. Heat Mass Transfer 2003; 4439–4451.
[20] J.T. Rogers, S.S. Goindi, Experimental laminar falling film [38] Y. Fujita, M. Tsutsui, Experimental investigation of falling
heat transfer coefficients on a large diameter horizontal tube, film evaporation on horizontal tubes, Heat Transfer—Jpn Res
Can J Chem Eng 67 (1989) 560–568. 27 (1998) 609–618.
[21] H. Kuwahara, A. Yasukawa, W. Nakayama, T. Yanagida, [39] J.J. Lorenz, D.T. Yung, Combined boiling and evaporation of
Evaporative heat transfer from horizontal enhanced tubes in liquid films on horizontal tube, Proceedings of the fifth ocean
thin film flow, Heat Transfer—Jpn Res 19 (1990) 83–97. thermal energy conversion (OTEC), Miami Beach, vol. 3
[22] W.H. Parken, L.S. Fletcher, V. Sernas, J.C. Han, Heat transfer 1978, p. VI-46-70.
through falling film evaporation and boiling on horizontal [40] G. Kocamustafaogullari, I.Y. Chen, Falling film heat transfer
tubes, J Heat Transfer 112 (1990) 744–750. analysis on a bank of horizontal tube evaporator, AIChE J 34
[23] Y. Tan, G. Wang, S. Wang, N. Cui, An experimental research (1988) 1539–1549.
on spray falling film boiling on the second generation [41] L.K. Brumfield, T.G. Theofanous, On the prediction of heat
mechanically made porous surface tubes, Proceedings ninth transfer across turbulent liquid films, J Heat Transfer 98
international heat transfer conference, Jerusalem, vol. 6 1990, (1976) 496–502.
p. 269–73. [42] V.P. Carey, Liquid–vapor phase change phenomena, An
[24] Y. Fujita, M. Tsutsui, Experimental and analytical study of introduction to the thermophysics of vaporization and con-
evaporation heat transfer in falling films on horizontal tubes, densation processes in heat transfer equipment, Taylor and
Proceedings of the 10th international heat transfer conference, Francis Books Ltd, New York, 1992, p. 105–6.
Brighton, vol. 6 1994, p. 175–80. [43] J.R. Thome, Condensation on external surfaces, Engineering
[25] Y. Fujita, M. Tsutsui, Evaporation heat transfer of falling databook III, 2004 chapter 7, http://www.wlv.com/products/.
films on horizontal tube. Part 2. Experimental study, Heat [44] M. Cerza, V. Sernas, Boiling nucleation criteria for a falling
Transfer—Jpn Res 24 (1995) 17–31. water film, Proceedings of the 23rd national heat transfer
[26] S.A. Moeykens, Heat transfer and fluid flow in spray evapo- conference, multiphase flow and heat transfer, Denver, CO,
rators with application to reducing refrigerant inventory. PhD vol. 47 1985 ASME HTD, p. 111–6.
Thesis, Iowa State University; 1994. [45] J.M. Gonzalez Garcia, J.M. Saiz Jabardo, W.F. Stoecker,
[27] S.A. Moeykens, M.B. Pate, Spray evaporation heat transfer Falling film ammonia evaporators. ACRC Technical Report,
of R-134a on plain tubes, ASHRAE Trans 100 (2) (1994) TR-33, University of Illinois, Urbana; 1992.
173–184. [46] D. Yung, J.J. Lorenz, E.N. Ganić, Vapor/liquid interaction and
[28] R. Armbruster, J. Mitrovic, Heat transfer in falling film on a entrainment in falling film evaporators, J Heat Transfer 102
horizontal tube, Proceedings of the national heat transfer (1980) 20–25.
conference, Portland, vol. 12 1995 [ASME HTD-vol. 314], [47] S.S. Kutateladze, I.I. Gogonin, V.I. Sosunov, The influence of
p. 13–21. condensate flow rate on heat transfer in film condensation of
[29] X. Hu, The intertube falling-film modes: transition, hysteresis stationary vapour on horizontal tube tube banks, Int J Heat
and effect on heat transfer. PhD Thesis, University of Illinois; Mass Transfer 28 (1985) 1011–1018.
1995. [48] R. Armbruster, J. Mitrovic, Patterns of falling film flow over
[30] S.A. Moeykens, W.W. Huebsch, M.B. Pate, Heat transfer of horizontal smooth tubes, Proceedings of the 10th international
R-134a in single-tube spray evaporation including lubrificant heat transfer conference, Brighton, vol. 3 1994, p. 275–80.
effects and enhanced surface results, ASHRAE Trans 101 (1) [49] V. Dhir, K. Taghavi-Tafreshi, Hydrodynamic transition during
(1995) 111–123. dripping of a liquid from underside of a horizontal tube,
[31] J.T. Rogers, S.S. Goindi, M. Lamari, Turbulent falling film ASME Paper 1981; [No 81-WA/HT-12].
flow and heat transfer on horizontal tube, Proceedings of the [50] X. Hu, A.M. Jacobi, The intertube falling film. Part 1. Flow
national heat transfer conference, Portland, vol. 12 1995 characteristics, mode transitions, and hysteresis, J Heat
[ASME HTD-vol. 314], p. 3–12. Transfer 118 (1996) 616–625.
[32] X. Zeng, M.-C. Chyu, Z.H. Ayub, Evaporation heat transfer [51] Y. Wei, A.M. Jacobi, Vapor-shear, geometric, and bundle-
performance of nozzle-sprayed ammonia on a horizontal tube, depth effects on the intertube falling-film modes, Proceedings
ASHRAE Trans 101 (1) (1995) 136–149. of the first international conference on heat transfer, fluid
[33] X. Hu, A.M. Jacobi, The intertube falling film. Part 2. Mode mechanics, and thermodynamics, vol. 1(1), Kruger National
effects on sensible heat transfer to a falling liquid film, J Heat Park, South Africa, 2002, p. 40–6.
Transfer 118 (1996) 626–633. [52] H. Honda, S. Nozu, Y. Takeda, Flow characteristics of
[34] J.V. Putilin, V.L. Podberezny, V.G. Rifert, Evaporation heat condensate on a vertical column of horizontal low finned
transfer in liquid films flowing down horizontal smooth and tubes, Proceedings of the ASME/JSME thermal engineering
longitudinally profiled tubes, Desalination 105 (1996) 165–170. joint conference, Honolulu, vol. 1, 1987, p. 517–24.
[35] R. Armbruster, J. Mitrovic, Evaporative cooling of falling [53] H. Honda, B. Uchima, S. Nozu, H. Nakata, E. Torigoe, Film
G. Ribatski, A.M. Jacobi / International Journal of Refrigeration 28 (2005) 635–653 653

condensation of R-113 on in-line bundles of horizontal finned [69] S.A. Moeykens, M.B. Pate, Effect of lubricant on spray
tubes, J Heat Transfer 113 (1991) 479–486. evaporation heat transfer performance of R-134a and R-22 in
[54] J.F. Roques, V. Dupont, J.R. Thome, Falling film transitions tube bundles, ASHRAE Trans 102 (1) (1996) 10–426.
on plain and enhanced tubes, J Heat Transfer 124 (2002) [70] X. Zeng, M.-C. Chyu, Z.H. Ayub, Performance of nozzle-
491–499. sprayed ammonia evaporator with square-pitch plain-tube
[55] X. Zeng, M.-C. Chyu, Z.H. Ayub, Ammonia spray evaporation bundle, ASHRAE Trans 103 (2) (1997) 68–81.
heat transfer performance of single low-fin and corrugated [71] T.B. Chang, J.S. Chiou, Spray evaporation of R-141b on a
tubes, ASHRAE Trans 104 (1A) (1998) 185–196. horizontal tube bundle, Int J Heat Mass Transfer 42 (1999)
[56] M.-C. Chyu, A.E. Bergles, Enhancement of horizontal tube 1467–1478.
spray film evaporators by structured surfaces, Adv Enhanced [72] R.A. Tatara, P. Payvar, Measurement of spray boiling refrige-
Heat Transfer 43 (1985) 39–47. rant coefficients in an integral-fin tube bundle segment simu-
[57] R.L. Webb, C. Pais, Nucleate pool boiling data for five lating a full bundle, Int J Refrig 24 (2001) 744–754.
refrigerants on plain, integral-fin and enhanced tube geo- [73] X. Zeng, M.-C. Chyu, Z.H. Ayub, Experimental investigation
metries, Int J Heat Mass Transfer 35 (1992) 1893–1904. on ammonia spray evaporators with triangular-pitch plain tube
[58] W. Nakayama, T. Daikoku, H. Kuwahara, T. Nakajima, bundle. Part I. Tube bundle effect, Int J Heat Mass Transfer 44
Dynamic model of enhanced boiling heat transfer on porous (2001) 2299–2310.
surfaces. Part II. Analytical modeling, J Heat Transfer 102 [74] X. Zeng, M.-C. Chyu, Z.H. Ayub, Experimental investigation
(1980) 451–456. on ammonia spray evaporators with triangular-pitch plain tube
[59] G.N. Danilova, V.G. Burkin, V.A. Dyundin, Heat transfer in bundle. Part II. Evaporator performance, Int J Heat Mass
spray-type refrigerator evaporators, Heat Transfer—Soviet Transfer 44 (2001) 2081–2092.
Res 8 (1976) 105–113. [75] Z.H. Liu, Q.Z. Zhu, Y.M. Chen, Evaporation heat transfer of
[60] V. Slesarenko, Investigation of heat transfer exchange during falling film on a horizontal tube bundle, Heat Transfer—Asian
sea water boiling in a horizontal thin film desalination plant, Res 3 (2002) 42–55.
Desalination 28 (1979) 311–318. [76] D. Barba, R. Di Felice, Heat transfer in turbulent flow on a
[61] R. Semiat, S. Sideman Moalen, D. Maron, Turbulent film horizontal tube falling film evaporator. A theoretical approach,
evaporation on and condensation in horizontal elliptical Desalinations 51 (1984) 325–333.
conduits, Proceedings of the sixth international heat transfer [77] I.Y. Chen, G. Kocamustafaogullari, An experimental study
conference, Toronto, vol. 4, 1978, p. 361–6. and practical correlations for overall heat transfer performance
[62] V.G. Bukin, G.N. Danilova, V.A. Dyundin, Heat transfer from of horizontal tube evaporator design. Heat transfer equipments
freons flowing over bundles of horizontal tubes that carry a fundamentals, design, applications and operating problems,
porous coating, Heat Transfer—Soviet Res 14 (1982) 98–103. ASME HTD 108 (1989) 23–32.
[63] J.J. Lorenz, D. Yung, Film breakdown and bundle-depth [78] V. Sernas, Heat transfer correlation for subcooled water films
effects in horizontal-tube, falling-film evaporators, J Heat on horizontal tubes, J Heat Transfer 101 (1979) 176–178.
Transfer 104 (1982) 569–571. [79] J.T. Rogers, Laminar falling film flow and heat transfer
[64] X. Zeng, M.-C. Chyu, Z.H. Ayub, Characteristic study of characteristics on horizontal tubes, Can J Chem Eng 59 (1981)
sprayed fluid flow in a tube bundle, ASHRAE Trans 100 (1) 213–222.
(1994) 63–72. [80] Y. Fujita, M. Tsutsui, Evaporation heat transfer of falling films
[65] M.-C. Chyu, X. Zeng, Z.H. Ayub, Nozzle-sprayed flow rate on horizontal tube. Part 1. Analytical study, Heat Transfer—
distribution on a horizontal tube bundle, ASHRAE Trans 101 Jpn Res 24 (1995) 1–16.
(2) (1995) 443–453. [81] S.S. Kutateladze, I.I. Gogonin, Heat transfer in film conden-
[66] S.A. Moeykens, M.B. Pate, The effects of nozzle height and sation of slowly moving vapor, Int J Heat Mass Transfer 22
orifice size on spray evaporation heat transfer performance (1979) 1593–1599.
for a low-finned, triangular-pitch tube bundle with R-134a, [82] W.M. Rohsenow, A method of correlating heat transfer data
ASHRAE Trans 101 (2) (1995) 420–433. for surface boiling liquids, J Heat Transfer 74 (1952) 969–976.
[67] S.A. Moeykens, B.J. Newton, M.B. Pate, Effects of surface [83] J.M. Saiz Jabardo, Estudo do mecanismo de evaporação em
enhancement film-feed supply rate, and bundle geometry on pelı́cula descendente. Report to National Council for Scientific
spray evaporation heat transfer performance, ASHRAE Trans and Technological Development (CNPq) of the Ministry of
101 (2) (1995) 408–419. Science and Technology of Brazil; 1996.
[68] S.A. Moeykens, J.E. Kelly, M.B. Pate, Spray evaporation heat [84] A.E. Bergles, W.M. Rohsenow, The determination of forced-
transfer performance of R-123 in tube bundles, ASHRAE convection surface-boiling heat transfer, J Heat Transfer—Ser
Trans 102 (2) (1996) 259–272. C 86 (1964) 365–372.

Potrebbero piacerti anche