Sunteți pe pagina 1din 13

Chemical Engineering and Processing 38 (1999) 683695

A corrugation geometry based model for efciency of structured


distillation packing
,
Z& . Olujic *, A.B. Kamerbeek
1
, J. de Graauw
Laboratory for Process Equipment, Delft Uni6ersity of Technology, Leeghwaterstraat 44, 2628 CA Delft, The Netherlands
Received 23 April 1999; accepted 23 April 1999
Abstract
Although the structured packing is a well established gasliquid contacting device, the understanding of its function is
insufcient and often leads to poor exploitation of the available phase separating potential. This is a consequence of a rather
supercial approach to modelling the packing performance through the years resulting in a lack of information on the nature and
extent of interaction between counter-currently owing gas and liquid phases and the micro and macro geometry of a rather
ordered structure with a pronounced ow discontinuity at the transition among packing elements. This paper addresses the
relation between the uid-dynamics imposed by packing geometry and the mass transfer efciency, and introduces a performance
prediction method which does not require packing specic constants to describe mass transfer coefcients of phases. 1999
Elsevier Science S.A. All rights reserved.
Keywords: Corrugated sheets; Distillation; Mass transfer efciency; Structured packing
www.elsevier.com/locate/cep
1. Introduction
Great advances have been made during the last 15
years in distillation technology, particularly in heavier
(bulk and ne) chemicals and air separations, by the
introduction of high efciency structured packing com-
posed mostly of corrugated (metal, stretched metal or
wire gauze) sheets. The geometry of a corrugated sheet
structured packing (CSSP) is highly ordered and en-
ables utilisation of largest possible surface areas per
unit volume. Thanks to its extremely large void fraction
(porosity) and low liquid hold up, CSSP has the poten-
tial for achieving high mass transfer efciency at a
relatively low pressure drop, which is not always
utilised fully in practice because of the detrimental
effect of liquid and gas ow maldistributions. There-
fore, its performance is very sensitive to malfunction of
column internals such as poorly designed or/and in-
stalled liquid distributors, liquid collecting devices, va-
pour distributor and support structures.
As suggested in the sketch shown in Fig. 1, CSSP is
installed as a bed of certain length/height and diameter,
consisting of a number of stacked elements (packing
layers) with heights ranging from 0.15 to 0.3 m, rotated
to each other usually by 90 to produce a large-scale
mixing effect for both gas and liquid at each transition
from layer to layer. Each packing element/layer is
equipped with wall wipers to avoid excessive wall ow.
Basically, each packing element/segment consists of a
number of corrugated sheets, each second with corruga-
tions inclined in the opposite direction. Standard corru-
gation inclination angle is 45 and often an angle of 60
is offered, for additional capacity gain. The most com-
mon surface area is around 250 m
2
/m
3
, and some
manufacturers offer surfaces from 100 to 750 m
2
/m
3
as
a standard. The exibility in this respect depends on the
design of the packing surface, which is the most distinc-
tive proprietary characteristic of a packing. With the
exception of Montz-pak B1 packing, all commercial
corrugated sheet structured packings have a regular
pattern of holes or other type of apertures in the
surface. This to some extent enables both phases to go

Dedicated to Professor Em. Dr-Ing. Dr h.c. mult. E.-U. Schlu n-


der on the occasion of his 70th birthday.

Based on a paper presented at the AIChE Annual Meeting,


November 1015, 1996, Chicago.
* Corresponding author. Tel.: +31-15-2786678; fax: +31-15-
2786975.
E-mail address: z.olujic@wbmt.tudelft.nl (Z& . Olujic)
1
Now with SHELL.
0255-2701/99/$ - see front matter 1999 Elsevier Science S.A. All rights reserved.
PII: S0255- 2701( 99) 00068- 9
Z&. Olujic et al. / Chemical Engineering and Processing 38 (1999) 683695 684
to the other side of the sheet, which is however limited
to the neighbouring channel only. The ow channels
are formed between tightly packed corrugation sheets
with oppositely oriented channels enabling lateral ow
parallel to sheet orientation only. As mentioned above,
the necessary (large-scale) mixing effect is achieved by
rotating subsequent element/layer to each other by 90,
which is, however accompanied by a considerable pres-
sure loss due to the abrupt change in gas ow direction
at each transition between packing layers. This compo-
nent, as well as that due to gasgas interaction at the
interface created at the crossings of gas ow channels,
represents a major contribution to total pressure drop
experienced by the gas or vapour at its zigzag way
upward in a packed bed [1].
Although the structured packing is considered to be
an easy to understand and to implement gasliquid
contacting device, there have been a number of major
failures in practice, particularly those in conjunction
with high pressure distillations, indicating a worrying
lack of understanding of the function of this rather
simple gasliquid contacting device. In general, the
CSSP is a lm ow-type device and it works properly as
long as the liquid moves as a continuous thin lm held
by surface tension. However, the surface and corruga-
tion designs of proprietary packings differ considerably
and this has a profound effect on packing performance.
Unfortunately, present models are not fundamental
enough, i.e. do not account properly for the effects
imposed by the geometry of CSSP. This became obvi-
ous at the beginning of our work and we have realised
that a major prerequisite for utilisation of the full
potential of CSSP is a thorough understanding of the
relationship between the macro and micro geometry
imposed hydrodynamics and the mass transfer process.
In a previous general paper [1], an outline was given
introducing an overall and a detailed approach to the
modelling of the hydraulics and separation performance
of the CSSP. The purpose of this paper is to discuss in
more detail the macro geometry related aspects of the
performance of CSSP, and to validate an overall mass
transfer performance prediction method which is
founded on these considerations.
2. Overall mass transfer efciency model
In design/redesign of packed distillation columns, it
is convenient to translate the results of equilibrium
stage calculations into packed height using HETP [2]
which represents a given height of a packed bed, h
pb
,
that can produce the concentration change that would
be produced by an equilibrium stage:
h
pb
=N HETP (1)
where N is the number of equilibrium stages (theoreti-
cal plates). HETP, i.e. the height of packing equivalent
to a theoretical plate, is a common measure of mass
transfer efciency of packings.
In order to establish a practical model for prediction
of HETP we have to employ a more fundamental
quantity, the height of an overall phase transfer unit,
which in distillation applications due to prevailing resis-
tance to mass transfer is related to the gas phase,
HTU
oG
.
For an idealised situation, where the equilibrium and
operating lines are straight, HETP and HTU
oG
are
related through the stripping factor, u, i.e. the ratio of
slopes of equilibrium and operating lines:
HETP=
ln u
u1
n
HTU
oG
(2)
with
u=
m
L
V

(3)
where m=dy/dx is local slope of the equilibrium line
and (L/V) the slope of the operating line.
The relationship between the overall height of the gas
phase transfer unit and individual lm transfer units
HTU
G
and HTU
L
, which is, according to generally
adopted two-lm theory, based on the concentration
driving force across the gas and liquid lms, is given by:
HTU
oG
=HTU
G
+u HTU
L
(4)
Certainly, the slope of the equilibrium line in con-
junction with the slope of the operating line plays an
important role in mass transfer calculations. For binary
and pseudo binary distillations with a constant value of
relative volatility, the slope of the equilibrium line
Fig. 1. Schematic illustration of a corrugated sheet structured packing
bed.
Z&. Olujic et al. / Chemical Engineering and Processing 38 (1999) 683695 685
Fig. 2. Comparison of predicted HETP values, by BRF and BS
methods, with experimental data from SRP (the base case: Montz-
pak B1-250).
The latter one, however requires one packing type
constant for prediction of interfacial area as well as
characteristic friction factors for each packing, and
these are related to mass transfer coefcients via liquid
hold-up, i.e. effective velocities of phases.
A comparison with experimental data for Montz-pak
B1-250 [7] shown in Fig. 2, where measured and pre-
dicted HETP values are plotted against the F-factor,
i.e. the square root of the supercial momentum ux of
the gas (vapour) phase, indicates a deviation of both
methods in trend and a signicant deviation of the
BS-model [5] in absolute values. The latter is repre-
sented by one curve, based on characteristic constants
for 45 packing obtained as an average of original
values given in Ref. [6] for the same packing with
smaller (B1-200) and larger (B1-300) specic surface
area. Since working expressions do not account for the
corrugation angle, this method cannot see the differ-
ence other than through corresponding constants deter-
mined experimentally. Obviously, an easy t can be
obtained by matching the experimental values which
will bring corresponding curves at the level of curves
calculated by the BRF-method. The BRF-method is
represented by two curves for each packing. This is
done to illustrate the difference ( :4%) caused by the
fact that the liquid mixing length damping coefcient
(0.9) in the expression for liquid phase mass transfer
coefcient appears in one reference in the denominator
[3] and in the other in the numerator [4]!
With respect to experimental evidence given in Fig. 2,
striking is the deteriorating effect of the F-factor on the
performance of B1-packing in the preloading region.
The performance improves beyond the loading point,
which is more pronounced with 45 packing, and this
lasts until the point of onset of ooding is reached. As
already mentioned, none of these methods follows the
trend observed with the efciency of B1-250 packing.
Table 1 summarises the differences in predictions of
compared methods, expressed by ratios of major
parameters at a characteristic F-factor.
The difference in estimated effective surface area is
quite large and, as illustrated in Ref. [1], both methods
assume a strong dependence of the effective surface
area on the F-factor, more pronounced in the case of
the BS-method. It should be noted that in this total
reux distillation case an increase in F-factor means a
corresponding increase in both gas and liquid ow
rates. Strikingly, the mass transfer coefcient correla-
needed for calculation of the stripping factor u may be
readily obtained from the slope of the equilibrium curve
(see Appendix A).
Individual phase transfer units of respectively gas
and liquid phase are dened as
HTU
G
=
u
Gs
k
G
a
e
(5a)
and
HTU
L
=
u
Ls
k
L
a
e
(5b)
where a
e
is the effective area, i.e. the surface area of the
gasliquid interface, and k
G
and k
L
are mass transfer
coefcients of gas and liquid phase, respectively.
Certainly, the accuracy of both HETP and rate-
based calculation methods depends directly on the ac-
curacy of correlations used for prediction of basic
design parameters, i.e. mass transfer coefcients of
phases and the effective surface (interfacial) area.
There are two generalised mass transfer predicting
methods, which have gained wider attention in the
literature: Bravo and coworkers [3,4] and Billet and
Schultes [5,6]. These methods are incorporated in most
commercial software owsheeting packages.
The method of Billet and Schultes [5,6], the BS-
method, represents an extension of the random packing
method of the same authors. As such, it requires pack-
ing specic constants for each type and size of struc-
tured packing, and therefore is less suitable than the
method by Bravo and coworkers [3,4], the BRF-
method, developed for corrugated sheet metal packings.
Table 1
Ratios of mass transfer parameters of Billet and Schultes (BS), and Bravo, Rocha and Fair (BRF) models
a
HETP HTU
oG
HTU
L
HTU
G
k
L
a
e
k
G
a
e
k
L
k
G
a
e
0.54 1.42 1.86 1.76 2.30 0.81 BS/BRF 0.63 0.63 0.70
a
Montz-pak B1-250, total reux distillation, cyclohexane/n-heptane, d=0.43 m, h
pb
=3.3 m, p=1.03 bar, F-factor=2 m/s(kg/m
3
)
0.5
.
Z&. Olujic et al. / Chemical Engineering and Processing 38 (1999) 683695 686
Fig. 3. Heights of transfer units as predicted for the base case by the
BRF and BS methods.
3. Delft model equations
3.1. Corrugation geometry-related parameters
A packing element/layer comprises a large number of
identical triangular ow channels created between cor-
rugated plates with opposite orientation of corruga-
tions. As illustrated in Fig. 4, which shows the basic
zigzag ow channel conguration and major corruga-
tion dimensions, two sides of the triangular cross-sec-
tion are occupied by walls. The third facing the
neighbouring sheet is open, creating an interface with
the open side of crossing gas ow channels from the
opposite sheet. In case of a packing with a corrugation
inclination angle of 45, the channels cross each other
at an angle of 90, which is accompanied by a substan-
tial pressure drop. The length of individual channels
depends on the element height, the corrugation angle
and the proximity of the wall. Channels ending at the
wall are shorter and this implies that in a column with
a diameter equal to or smaller than the packing element
height, the gas ow makes relatively more bends than
in the case of larger diameter columns. As clearly
demonstrated elsewhere [10], this effect is enhanced by
an enlarged liquid load in the wall zone. It should be
noted that making sharp bends at each transition be-
tween packing layers (90 in the case of a corrugation
angle of 45) introduces another major source of pres-
sure loss in the gas phase, partly enhanced due to
inevitable entrance effects. Namely, the gas ow chan-
nels are relatively short, and the transition between
packing elements actually represents a ow discontinu-
ity for both phases. Establishing the developed ow
prole requires a considerable length under laminar
ow conditions, which is the case at low gas loads in a
packed bed.
As stated in our model for hydraulics of structured
packing [1], the gas ow can be considered as a contin-
uous zigzag ow through a triangular channel of con-
stant dimensions (Fig. 4). The total length of a typical
zigzag ow channel is equal to the product of the
number of packing elements (layers) in a bed and the
tions of the BS-method produce much higher values
than those of the BRF-method, which largely compen-
sates for smaller effective surface area. The correspond-
ing HTU curves are shown in Fig. 3.
Also, it should be noted that according to experimen-
tal evidence the best mass transfer efciency is achieved
at the low end of the F-factor, which is in contradiction
with the trend in effective surface area curves. This
indicates that the nature of the mass transfer process is
not reproduced adequately. A more recent correlation
for effective surface area [8], developed on the basis of
absorption experiments carried out with Mellapak
250Y, indicates a slightly less pronounced loss of sur-
face area at the low end, but generally it does not differ
considerably from the BRF method. As illustrated in
Ref. [1], our detailed liquid distribution studies de-
scribed thoroughly in Ref. [9] give a much more opti-
mistic picture in this respect, which is certainly the case
with pilot diameter columns and high quality of initial
liquid distribution.
Another factor, which may play an important role in
the absolute sense, is the degree of approach to reality
from the hydraulic point of view. The major parameter
in this respect is certainly the hydraulic diameter of the
gas ow channel. The BS-method relies on a general
denition relating packing porosity and specic surface
area: d
hG
=4m/a, that gives a value close to the corruga-
tion side length, s, which is taken as the hydraulic
diameter in the BRF-method. Both values differ consid-
erably from the hydraulic diameter of a gas ow chan-
nel with a triangular cross-section.
This and other discrepancies with respect to the
actual situation, and the need to account properly for
observed effects of corrugation geometry, have moti-
vated this effort towards establishing a more general
model for mass transfer efciency of structured
packings.
Fig. 4. Basic geometry and dimensions of triangular, gasliquid ow
channel.
Z&. Olujic et al. / Chemical Engineering and Processing 38 (1999) 683695 687
length of a straight gas ow channel in a packing
element. The latter is dened as the ratio of the packing
element height, h
pe
, and the sine of the corrugation
inclination angle,h:
i
G,pe
=
h
pe
sin h
(6)
There are two standard corrugation angles preferred
in practice, 45 or 60 with respect to the horizontal axis.
The hydraulic (equivalent) diameter of a triangular
gas ow channel, specied by corrugation height, h,
corrugation side length, s, and corrugation base width
or pitch, b, with corrugation sides covered by a thin
liquid lm of constant thickness, l, is
d
hG
=
(bh2ls)
2
bh
bh2ls
2h

2
+
bh2ls
b

2
n
0.5
+
bh2ls
2h
(7)
The V-shaped fraction of the cross-section of the
triangular gas ow channel occupied by the liquid lm
is dened as
=
2s
b+2s
(8)
where s and b stand for the lengths of the corrugation
side and ow channel (corrugation) base, respectively.
The corrugation height to base ratio of most conven-
tional packings is around 0.5. Depending on the surface
design, the thickness of sheet metal material ranges
from 0.1 to 0.2 mm. Thanks to this, the porosity of
sheet metal packings is extremely high, between 0.95
and 0.99. One should note that this is certainly not the
case with corrugated sheet packing made of plastic,
ceramics or glass.
Effective gas and liquid velocities respectively are
dened as [3,4]
u
Ge
=
u
Gs
(mh
L
) sin h
(9)
and
u
Le
=
u
Ls
m h
L
sin h
(10)
where m is the void fraction (porosity) of the packing,
and h
L
is the liquid hold-up.
If the liquid is well distributed and there is no
excessive entrainment, which is a reasonable assump-
tion for the purpose of modelling packing hydraulics in
the preloading region, then the liquid hold-up follows
simply from the product of the nominal packing surface
area, a, and the mean liquid lm thickness, l:
h
L
=la (11)
It is known from experimental observations that
liquid hold-up is not affected signicantly by gas ow
in the preloading range [11]. The same has been ob-
served with counter-current annular ow in pipes [12].
This implies that the liquid lm thickness can be deter-
mined from correlations developed for liquid lms in
the absence of counter-current gas ow. The ow of
thin liquid lms over packing surfaces is assumed to be
laminar, and the mean liquid lm thickness can be
estimated using the well known Nuelt formula for
falling lms adapted to the inclined wall situation as
encountered in our case:
l=
3v
L
v
Ls
z
L
g a sin h

1
3
(12)
where v
L
is the liquid viscosity, z
L
is liquid density and
g is gravitational acceleration. Liquid hold-up measure-
ments have conrmed that Eq. (11) in conjunction with
Eq. (12) reproduces the relation between operating
hold-up and liquid load satisfactorily for the purposes
of hydraulic calculations [1].
In general, inuenced by gravity, the liquid will tend
to ow at an angle steeper than the corrugation angle.
Eq. (12) accounts for this, and for a ow along a 45
inclined channel, it will give a lm thickness approxi-
mately 12% larger than that of a falling lm.
3.2. Gas phase mass-transfer coefcient
Within the operating range of a distillation column,
vapour (gas) ow may be characterised as laminar to
slightly turbulent, i.e. it mainly covers the ow regime
transition. Therefore we will assume here that over
typical operating (F-factor) range, the overall gas phase
mass transfer coefcient can be represented as an aver-
age of two individual, laminar and turbulent ow
contributions:
k
G
=k
G,lam
2
+k
G,turb
2
(13)
with
k
G,lam
=
Sh
G,lam
D
G
d
hG
(14a)
and
k
G,turb
=
Sh
G,turb
D
G
d
hG
(14b)
where D
G
and D
L
stand for the diffusion coefcient of
gas and liquid phase, respectively.
Fig. 2 indicates a better mass transfer performance at
low F-factors. This may be considered as a consequence
of the variation in the degree of mass transfer enhance-
ment (associated with prevailing ow regime) due to
entrance effects imposed by abrupt change in ow
direction at each transition between packing elements.
In other words, we assume here that both gas ow and
mass transfer proles are partly undeveloped because of
entrance effects associated with the ow through a
Z&. Olujic et al. / Chemical Engineering and Processing 38 (1999) 683695 688
relatively short gas ow channel as is the case within a
packing element. If we assume that the extent of mass
transfer enhancement will be equal to that observed
with heat transfer in laminar and turbulent regimes
under similar conditions, we can make use of character-
istic dimensionless expressions developed for heat trans-
fer applications. According to VDI-Warmeatlas [13],
the following Sherwood number expressions can be
used for laminar and turbulent ow regime,
respectively:
Sh
G,lam
=0.664Sc
G
1/3

Re
Gvr
d
hG
l
G,pe
(15)
and
Sh
G,turb
=
Re
Grv
Sc
G
x
GL

8
1+12.7
x
GL

8
(Sc
G
2/3
1)

1+
d
hG
l
G,pe

2/3
n
(16)
with gas phase Schmidt number as:
Sc
G
=
v
G
z
G
D
G
(17)
and (d
hG
/l
G,pe
) as the ratio of hydraulic diameter and
the length of gas ow channel within a packing ele-
ment. Re
Grv
represents a gas (vapour) phase Reynolds
number based on relative velocity expressed as
Re
Grv
=
z
G
(u
Ge
+u
Le
)d
hG
v
G
(18)
Since the friction factor curve for a triangular ow
channel differs slightly from that for a circular tube
[14], we will assume here the validity of the well known
Colebrook and White expression, which covers the
whole range of hydraulic roughness. Instead of the
original implicit expression, we will make use of a very
accurate explicit approximation [15]:
x
GL
=

2 log

l
d
hG

3.7

5.02
Re
Grv
log

l
d
hG

3.7
+
14.5
Re
Grv

2
(19)
in which the relative roughness is dened as the ratio of
the absolute roughness of the liquid lm surface, which
is taken to be approximately equal to lm thickness
[16,17], and the hydraulic diameter for gas ow.
The predicted mass transfer coefcients are shown in
Fig. 5 as a function of F-factor and operating pressure,
for the base case, B1-250 packing with corrugation
angle of 45 and 60, respectively. The slope is slightly
lower and the absolute values considerably lower than
that provided by the BS- and BRF methods. In other
words, our model assumes a more pronounced resis-
tance to mass transfer in gas phase. However, it should
Fig. 5. Comparison of gas phase mass transfer coefcient correlations
of the Delft, BRF and BS methods.
be noted that relatively larger values of the mass trans-
fer coefcient produced by the BS- and BRF methods
as well as corresponding slopes correspond with the
slope of their effective surface area curves, which ulti-
mately leads to a smoothening effect in the slope of
HTU
oG
curves.
3.3. Liquid phase mass-transfer coefcient
The mass transfer resistance in liquid phase is consid-
ered as nearly negligible in many distillation applica-
tions. It is a general belief that it is quantied properly
using a penetration theory-based expression. We will
use the same expression as proposed by Bravo et al. [3].
However, instead of the corrugation side, s, our charac-
teristic length of the liquid ow path will be equal to
the hydraulic diameter of the triangular ow channel
dened by Eq. (7):
k
L
=2
D
L
u
Le
y 0.9 d
hG
(20)
It should be noted that recently some attempts have
been made [1820] to develop correlations more suit-
able for structured packings. However, it is doubtful
whether this may be considered as an improvement
regarding the uncertainties related to the real magni-
tude of k
L
values. For instance, Shetty and Cerro [20]
recently derived a theoretically-based correlation which
seems to produce liquid phase mass transfer coefcients
which are greater by a factor of two than the BRF
method.
Fig. 6 shows calculated values of mass transfer coef-
cient of the liquid phase as a function of the F-factor at
different operating pressures. Obviously, the effect of
both operating pressure and the F-factor on k
L
is
signicant, mostly because of correspondingly large
variations in the liquid load. The latter is practically
negligible in the case of the BRF-method. The atmo-
spheric pressure curve of the Delft method lies above
those predicted by the BRF- and BS methods, which
Z&. Olujic et al. / Chemical Engineering and Processing 38 (1999) 683695 689
implies a relatively smaller resistance to mass transfer
on the liquid side, which is in accordance with the trend
suggested in Ref. [20].
3.4. Effecti6e surface area
The size of effective surface area is an ongoing issue
in all elds of application of packed beds, and as
suggested earlier, in BS- and BRF-methods it appears
to be quite unrealistic. Namely, both methods use a
correlation based on considerations related to be-
haviour of random packings, in which case the effective
wetting of the surface appeared to be a strong function
of both liquid and gas loads. Taking corrugated sheet
metal as a typical example, both increased liquid load
and low surface tension will encourage a more thor-
ough wetting of material, but not to such an extent as
with random packings. The most attened curves pub-
lished so far are those by Weimer and Schaber [21]
obtained from absorption of atmospheric CO
2
in caus-
tic solutions. For metal Mellapak 250Y effective sur-
face areas were in the range of 85%, at low liquid loads
(15 m
3
/m
2
h) to 95% of the nominal surface area at high
liquid load (30 m
3
/m
2
h), and more conservative values
were obtained with KOH than NaOH. Also, the in-
creasing gas ow appeared to have some minor interfa-
cial area enhancing effect.
A more profound effect is that of construction mate-
rial. For instance, as shown in Ref. [21], only 5570%
of the surface area of plastic Mellapak 250Y proved to
be covered by water under the same operating condi-
tions. A short but very informative account on wet-
tability of packing construction materials can be found
in a recent book by Stichlmair and Fair [22].
Liquid spreading quality will also be affected to a
great extent by the design of both packing surface
(texture of material and the size and form of apertures)
and corrugation (sharp or rounded ridges and crimp
angle). This is illustrated in Fig. 7 which shows effective
area as a function of liquid load as observed in a study
Fig. 7. Wetting curves of Montz-pak B1-250 and Ralu-pak 250YC as
measured by Stoter [9].
on CO
2
absorption in diluted aqueous solution of
NaOH [9] with B1-250 and Ralu-pak 250YC packings.
With Montz-pak B1-250, a 100% utilisation of available
surface area was obtained at liquid loads above 10
m
3
/m
2
h. This was not case with Ralu-pak 250YC
which at maximum liquid load reached approximately
94%. The observed difference in effective areas, more
pronounced at the lower liquid loads, can be explained
on the basis of liquid spreading behaviour of these two
packings with different surface design. Namely, in the
case of Montz-pak, a packing with a continuous shal-
low embossed surface, the liquid easily reaches corruga-
tion ridge and at each crossing of corrugations liquid is
transferred to neighbouring sheets with corrugations
inclined in the reverse direction. In this way a limited
lateral transport is achieved, resulting in a thorough
spreading and liquid mixing within the height of one
packing element. On the other hand, the Ralu-pak,
with slits on one corrugation side opened to allow
liquid to ow to other side of the sheet, proved to be a
typical channel, i.e. rivulet ow packing, enabling max-
imum lateral transport of liquid, however at the cost of
very limited spreading over the packing surface and no
liquid mixing at all within a packing element. In other
words, the liquid spreading of Montz-pak B1-250 is
narrow but thorough, and that of Ralupak-250YC of
maximum extent but covers a minimum of packing
surface. Consequently the latter is less efcient but also
much less sensitive to any kind of severe initial liquid
maldistribution [23,24].
The loss of effective surface area observed with
Montz-pak B1-250 at liquid loads below 5 m
3
/m
2
h is
attributed to a change in liquid spreading behaviour
caused by the liquid ow rate per ow channel drop-
ping below a minimum at which small rivulets are
formed incapable of reaching the corrugation ridges,
thus forcing the liquid to ow along the channel valley
only. This leads to a signicant reduction in effective
surface area and is less pronounced in the case of
organic liquids than water.
Fig. 6. Comparison of liquid phase mass transfer coefcient corella-
tions of the Delft, BRF and BS methods.
Z&. Olujic et al. / Chemical Engineering and Processing 38 (1999) 683695 690
In order to get to a reasonable approximation of the
situation encountered in total reux experiments used
to validate the mass transfer efciency model, we as-
sumed that all packings tested will exhibit the same
wetting behaviour. The experimental curve for Montz-
pak B1-250 can be reproduced by a simple exponential
expression and the following relation can be used to
describe the effect of liquid load on the size of effective
surface area:
a
e
=a
p
(1V)

1+
A
u
Ls
B

(21)
where d ( ) stands for the void fraction of the packing
surface, i.e. the fraction of the surface area occupied by
holes (0.1 for Montz-pak BSH, and other packings with
a pattern of holes like Flexipac and Mellapak). A and B
are the packing type and size-dependent constants (A=
0.000002143, B=1.5, for Montzpak B1-250).
Fig. 8 shows the calculated effective surface area of
Montz-pak B1-250 as a function of F-factor for three
different operating pressures. For comparison, curves
calculated by BRF- and BS-methods for the atmo-
spheric pressure case are also included. To get an
impression of the variation in liquid load over the
whole range of operating pressures and F-factors, the
corresponding liquid loads for the cyclohexane/n-hep-
tane system are summarised in Table 2.
It should be noted that with the cyclohexane/n-hep-
tane system even at the lowest pressure, the liquid loads
are high enough to ensure a nearly complete active
wetting of Montz-pak B1-250 at a F-factor above 1.5
m/s(kg/m
3
)
0.5
, i.e. in the range of concern of packed
column designers. Initial liquid distribution has been
provided by a high performance narrow trough distrib-
utor with an equivalent of 145 drip points per square
meter, to satisfy with one distributor the required qual-
ity of initial distribution density for both 250 and 400
m
2
/m
3
packings.
Table 2
Liquid loads as a function of F-factor as encountered at three
operating pressures employed in tests carried out with the cyclohex-
ane/n-heptane system in a 0.43-m i.d. column of the Separation
Research Program at the University of Austin, TX
u
Ls
(m
3
/m
2
h) F-factor (m/s(kg/m
3
)
0.5
)
p=0.33 bar p=4.14 bar p=1.03 bar
1.49 5.82 2.71 0.25
2.99 5.41 0.50 11.63
5.98 10.82 1.00 23.26
1.5 34.89 8.97 16.23
2.00 21.64 11.96 46.52
2.50 27.06 14.94 58.15
32.45 69.78 17.93 3.00
4. Results and discussion
For validation of the Delft model, total reux distil-
lation data obtained with various types and sizes of J.
Montz packings were used, available in the data base
formed jointly by our laboratory and Separation Re-
search Program at the University of Texas at Austin [7].
Table 3 contains geometric features of packings consid-
ered in this study. The process conditions and corre-
sponding physical properties are given in Table 4.
Fig. 9 shows the effect of operating pressure on the
efciency of B1-250 packing. Obviously, there is an
improvement in efciency with increasing pressure,
however, as expected, at the cost of a considerable
reduction in the capacity. As mentioned already, the
measured curves indicate a performance deteriorating
trend with increasing F-factor, which is particularly
strong around the loading point in the case of highest
pressure. This sharp deterioration in efciency around
the loading point reminds us of the well known ef-
ciency hump effect [25,26], which appeared to be a
major hindrance for successful application of structured
packing in high pressure distillations. It should be
noted that, according to Table 4, the liquid load at this
point is in the range of liquid loads encountered in such
applications. Our knowledge of this phenomenon is still
insufcient, both backmixing in gas [27] and liquid [25]
phases may be considered to be a major problem in this
case.
Calculated curves reproduce the observed trend very
well and can be considered as reasonably accurate. The
method is designed actually to model the mass transfer
in the preloading region, and in the loading region it is
mainly on the safe side, from a designers point of view.
The improvement in performance experienced above
the loading point is the consequence of an abrupt
change in the hydraulic regime, with increased liquid
hold-up at transitions between packing layers and gas
phase breaking through liquid curtains. This is accom-
panied by creation of a signicant fraction of liquid
Fig. 8. Interfacial areas at various pressures, according to the Delft
model, including atmospheric pressure curves predicted by the BRF
and BS methods.
Z&. Olujic et al. / Chemical Engineering and Processing 38 (1999) 683695 691
Table 3
Geometrical characteristics of J. Montz packings employed in this study
m () Packing h
pe
(m) a
p
(m
2
/m
3
) h (m) b (m) s (m)
0.98 0.197 B1-250 0.0120 244 0.0225 0.0165
B1-250.60 0.98 245 0.211 0.0120 0.0223 0.0164
0.96 0.197 0.0074 394 0.0140 B1-400 0.0103
0.96 0.215 0.0074 B1-400.60 0.0143 390 0.0103
0.97 0.194 0.0074 378 0.0151 BSH-400 0.0106
0.97 0.215 0.0074 0.0148 BSH-400.60 0.0105 382
droplets, thus increased effective area. Since this condi-
tion of intensive interaction between phases is limited to
the region close to the transition between packing ele-
ments, the net result is an improvement in mass transfer
efciency, which however diminishes with gas load
approaching the point of onset of ooding. Beyond this
point, efciency decreases sharply and is generally more
than halved when the point of hydraulic ooding is
reached (usually around a pressure drop of 10 mbar/m).
It should be noted that packings included in this study
exhibited the behaviour mentioned above, which, as
suggested in Fig. 2, appeared to be less pronounced in
the case of a 60 corrugation angle.
In general, the situation in the loading region is so
complex that there is a question whether it can be
modelled properly in a sophisticated way. An addi-
tional difculty is that packing surface and corrugation
design may have a signicant inuence on the nature of
the gasliquid interaction and consequently the ef-
ciency curve pattern in the loading region. With respect
to potential for capacity increase, this has been demon-
strated clearly in a most recent paper by Billingham
and Lockett [28].
Regarding the model predictions, the fact that the
calculated curve for atmospheric pressure lies above
that for vacuum conditions is striking. As shown in
Figs. 5 and 6, the curves of corresponding k
G
and k
L
values follow the pressure, with the largest values of k
G
in a vacuum and k
L
at highest pressure, respectively.
According to Fig. 8, effective surface areas differ
slightly at an F-factor larger than 1, and below that
value the difference is not so large as to have signicant
consequence for the efciency curve. Corresponding
heights of transfer units are shown in Fig. 10. Obvi-
ously there is nothing strange in the pattern of HTU
L
curves, as well as with HTU
G
and HTU
oG
curves at
highest pressure. Interestingly, HTU
G
and consequently
HTU
oG
curves of vacuum and atmospheric pressure
cross each other, and below an F-factor of approxi-
mately 1.5 the atmospheric pressure curves are above
the vacuum curves. This indicates that a somewhat
larger reduction in the effective surface area should be
anticipated for vacuum than suggested in Fig. 8. With
such a correction, the vacuum HETP curve would
follow the atmospheric curve, however at a slightly
larger value. However, the difference in absolute values
is smaller than the difference in corresponding stripping
factors which means that something else is needed to
get corresponding HETP curves to lie in proper order.
For the time being, it is not necessary to insist on this,
because this anomaly is not affecting the overall accu-
racy of our model profoundly.
Fig. 11 gives an impression of the accuracy of the
Delft method in the case of B1-250 packing, also in-
cluding the effect of the corrugation angle. Based on
this a parallel can be drawn with performance of estab-
lished BS- and BRF-methods (Fig. 2). As already
shown in Fig. 9, the predicted curve of 45 packing lies
above the measured one, roughly 10% higher, which
can be considered as a not too conservative value. On
the other hand, the 60 packing curve matches the
measured curve nearly perfectly and the larger value in
the loading range is on the safe side. Obviously, the
Table 4
Physical properties of the cyclohexane/n-heptane system, as encoun-
tered at bottom conditions in a total reux distillation column of
Separations Research Program at the University of Texas at Austin
Pressure: Physical property
0.33 bar 4.14 bar 1.03 bar
Temperature (C) 61 97 154
561 657 Liquid density 625
(kg/m
3
)
1.6110
4
2.9710
4
4.3110
4
Liquid viscosity
(Pa s)
9.1710
9
Liquid diffusivity 2.7210
9
4.4410
9
(m
2
/s)
13.14 Vapour density 1.19 3.53
(kg/m
3
)
6.9410
6
7.7810
6
9.1710
6
Vapour viscosity
(Pa s)
1.3910
6
Vapour diffusivity 11.410
6
4.1710
6
(m
2
/s)
0.008 0.017 0.014 Surface tension
(N/m)
1.86 1.64 1.42 Relative volatility
()
1.5 1.35 1.21 Stripping factor
()
0.51 0.53 0.50 Schmidt number
()
Z&. Olujic et al. / Chemical Engineering and Processing 38 (1999) 683695 692
Fig. 9. Comparison of calculated and measured HETP values, with
illustration of the operating pressure effect.
Fig. 11. Comparison of predicted HETP values for B1-250 with
experimental data from SRP, with illustration of the corrugation
angle effect.
trends in the preloading range are matched perfectly,
and this conrms the validity of the term accounting
for enhancement in mass transfer due to entrance (short
channel) effects, which plays a signicant role at low
gas loads.
Certainly, the difference in predicted HETP values is
signicantly smaller than the measured one. The same
can be concluded from Figs. 12 and 13 which show the
HETP curves obtained with larger surface area, B1-
and BSH-type packings. In both cases there is no good
agreement of our model with experimental data, partic-
ularly in the case of 60 packings. Namely, in the
preloading range, the measured values differ between
35 and 40% and the calculated ones by only 15%. The
reduction in the effective gas velocity and consequently
the pressure drop due to gasliquid interaction is
around 18%, which means that other factors may play
a role here, most probably a signicant difference in the
extent of the loss of effective surface area. It should be
noted that effective surface area has been calculated
using the same model as for B1-250 packing. This looks
to be a rather optimistic picture of wetting behaviour of
large surface area packings. Namely, by reducing the
effective surface area of 45, B1-400 packing by 20%
and that of 60 packing by 25% would lift the calcu-
lated HETP curves to the level of measured ones. A
similar effect was obtained with BSH packings. This
degree of reduction in effective area can be considered
as a realistic one, and a simple point source study on
spreading behaviour, as described in Ref. [9], could help
to clarify the uncertainties about the wetting behaviour
of packings with larger surface area.
Another consequence of a change in the corrugation
angle could be the effective angle of liquid ow. At
higher liquid loads it will certainly be larger than the
corrugation angle, which is taken as the basis angle in
the model calculations. This is a model parameter
which can be manipulated accordingly. Fig. 12 shows
the effect of increasing the liquid ow angle by 15 with
respect to the corrugation angle. As indicated by the
dotted line, this resulted in a roughly 5% improvement
with respect to the original curve.
A distinctive characteristic of BSH packing is that it
contains a regular pattern of holes covering roughly
10% of the surface area. This loss of surface area is
accounted for in Eq. (21) by the corresponding void
fraction, d=0.1. However, it should be noted that
there is a much more open surface area in this case,
because BSH is an expanded (stretched) metal packing,
with the open area nearly equal to the area occupied by
metal. Our liquid spreading experiments have shown
that organic liquids penetrate to some extent through
Fig. 12. Comparison of predicted HETP values for B1-400 with
experimental data from SRP, with illustration of corrugation angle
effect. Fig. 10. Effect of operating pressure on predicted HTU values.
Z&. Olujic et al. / Chemical Engineering and Processing 38 (1999) 683695 693
Fig. 13. Comparison of predicted HETP values for BSH-400 with
experimental data from SRP, with illustration of corrugation angle
effect.
between corrugation geometry imposed uid-dynamics
and mass transfer performance of structured packings.
5. Concluding remarks
A corrugated sheet geometry based-approach to
modelling the mass transfer efciency of structured
packings has been introduced and validated. With the
corresponding hydraulic diameter of the triangular gas
ow channel, the expressions for mass transfer coef-
cient produce the same trends but respectively lower
values of gas phase and higher values of liquid phase
coefcients than known methods. The mass transfer
coefcient expression also accounts for entrance effects
and gasliquid friction at the interface, and does not
include any packing-specic constant. Predictions are
generally in good agreement with experimentally ob-
tained values and trends. All geometry-related parame-
ters are adjustable and may be manipulated accordingly
to analyse the performance and evaluate potential for
further improvements.
With a more appropriate wetting function, the pro-
posed method will ensure accurate predictions of per-
formance of corrugated sheet structured packings with
large specic surface area.
Obviously, a reliable prediction of the effective sur-
face area under given conditions is the key to the
success of a generalised packing performance prediction
method. This has been recognised as the weakest point
in the approach and a promising solution is expected to
be the generation of characteristic wetting functions
from predictions of a detailed liquid distribution model
[9,29]. This approach will be evaluated and compared
with that proposed by Shetty and Cerro [20] and the
Hanley et al. [30]. The latter is a percolation model
accounting also for phenomena occurring in the load-
ing region.
Another obvious area for improvement is adoption
of a more sophisticated approach to the description of
the liquid phase mass transfer coefcient, and the cor-
rugation geometry-based approach by Shetty and Cerro
[20] looks to be very promising in this respect.
6. Nomenclature
a
p
packing surface area, m
2
/m
3
a
e
effective surface area, m
2
/m
3
b corrugation base length, m
D
G
gas phase diffusion coefcient, m
2
/s
D
L
liquid phase diffusion coefcient, m
2
/
s
column diameter, m d
c
d
hG
hydraulic diameter for the gas phase,
m
g gravity acceleration, m/s
2
these rather small apertures in the surface [29], and that
the size of the openings in the surface should be kept
small enough to avoid a quite detrimental effect on
liquid spreading.
The thin dotted lines in Fig. 13 indicate the effect of
an increase in the open surface area. An additional 5%
of open surface will result in a roughly 11% increase in
the corresponding HETP values of the 45 packing and
this would lead to a near-perfect agreement with the
measured curve. In case of the 60 packing, a more
pronounced reduction in effective surface area is needed
( \25%) to get close enough to the measured curve.
Also the trend of measured BSH curves is repro-
duced in a fairly good manner. There is a less pro-
nounced mass transfer enhancement effect in this case,
which is in accordance with expectations based on the
fact that the length-to-diameter ratio in this case is
much larger than in the case of 250 m
2
/m
3
packings.
Strikingly, Fig. 12 indicates that the performance of
B1-400 packing is practically insensitive to the F-factor
up to the loading point, and therefore the prediction of
our method, on the lower end of F-factor, is too
optimistic. This region however is of less practical
importance. On the other hand, Fig. 12 indicates the
existence of a small efciency hump in the case of 45
packing, which is rather unexpected regarding the liq-
uid load at this point. Also, the improvement in perfor-
mance of BSH packing behind the loading point is
quite large, particularly in the case of 60 packing. This
may be considered as a consequence of the open surface
area and its interaction with the liquid ow. This may
also be the reason for the somewhat larger capacity of
BSH packing.
This and other not easily explainable manifestations
of packing performance experienced by carrying out
total reux distillation experiments in a 0.43-m i.d.
column demonstrate that our general knowledge in this
respect is still insufcient. Certainly further, more de-
tailed experimental and modelling efforts are needed to
improve our understanding of the nature of interaction
Z&. Olujic et al. / Chemical Engineering and Processing 38 (1999) 683695 694
corrugation height, m h
liquid hold-up, h
L
height of the packed bed, m h
pb
height of the packing element, m h
pe
k
G
gas phase mass transfer coefcient,
m/s
k
L
liquid phase mass transfer coefcient,
m/s
L liquid phase molar ow rate, kmol/s
length of the gas ow channel in a l
G,pe
packing element, m
slope of the equilibrium line, m
Re
Ge
effective gas phase Reynolds
number,
Re
Grv
relative velocity Reynolds number,
Sc Schmidt number,
Sherwood number, Sh
corrugation side length, m s
effective gas velocity, m/s u
Ge
supercial gas velocity, m/s u
Gs
effective liquid velocity, m/s u
Le
u
Ls
supercial liquid velocity, m/s
V vapour (gas) phase molar ow rate,
kmol/s
mole fraction of the light key com- x
lk
ponent in liquid,
mole fraction of the light key com- y
lk
ponent in vapour,
Greek letters
h corrugation inclination angle,
relative volatility of the light key h
lk
component,
l liquid lm thickness, m
m packing porosity,
fraction of the triangular ow chan-
nel occupied by liquid,
u stripping factor (ratio of slopes of
equilibrium and operating lines),
v
G
viscosity of gas, Pa s
viscosity of liquid, Pa s v
L
z
G
density of gas, kg/m
3
density of liquid, kg/m
3
z
L
gasliquid friction factor, x
GL
d packing surface void fraction,
Indices
gas phase G
liquid phase L
laminar ow lam
turbulent ow turb
Acknowledgements
The authors would like to thank J. Montz GmbH for
permission to use and publish experimental results. P.
Le Haen, M. Behrens and P. Saraber, graduate students
at TU Delft have critically evaluated the method and
indicated the potential for further improvements.
Appendix A. The slope of equilibrium line
For a binary/pseudo-binary system, the yx equi-
librium curve is dened as
y
lk
=
h
lk
x
lk
1+(h
lk
1)x
lk
(A1)
where h
lk
is the average value of relative volatility of
light component relative to heavy component, and y
lk
and x
lk
are mole fractions of light component in the gas
phase and the liquid phase, respectively. In the case of
multicomponent mixtures the indices lk and hk are
related to light and heavy key components, respectively.
The slope of the yx equilibrium curve is
m=
dy
lk
dx
hk
=
h
lk
[1+(h
lk
1)x
lk
]
2
(A2)
It should be noted that m may vary considerably
along the column, depending on the curvature of the
equilibrium line in rectication and stripping sections.
Certainly, most pronounced variations are expected in
the total reux distillation where the stripping factor is
identical to the slope of the equilibrium line. Also, the
curvature of the equilibrium curve itself is considerable
in the case of test systems with relative volatility above
1.5. To avoid errors, i.e. to account properly for the
change in relative volatility with composition, the given
range of measured compositions should be divided into
the appropriate number of segments.
References
[1] Z. Olujic, Development of a complete simulation model for
predicting the hydraulic and separation performance of distilla-
tion columns equipped with structured packings, Chem.
Biochem. Eng. Q. 11 (1997) 3146.
[2] J.D. Seader, E.J. Henley, Separation Process Principles, Wiley,
New York, 1998.
[3] J.L. Bravo, J.A. Rocha, J.R. Fair, A comprehensive model for
the performance of columns containing structured packings,
ICHEME Symp. Ser. 128 (1992) 489507.
[4] J.A. Rocha, J.L. Bravo, J.R. Fair, Distillation columns contain-
ing structured packings: A comprehensive model for their perfor-
mance. 2. Mass-transfer model, Ind. Eng. Chem. Res. 35 (1996)
16601667.
[5] R. Billett, M. Schultes, Predicting mass transfer in packed
columns, in: Packed Column Design and Analysis, Ruhr Univer-
sity Bochum, Bochum, 1989.
[6] R. Billet, Packed Towers in Processing and Environmental Tech-
nology, VCH, Weinheim, 1995.
[7] A.F. Seibert, J.R. Fair, Z. Olujic, Inuence of corrugation
geometry on the performance of structured packings: An experi-
Z&. Olujic et al. / Chemical Engineering and Processing 38 (1999) 683695 695
mental study, Proceedings of Distillation Topical Conference,
AIChE Spring National Meeting, Houston, TX, March 1418,
1999.
[8] E. Brunazzi, G. Nardini, A. Paglianti, L. Petarca, Interfacial
area of Mellapak packing: Absorption of 1.1.1-trichlorethane by
Genosorb 300, Chem. Eng. Technol. 18 (1995) 248255.
[9] C.F. Stoter, Modelling of Maldistribution in Structured Pack-
ings: From detail to column design, D.Sc. Thesis, Delft Univer-
sity of Technology, 1993.
[10] Z. Olujic, Effect of column diameter on pressure drop of a
corrugated sheet structured packing, Trans. IChemE 77 (Part A)
(1999).
[11] H.-J. Verschoff, Z. Olujic, J.R. Fair, A general correlation for
predicting the loading point of corrugated sheet structured pack-
ings, Preprints of the AIChE Topical Conference on Separation
Science and Technology, Los Angeles, CA, November 1621,
1997, Vol. I, 7580.
[12] H. Imura, H. Kusuda, S. Funatsu, Flooding velocity in a
count-current annular two-phase ow, Chem. Eng. Sci. 32 (1977)
7987.
[13] VDI-Warmeatlas, Verlag des Vereins Deutscher Ingenieure, Du s-
seldorf, 1974.
[14] H. Richter, Rohrhydraulik, Springer, Berlin, 1958.
[15] Z. Olujic, Compute friction factors fast for ow in pipes, Chem.
Eng. 88 (25) (1981) 9193.
[16] W.H. Henstock, T.J. Hanratty, The interfacial drag and the
height of the wall layer in annular ows, AIChE J. 22 (1976) 990.
[17] S.C. Yao, N.D. Sylvester, A mechanistic model for two-phase
annular-mist ow in vertical pipes, AIChE J. 33 (1987) 1008
1012.
[18] M. Henriques de Brito, U. von Stoockar, P. Bomio, Predicting
the liquid phase mass transfer coefcient k
L
for the Sulzer
structured packing Mellapak, ICHEME Symp. Ser. 128 (1992)
B137B143.
[19] E. Brunazzi, A. Paglianty, Liquid-lm mass transfer coefcient
in a column equipped with structured packing, Ind. Eng. Chem.
Res. 36 (1997) 37923799.
[20] S. Shetty, R. Cerro, Fundamental liquid ow correlations for the
computation of design parameters for ordered packings, Ind.
Eng. Chem. Res. 36 (1997) 771783.
[21] T. Weimer, K. Schaber, Absorption of CO
2
from the atmosphere
as a method for the estimation of effective interfacial areas in
packed columns, ICHEME Symp. Ser. 142 (1997) 417427.
[22] Stichlmair, J.R. Fair, Distillation Principles and Practice, Wiley,
New York, 1998.
[23] Z. Olujic, F. Stoter, J. de Graauw, Performance evaluation of
structured packingsLarge versus small scale, Preprints of the
AIChE Separations Division Topical Conference on Separations
Technologies, Miami, Nov. 26, 1992, pp. 201207.
[24] R. Potthoff, Maldistribution in Fu llko rperkolonnen, Fortschrit-
tberichte VDI, Nr. 294, VDI-Verlag, Du sseldorf, 1992.
[25] F.J. Zuiderweg, Z. Olujic, J.G. Kunesh, Liquid backmixing in
structured packing in high pressure distillation, ICHEME Symp.
Ser. 142 (1997) 865872.
[26] C.W. Fitz, J.G. Kunesh, A. Shariat, Performance of structured
packing in a commercial-scale column at pressures of 0.0227.6
bar, Ind. Eng. Chem. Res. 38 (1999) 512518.
[27] J.L. Nooijen, K.A. Kusters, J.J.B. Pek, The performance of
packing in high pressure distillation applications, ICHEME
Symp. Ser. 142 (2) (1997) 885897.
[28] J.F. Billingham, M.J. Lockett, Development of a new generation
of structured packing for distillation, Presented at the 1998
AIChE Annual Meeting, Miami, FL, Nov. 1620, 1998.
[29] Z. Olujic, A. Roelofse, F. Stoter, J. de Graauw, LDESP: A
simulation and optimisation environment for structured pack-
ings, ICHEME Symp. Ser. 142 (1997) 949960.
[30] B. Hanley, B. Dunbobbin, D. Bennett, A unied model for
countercurrent vapor/liquid packed columns. 2. Equations for
the mass-transfer coefcient, mass transfer area, the HETP, and
the dynamic liquid holdup, Ind. Eng. Chem Res. 33 (1994)
12221230.
.

S-ar putea să vă placă și