Sunteți pe pagina 1din 9

1632 Ind. Eng. Chem. Res.

2002, 41, 1632-1640

Distillation Sieve Trays without Downcomers: Prediction of


Performance Characteristics
J. Antonio Garcia† and James R. Fair*
Department of Chemical Engineering, The University of Texas at Austin, Austin, Texas 78712

A large amount of performance data on larger-scale trays without downcomers has become
available recently, and the data have been examined with a view toward understanding and
modeling the contacting mechanisms under distillation conditions. The objective of this paper
is to describe new predictive models for efficiency, pressure drop, and flooding of trays without
downcomers and to show how they can be applied to distillation separations. Such an effort has
not been reported previously. The contacting devices are assigned the generic name of dual-
flow trays. The models are shown to give reasonable estimates of the performance of these trays
in larger columns with open areas in the range of 15-25%, hole diameters in the range of 12-
25 mm, and tray spacings in the range of 0.3-0.6 m.

Bubble trays for distillation columns normally involve The published reports cover tests in distillation columns
liquid flowing between trays through connecting down- with diameters as large as 2.5 m (8.2 ft). Test data for
comers. Such contacting devices are of the cross-flow a 1.0-m (3.3-ft) column containing turbogrid trays have
type, and vapor flows only through dispersers on the been presented by researchers at the Institute of Process
trays such as holes, valves, or bubble caps. Less well- Fundamentals in Prague.3-6 The FRI work was con-
known and -specified are perforated trays without ducted in a 1.2-m (4.0-ft) research column, using several
downcomers, wherein the liquid and vapor flow coun- hydrocarbon systems plus water/alcohol and water/
tercurrently through the same tray openings. These are steam. Both Shell and FRI confirmed that the slotted
often called dual-flow trays, but in specialty forms, they openings of the turbogrids had equivalent performance
have other names such as turbogrid trays and ripple characteristics to the round holes of dual-flow trays if
trays. The devices are used for special services, espe- slots with equivalent diameters were used. The results
cially when openings of a cross-flow tray might foul. of the FRI tests have been released to the public and
Their less general use appears to have derived from an are now available from Oklahoma State University.7
expected narrow operating range of high efficiency, as Very little has been done to model the mass transfer
well as a general unavailability of design models that performance of dual-flow trays. In the 1950s and 1960s,
can enable reliable prediction of their performance. work in Europe with very small columns led to attempts
In the mid- to late-1950s Fractionation Research, Inc. at dealing with the hydraulics of the trays, but serious
(FRI), undertook an extensive research program on scale-up problems were anticipated, and no further work
dual-flow trays, obtaining performance data for several transpired that would aid the commercial-scale de-
different systems but not providing a general, funda- signer. The most recently reported attempt to model
mental correlation of the data for pressure drop, ef- dual-flow tray efficiency, by Xu et al.,8 appeared in 1994.
ficiency, and flooding. Meanwhile, Shell completed an They studied methanol/water and methanol/2-propanol
extensive program of testing its turbogrid trays, similar distillations in a 0.3-m (1.0- ft) column and provided
in characteristics to dual-flow trays. Recently, FRI valuable insights into the mass transfer relationships
released all of its dual-flow test results, and these data involved. However, they made no attempt to combine
form a major basis for the modeling in the present their data with those of others to arrive at a more
paper. Shell has released none of its turbogrid test generalized treatment of dual-flow tray efficiency.
results. Perhaps a key finding, although nonquantitative, has
Dual-flow trays have been applied to many situations been that the insertion of dual-flow trays as replace-
in which a broad operating range (high turndown ratio) ments for cross-flow trays dramatically alleviates prob-
is not essential. In its range of application, the tray lems with severe fouling. According to Baird,9 his
provides a very high mass transfer efficiency with low company has “successfully supplied our DualFlo trays
capital investment. Importantly, the application of such to over 70 installations where solids or scaling repre-
devices to fouling systems has been eminently success- sented a potential operating problem. These trays,
ful, the alternating vapor-liquid passage through the which range from 6 in. to 7 ft (0.15 to 2.1 m) in diameter,
holes providing a self-cleaning action. stay cleaner longer than other designs and are easier
to remove and clean”. Such comments have encouraged
Previous Work designers to use the devices, termed “self-cleaning
trays”, without adequate predictive model support.
Very few of the results of Shell’s investigations of
turbogrid trays have appeared in the open literature.1,2 Modeling Approach

* To whom correspondence should be addressed. E-mail: Figure 1 shows a schematic of a dual-flow tray. Vapor
fair@che.utexas.edu. and liquid flow countercurrently, using the same open-
† Current address: DuPont Engineering Technology, Bran- ings, and a volume of aerated liquid called the froth zone
dywine 8226, 1007 Market Street, Wilmington, DE 19898. is generated above the tray perforations. It is in this
10.1021/ie010326w CCC: $22.00 © 2002 American Chemical Society
Published on Web 02/08/2002
Ind. Eng. Chem. Res., Vol. 41, No. 6, 2002 1633

Figure 1. Schematic of a dual-flow tray showing mass transfer


zones.

Figure 3. Efficiency of stamped turbogrid trays, benzene/toluene,


atmospheric pressure. Open area ) 14%, tray spacing ) 500 mm,
hole diameter ) 12 mm, total reflux. Data from ref 2.

Figure 2. Efficiency of stamped turbogrid trays, methanol/water,


1.0-m column, atmospheric pressure, total reflux. Data from ref
5.

zone that the majority of the mass transfer is presumed Figure 4. Efficiency and pressure drop of dual-flow trays,
to occur. The space above the froth, called the spray cyclohexane/n-heptane, tray spacing ) 0.61 m, hole diameter )
zone, is available for additional mass transfer, and this 12.7 mm, column diameter ) 1.2 m, pressure ) 1.63 atm, total
zone differs in performance from the equivalent zone of reflux. FRI data.
a cross-flow tray.
Observations of operating dual-flow trays reveal a range of 75-90% of the maximum throughput, i.e.,
dynamic contacting process in which a given opening where the efficiency becomes very low and the pressure
alternately passes vapor and liquid. This accounts for drop becomes very high. Another important objective
the self-cleaning character of the device. At any instant, of the present effort is to develop a generalized relation-
a certain fraction of the holes are actively passing vapor, ship for predicting this maximum (“flooding”) condition.
either as jets or as bubbles. Efficiency profiles (Figure Test conditions in the database (Table 1) include the
2) show a strong modification of the hole dynamics to conditions for peak efficiency. The database covers a
the extent that there is less self-cleaning potential at broad range of physical properties.
lower loadings and clearly less froth volume to accom- Flooding. The FRI test data (see Table 1) include
modate mass transfer needs. At very high loadings, measurements of the near-flood condition. As defined
liquid entrainment occurs, thereby reducing the mass by FRI, the flood point is evidenced by a significant
transfer efficiency. Figures 2 and 3 (refs 2 and 5) show increase in liquid holdup (pressure drop). The “true
the sharp efficiency profiles characteristic of dual-flow flood” with essentially no separation is at a loading
trays. An effect of column diameter is indicated in slightly higher than the reported flood point. The flood
Figure 3. Figure 4 shows profiles of both efficiency and data have been correlated in the same semiempirical
pressure drop for representative FRI tests.3 For all of manner used successfully for cross-flow type trays,10 as
the figures, it is apparent that a “peak efficiency” exists, shown in Figure 5. The plotting method provides a very
and in the present paper, an attempt is made to model good fit of the data, and this leads to some insights
this efficiency. Depending on the propensity for liquid regarding the mechanics of contacting. The capacity
entrainment upward, the peak appears to be in the factor, Csb, has been normalized to a tray spacing of 0.61
1634 Ind. Eng. Chem. Res., Vol. 41, No. 6, 2002

Table 1. Database, Tests on Dual-Flow Trays


hole Fs peak
pressure TS % diam flood efficiency
refa system (kPa) (m) open (mm) Ep/Fs [m/s (kg/m3)0.5] (% flood)
FRI C6/C7 28 0.61 13.5 12.7 79/1.98 2.20 90
FRI C6/C7 28 0.61 18.9 25.4 75/2.26 2.59 87
FRI C6/C7 28 0.61 25.6 25.4 63/2.56 2.93 88
FRI C6/C7 166 0.61 13.5 12.7 83/1.77 2.01 88
FRI C6/C7 166 0.30 19.1 12.7 55/1.89 1.98 96
FRI C6/C7 166 0.91 19.1 12.7 57/3.11 3.39 92
FRI C6/C7 166 0.61 18.9 25.4 81/2.20 2.34 94
FRI C6/C7 166 0.61 25.6 25.4 55/2.60 2.89 90
FRI C6/C7 345 0.61 18.9 25.4 98/2.01 2.34 86
FRI C6/C7 345 0.91 29.3 11.9 67/2.81 3.17 88
FRI C4 1138 0.61 13.5 0.5 121/0.83 1.43 58
FRI C4 1138 0.30 19.1 1.0 90/1.29 1.46 88
FRI C4 1138 0.61 18.9 1.0 115/1.13 1.77 64
FRI C4 1138 0.61 25.6 1.0 92/1.62 1.73 94
FRI C4 1138 0.91 29.3 0.47 96/1.89 2.07 91
FRI C4 2073 0.61 13.5 0.50 114/0.67 0.76 89
FRI C4 2073 0.61 29 0.47 91/0.91 1.01 90
FRI C4 2756 0.61 13.5 0.50 150/0.43 0.52 81
FRI C4 2756 0.61 29 0.47 87/0.79 0.88 90
FRI C4 3456 0.61 13.5 0.50 >180/0.24 0.33 74
FRI IPA/H2O 101 0.61 19.1 0.5 75/2.34 2.46 95
FRI xylenes 16 mm 0.61 12.9 12.7 81/2.01 3.07 65
FRI xylenes 16 mm 0.61 17.8 12.7 63/2.01 ∼3.3 61
FRI C8/C10 10 mm 0.61 12.9 12.7 72/2.81 (3.78) (74)
FRI C8/C10 10 mm 0.61 17.8 12.7 42/2.56 ∼3.2 (81)
5 MeOH/H2O 101 0.41 10.5 b 95/1.27 1.87 68
FRI MeOH/H2O 101 0.41 14.2 b 82/1.92 2.66 72
FRI MeOH/H2O 101 0.41 18.2 b 82/1.98 2.78 71
FRI MeOH/H2O 101 0.41 23.6 b ∼53/∼2.3 NA NA
2 Bz/Tol 101 0.51 14 12.7 87/1.63 2.44 67
FRI Bz/Tol 101 0.51 14 12.7 80/1.96 2.62 75
a Column diameters: FRI, 1.2 m; ref 5, 1.0 m; ref 2, 0.45 and 2.50 m. b Slots, 4 mm × 150 mm.

that, at lower loadings, there is a distinct loss of


efficiency, much like the case for cross-flow sieve trays
operating under conditions of weeping and dumping or
the case for packed columns with poor liquid distribu-
tions. Observations indicate that, in this region, a
definite loss of froth height occurs, fewer holes have
intermittent vapor and liquid flow, and the residence
time of vapor contacting liquid is shorter. The loss of
efficiency for weeping/dumping of cross-flow trays has
been interpreted in an approximate way by adapting
the Colburn11 relationship for liquid entrainment. The
same approach will be used for dual-flow trays, as
discussed later.
Liquid Entrainment in Vapor. The experimental
curves of Figure 2 are representative of the shapes of
the efficiency-velocity relationships found for all tests
of dual-flow trays. To use these curves as an example,
the efficiency at flow rates higher than the peak value
are subject to the same type of entrainment analysis
as used for cross-flow trays.10,12 The peaks occur at
Figure 5. Flooding capacity of dual-flow trays based on FRI data, about 1.10 and 1.85 m/s superficial velocity for the
1.2-m column diameter.
10.5% open and 18.2% open trays, respectively. At
higher loadings, the effect of entrainment on efficiency
m (24 in.). Thus, the 0.30-m spacing has about 81%, and can be approximated by a modification10 of the Colburn
the 0.91-m spacing about 116%, of the capacity of the relationship
0.61-m spacing.
The effect of open area on flooding is evident but not 1
well understood. Apparently, the higher hole velocity Ew/Ep ) (1)
1 + EpΨ/(1 - Ψ)
for lower open area results in jetting, which increases
liquid entrainment. A small effect of hole size on
flooding is indicated, with 25-mm holes having about or
6% less capacity than the more standard 12.5-mm holes.
1-φ
This difference might be more apparent than real, but Ψ) (2)
it could be a result of the longer jet length for the larger φEp + (1 - φ)
holes.
Efficiency at Lower Loadings. The profiles in where φ ) Ew/Ep, Ew is the “wet” efficiency, Ep is the
Figures 2-4 are typical of dual-flow trays, and it is clear “dry” efficiency at the peak, and Ψ is the ratio of the
Ind. Eng. Chem. Res., Vol. 41, No. 6, 2002 1635

Figure 7. Effect of hole diameter on tray efficiency and pressure


drop of dual-flow trays. Cyclohexane/n-heptane at 1.63 atm and
total reflux. Tray spacing ) 0.61 m, open area ) 19.1%. FRI data.

drop curve, indicating the onset of loading where higher


vapor velocities cause increased liquid holdup. This is
analogous to loading in packed columns.
For a pressure balance to be maintained on a given
dual-flow tray

ht ) hgd + hL ) hL′ - hLd (4)

Figure 6. Fractional entrainment of liquid in vapor. Methanol/ where ht is the total pressure drop across the tray,
water, 1.0-m-diameter column. Data from ref 5 and Figure 2. The
superimposed curve for cross-flow sieve trays is taken from ref
height of clear liquid; hgd is the pressure drop for vapor
12. passing through fraction of holes x; hL is the residual
pressure loss for vapor flowing through the froth; hLd is
moles of liquid entrained to the total downflow including the pressure loss for liquid flowing through 1 - x
entrainment recycle. fraction of holes; and hL′ is the liquid head to force liquid
Representative values of Ψ calculated from the ex- through 1 - x fraction of holes.
perimental data of Figure 2 are shown in Figure 6. Also Equation 4 implies that hL′ is greater than hL, because
shown is a composite curve for cross-flow sieve tray liquid and vapor flow in opposite directions through the
entrainment, taken from ref 12. It appears that similar holes. The dynamics of the dual-flow tray are such that
mechanisms for entrainment prevail for the two tray momentum in the liquid is converted to static head to
types. Tentatively, the generalized curves for sieve trays allow liquid to flow downward against the static pres-
can be used for estimating the liquid entrained by the sure gradient. Because the mechanisms of flow are so
rising vapor of dual-flow trays. This assumes that the poorly understood, we will concentrate on the gas flow
vapor breaks through a dual-flow tray froth in the same portion of eq 4
manner that it breaks through a cross-flow sieve tray
froth. A more quantitative assessment of the entrain- ht ) hdG + hL (5)
ment effect must await experimental data; however,
prudent designs call for an approach to flooding not with
greater than 80-85%, close to the peak efficiency region,
where the entrainment effect is minor. Ux2FL
Low Loading Effects. To interpret the loss of hdG ) (6)
efficiency at lower loadings, an “entrainment-in-reverse” 2gCv2FG
approach is used, even though there is no recycle due
to entrainment. To take an example from Figure 2, for where Ux is the linear velocity of vapor through x
a superficial velocity of 1.10 m/s (1.10/1.85 or about 60% fraction of total holes and Cv is the orifice coefficient, a
of the peak loading), the efficiency is 58%, or 70% of function of the hole diameter and effective open area.
the peak value. Using an adaptation of eq 2, one obtains Orifice Coefficient. For flow through holes in a dual-
flow tray, the correlation developed originally for sieve
1 - φ′ 1 - 0.70 trays by Leibson et al.,13 presented originally in graphi-
Ψ′ ) ) ) 0.34 (3)
φ′Ep + (1 - φ′) 0.70(1 - 0.70) cal form, can be represented analytically by

where φ′ is the fraction of the peak efficiency (E/Ep) Cv ) 0.74(Ah/Aa) + exp[0.29(tt/dh) - 0.56] (7)
and Ψ′ is a correlating parameter. Appropriate values
of Ψ′ thus enable the total efficiency curve to be where tt is the tray thickness, dh is the hole diameter,
established. and the hole area Ah is based on the holes passing vapor.
Pressure Drop. Representative pressure drop data Orifice coefficients obtained for the tray geometries
for dual-flow trays of two open areas are shown in considered in this work (Table 1) ranged from 0.58 to
Figure 4, and Figure 7 shows the effect of hole size on 0.72. By comparison, for a sieve tray with 12% hole area,
pressure drop (as well as on efficiency). In the region of 12-mm holes, and a tray thickness of 1.6 mm, the orifice
peak efficiency, there is an evident break in the pressure coefficient has a value of 0.80. For the two hole sizes of
1636 Ind. Eng. Chem. Res., Vol. 41, No. 6, 2002

Figure 7, the difference in orifice coefficient accounts


for about one-third of the difference in total pressure
drop.
Liquid Holdup. Earlier work by Kotschering et al.14
provided a correlation for equivalent clear liquid height
on dual-flow trays, which was modified later by Xu et
al.8 The Xu modification was adapted to the FRI data
in the present work as follows

(LML)n[US(FG/FL)0.5]b2
hL ) b1 (8)
FL(Ah/Aa)b3 (tt/dh)0.42

with regressed values of the constants b1 ) 0.01728, b2


) 1.0, and b3 ) 1.50. In the original work, the exponent
n is obtained from the expression Figure 8. Fraction of holes passing vapor, cyclohexane/n-heptane,

()
1.63 atm. Tray spacings, open areas, and hole diameters as
Ah 1.5
indicated. FRI data.
n ) 4.3 (9)
Aa

Fraction of Holes Passing Vapor. Because there are


no measurements of this parameter, it must be deduced
from measured pressure drop data for the entire tray
(eq 5). After correcting for liquid holdup hL by eq 8, eq
6 is used

Ux2FL
hdG ) ht,meas - hL ) (10)
2gCv2FG

from which

x
Q FL
χ) (11)
Ah (ht,meas - hL)(2gCv2FG)

where Q is the volumetric flow rate of vapor.


Even though this approach is semiempirical, it is
useful for predicting tray pressure drops from the
number of active holes. The fraction χ has been cor-
related as a function of hole area, tray spacing, and
approach to flood using the FRI data as Figure 9. Parity plot for the experimental and calculated
pressure drops of dual-flow trays. Cyclohexane/n-heptane, i-

( )( )
Ah/At
[ (|% floodC - B|)]
butane/n-butane, and 2-propanol/water (136 points).
0.8 0.2
TS
χ)A exp -0.35
0.2 0.610
(12)
where TS is the tray spacing in meters and A, B, and C
are the regression correlation constants (A ) 0.4668, B
) 90, C ) 45).
A representative plot of fractional holes passing vapor
is shown as Figure 8. Although the trend with loading
might seem counterintuitive, until direct measurements
are made, the relationships will remain useful for the
design process, as shown later in a comparison of
calculated and measured overall pressure drop data.
Pressure Drop Model Analysis. The approach
outlined above was used to develop the parity plot in
Figure 9. The plot is based on all of the total reflux FRI
Figure 10. Example comparison of predicted and experimental
experimental data. The mean absolute deviation is pressure drops for dual-flow trays. Cyclohexane/n-heptane, total
23.0%, and the average deviation is 0%. Of the 136 reflux, 1.63 atm. Column diameter ) 1.2 m, tray spacing ) 0.91
points, 13 lie above the (25% range. Most of these m, open area ) 19%, hole diameter ) 12.7 mm. FRI data.
“outside” points derive from near-flooding conditions,
where measurements are not precise. If the water/2- The efficiency curve has been included to show the
propanol system is excluded, leaving only hydrocarbon efficiency peak at the pressure drop curve break.
systems, only 9 of the 122 points are outside the (25% Froth Height. A two-phase porosity relationship for
range. In general, pressure drop is underpredicted in dual-flow trays was developed by Mahendru and Hackl15
the near-flooding zone. A representative performance and used by Xu et al.8 to predict liquid holdup for a 300-
plot showing the pressure drop fit is given in Figure 10. mm (11.8-in.) column operated with the methanol/water
Ind. Eng. Chem. Res., Vol. 41, No. 6, 2002 1637

and methanol/2-propanol systems. The porosity  (vol- Table 2. Comparison of Calculated and Measured
ume fraction of vapor in the froth) determined by Xu et Efficiency for Dual-Flow Trays
al. is hole efficiency,

[ ]
pressure TS % diam peak, calculated
-0.2 refa system (kPa) (m) open (mm) meas (%)b
Us2FG
 ) 1.0 - 0.0946 (13) FRI C6/C7 28 0.61 13.5 12.7 79 43
ghLFL FRI C6/C7 28 0.61 18.9 25.4 75 53
FRI C6/C7 28 0.61 25.6 25.4 63 54
Accordingly, the froth height for dual-flow trays, hf, can FRI C6/C7 166 0.61 13.5 12.7 83 67
be calculated from the porosity and equivalent clear FRI C6/C7 166 0.61 19.1 12.7 81 69
FRI C6/C7 166 0.91 19.1 12.7 57 71
liquid height FRI C6/C7 166 0.61 18.9 25.4 81 74
FRI C6/C7 166 0.61 25.6 25.4 55 70
hL FRI C6/C7 345 0.61 18.9 25.4 98 75
hf ) (14) FRI C6/C7 345 0.91 29.3 11.9 67 71
(1 - ) FRI C4 1138 0.61 13.5 12.7 121 77
FRI C4 1138 0.30 19.1 25.4 90 71
As a note of caution, eqs 10 and 11 should be used only FRI C4 1138 0.61 18.9 25.4 115 77
in the vicinity of the peak efficiency, i.e., where the froth FRI C4 1138 0.61 25.6 25.4 92 75
is well-developed and stable. At low loadings, the liquid FRI C4 1138 0.91 29.3 11.0 96 74
on the tray is not well-aerated. FRI IPA/H2O 101 0.61 19.1 12.7 75 69
FRI xylenes 16 mm 0.61 12.9 12.7 81 60
FRI xylenes 16 mm 0.61 17.8 12.7 63 62
Mass Transfer Efficiency FRI C8/C10 10 mm 0.61 12.9 12.7 72 72
alcohols
Spray Zone. Previous models (e.g., Xu et al.8) follow FRI C8/C10 10 mm 0.61 17.8 12.7 42 86
earlier work with cross-flow trays and are based on the alcohols
assumption that essentially all of the interphase trans- 5 MeOH/H2O 101 0.41 14.2 c 82 80
fer occurs within the liquid-continuous froth immedi- FRI MeOH/H2O 101 0.41 18.2 c 82 80
ately above the tray floor. However, as pointed out 2 Bz/Tol 101 0.51 14 12.7 87 60
earlier, there can be significant mass transfer in the a Column diameters: FRI, 1.2 m; ref 5, 1.0 m; ref 2, 0.45 m.

spray zone above the froth, where all of the downflow b Contribution of spray zone not included. c Slots, 4 mm × 150 mm.
liquid comes into contact with the rising vapor (thereby
differing from conventional cross-flow trays). Because model of Garcia and Fair16 was selected as the most
it is likely that complete horizontal mixing takes place recent and extensively tested sieve tray model available.
in the froth (in the region of peak efficiency), the To employ the model, one must utilize a column
additional mass transfer should be taken into account. diameter such that the active (bubbling) area is equiva-
Because the froth is well-mixed, efficiency in the froth lent (after considering downcomer area) to the total
cannot exceed 100% and because FRI obtained efficien- cross-sectional area of the dual-flow tray.
cies as high as 150% (Table 1), it is clear that, for some The correspondence between modeled and measured
systems, mass transfer in the spray zone can be peak efficiencies for each system is shown in Table 2.
significant. The liquid flowing through this zone has In almost all cases, the predicted value is lower than
been observed to flow primarily as streams, with the observed value, lending some credence to the
breakup into droplets for low-surface-tension systems importance of mass transfer in the spray zone. If a
such as i-butane/n-butane. The resulting additional multiplier of 1.2 is used to account for mass transfer in
interfacial area is available for mass transfer. the spray zone, better agreement is obtained, as shown
There appear to be no published data for spray zone in Figure 11. This multiplier is empirical, based on the
mass transfer. Studies by the Separations Research earlier discussion, and the modeled efficiency for the
Program have included empty-column contacting of froth zone is more fundamental, although more conser-
rising vapor with liquid fed through a conventional vative. Clearly, the model prediction must be discounted
distributor (430 streams/m2) normally used for packed at lower loadings. The low-load correction is empirical,
column work. A maximum of one theoretical stage was based on experimental observations. The ratio of cor-
obtained at contacting heights in the range of 3.5 m. rected efficiency to peak efficiency can be obtained
Clearly, for representative heights of the spray zone in through a rearrangement of eq 3
dual-flow trays (less than a tray spacing), one might
E 1
expect no more than 0.1-0.2 theoretical stages, depend- φ′ ) ) (15)
Ep Ψ′
ing on conditions. At present, any correction with a mass 1 + Ep
transfer model to account for spray zone transfer must 1 - Ψ′
be regarded as empirical. Analysis of the FRI data Values of Ψ′ calculated from experimental data are
indicates that up to 0.2 theoretical stages can be shown in Figure 12. The scatter indicates that this
achieved in the spray zone, depending on the system simplistic approach might neglect some geometric and
and flow conditions. It is prudent, considering the mechanistic effects. However, prudent tray specifica-
present level of knowledge, to neglect the mass transfer tions call for hole sizes in the 12.7-25.4-mm range and
contribution of the spray zone. open areas in the 15-20% range. The lines in the
Froth Zone. Sieve tray mass transfer in the froth summary plot (Figure 12d) can be adapted to most
zone has been modeled successfully for cross-flow sieve practical situations.
trays. If we assume that mass transfer behavior in a
dual-flow froth can be treated the same as in a cross- Tray Geometry Variables
flow froth, as mentioned earlier, then we can estimate
the value of the dual-flow efficiency using models In the present study, emphasis has been placed on
developed for cross-flow sieve trays. The point efficiency dual-flow trays with 15-20% open area, 12.7-mm holes,
1638 Ind. Eng. Chem. Res., Vol. 41, No. 6, 2002

and tray spacings of 0.5-0.7 m. The reason is that


design studies have shown that, for most applications,
this is an economical combination of tray geometric
variables. For example, Figure 5 shows the capacity
advantage of large open area and high tray spacing.
However, when mass transfer efficiency and pressure
drop are added to the comparison, the extreme cases of
geometry might not be optimum. Still, this is a matter
for detailed analysis of individual cases.
Figure 7 shows the effect of hole diameter on ef-
ficiency and pressure drop for the cyclohexane/n-hep-
tane system at 1.63 atm and 19.1% open area. The
larger holes have a higher efficiency but also a higher
pressure drop. The effect of tray spacing on pressure
drop is minor if operation is well below the flood point.
Tray spacing can affect the efficiency in that it can
provide extra volume for mass transfer in the spray
zone.
Finally, mention should be made of tray diameter
effects. Figure 3 shows a difference between 0.45- and
2.50-m-diameter trays, more in capacity than in ef-
Figure 11. Parity plot, predicted vs observed peak efficiency for ficiency. The importance of maintaining a good liquid
dual-flow trays.

Figure 12. Low-loading discount factor Ψ′ for representative experimental data: (a) 12.7-mm holes, 18-20% open area; (b) 25.4-mm
holes, 18-20% open area; (c) 12.7-mm holes or smaller, 13-15% open area; (d) summary plot.
Ind. Eng. Chem. Res., Vol. 41, No. 6, 2002 1639

distribution has been mentioned (spray nozzles are often Acknowledgment


used to add reflux to the top tray). Zuiderweg et al.2
observed generally higher capacities and lower pressure This paper would not have been possible without the
drops for larger tray diameters, suggesting that the generous provision of large-scale performance data by
methods given here, largely based on a column diameter Fractionation Research, Inc. (FRI). The authors extend
of 1.2 m, are somewhat conservative. However, reports thanks to FRI, as well as to the Separations Research
of poor dual-flow tray performance for larger diameters Program at the University of Texas and to Koch-Glitsch,
and very low flow parameters (as in high-vacuum Inc., for providing support for the modeling and analysis
columns) must be recognized. It is possible that very effort.
high volumetric ratios of vapor to liquid create a
tendency for the vapor flow to tend toward the center Nomenclature
portion of the tray, leading to reduced tray efficiency. A, B, C ) constants in eq 12
Thus, the findings described in this paper should be Aa ) tray active area, m2
used with great caution for the low values of the flow Ah ) tray hole area, m2
parameter LML/VMVxFG/FL. At ) total column cross-sectional (superficial) area, m2
Ax ) area of holes open to vapor at any instant ) χAh, m2
Summary and Conclusions b1, b2, b3 ) constants in eq 8
Csb ) Souders-Brown capacity parameter, m/s
We have proposed a rational method for the analysis Cv ) orifice coefficient, eq 6
and design of dual-flow tray distillation columns. The dh ) hole diameter, mm or m
approach is based on a large number of performance Ep ) peak efficiency, fractional
tests under distillation conditions in a large-diameter Ew ) efficiency corrected for liquid-in-vapor entrainment,
column. No rational model useful for commercial-scale fractional
dual-flow tray column design or analysis has been Fs ) vapor F factor based on superficial area, UsFG0.5, m/s
published previously. We have made suggestions for (kg/m3)0.5
possible mechanisms of phase contacting in the liquid- g ) gravitational constant, m/s2
continuous froth as well as in the vapor-continuous h ) pressure loss, m of tray liquid
spray zone. There is a clear need for additional studies, hdG ) drop for vapor flow through orifices
both experimental and theoretical. hL ) residual drop through two-phase mixture
The following sequence of steps should be followed in hLD ) drop for liquid flow through orifices
designing or rating dual-flow columns: (1) The column hL′ ) head required to force liquid through the orifices
diameter should be determined on the basis of flooding ht ) total pressure drop for a dual-flow tray
limits and a prudent approach to flooding. (This is the ht,meas ) measured total pressure drop for a dual-flow tray
same procedure used for other contacting devices such hf ) height of froth on the tray, m
as cross-flow trays and packings.) An approach to L ) liquid flow, kg mol/s
flooding of about 75-80% is in the range of peak m ) slope of the equilibrium curve
n ) exponent in eq 8
efficiency (Table 1). The effect of entrained liquid in the
ML, MG ) molecular weights of liquid and vapor, respec-
vapor can be used to pinpoint the flooding approach as
tively
well as the peak efficiency. (2) The peak efficiency
Q ) volumetric flow rate of vapor, m3/s
should be modeled by an adaptation of a froth-contacting
tt ) tray metal thickness, mm
model for cross-flow sieve trays. To use the model, one
TS ) tray spacing, m
needs to know the fraction of holes that are active.
U ) vapor velocity, m/s
Because the model provides only the efficiency in the
Us ) superficial velocity
froth, enhancement of this efficiency can occur through
Ux ) velocity through x fraction of total number of holes
added mass transfer above the froth (spray zone). On
V ) vapor flow, kg mol/s
the basis of the FRI measurements, the enhanced
χ ) fraction of total holes passing vapor at any instant
efficiency can be as much as 1.2 times the modeled
efficiency, but until more definitive studies are made, Greek Letters
a factor of 1.0 is recommended. (3) The efficiency below  ) void fraction in the froth
the peak point should be obtained by an empirical λ ) ratio of slopes, equilibrium curve to operating line )
correction to the peak efficiency. Tentative plots have mV/L
been provided to enable an estimation of the effect of FL ) liquid density, kg/m3
loading on efficiency. (4) The pressure drop should be FG ) vapor density, kg/m3
calculated on the basis of conventional equations, taking φ ) ratio of wet efficiency (with entrainment) to dry (peak)
into account the fraction of the total holes passing vapor efficiency
and the porosity of the froth. φ′ ) ratio of lower loading efficiency to peak efficiency
This is a pioneering attempt at the modeling of dual- Ψ ) efficiency discount factor for entrainment in vapor
flow tray performance. Future studies under distillation Ψ′ ) efficiency discount factor for lower loadings
conditions might require experimental techniques not
yet fully developed, e.g., X-ray tomography to show froth Literature Cited
quality changes and the dynamic activity at the holes.
Because such work has not yet been effective for the (1) Zuiderweg, F. J.; Verburg, H.; Gilissen, F. A. H. Comparison
more conventional devices, one might not expect dual- of fractionating devices. Proc. Int. Symp. Distill. 1960, 201.
(2) Zuiderweg, F. J.; de Groot, J. H.; Meeboer, B.; van der Meer,
flow technology advances soon. Perhaps the work re- D. Scaling up distillation plates. Proc. Int. Symp. Distill. 1969, 5,
ported here will help reactivate interest in a device that 78.
is efficient and inexpensivesand that can operate well (3) Huml, M.; Standart, G. The hydraulics of large turbogrid
under fouling situations. trays. Brit. Chem. Eng. 1966, 11 (11), 1370.
1640 Ind. Eng. Chem. Res., Vol. 41, No. 6, 2002

(4) Kastanek, F.; Huml, M.; Braun, V. Measuring the efficiency (12) Fair, J. R. Entrainment-efficiency effects on distillation
of a column of one metre diameter. Proc. Int. Symp. Distill. 1969, sieve trays. Presented at the AIChE Annual Meeting, Chicago,
5, 100. IL, Nov 15, 1996.
(5) Kastanek, F.; Rylek, M. Turbogrid tray efficiencies. Collect. (13) Leibson, I.; Kelley, R. E.; Bullington, L. A. How to design
Czech. Chem. Commun. 1970, 35, 3367. perforated trays. Pet. Ref. 1957, 36 (2), 127.
(6) Kastanek, F.; Standart, G. Efficiency of selected types of (14) Kotschering, N. A.; Oleski, W. M.; Dilman, W. W. Inves-
large distillation trays at total reflux. Sep. Sci. 1967, 2, 439. tigation of dualflow tray behavior under distillation conditions.
(7) Fractionation Research, Inc., Stillwater, Oklahoma. Re- Khim. Prom. 1960, 7, 591.
search Progress Reports, 1955-1958. Available from Oklahoma (15) Mahendru, H. L.; Hackl, A. Contribution to the design of
State University. sieve trays without downcomers. Inst. Chem. Eng. Symp. Ser. 56
(8) Xu, Z. P.; Afacan, A.; Chuang, K. T. Efficiency of dualflow 1979, 3.2/35-47.
trays in distillation. Can. J. Chem. Eng. 1994, 72, 607. (16) Garcia, J. A.; Fair, J. R. A fundamental model for the
(9) Baird, J. L. Letter to the editor. Chem. Eng. Prog. 1999, 95 prediction of distillation sieve tray efficiency. Ind. Eng. Chem. Res.
(5), 7. 2000, 39, 1809 (part 1), 1818 (part 2).
(10) Fair, J. R. How to predict sieve tray entrainment and
flooding. Petro/Chem Eng. 1961, 33 (10), 45. Also, recent editions Received for review April 12, 2001
of Perry’s Chemical Engineers’ Handbook; McGraw-Hill: New Revised manuscript received December 13, 2001
York. Accepted December 17, 2001
(11) Colburn, A. P. Effect of entrainment on plate efficiency in
distillation. Ind. Eng. Chem. 1936, 28, 536. IE010326W

S-ar putea să vă placă și