Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

aration was obtained using a membrane containing 20% Discussion

sulfolane. The behavior of the sulfolane plasticized vinylidene flu-


The membrane behavior in the lower sulfolane concen-
oride membrane can probably be adequately explained by
tration range probably results from a complicated SO2
the high solubility of SO2 in sulfolane. It would be inter-
solubility effect which changes with pressure and the
amount of sulfolane in the membrane in a nonlinear man- esting to compare the premeation data with quantitative
data on the solubility of SO2 and N2 in sulfolane at vari-
ner. The decreased membrane selectivity and high flux at
ous temperatures and pressures but unfortunately these
sulfolane compositions greater than 12.8% probably is due
data are not available. The membrane behavior appears
to over-plasticization which causes pin-hole leaks.
to be very similar to the permeability of organic vapors
An addition level of 8.2% sulfolane was chosen for fur-
ther parameter studies. through plastic films which show a complicated depen-
dence on pressure and concentration owing to a strong in-
Effect of Permeator Temperature teraction between the solute and membrane (Li, et ai,
Figure 2 shows the effect of cell temperature on mem- 1965). This is in contrast to permanent gases which have
brane selectivity and flux using a membrane containing permeation coefficients independent of pressure. Thus, a
8.2% sulfolane and a feed containing 6% SO2. For the runs high flux and separation factor results at high pressures
made at 14° a container of ice was placed in the constant and high SO2 concentrations where saturation of the gas
temperature enclosure to reduce the temperature below with SO2 is approached. Since the diluent is a permanent
room temperature. As can be seen, the separation factor gas the behavior of the system is probably more compli-
decreased rapidly with increasing temperature while there cated than vapor permeation, however.
was a nearly linear increase in flux. This response is prob- The long-term performance of the plasticized mem-
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ably due to a decrease in the solubility of SO2 in the plas- brane has not been determined and probably deterioration
ticizing medium with increasing temperature coupled with a decrease in selectivity and flux with time could be
with an increase in the diffusion coefficient. expected. From this standpoint the lower temperature
Effect of Feed Composition would probably be better or, it may be possible to use a
Downloaded via AALTO UNIV on November 27, 2022 at 06:48:32 (UTC).

heavier sulfone as a plasticizer to advantage.


The effect of feed composition on flux, permeate com-
position, and separation factor is shown in Figure 3. These Literature Cited
data were taken at room temperature using membranes Kirk, R. E., Othmer, D. F., Ed., “Encyclopedia of Chemical Technology,"
containing 8.2% sulfolane. Data were not obtained using Vol. 19, 2nd ed, Interscience, New York, N. Y., 1969, p 252.
LI, N. N., Long, R. B., Henley, E. J., Ind. Eng. Chem., 57, 18 (1965).
the 12.5% SO2 feed material and 500 psig cell pressure be- McCandless, F. P., Ind. Eng. Chem., Process Des. Develop., 11, 470
cause the gas mixture becomes saturated in SO2 below (1972)
.

this pressure. McCandless, F. P., Ind. Eng. Chem., Process Des. Develop., 12, 354
(1973)
.

As can be seen, the permeation rate and separation fac- Stern, S. A. Garels, P. J., Sinclair, T. F., Mohr, P. H., J. Appl. Polym.
tor are highly dependent on feed composition and this Sci.. 7, 2035 (1963).

probably indicates a nonlinearity in the concentration de- Received for review June 6, 1973
pendence of the permeability coefficient. Accepted September 19, 1973

Drop Size Distributions Produced by Turbulent Pipe Flow of Immiscible Fluids through
a Static Mixer

Stanley Middleman
Chemical Engineering Department. University of Massachusetts, Amherst, Massachusetts 01002

Data are presented for six organic liquids of viscosities ranging from 0.6 to 26 cP and interfacial ten-
sions ranging from 5 to 46 dyn/cm dispersed in water as the continuous phase. The effect of mixer
pitch and number of mixing elements is illustrated. Some aspects of Kolmogoroff’s theory provide a
basis for correlation of the data.

When a mixture of two immiscible fluids is subjected to The Sauter mean is a particularly useful average since the
pipe flow, a dispersion is created which can be character- interfacial area per unit volume can be obtained directly
ized by a “drop-size distribution function,” ( ). From from
(£)) it is possible to calculate various average drop sizes.
For example, the Sauter mean diameter may be defined Av =
6<t>/D32 (2)
as
where is the volume fraction of dispersed phase.
Z)32
=
f
"0
( )) dD/ f
"0
D2 () dD (1) The drop-size distributions produced by turbulent pipe
flow have been the subject of theoretical and experimental

78 Ind. Eng. Chem., Process Des. Develop., Vol. 13, No. 1, 1974
The energy dissipation rate é is a local quantity that
undoubtedly varies spatially across the pipe diameter.
Nevertheless, to make progress, we will relate e to macro-
scopic variables by assuming that
=
a2(QAP/pAL) (8)
The term QAP is simply the rate of energy input into the
fluid, and pAL is the mass of fluid in the pipe of cross sec-
tional area A and length L. In eq 8 we let p be the contin-
Figure 1. The Kenics static mixer. uous phase density, since we will only consider fairly low
volume fraction dispersions. It is convenient to replace
studies by Hinze (1955), Paul and Sleicher (1965), Sleich- Q/A with the average velocity Vin subsequent equations.
er (1962), Collins and Knudsen (1970), and Hughmark If eq 8 is substituted into eq 7 one finds, after some al-
(1971), the results of which make it possible to predict the gebraic rearrangement, that
behavior of such systems from a knowledge of physical
properties and macroscopic flow variables. Our work fo-
() =
[( / „) \\>3'5/2·5] (9)
cusses on the production of dispersions by flow through a
In eq9 the pipe diameter Do has been introduced. The
pipe containing an in-line motionless mixer, the Kenics Weber number is defined as
static mixer.
The Kenics static mixer is shown in Figure 1. It consists A\Ve
=
pV2DJo (10)
of a series of stationary elements, fixed relative to the
The friction factor is
pipe wall, which divert the flow field and cause the mix-
ing action. The mixing elements are rectangular, and split f =
DjAPI2pVlL (11)
the pipe cross section into two semicircular sections. Each
element is twisted through 180°, and alternate left- and Since eq 9 indicates only what is a function of, all nu-
merical coefficients in front ofD/Do have been dropped.
right-hand twists are fixed in series down the pipe axis.
The dependence of the drop size distribution on mixer If eq 9 is used in eq 1, it follows immediately that
geometry has been examined, and the results are consid- D,2/D0 =
CAWe-3'5/-2'5 (12)
ered in the light of a simple hydrodynamic treatment.
a result which can be subjected to experimental verifica-
Theory tion. In pipe flows, the friction factor is a weak function of
We begin with the idea that the turbulent flow field Reynolds number for Nne > 3000, and a reasonable ap-
may be characterized by an energy spectrum function proximation is seen, e.g., in Bird, et al. (1960), to be
E(k), defined so that E(k) dk is the energy per unit mass f ~
NRe~i4 (13)
of continuous phase associated with turbulent fluctuations
of wave number k to k + dk. A drop of diameter D is sta- Thus we see that
bilized by surface energy of magnitude 4 2 . The drop is
D32/D0 =
CA\Ve~3 «AV 10
(14)
subject to disruptive energy, associated with turbulent
fluctuations, of magnitude 1kirDipt, where The point of displaying the Reynolds number explicitly is
to indicate the very weak dependence of drop size on
c =
E(k) dk (3) Reynolds number at constant Weber number. This depen-
*T/£> dence would very likely be masked by experimental error
in most cases.
Only energy in turbulent eddies of length scales smaller
than D (wave numbers k > I/O) is considered, on the If we wish to compare dispersions made in a circular
grounds that larger eddies merely transport a drop, but do pipe to dispersions made in the same pipe, but one con-
not disrupt it. taining a motionless mixer, then eq 12 suggests
We assume that ( ) should be some function of the (/o/f m)" = 0
( 32)µ/( :12) (15)
ratio of turbulent energy to surface energy, and write
at constant flow rate, and assuming equal physical prop-
erties. Friction factors in a Kenics static mixer have been
VeTrO^f E(k) determined over a wide range of Reynolds numbers and
() = _« ID
4 2 mixer geometries, and hence it should be possible to
subject eq 15 to experimental test. It is worth noting at
To proceed, it will be assumed that the gross features of this point, however, that fo/fta is typically of the order of
the turbulent energy spectrum are governed by Kolmogo- 10-2 for Ars > 3000, and so one may expect to reduce the
roff’s theory of turbulence (see Hinze, 1955, 1959), which drop size by an order of magnitude through the use of the
gives static mixer.
E(k) =
a1e2/3fc_5/3 (5) Experimental Methods
in the inertial subrange (to be defined subsequently) where The continuous phase was water in all cases reported
« is the energy dissipation rate per unit mass and «i is here. The dispersed phase fluids are listed in Table I
some undetermined coefficient. It follows that along with the pertinent physical properties.
Figure 2 shows the layout of the experimental system.
ly
f .. n
E(k) dk =
3/2a1e2'3D2'3 (6) Static mixers varying in pitch (Le/D) and number of ele-
ments in series were placed inside glass pipes. The dis-
and persed phase was introduced through a small glass tube
which entered the pipe through a T branch about one di-
«iP~2, 3^Z)5 ameter upstream of the first element of the mixer. On the
( )) = 3
(7)
16ae ] outside of the pipe, at the position of the last element of

Ind. Eng. Chem., Process Des. Develop., Vol. 13, No. 1, 1974 79
dispersed
phase
inlet
f

inlet ;

camera

Figure 2. Schematic of system for production and measurement


of drop sizes.

Table I. Properties of Dispersed Phase Liquids at 25° 1 10 io2 io3 io4

p, g ml µ', cP s, dyn/cm Nwe

Anisóle 0.99 1.0 26 Figure 3. Sauter mean drop diameter for the benzene-water sys-
tem.
Benzene 0.87 0.6 40
Benzyl alcohol 1.0 5.0 5
Cyclohexane 0.76 0.8 46
Oleic acid 0.90 26.0 16
Toluene 0.87 0.6 32

the mixer, a small rectangular transparent plastic box was


placed. When filled with water the box made it possible to
photograph the dispersion flowing beyond the last element ~
D0
=
1/2" I'
with minimal optical distortion due to curvature of the Cl i oleic acid - -1

pipe walls. benzyl alcohol


~

I q-2
Photographs were taken with a 35-mm Pentax camera, benzene C ¿

toluene 3

using a 135-mm lens on a bellows rack. High-speed flash cyclohexane


was provided by an EGG Model 549 microflash of O.S^sec r anisóle ·
duration. The negatives were placed in 35-mm slide
frames and projected, for measurement, onto sheets of I 10 io2 io3 io4
white paper. N we
The volume flow rate was determined by collecting the
effluent from the pipe, over a known time interval, in a Figure 4. Sauter mean drop diameter for six different fluids.
graduated cylinder. After the dispersion separated into
two distinct layers in the cylinder, it was possible to de-
/o)2'8 = 6.2. According to eq 15, then, we should observe
termine the volume fraction of dispersed phase.
(DszIo/IDszXm = 6.2. The observed value is 7.3.
The drop diameter measurements were classified into 'For all'of the empty pipe data shown the /m,/o ratio is
size groups of 10- or 20-µ intervals, and the Sauter mean
100, and the drop size ratio should be constant. This does
diameter was calculated from these discrete data using not appear to be so even over the narrow range displayed,
__
M and this may indicate a failure of this aspect of the
D.i 2
=
Xn,I)/
i-l
/J2n,Dr
i-l
(16) model. Equation 8 is the most likely source of this dis-
crepancy. Even so, these results are not in bad agreement
The number of drops counted, M, was usually of the order with the general features of this model.
of 100 to 200. Since our goal was not to study drop size behavior in an
Results empty pipe, we have not pursued this point further. In-
stead, we turn to a more detailed study of behavior in the
Figure 3 shows data obtained in the benzene-water sys- Kenics static mixer.
tem in two different diameter pipes containing static mix- Effect of Dispersed Phase Viscosity. Figure 4 shows
ers. We find it convenient, and apparently acceptable all of the data obtained in 1 and 0.5-in. static mixers of
within experimental accuracy, to plot D32/D0 as a func- constant pitch (Le/D 1.5) and constant number of ele-
=

tion only of the Weber number. Since the Reynolds num- ments (ne 21). The properties of the six fluids studied
=

ber range is only about an order of magnitude for the data are listed in Table I. We note that the data for dispersed
shown, the tenth-power dependence predicted by eq 14 is phase fluids of viscosity comparable to that of the contin-
too weak to make itself felt. This is likely to be the case in uous phase (water) are grouped together, while data for
any practicable application of a correlation of the type benzyl alcohol (µ' 5 cP) and oleic acid (µ'
= 26 cP) are =

shown in Figure 3. separated from the main body of the data, at higher
However, in going from a static mixer to an empty pipe, Weber numbers, and indicate that larger mean drop sizes
at constant Weber number and Reynolds number, there is are produced.
a large change in friction factor, and eq 15 suggests that a Qualitatively, such a result is to be expected, since a
significant effect on drop size should be observed. A limit- high viscosity in the dispersed phase retards disruption of
ed amount of data were taken in an empty pipe, and these the drop. In writing eq 4 it was assumed, however, that
are shown in Figure 3. As expected, the drop size is con- only surface energy stabilizes the drop. Empty pipe data
siderably different in the two cases. At a Weber number show similar effects, and Hinze (1955) and Collins and
of 100, for example, the Reynolds number in the 0.5-in. Knudsen (1970) suggest methods of correlating the data to
pipe is 6200, and the ratio of friction factors, fu/fo, is ob- account for dispersed phase viscosity. The methods do not
served to be /m//o 100 from which it follows that (/m/
=
work well with our data.

80 Ind. Eng. Chem., Process Des. Develop., Vol. 13, No. 1, 1974
1

I 10 io2 io3
Nw.
Figure 7. Effect of number of elements on drop size.

Figure 5. Effect of volume fraction on drop size.

Figure 8. Effect of number of elements on drop size at a fixed


Weber number.

oleic acid
0=1/2”
V \|.n
X'¿^-
^^

10
I 10

V (crrvsec"
Figure 6. Effect of pitch (Le/D) on drop size.

Effect of Volume Fraction of Dispersed Phase. The


data shown in Figures 3 and 4 were all taken at low vol-
100 D (cm)
ume fractions, , in the range of 0.5 to 1%. It was as-
sumed. in writing eq 4, that drop size is determined only Figure 9. Drop size distribution curve from a typical run.

by factors affecting drop break-up, and that once a drop is


formed it will either remain at that size or, possibly, be Effect of Mixer Pitch (Le/D). Figure 6 shows the effect
broken by subsequent interaction with a “high-energy” of pitch on drop size for oleic acid-water dispersions. At
eddy. This idea ignores the possibility of coalescence of constant flow rate, the friction factor is observed to de-
drops as a factor in determining drop size. Coalescence crease with increasing pitch. For example, in the range of
will be promoted by two factors: both a high volume frac- flow rates illustrated here, /2.0//1.0 5-7. Equation 15,
=

tion as well as turbulent mixing promote collision between then, suggests that the ratio of mean drop sizes should be
drops. The final drop size distribution, then, represents about a factor of 2. The data show some scatter, and the
equilibrium for a dynamic process which balances drop separation of the upper and lower sets is more like a fac-
break-up (essentially a Weber number dominated process) tor of 1.5 than 2. Still, the results are in general agree-
against drop coalescence (a volume fraction and Reynolds ment with expectations.
number related phenomenon). Effect of Number of Elements (ne). Figure 7 shows
Figure 5 shows data for the benzene-water system in a data obtained in the benzene-water system. Three mixers,
0.5-in. static mixer. Drop sizes were obtained as a func- of the same geometry but differing in the number of ele-
tion of volume fraction , with Reynolds number as a pa- ments ne, were used. As could be anticipated, longer ex-
rameter. In this case (of constant Do and physical proper- posure to the mixing action leads to reduced drop size.
ties) the Reynolds and Weber numbers are related by Awe Figure 8 shows these data, at a fixed Weber number, plot-
=
2.5 X 10-6ARe2. ted as a function of ne. The case ne 0 is simply the =

For relatively low Reynolds numbers (Are < 2250) empty pipe. It would appear that ten elements are suffi-
which correspond to very low Weber numbers (Awe < 12) cient to produce the equilibrium drop size distribution.
we see the effect of coalescence in Figure 5. At high Reyn- Drop Size Distribution. While the Sauter mean is a
olds numbers, which correspond to relatively high Weber characteristic measure of the drop sizes produced by the
numbers, the increase of drop size with volume fraction is mixer, it fails to give any indication of the range of sizes
very slight, indicating that drop break-up dominates the produced. Figure 9 shows a typical distribution curve for
process. the volume fraction fv, defined in such a way that

I rid. Eng. Chem., Process Des. Develop., Vol. 13, No. 1, 1974 81
f,.(D,) dD, =
nl) I) (17)
1=1

fv is a discrete distribution function, since the drop size


measurements were classified into discrete groups. It is
the discrete analog of the continuous function ). If the
size interval for classification is small compared to the
range (which is true in our cases) then it is reasonable to
present fv as a continuous distribution, as shown in the
figure.
The bimodal character in the neighborhood of the mean
is statistical noise associated with the fact that only a few
hundred drops were counted. It does not appear in most of
our fv curves. The small peak at low drop size may be IOO D (cm)
real. It appears in a high proportion of our distribution Figure 10. Cumulative distribution curve for data of Figure 9.
curves. It is associated with the fact that when a “moth-
er” drop breaks, it often produces two major (smaller)
“daughter” drops, and a much smaller “satellite” drop
(Collins and Knudsen, 1970). The satellite drop probably
forms from the “neck” which attaches two daughter drops
just before they separate from the mother drop.
Figure 10 shows the data of Figure 9 replotted as a cu-
mulative distribution function, Fv(Dk), defined as

FviDk) =
/() (18)

Fv(D) is simply the total volume fraction associated with


drops of diameter less than some size D. From Figure 10
one can determine that 70% of the volume of the dispersion
is associated with drops whose diameters lie within ±20%
of the Sauter mean. In this sense, then, the drop-size dis-
tribution is relatively narrow.
Figure 11 shows cumulative distribution data for a large
portion of our data. When the drop size is normalized to
the mean, D32, for each individual set of data, all of the
data fall on' a single curve. Any fv peaks associated with
satellite drops are “lost” in experimental scatter when the
data are plotted for several runs of several different dis-
Figure 11. Cumulative distribution data for several fluids and
persed phases. This suggests (as in fact is apparent when flow rates.
all individual /v curves are examined) that the satellite
peak does not occur in all data, and that its position is
not always at the same fraction of D32. boundaries, as in the case here, will be inhomogeneous. In
The linearity of Figure 11, since'it is plotted on “normal fact, there is strong evidence, described by Hughmark
(1971), that most of the drop break-up in an empty pipe
probability” coordinates, suggests that the drop-size dis-
occurs in the wall region.
tributions produced in the Kenics static mixer are roughly
normally distributed about the Sauter mean. A fit of the Isotropy relates to the question of whether parameters
data gives the distribution as such as the fluctuating components of the flow depend on
the coordinate directions. Again, in a pipe flow, it is un-
likely that isotropic turbulence is achieved except far from
solid surfaces. Nevertheless, Kolmogoroff’s ideas provide
the only simple vehicle for producing a semiquantita-
In interphase transport process a major factor is the
tive theory. The equations developed above are, at least,
manner in which interfacial area is distributed over the
testable through examination of experimental data. We
drop sizes. A distribution function for area, fA(D), is de- conclude that several features of our results are consis-
fined so that /a dD is the fraction of interfacial area asso-
ciated with drops in the size range D to D + dD. fA is re- tent, in a general way, with the predictions of the theory.
In cross section, a circular pipe containing a static
lated to fv by
mixer has a boundary which can be considered to be two
f A(D/D32) =
(D32/D)fv(D/D32) (20) semicircles sharing a common diameter. The hydraulic
radius of such a boundary, when defined as the ratio of
cross sectional area to wetted perimeter, is just Rh = x/iDo
Discussion (1 + 2/ )™1. One might be tempted to use a hydraulic ra-
In the development of the theory for this system several dius either as a more appropriate length scale than D0 it-
ideas of Kolmogoroff, related to the manner in which self, or in defining a Reynolds number for consideration of
energy is distributed over wave numbers in a turbulent some aspects of the dynamics of the flow field. Neither
flow, were used. It is appropriate at this point to question approach sheds much light on the results presented here.
the applicability of Kolmogoroff theory to flow in a Ken-
s If one examines the treatment of frictional effects for
ics static mixer. One restriction of the theory is that the flow in a noncircular cross section, along the lines given
flow field should be both homogeneous and isotropic. Ho- for example in Bird, et al. (1960), it is apparent that the
mogeneity relates to the absence of spatial variations of concept of hydraulic radius is of value only when most of
turbulent properties of the flow. Any shear flow with the friction producing flow is axial. If the elements of the

82 Ind. Eng. Chem., Process Des. Develop., Vol. 13, No. 1, 1974
static mixer were straight, instead of helical, the mean olds numbers in the neighborhood of 104, the drop size is
flow would indeed be axial. The helical nature of the de- comparable to .
vice, however, sets up very strong secondary flows which
are radially directed, and which contribute significantly to Nomenclature
the dynamics of the flow field. We find, then, that consid- A = cross sectional area of pipe
eration of the hydraulic radius is no aid in elucidating Av = area per unit volume, cm-1
these results. D = diameter of a drop, cm
The conditions of applicability of eq 5 require consider- Do = pipe diameter, cm
ation of eddies of a size that is very small with respect to 32 = Sauter mean diameter, cm
E(k) =
energy spectrum function, cm3/sec2
macroscopic dimensions, such as pipe diameter, but large f = friction factor
in comparison to the “dissipation scale,’’ , which is given /a(D) dD =
area fraction of drops of size D to D + dD
by Hinze (1959) as fv(D) dD = volume of fraction of drops of size D to D +
dD
=
( /3 )1/4 (2D FV(D) = cumulative distribution function, cm-1
where v is the kinematic viscosity of the continuous phase.
k = wave number, cm-1
This restriction is equivalent to the requirement that Le = axial length of mixer element, cm
N-Re = pVDo/µ, Reynolds number
« 32 « D0 (22) Nwe = pV^Do/a, Weber number
ne = number of mixer elements
which is true for most of the data presented here. /L = pressure drop per unit length of pipe, dyn/cm3
Under conditions of extremely high-energy input it is Q = volume flow rate in pipe, cm3/sec
possible to produce drops of a size comparable to or Greek Letters
smaller than . Such drops would be in a dynamic regime < = turbulent energy per unit mass, cm2/sec2
known as the viscous subrange, for which eq 5 must be re- ( = local energy dissipation rate per unit mass, cm2/sec3
placed by =
dissipation microscale, cm
µ(µ') = continuous (dispersed) phase viscosity, g/cm sec
E(k) =
a3vtU3k~v3 (23) v = kinematic viscosity of continuous phase, cm2/sec
If the analysis is carried through as in the case for the in- p =
density of continuous phase, gm/cm3
ertial subrange, one finds eventually that
= interfacial tension, g/sec2
= volume fraction of dispersed phase
D32/D0
=
µ / 3/-1 (24) Subscripts
dimensionless group easily seen to be the 0 =
empty pipe
where pV/ is a
same as Nwe/Nne· Equation 24, by comparison to eq 12,
=
pipe with mixer
suggests a much stronger dependence of D32 on flow rate, Literature Cited
and on friction factor.
Bird, R. B,, Stewart, W. E., Lightfoot, E. N., “Transport Phenomena,"
The magnitude of the dissipation scale can be esti- Wiley, New York, N. Y., 1960, pp 183-188.
mated using eq 8 in eq 21, and taking «2 1. The result
= Collins, S. B., Knudsen, J. G., AlChEJ., 16, 1072 (1970).
Hinze, J. 0., “Turbulence," McGraw-Hill, New York, N. Y., 1959, pp
is 183-186.
Hinze, J. O., AlChEJ., 1,289 (1955).
T?/D„ =
iV Re-3·''4/'—1/4 (25) Hughmark, G. A., AlChEJ., 17, 1000 (1971).
Paul, . I., Sleicher, C. A. Jr., Chem. Eng. Sci., 20, 57 (1965).
At a Reynolds number of 104, in a Kenics static mixer, we Sleicher, C. A. Jr., AlChEJ., 8, 471 (1962).
estimate /Do = 10-3. From Figure 3 we can see that at Received for review June 19, 1973
the highest Weber numbers, which correspond to Reyn- Accepted September 26, 1973

Ind. Eng. Chem., Process Des. Develop., Vol. 13, No. 1, 1974 83

You might also like