Sei sulla pagina 1di 10

Analysis of the factors inuencing dye

uptake on jet dyeing equipment


Martin Ferus-Comelo
Department of Colour and Polymer Chemistry, University of Leeds, Leeds LS2 9JT, UK
Email: m.ferus-comelo@leeds.ac.uk
Received: 30 March 2006; Accepted: 12 July 2006
Machine parameters that possibly inuence the dye-uptake rate on textile jet dyeing equipment are
reviewed in this study. Some of these parameters can be adjusted on a pilot-scale jet machine so as to
reproduce bulk conditions, whereas others cannot. It was impossible, for example, to simultaneously
achieve appropriate values for the nozzle ow velocity and the dyebath circulation time. The fabricliquor
interchange in the storage area of the machine was also different on pilot scale and on bulk scale.
Experiments on a pilot machine with a direct dye on cotton yielded, for most cases, no inuence of
fabric speed and dyebath ow rate on the dye-uptake rate despite high substantivity conditions. Only
when the ow rate dropped below 50 dm
3
min
)1
kg
)1
, the exhaustion rate diminished approximately
proportionally to the square root of the ow rate, probably because of dye depletion in the storage
chamber. On the pilot machine, the jet nozzle did not appear to contribute to the overall fabricliquor
interchange. Implications for bulk jet dyeing machines are discussed.
Introduction
An important objective of sample colorations in a textile
dyehouse laboratory is to simulate production-scale
conditions. While typical laboratory equipment, e.g. a
rotating beaker-type machine, is useful for colour
matching purposes, a more detailed process analysis in
the laboratory might require machines that operate more
closely like those in bulk production. One important
question in this context is if pilot-scale machines can be
used to accurately predict the dye exhaustion of bulk
dyeing machines.
The inuence of the liquor ow rate on the exhaustion
rate, initially established on package dyeing machines,
has been known for a long time [1] and has been
conrmed several times since then, also for other types of
dyeing equipment [24]. There are at least three possible
factors, which alone or in combination, can explain this
sensitivity of the dye-uptake rate. All three factors
contribute to a reduced dye supply at the bre surface,
thereby possibly leading to a process in which the rate-
determining step is no longer the diffusion of the dye
inside the bre (lm diffusion-controlled process) but the
diffusion of the dye from the dyebath to the bre surface
(liquid-diffusion limited). All three factors also gain in
importance as the uid velocity drops.
In the rst case, it is known that as the uid velocity
drops, a diffusional boundary layer of increasing thickness
develops at the substrate surface which reduces mass-
transfer to the individual bre [5]. In the second case, the
primary cause is a much diminished uid velocity near the
bre surface compared with the macroscopic average
dyebath ow rate. This is caused by ow resistance in the
bre assembly (microscopic depletion) [6,7]. The resistance
can occur, for example, in the intra-yarn pores [7] or at the
intersection of yarns in assemblies [8]. This phenomenon is
not to be confused with boundary layer effects which,
when ow direction is perpendicular to the bre axis, vary
little with ow rate [9]. Thirdly, the dyebath depletes as it
traverses the package or fabric pile, thus lowering the dye
supply, and consequently the uptake rate of subsequent
substrate layers (macroscopic depletion) [1013]. In the
third case, it is the macroscopic dye supply within the
entire package or pile, as opposed to the dye supply with
the yarn in the second case, which limits the overall uptake
rate. Thus, the second phenomenon can arise in a single
fabric layer, or even a single yarn, whereas the third
phenomenon requires the existence of several layers of
fabric or yarn. In this study, the term liquid diffusion
effects is used when reference is made to all three factors,
without making specic reference to either factor.
In a jet dyeing machine, fabric and dyebath are
circulating. A fabric circulation in a jet dyeing machine
may be divided into the following three phases:
Passage through the jet nozzle and, possibly, an attached
tube: In the nozzle, fabric, water and air-ows meet to
create a situation of high fabricliquor interchange. The
nozzles design and the uid velocity determine how
much kinetic energy is transferred to the fabric. Fabric
propulsion on bulk machines, especially at higher fabric
speeds, is mainly achieved by the impulse transfer in the
jet with the reel providing only a minor contribution
[14]. In order to ensure adequate propulsion of the fabric,
the liquor speed in the nozzle must be higher than the
fabric speed.
Retention, normally partially ooded, in the storage
chamber: In the storage chamber, there is some fabric
liquor interchange as the liquor travels over and
through the packed fabric to the suction side of the
main circulation pump, from where the dyebath is
propelled back into the nozzle. On modern low liquor
ratio machines, which allow the processing of cotton at
liquor-to-goods ratio of around 5:1, the dyebath level is
so low that only a fraction of the fabric in the storage
chamber is submerged. Additional dye transfer occurs
between the surfaces of the wet, compressed fabric
layers.
Acceleration from the storage chamber over the reel
and towards the nozzle: On its way from the storage
doi: 10.1111/j.1478-4408.2006.00040.x
2006 The Author. Journal compilation 2006 Society of Dyers and Colourists, Color. Technol., 122, 289297 289
chamber, where the speed is zero or almost zero, to the
reel, the fabric is accelerated quickly to its maximum
speed. This acceleration and the gentle squeezing effect
of the reel reduce the amount of water in the fabric
rope, which in the case of cotton, for example, is
around three times the fabric weight [15]. Both actions
contribute to a better overall fabricliquor interchange
as mechanical agitation assists dye penetration and
because more water is subsequently replenished in the
nozzle.
Based upon this general description of the fabric path in
the machine, it would be expected that the following
machine parameters could inuence the dye-exhaustion
rate.
Jet nozzle size and design
The water exiting the nozzle hits the fabric rope at an
angle, but interrupting the dyeing process after one or
two fabric circulations showed that, on the pilot-scale
equipment described in the Experimental section, there
was nevertheless only a little dye penetration of the rope.
Therefore, conditions in the jet nozzle could be described
in a rst approximation as a cylinder or, even simpler, a
sheet (the fabric rope) immersed in a ow parallel to its
axis. A suitable and well-established model describes the
mass-transfer from the solution to the bre with the help
of a diffusional boundary layer, which increases in
thickness with the length of the cylinder/sheet and which
decreases in thickness with increasing uid velocities [5].
As the thickness of the boundary layer increases, dye
transfer from the solution to the substrate reduces and,
possibly, so does the dye-exhaustion rate.
For a sheet immersed in a ow parallel to its axis, uid
velocity and average boundary layer thickness can be
correlated using Eqn 1. The term average boundary layer
thickness is used to indicate that d
D
is a function of the
plate length, k, and that d
D
is averaged over k (denitions
of the symbols used in this study, along with their
corresponding units, are shown in Appendix.
d
D
1:47
D
W
m

1=3

mk
v
fl
s
1
It is therefore conceivable that a higher liquor ow rate
in the nozzle and/or a more intensive fabricliquor
interchange, e.g. by attaching a tube to the nozzle which
increases the time of proximity between fabric and
dyebath [16], could reduce the boundary layer thickness
and increase the exhaustion rate.
Fabricliquor interchange in the storage chamber
Conditions in the storage chamber resemble a plug ow
situation as the liquor from the jet nozzle ows over and
through the fabric pile, i.e. by owing perpendicularly to
the bre axis. However, there are important deviations
from the plug ow model. For example, a signicant but
unknown portion of the liquor from the jet nozzle
bypasses the pile and ows directly to the suction side of
the main circulation pump. Furthermore, most of the
fabric in the storage chamber is not submerged, neither
on the pilot-scale machine and nor on the production
equipment, as the liquor level is so low. Nevertheless, the
plug ow model can serve to illustrate that both
microscopic depletion and macroscopic depletion could
inuence exhaustion kinetics.
First, there is an experimental evidence that mass-
transfer to the individual bre in the fabric may be
signicantly reduced compared with the individual bre
caused by microscopic depletion [8,17,18]. As a result,
dye supply may be higher than dye uptake by the bre
only in the easily accessible outer portions of the yarn.
Secondly, as a result of macroscopic depletion in the
fabric pile consisting of several fabric layers, dye
concentration continues to decrease as the dyebath ows
through the layers and as dye is taken up by each layer
so that dye supply is reduced for inner, more remote
layers of the fabric pile [1013].
The exhaustion rate in the storage chamber could thus
drop with lower dyebath ow rates as the liquid travels
more slowly over and through the piled-up fabric, leading
to a liquid diffusion-controlled process because of a
reduced dye supply. Also, the packing density of the
fabric presumably plays a role. At higher densities, the
bre surface becomes less accessible [6], but on the other
hand, higher pressure increases the dye transfer between
the wet, compressed fabric layers and the degree of
de-watering is also higher. The latter contributes to a
better fabricliquor interchange by increasing the pick-up
of fresh dyebath in the nozzle.
Fabric circulation frequency
The possible inuence of the fabric circulation frequency is
related to the fact that the fabric loses some of the adhering
water on its way from the storage chamber to the nozzle.
Experimental work on winch dyeing machines has shown
that the amount of dye on the bre surface decreases on the
way from the storage chamber to the reel [19]. The main
reason for this decrease is gravitational drainage of the
dyebath from the fabric once the fabric has left the dyebath
at the bottom of the machine [15]. As it is the amount of
water adhering to bre surface that is reduced, it is
accurate to refer to a reduction in the dye amount (g
dye kg
)1
bre), which may or may not go hand in hand
with a reduction in the dye concentration (g dye dm
)3
water). When fresh dye solution is supplied in the nozzle,
the dye amount rises again. A similar effect has been found
for continuous contacting treatment with squeezing rollers
[20]. These experiments showed that the fabricuid
interaction is limited to the moments immediately before
and after the squeezing when the dyebath is forced out of
the fabric or replenished.
This effect is qualitatively depicted in Figure 1. It
shows a decreasing bulk dye solution concentration,
arbitrarily chosen to diminish linearly with time. The dye
solution concentration at the bre surface will be
identical to the bulk concentration until the fabric is
lifted out of the dyebath if it is assumed that no liquid
diffusion effects exist and if dye amount gradients in the
substrate rope are ignored. The amount of dye adhering
to the bre surface then drops, as indicated by the grey
line, until the solution is replenished in the nozzle. At
half the fabric speed, the drop in the amount of dye
occurs only half as often, but its duration is twice as long
Ferus-Comelo Dye-uptake rate on textile jet dyeing equipment
290 2006 The Author. Journal compilation 2006 Society of Dyers and Colourists, Color. Technol., 122, 289297
and the drop is also deeper (dotted line). The lower the
fabric circulation frequency, it could therefore be argued,
the lower the average amount of dye at the bre surface
and the lower the dye-uptake rate in the case of a liquid
diffusion-controlled process.
Dyebath circulation frequency
Higher dyebath ow rates lead to higher uid velocities
in the nozzle and in the storage chamber and, therefore,
to a better dye supply in the inner layers of the fabric
rope, possibly leading to a higher exhaustion rate.
Liquor ratio
The liquor ratio is especially signicant in cellulosic
dyeing when it can strongly inuence the rate of
exhaustion [21] which tends to increase as the liquor
ratio decreases.
To the authors knowledge, the effect that dyebath and
fabric velocities of jet dyeing machinery have on
exhaustion kinetics has never been examined
systematically. It is the intention of this study to provide
some answers in this context including semiquantitative
explanations for the observed inuences.
Experimental
Dyeings were carried out on a pilot-scale jet dyeing
machine type Mathis JFO with a jet nozzle diameter of
55 mm and an average fabric load of 0.85 kg. The fabric
was stored in a freely rotating, perforated stainless steel
drum that prevented direct contact between the fabric
and the heated vessel wall. The fabric piled up in the
drum was only partially submerged by the dyebath level
at the employed liquor-to-goods ratio of 8:1. The fabric
speed could be continuously adjusted between 0 and
30 m min
)1
. Fabric circulation times were recorded with
the help of a magnetic detector installed next to the
circulation chamber. The dyebath ow rate was
adjustable between 0 and 200 dm
3
min
)1
and was
monitored by a magnetic ow meter in the circulation
line. Dyebath samples were taken during the process via
a sampling port in the circulation line.
For the experiments, a reactant xable dye, Optisal
Yellow 2RL from Clariant (CI Direct Yellow 162; 1) was
used in its commercially available form. It is a metal-free
azo dye with low migration tendency [Society of Dyers
and Colourists (SDC) class B/C]. The dye has a molecular
weight of 1237 g mol
)1
and possesses four sulphonic acid
groups and two azo bonds. The impurities, mostly
sodium sulphate, constituted 56% of the commercial dye.
Sodium chloride was used as electrolyte.
All the experiments took place at 65 C at 0.45% owf
dye concentration and with 10 g dm
)3
salt. These
conditions create a high substantivity environment,
which is most favourable for the detection of liquid
diffusion inuences on the dye-uptake rate. The fabric
was a pure cotton jersey with a dry weight of 136 g m
)1
which had previously undergone an alkaline scouring
treatment on a bulk machine in order to minimise lot-to-
lot differences. All the salt was added rst and
subsequently dye addition occurred by using either a
pressurised 0.5 dm
3
tank or a 1 dm
3
open tank which
connects to the main circulation line via a gear pump.
Times were recorded from the moment when the dye was
added. Samples were normally taken after 3, 5, 15, 30,
45, 60, 90 and 120 min. In some instances, the intervals
were changed to 3, 5, 7, 10, 15, 30, 60 and 120 min.
Dye exhaustion was determined indirectly via analysis
of the dyebath. The dyebath absorbance was measured at
410 nm on a K-Tron Uvicon 860 dual beam
spectrophotometer using cells of 10 mm pathlength and
then converted into a concentration based upon a
previously determined calibration curve. Three repeat
dyeings yielded an average variation in the exhaustion
value of <1%.
Initial tests had shown that, after the dye addition, a
small percentage of the dye remained in the tanks and
their adjacent pipes. In the calculations, these dye losses
were taken into account.
1
N N
N H
N
H
3
C
O
H
N
N
N N
Cl
N
N
N
N
Cl
N
H
N
N
N H
CH
3
O
NaO
3
S
SO
3
Na
SO
3
Na
NaO
3
S
Nozzle
passage
Drop before
nozzle at
half fabric
Drop before
nozzle
Solution in
bulk
1 fabric revolution
at half fabric speed
1 fabric
revolution
Time
D
y
e

a
m
o
u
n
t
/
d
y
e

c
o
n
c
e
n
t
r
a
t
i
o
n

Figure 1 Inuence of fabric circulation frequency on exhaustion
rate
Ferus-Comelo Dye-uptake rate on textile jet dyeing equipment
2006 The Author. Journal compilation 2006 Society of Dyers and Colourists, Color. Technol., 122, 289297 291
Results and Discussion
Simulation of production-scale jet dyeing conditions on
pilot scale
Before the pilot-scale dyeings were carried out, attempts
were made to establish the pilot-scale machine
parameters that would most appropriately scale down
those found in bulk production.
Liquor ratio
Achieving the same low liquor ratio as on production
machines is a major challenge for pilot-scale jet machine
designers [22]. The main reason for the difculty lies in
the higher proportion of dye liquor contained in the pipe-
work of the pilot-scale machine. The pipe size cannot be
scaled down proportionally because the pilot plants are
usually specied to run normal width fabric. Therefore,
the nozzle and also the attached pipe have to be of a
similar size as that of the production unit. The pilot-scale
unit was, for the fabric type and load used in the tests
and at the specied ow rates, capable of dyeing at a
minimum liquor ratio of 8:1, which is equal to or slightly
higher than the liquor ratio employed by a typical
modern cotton jet dyehouse.
Fabric circulation rate
Bulk-scale jet dyeing machines typically run at speeds
between 150 and 600 m min
)1
, resulting in a fabric rope
circulation time of 60180 s, depending on the length of
fabric rope. The 0.85 kg of fabric used on the pilot-scale
machine had a length of 6.25 m. Therefore, if the same
fabric circulation times were to be achieved on the pilot-
scale jet, a fabric speed of between 2 and 6 m min
)1
would seem appropriate.
Dyebath circulation rate
Dyebath circulation times for a modern production jet
dyeing machine vary between 60 and 120 s, although
values of up to 360 s are possible [23]. With a load of
0.85 kg on the pilot-scale jet and a liquor ratio of 8:1, the
average dyebath volume was calculated to be around
7 dm
3
. If the same dyebath circulation times were to be
achieved on pilot scale, the dyebath ow rate would have
to be between 1 and 7 dm
3
min
)1
.
Fabricliquor interchange in storage chamber
It is impossible to replicate the situation of a bulk
machine because its package density is much higher. The
fabric in the pilot-scale machine should therefore be more
accessible for the dyebath, but this effect might be offset
by a reduced dye transfer between fabric layers. The
lower compression and the much-reduced acceleration in
the case of the pilot-scale unit lead to less de-watering of
the fabric before it enters the nozzle. The overall
inuence of these different parameters is unknown and it
would appear that the interchange in the storage chamber
of the pilot-scale machine could be smaller, equal or
greater than that of a production machine.
Nozzle uid velocity and fabricliquor interchange
The pilot-scale machine came with a 55 mm diameter
circular nozzle, having a length of 75 mm and a slit
width for the liquor of around 3 mm. As the slit area of
the nozzle is known, the average liquor speed in the
nozzle can be calculated as a function of the dyebath
ow rate (Table 1).
Production jets are reported to have a liquor speed in
the nozzle between 200 m min
)1
on soft ow machines
and up to 1400 m min
)1
on pure jets [24], i.e. between
1.5 and 3 times higher than the fabric speed. The liquor
in the pilot jet nozzle would therefore have to reach a
speed of between 3 and 18 m min
)1
for fabric speeds of 2
and 6 m min
)1
, respectively.
Altogether, it would therefore appear that the pilot
machine settings listed in Table 2 appropriately scale
down production equipment conditions. There are,
however, two important differences between the pilot jet
and the production jet.
First, the jet of the pilot machine provides much less
force to propel the fabric than a production unit. In any
jet dyeing equipment, the fabric rope is transported by
the combined action of the winch reel and the jet nozzle.
How much each of them contributes to the overall force
depends on the machine design and the running
conditions. The total force required to transport the fabric
can be divided into two parts [14]: the force necessary to
lift the wet fabric from the dyeing chamber to the highest
point in the machine (F
l
) and the force necessary to
accelerate the fabric (F
a
; Eqn 2). The force exerted by the
winch reel, F
r
, can be approximated using Eqn 3.
F
R
F
l
F
a
mgh
l
mv
2
2
F
r
F
d
e
la
mgh
d
e
la
3
The force applied by the jet, F
j
, can be calculated if the
fabric rope is idealised as a pipe with a fairly rough
surface and with liquid owing around it. In this case,
the force is entirely due to friction losses at the rope
surface. Common pipe friction models, such as the one
developed by Nikuradse [14], can be used if it is further
assumed that the friction loss is the same for a uid ow
around the outside of a pipe (situation in jet nozzle) as
for a uid ow inside a pipe (normal tube; Eqn 4) [14].
F
j

v
fl
v
2
lqpd
8 2log d=k 1:138
2
4
Based on Eqns 24, the mechanics of the fabric
transport on a typical bulk machine can be compared
with the situation on the pilot-scale machine. The
calculations in Table 3 assume for both machines a wet
fabric mass, m, of 560 g m
)1
(i.e. approximately 300%
Table 1 Liquor speed in jet nozzle
Dyebath ow rate (dm
3
min
)1
) Liquor speed in nozzle (m min
)1
)
1 1.8
7 12.3
200 350
Ferus-Comelo Dye-uptake rate on textile jet dyeing equipment
292 2006 The Author. Journal compilation 2006 Society of Dyers and Colourists, Color. Technol., 122, 289297
pick-up), a rope diameter, d, of 0.04 m, a friction factor,
l, of 0.1 and a fabric roughness, k, of 0.003 m.
When the forces of winch and jet are added, the sum
is more than the total force calculated by Eqn 2 because
of inaccuracies in the underlying assumptions.
Nevertheless, the results are qualitatively instructive.
While on production machines, only around a third of
the total force required to move the fabric is contributed
by the winch, on the pilot-scale machine the winch itself
is almost entirely responsible for the fabric transport.
The much weaker force exerted by the pilot-scale jet is
mainly due to the lower uid velocity in the nozzle and
the lack of an attached tube, which could not be
installed because of space constraints point to a lower
fabricliquor interchange compared with the production
machine.
A second important difference is that the correctly
scaled-down dyebath ow rate leads to a signicantly
bigger diffusional boundary layer, d
D
, despite a dye
supply rate which is comparable with bulk conditions
when related to fabric mass, as expressed in the specic
dyebath ow rate in Table 2, on the pilot machine. The
bigger the d
D
is, the higher the likelihood that the system
is controlled by liquid diffusion, i.e. by dye diffusion
through a boundary layer. The ratio of the d
D
values for
the pilot and the bulk machine for a given dyeing system
can be related to the uid velocity, v

, using Eqn 5.
d
D;pilot
d
D;bulk
/
v
n
fl;bulk
v
n
fl;pilot
5
In Eqn 5, the exponent n is an empirically determined
coefcient that ranges normally between 0.33 and 1
[25,26]. The results for the ratio of pilot-scale to bulk d
D
values (v
,pilot
2 m min
)1
, v
,bulk
225 m min
)1
), when
the settings of Table 2 are used, were found to vary
between 5 (for n 0.33) and 113 (for n 1). It is
therefore clear that different ow regimes in pilot and
bulk machine result in signicant differences of the
boundary layer thickness and, possibly, also of the dye-
uptake rate.
Consequently, although the cycle times for both
dyebath and fabric might be appropriately scaled down,
the kinetics of the dye uptake could change from a lm
diffusion-controlled system in the bulk machine to a
liquid diffusion-controlled system in the pilot machine.
For both reasons, the reduced fabricliquor interchange
and the bigger diffusional boundary layer in the nozzle, it
might therefore be better to adjust the ow rate on the
pilot machine such that the absolute speed difference
between liquid and fabric in the jet nozzle is
approximately equal to the difference on the bulk
machine, i.e. 75 m min
)1
under low dynamic and
800 m min
)1
under high dynamic conditions. This leads,
however, to a much reduced dyebath circulation time
(Table 4).
The uid velocity in the nozzle of the pilot machine
reached a maximum of 350 m min
)1
, well below the very
Table 2 Machine settings
a,b
Machine
Fabric
circulation time
(s per revolution)
Fabric
speed
(m min
)1
)
Dyebath
circulation time
(s per revolution)
Dyebath
ow rate
(dm
3
min
)1
)
Specic dyebath
ow rate
(dm
3
kg
)1
min
)1
)
Nozzle liquor
speed (m min
)1
)
Low dynamic 180 P: 2
B: 150
360 P: 1
B: 80
P: 1.3
B: 1.3
P: 2
B: 225
High dynamic 60 P: 6
B: 400
60 P: 7
B: 440
P: 8
B: 8
P: 12
B: 1200
a P, pilot-scale; B, bulk
b Dyebath ow rates are calculated for a liquor ratio of 8:1 and a fabric weight of 136 g m
)1
Table 3 Forces on fabric rope in pilot-scale and bulk machine
Method Winch (N) Jet (N) Total (N) Winch (%)
Pilot scale
a
1.8 0.0 1.7 106
Bulk
b
9.6 25 28 34
Equation 3 4 2 3/2
a Parameter values (pilot scale): l 0.05 m, a 90, v 0.1 m s
)1
, v


0.2 m s
)1
, h
l
0.3 m, h
d
0.3 m
b Parameter values (bulk): l 0.5 m, a 180, v 6 m s
)1
, v


12 m s
)1
, h
l
1.5 m, h
d
1.5 m
Table 4 Modied pilot-scale settings for the simulation of bulk-scale jet dyeing conditions
a
Machine
Fabric
circulation
time (s per
revolution)
Fabric
speed
(m min
)1
)
Dyebath
circulation
time (s
per revolution)
Dyebath ow
rate (dm
3
min
)1
)
Specic dyebath
ow rate
(dm
3
kg
)1
min
)1
)
Nozzle liquor
speed (m min
)1
)
Low dynamic 180 P: 2
B: 150
P: 10
B: 360
P: 40
B: 80
P: 50
B: 1.3
P: 70
B: 225
High dynamic 60 P: 6
B: 400
P: 2
B: 60
P: 200
B: 440
P: 250
B: 8
P: 350
B: 1200
a P, pilot-scale; B, bulk
Ferus-Comelo Dye-uptake rate on textile jet dyeing equipment
2006 The Author. Journal compilation 2006 Society of Dyers and Colourists, Color. Technol., 122, 289297 293
high speeds possible in certain types of production units.
Nevertheless, it was possible to simulate the hydro-
dynamic characteristics of gentler bulk jets. As the fabric
speed was much less than the liquid speed in the nozzle,
it would be expected that changes in the fabric speed, e.g.
from 2 to 6 m min
)1
, would not have a signicant
inuence on the thickness of the boundary layer. It was
conceivable, however, that changes in the liquor speed in
the jet nozzle, e.g. from 70 to 350 m min
)1
, could have an
inuence on the exhaustion kinetics.
While the liquor speed in the nozzle of the pilot
machine could be adjusted to approximate the production
jet, the fabricliquor interchange in the pilot nozzle
probably remained nevertheless weaker due to the lack of
an attached tube. For the storage chamber, a comparison
between pilot and bulk is more complex. If the specic
dyebath ow rate is used to estimate the likelihood of
macroscopic depletion in the storage chamber, overall
dye supply was around six times higher on the pilot
machine (50 dm
3
kg
)1
min
)1
on pilot compared with
8 dm
3
kg
)1
min
)1
on bulk machine; Table 5). This
comparison, however, could be misleading because for
microscopic depletion, the velocity through the fabric pile
is decisive which, in contrast to the ow rate, depends on
the geometry of the storage chamber, i.e. its surface area-
to-length ratio. A long and narrow storage chamber could
thus lead to higher velocities, and a better microscopic
dye supply, than on the pilot machine, despite a lower
specic dyebath ow rate.
Effects of various parameters on dye-uptake rate
When the pilot machine settings corresponded to scaled-
down production machine settings, variations over a
fairly wide range in fabric and liquor speed did not affect
the uptake rate (Figure 2).
When the dyebath ow rate was further reduced, the
exhaustion rate started to drop (Figure 3). At ow rates
lower than approximately 30 dm
)3
min
)1
, the nozzle
action reduced visibly. Instead of providing a full, even
water jet, most of the water exited the nozzle at the
bottom and barely touched the fabric rope.
Unfortunately, the experimental error of estimating the
exhaustion value increased at low ow rates because the
main centrifugal pump of the machine became
susceptible to minor pressure variations in the circulation
line introduced by pressurised additions of dye to the
main vessel. Nevertheless, the data indicate a clear
tendency.
In order to nd out whether the reduced dye-uptake
rate was caused by a diminished fabricliquor interaction
in the nozzle, the nozzle diameter was increased from 55
to 90 mm. At the larger diameter, the fabric rope did not
touch the nozzle walls because its diameter was much
smaller so that fabricliquor interaction was reduced to a
minimum, even at ow rates of 40 dm
)3
min
)1
and more.
Surprisingly, the exhaustion rate did not decrease
0
20
40
60
80
100
0
Time, min
E
x
h
a
u
s
t
i
o
n
,

%
40 dm
3
min
1
2 m min
1
60 dm
3
min
1
6 m min
1
80 dm
3
min
1
6 m min
1
80 dm
3
min
1
9 m min
1
120 90 60 45 30 15 5 3
Figure 2 Inuence of fabric and liquor speed under scaled-down
bulk conditions; exhaustion values in percentage of equilibrium
values, legend quotes dyebath ow rate and fabric speed
0
20
40
60
80
100
0
Time, min
E
x
h
a
u
s
t
i
o
n
,

%
10 dm
3
min
1
20 dm
3
min
1
30 dm
3
min
1
40 dm
3
min
1
120 60 30 15 10 7 5 3
Figure 3 Inuence of liquor speed at low levels of dyebath ow
rate
0
0 3 5 7 10 15 30 60 120
20
40
60
80
100
Time, min
E
x
h
a
u
s
t
i
o
n
,

%

9040 dm
3
min
1
5540 dm
3
min
1
Figure 4 Inuence of nozzle diameter; legend quotes nozzle
diameter and dyebath ow rate
Table 5 Critical uid velocity
a
Parameter Symbol
b
Value Source
Diffusion coefcient D
W
1.7 10
)11
m
2
s
)1
[32]
Fabric speed m 4.4 10
)7
m
2
s
)1
[33]
Plate length k 0.05 m Own estimate
Boundary layer
thickness
d
D
3 10
)5
m [5]
Fluid velocity v
,crit
3.6 m min
)1
a The parameter k was estimated using the effective length of the jet
nozzle, i.e. 5 cm
b See Appendix
Ferus-Comelo Dye-uptake rate on textile jet dyeing equipment
294 2006 The Author. Journal compilation 2006 Society of Dyers and Colourists, Color. Technol., 122, 289297
(Figure 4). This indicated that the fabricliquor
interchange in the storage area was sufcient at specic
dyebath ow rates above 50 dm
3
kg
)1
min
)1
. The drop in
the exhaustion rate illustrated in Figure 3 would then be
mainly the consequence of a diminished dye supply in
the storage chamber.
The importance of the storage chamber for the overall
dye-uptake rate would also explain why the fabric speed,
as illustrated in Figure 2, did not inuence the exhaustion
rate. If dye supply occurred mainly in the nozzle, it would
be expected that, as illustrated in Figure 1, a higher fabric
speed also led to a higher uptake rate. If the dye supply
were essentially uniform in the entire machine, the
inuence of the fabric circulation frequency would vanish,
as was found in the experiments.
The fact that increases in the ow speed beyond
40 dm
)3
min
)1
did not increase the uptake rate implied
that any liquid diffusion inuence was eliminated.
Consequently, the exhaustion rate was lm-diffusion
limited, or, in other words, it was solely determined by
the dye demand from the bre, not by the dye supply
from the dyebath. In separate experiments, not reported
here, on a rotating beaker-type laboratory machine with
identical substrate and recipe, it was found that the
inuence of the rotation speed on the uptake rate also
disappeared beyond a minimum rotation frequency. It
would therefore be expected that both types of equipment
would yield the same exhaustion curve. Figure 5 indeed
conrms this prediction, within the margins of the
experimental error.
Qualitative mathematical analysis
For a qualitative or semiquantitative mathematical
analysis and explanation of why the exhaustion rate
dropped at lower ow rates, the varying ow conditions,
depending on the location inside the machine, require a
multitude of (sub)models. In a rst, crude analysis, it may
be sufcient to distinguish between conditions in the jet
nozzle and those found in the storage chamber.
Jet nozzle
Boundary layer effects are negligible for boundary layer
thicknesses of around 30 lm and less [5]. Rearrangement
of Eqn 2 for the mainstream velocity yields Eqn 6.
Equation 6 may be used to estimate the critical uid
velocity, v
,crit
, below which the exhaustion rate would be
expected to drop signicantly because of boundary layer
formation (Table 5).
v
fl
1:47
2
D
W
m

2=3
mk
d
2
D
6
Table 5 shows that boundary layer effects would be
expected for ow rates of less than around 3.6 m min
)1
,
less than the jet nozzle ow velocities during the
experiments (Table 4). It would therefore be likely that
boundary layer effects at the outside of the fabric rope in
the jet nozzle were insignicant. It should be noted that,
nevertheless, the inner parts of the fabric rope were
supplied with little or no dye at all in the nozzle
(macroscopic depletion).
Storage chamber
As illustrated in Figure 4, the exhaustion rate after 3 min
fell by a little more than factor 2 when the dyebath ow
rate reduced by a factor of 4, leading to the following
relationship between the uid velocity and the maximum
exhaustion rate (Eqn 7).
DExh
max
Dt
/

Q
p
7
As, according to the analysis above, the storage chamber,
and not the jet nozzle, was the location in the pilot-scale
machine where most of the fabricliquor exchange took
place, the explanation for relationship expressed in Eqn 7
should be found in this machine section.
In the pilot-scale machine, only around half of the
liquid passed over and through the fabric pile of
the storage chamber. Using the information that the
exhaustion rate in the experiments started to drop below
a ow rate, Q, of 40 dm
3
min
)1
and with a cross-section
area of the fabric pile of around 0.04 m
2
, the average
uid velocity, v
,C
, in the storage chamber was around
0.5 m min
)1
(Eqn 8).
v
fl;C

Q
2
1
A
f;C
8
A relationship of the kind expressed in Eqn 7 would be
expected if boundary layer effects according to the plane
sheet model are assumed [27]. However, as this model
assumes a ow parallel to the (bre) axis and laminar
ow, both of which assumptions are probably invalid in
the storage chamber, the agreement is probably merely
coincidental. For ow perpendicular to the bre axis, on
the other hand, boundary layer thickness varies very little
with uid velocity [9].
A more suitable model to quantify the inuence of the
liquor ow rate explains the reduced uptake rate by a
dye supply limitation to the individual bre, i.e.
microscopic depletion. Theoretical considerations [9,28]
and experiments [29] lead to the conclusion that mass-
transfer from the dyebath to the bre surface is related to
the ow rate according to Eqn 9, when the exponent n
varies between 0.3 and 0.5.
Laboratory
Pilot-scale
0
0 3 5 7 10 15 30 60 120
20
40
60
80
100
Time, min
E
x
h
a
u
s
t
i
o
n
,

%
Figure 5 Inuence of equipment type (laboratory or pilot scale)
on the exhaustion
Ferus-Comelo Dye-uptake rate on textile jet dyeing equipment
2006 The Author. Journal compilation 2006 Society of Dyers and Colourists, Color. Technol., 122, 289297 295
DExh
max
Dt
/ Q
n
9
As model assumptions are fairly appropriate for the
situation in the storage chamber (ow vertical to the bre
axis, no requirement of laminar ow), agreement between
Eqns 7 and 9 lends some support to the conclusion that a
reduced mass-transfer coefcient to the individual bre,
or microscopic depletion, played an important role.
A second effect, macroscopic depletion, i.e. dye
depletion because of dye uptake by subsequent layers of
fabric, may additionally have contributed to the observed
drop of the exhaustion rate. If it is assumed that diffusion
in the fabric pile is negligible compared with convective
transport, and if it is further assumed that convective
transport is the rate-limiting step during dye uptake, then
it can be shown that mass-transfer to the fabric pile is
proportional to the ow velocity [11]. Although,
according to Eqn 7, in the present experiments mass-
transfer was proportional to the square root of the ow
velocity, dye depletion effects almost certainly played a
part, as stopping the process after the rst one or two
circulations revealed that the inner sections of the fabric
rope were still largely free of dye.
Bulk-scale conditions
The experiments on the pilot-scale machine indicated
that the jet nozzle did not contribute signicantly to the
overall fabricliquor interchange and therefore it did not
inuence the dye-uptake rate, either. This situation is
probably different on a bulk-scale machine because
higher liquid speeds in the nozzle and an attached tube
ensure better dye supply to the rope interior. Separate
tests on bulk jet dyeing machines, carried out in a
different context and not reported here, indeed suggest
that the nozzle on a bulk machine contributes noticeably
to the fabricliquor interchange.
In the storage chamber, much depends on the ow rate
and ow velocity in the fabric pile as they determine the
dye supply rate, and thus the mass-transfer rate to the
bres. Based on the ow rate and storage chamber
dimensions of a standard bulk machine, the author
calculated a ow velocity of around 1.6 m min
)1
[30]. As
liquid diffusion effects were observed on the pilot
machine below 0.5 m min
)1
, it follows that on bulk
machines of standard design microscopic depletion
probably does not exist in the storage chamber, unless
accessibility to the bre surface is dramatically reduced
because of fabric construction or compression. Due to the
much longer pathway across the pile on the bulk machine
and due to the lower ow rate per kilogram of fabric, on
the other hand, the risk of macroscopic depletion seems to
be signicantly higher on big machines.
It has to be additionally considered that processes in
industry usually have a much lower uptake rate than the
processes used here because the latter are controlled by
temperature and/or by gradual salt addition, thereby
reducing the possible impact of liquid diffusion. The
present investigation therefore does not permit a
conclusion whether or not liquid diffusion effects play a
role under industry typical dye-uptake rates. The author
is not aware of any published investigation into the
inuence of fabric and liquor speed on the exhaustion
rate of production jets, but practical experience in
industry [8] and model calculations [10,11,31] indicate
that such an inuence is negligible as long as the
exhaustion rate is controlled.
Conclusions
Liquor ratio, dyebath and fabric circulation times as well
as fabricliquor interchange conditions in the jet nozzle
and in the storage chamber are parameters that
characterise a jet dyeing machine in terms of exhaustion
behaviour. The parameters on the pilot-scale jet could be
adjusted to match three of the ve parameters of bulk
machines. These included the liquor ratio, the fabric
circulation frequency and one of the following two:
dyebath circulation frequency or nozzle uid speed. If the
dyebath circulation time was adjusted to match bulk
conditions, then the uid velocity in the pilot jet was
much smaller. On the other hand, if the uid velocity
was increased, then the dyebath circulation time was
shorter than on industrial machines. From a machine
design point of view, the situation could be improved if
nozzles on the pilot-scale machine were designed with a
smaller slit width, yielding higher velocities at the same
dyebath circulation rate. Even at matching uid
velocities, fabricliquor interchange in the nozzle of the
pilot machine remained weaker than on a bulk machine.
Conditions in the storage chamber were very different,
but competing inuences make a prediction of a
benecial or adverse impact on exhaustion kinetics
difcult. It is thought that, however, when bulk-like
nozzle uid velocities were employed, macroscopic dye
supply in the storage chamber of the pilot machine was
better, due to a higher average specic ow rate, but
microscopic dye supply might have been worse because
of a lower ow velocity.
Dyeings carried out under high substantivity
conditions on the pilot-scale machine showed that the
exhaustion rate was unaffected by variations in fabric
and dyebath ow speeds over a fairly wide range,
matching the exhaustion rate of a laboratory beaker-
type machine. As the exhaustion rate was found, under
scaled-down industrial settings, to be independent of
fabric and liquor speeds, the dye supply at the bre
surface must have been at least equal to the amount
taken up by the bre at any moment in the entire
machine. If it had been different, i.e. if the dye supply
had been, for example, reduced in the storage chamber,
the fabric speed would have had an inuence on the
overall exhaustion rate because the more frequently the
fabric would have passed through the nozzle, the more
often the dye would have been replenished. Therefore,
interestingly, in the pilot machine, the dye supply to
the bre surface in the storage chamber alone, i.e.
without jet nozzle, seems to have been sufcient and,
as a result, the exhaustion rate was found to be
independent of the fabric circulation rate. In other
words, the jet nozzle on the pilot-scale machine did not
contribute signicantly to the overall fabricliquor
interchange.
Ferus-Comelo Dye-uptake rate on textile jet dyeing equipment
296 2006 The Author. Journal compilation 2006 Society of Dyers and Colourists, Color. Technol., 122, 289297
The fact that the fabric rate did not appear to play a
role in the exhaustion kinetics marks a difference to the
older winch dyeing machine. The difference was most
likely due to a better fabricliquor interchange in the
storage chamber of the jet machine because the dyebath
was owing more rapidly over and through the stacked
fabric. On a winch machine, on the other hand, a slow
moving fabric rested in a non-circulating bath, thus
creating a fairly static situation.
When the specic dyebath ow rate dropped to below
50 dm
3
kg
)1
min
)1
, the exhaustion rate diminished. This
was explained by liquid diffusion effects in the storage
chamber, which became apparent when the uid velocity
was less than around 0.5 m min
)1
.
For future research, it would be clearly benecial if
better hydrodynamic models were available to describe
the liquor ow and fabricliquor interchange in the
nozzle and the storage chamber of a jet dyeing machine.
Additionally, it would be interesting, not the least from
an industrial application point of view, to analyse the
inuence of parameters examined here on dye
unlevelness of the jet-dyed fabric.
Acknowledgements
The author would like to thank Dr J H Nobbs and Prof J
Carbonell and the referees, especially one of them who
made very useful suggestions for a more balanced data
interpretation, for their contributions, and the Worshipful
Company of Clothworkers as well as the Worshipful
Company of Dyers, both in London, for their help in
nancing the pilot-scale jet dyeing machine.
References
1. N Armeld, J.S.D.C., 73 (1947) 381.
2. P Alexander and R F Hudson, Text. Res. J., 20 (1950) 481.
3. T Vickerstaff, The Physical Chemistry of Dyeing, 2nd Edn
(London: Oliver & Boyd, 1954) 144.
4. N Etters and S J English, Text. Chem. Colorist, 20 (1988) 21.
5. R McGregor and R H Peters, J.S.D.C., 81 (1965) 393.
6. R McGregor and R H Peters, J.S.D.C., 81 (1965) 429.
7. J H Brooks, J.S.D.C., 88 (1972) 184.
8. F Hoffmann, Textilveredlung, 70 (1989) 381.
9. A Kretschmer, Textil-Praxis Int., 45 (1990) 265.
10. J Boulton and J Crank, J.S.D.C., 68 (1952) 109.
11. F Hoffmann and P-F Mueller, J.S.D.C., 95 (1979) 178.
12. R W Burley, J R Flower and I D Rattee, J.S.D.C., 85 (1969)
187, 193.
13. R W Burley, J R Flower and I D Rattee, J.S.D.C., 87 (1971) 278.
14. B Boehnke, Melliand Textilber., 79 (1998) 346.
15. M R Fox, J.S.D.C., 84 (1968) 401.
16. J Vernazza, Textilveredlung, 9 (1974) 151.
17. R McGregor, A E Nounou and R H Peters, J.S.D.C., 90
(1974) 246.
18. H Sasaki, E Yanai and H Arakai, Text. Res. J., 63 (1993) 614.
19. D H Wyles, in Engineering in Textile Coloration, Ed. C
Duckworth (Bradford: SDC, 1983) 1.
20. R Burley, Text. Manuf., 96 (1970) 332.
21. M M Breuer and I D Rattee, The Physical Chemistry of Dye
Adsorption (London: Acedemic Press, 1974) 61.
22. S S Smith, in Engineering in Textile Coloration, Ed. C
Duckworth (Bradford: SDC, 1983) 443.
23. H P Rouette, Lexikon der Textilveredlung, Tome II (Dulmen:
Laumann Verlag, 1995) 922.
24. J Cegarra, P Puente and J Valldeperas, The Dyeing of Textile
Materials (Biella: Texilia, 1992) 215.
25. V G Levich, Physicochemical Hydrodynamics (Englewood
Cliffs, NJ: Prentice Hall, 1962).
26. P Alexander, D Gough and R F Hudson, Trans. Faraday
Soc., 45 (1949) 1058.
27. R McGregor, R H Peter and K Varol, J.S.D.C., 86 (1970) 442.
28. F Y Telegin, J.S.D.C., 114 (1998) 49.
29. R H Peters, Textile Chemistry, Vol. 3 (Oxford: Elsevier,
1975) 771.
30. F Kirchner, personal communication (July 2006).
31. M Vosoughi, PhD Thesis (Edinburgh: Heriot-Watt
University, 1993).
32. M K Inglesby and S H Zeronian, Dyes Pigm., 50 (2001) 3.
33. R C Weast, M J Astle and W H Beyer, Handbook of
Chemistry and Physics (Boca Raton, FL: CRC, 1988).
Appendix Denition of symbols used in this study
a Angle described by fabric as it passes around the winch
d
D
Boundary layer thickness (m)
k Plate length (m)
l Friction factor
q Density uid (kg m
)3
)
m Kinematic viscosity of water (m
2
s
)1
)
A
f,C
Cross-section fabric pile in storage chamber (m
2
)
d Diameter fabric rope (m)
D
W
Diffusion coefcient of the dye in water (m
2
s
)1
)
D(Exh)
max
/Dt Maximum exhaustion rate (% min
)1
)
F
a
Force necessary to accelerate the fabric (N)
F
d
Force of wet fabric at exit of winch (N)
F
l
Force necessary to lift the wet fabric from the dyeing chamber to the highest point in the machine (N)
F
j
Force applied by the jet (N)
F
r
Force exerted by the winch reel (N)
g Gravitational constant (m s
)2
)
h
d
Dropping height of fabric from winch to storage chamber (m)
h
l
Lifting height of fabric (m)
k Roughness rope surface (m)
l Effective length of nozzle (m)
m Mass of wet fabric (kg m
)1
)
Q Flow rate (m
3
s
)1
)
v Fabric speed (m s
)1
)
v

Fluid velocity (m s
)1
)
v
,C
Fluid velocity in storage chamber (m s
)1
)
Ferus-Comelo Dye-uptake rate on textile jet dyeing equipment
2006 The Author. Journal compilation 2006 Society of Dyers and Colourists, Color. Technol., 122, 289297 297

Potrebbero piacerti anche