Dep Kraan 20050912
Dep Kraan 20050912
Dep Kraan 20050912
Proefschrift
door
Samenstelling promotiecommissie:
Dr. ir. G.F. Woerlee heeft als begeleider in belangrijke mate aan de totstandkoming
van dit proefschrift bijgedragen.
Dit werk is financieel ondersteund door de ministeries van EZ, OCenW en VROM
via het EET programma van SenterNovem, project EET K99101.
ISBN-10: 9090199128
ISBN-13: 9789090199122
Printed by Febodruk B.V., Enschede
All rights reserved. No part of the material protected by this copyright notice may be reproduced or utilised in any
form or by any means, electronic or mechanical, including photo copy, recording or by any information storage
and retrieval system, without prior permission from the publisher.
Voor mijn ouders
en voor Uschi
Summary
Silk, wool, nylon and polyester were dyed successfully in scCO2 with two non-
polar dyes containing vinylsulphone or dichlorotriazine reactive groups. It was
found that the presence of a small amount of water increased the coloration of wool
and nylon significantly and that silk only reacted with the dyes when water was
added. This was the first time that silk was dyed successfully in scCO2. Maximum
color depths were observed when both the scCO2 and the textiles were saturated
with water.
Because no generally accepted procedure exists for the design of heat exchangers
for supercritical fluids, this was investigated with computational fluid dynamics
simulations. Heat transfer coefficients were determined for different temperatures,
pressures and Reynolds numbers, for CO2 flowing up or down through a vertical
pipe.
A technical-scale, 100-litre dyeing machine was designed and built to test polyester
beam dyeing in scCO2 at 300 bar and 120ºC. A new type of pressure vessel was
used, consisting of a steel liner with carbon fibers wound around to take up the
radial forces and a yoke construction for the axial forces. This configuration lowers
the investment cost but also the operating cost, because the amount of steam
required to heat the vessel is lower than for a completely steel vessel. Furthermore,
because the carbon fiber vessel requires less heating due to the low heat capacity of
the carbon fibers, the process time is shortened. To circulate the CO2 with the
dissolved dye through the textile, a low-pressure centrifugal pump was designed
for service in scCO2 and placed inside the dyeing vessel.
Also a commercial-scale 1000-liter supercritical dyeing machine was designed, for
treating 300 kg polyester while recycling all dye and 96% of the CO2. An
economical analysis showed that, although the purchase cost for a supercritical
machine is higher (500 k€) than for an aqueous machine (100 k€), the operating
cost is lower (0.35 instead of 0.99 € per kg polyester). This is caused by the higher
rate of dyeing and by the simpler dye formulations that can be used in scCO2. The
overall result is a 50% lower process cost for the supercritical process.
Zijde, wol, nylon en polyester zijn met goed resultaat in scCO2 geverfd met twee
apolaire kleurstoffen met de reactieve groep vinylsulfon respectievelijk
dichloortriazine. Er werd vastgesteld dat de aanwezigheid van een kleine
hoeveelheid water de aankleuring van wol en nylon significant deed toenemen en
dat zijde alleen reageerde met de kleurstof wanneer water werd toegevoegd. Dit
was de eerste keer dat zijde in scCO2 is geverfd met een goed resultaat. Maximale
kleurdiepten werden gevonden wanneer zowel het scCO2 als het textiel waren
verzadigd met water.
Het niet-reactief verven van polyester werd experimenteel onderzocht in een nieuw
40-liter drukvat. De verzadigingsaankleuring nam toe en de verdelingscoëfficiënt
nam af met de temperatuur en dichtheid van de scCO2. De adsorptie van kleurstof
aan het polyester was exotherm en volgde Nernst adsorptie. Aangezien de waarden
van de verzadigingsaankleuringen, de temperatuurafhankelijkheid van de
verdelingscoëfficiënt en ook het adsorptiemechanisme hetzelfde waren als in
waterverven, werd geconcludeerd dat dispers verven in scCO2 hetzelfde
thermodynamische gedrag vertoont als in water.
Omdat geen algemeen geaccepteerde procedure bestond voor het ontwerp van
warmtewisselaars voor superkritische fluïda, werd dit onderzocht met
“computational fluid dynamics” simulaties. Warmteoverdrachtscoëfficiënten
werden bepaald voor verschillende temperaturen, drukken en Reynoldsgetallen,
voor omhoog- en omlaagstromende CO2 in een verticale pijp.
1 Introduction 1
Epilogue 131
Dankwoord 133
Publications 137
Chapter 1
Introduction
The importance of the textile processing industry is clearly illustrated by the world
production in 2003, being 52 Mton. The two main fiber types are polyester (22
Mton) and cotton (20 Mton), the remainder is accounted for by wool, silk, nylon,
viscose and some specialty fibers [1].
Virtually all of this textile is dyed, using worldwide 0.7 Mton of dye powder, with
a total value of 19 billion € [2]. Approximately 5 to 10 % of the non-reactive dyes
are lost with the wastewater, for reactive dyes this is even 50 %. The resulting
pollution lies between 40 and 300 liter per kg textile [3], giving a total amount of 2
to 20 billion cubic meters of wastewater annually, which indicates the seriousness
of this environmental problem.
Although other chemicals, such as salts and dispersing agents, are used in the
dyeing process, the most problematic pollutant is the dye itself. Inherent to their
purpose, dye molecules are designed to be resistant to degradation by light, water
and many chemicals [4]. Therefore, treatment in municipal water purification
facilities can not decolorize dye house effluents. Instead, the wastewater has to be
treated at the site.
1
production of toxic sludge, which has to be disposed of or incinerated. All other
techniques to purify the water are characterized by either a high process cost or a
production of toxic waste. Since the treated water has to satisfy ever more stringent
environmental regulations, the end-of-pipe solutions become increasingly
expensive for the textile industry. It is therefore that a solution at the source of the
problem, i.e. replacement of water as a dyeing solvent, is preferable.
In the sixties and seventies, this consideration has led to some research on the use
of chlorinated hydrocarbons, mainly perchloroethylene, as dyeing media [6]. This
so-called solvent dyeing has, however, a major disadvantage in the toxicity of the
solvents. The technique has not been applied in industry and all research on the
subject has been stopped.
When a solid is heated, the thermal motion of the molecules increases, the solid
melts and a liquid and a vapor phase are formed. In figure 1.1 these three states of
matter are graphically presented. When a vapor below its critical temperature is
compressed, it condenses when the vapor-liquid equilibrium line is crossed. Above
the critical temperature however, the thermal energy of the vapor molecules is so
high, that condensation is no longer possible, no matter how much the pressure is
increased. The vapor-liquid equilibrium line ends at the critical temperature. When
a fluid is above its critical temperature and the corresponding critical pressure, it
can not be regarded as a vapor or a liquid and it is referred to as a supercritical
fluid.
The most widely used supercritical fluid is carbon dioxide, because it combines a
relatively mild critical point with non-flammability, non-toxicity and a low price.
Because of its green and safe character, it is the best supercritical solvent for textile
2
p (bar)
supercritical
fluid
solid liquid
74 bar
critical point
vapor
triple point
31ºC T (ºC)
To allow solubilization of low vapor pressure compounds such as the solid dyes,
both the density and the temperature of CO2 need to be sufficiently high, in the
order of 600 kg/m3 and 120ºC respectively. To obtain this set of conditions, the
pressure needs to be around 300 bar, so that both the temperature and pressure are
far above the critical point (31ºC, 74 bar).
The main advantage of using scCO2 instead of water in a dyeing process, is the
easy separation of the CO2 and the unused dye that remains after the dyeing
process. Depressurization leads to precipitation of the excess dye and gives clean,
gaseous CO2, so that both compounds can be recycled and no waste is generated.
Furthermore, after the dyeing process, the textile does not need an energy-intensive
drying step as it does in aqueous dyeing.
The above advantages are consequences of the high vapor pressure of CO2, but also
the physical properties that are characteristic of all supercritical fluids facilitate the
dyeing process: The intense thermal motion of the molecules and the low viscosity
3
result in a high diffusivity in the fluid, facilitating the mass transport of the dye
towards the fibers.
Additional advantages of scCO2 exist for the case of polyester dyeing. Because
scCO2 is a non polar solvent, no dispersing agents are needed when non polar dyes
are used. This means that simpler dye formulations can be used than in aqueous
polyester dyeing where dispersing agent makes up around 50% of the dye powder.
Another advantage specifically for polyester is that under supercritical conditions
the CO2-molecules penetrate and swell the polymer. This plasticizes the textile
fibers and increases diffusion coefficients of dyes inside the polyester with one
order of magnitude [9], relative to aqueous dyeing.
1.4.1. Process
In supercritical textile dyeing, the choice of dye depends on the dyeing mechanism
which, in its turn, depends on the type of fiber that is to be dyed. For non polar
textiles, such as polyester, non polar, non reactive dyes are dissolved into the
supercritical phase, transported to the fiber and adsorbed to the surface. Finally, the
dye molecules diffuse into the CO2-swollen polymer matrix, where they are bound
to the polyester molecules by physical attraction, mainly dispersion forces. Upon
depressurization, the CO2-molecules exit the shrinking fibers and the dye
molecules are retained. This dyeing mechanism was suggested by Saus et al. [10]
and subsequently confirmed by Tabata et al. [11]. Because these non polar dyes are
also used in aqueous dyeing processes, where they are dispersed in the water, they
are generally called disperse dyes.
The dyeing mechanism and therefore the type of dye is different for natural
textiles. Since these are polar due to the presence of hydroxyl groups (cotton) or
amino groups (wool, silk), they have no physical attraction towards disperse dyes.
To bind a dye molecule to such a fiber, a chemical bond is needed to realize
sufficient fastness, for which reason reactive dyes have to be used.
4
As is the case for polyester, for natural textiles the currently available dyes that are
used in water can be used in scCO2. However, these reactive dyes are polar and
therefore not soluble in scCO2. It has been shown by Jun et al. [12] that it is
possible to dye natural fibers with these polar dyes in scCO2 by using reverse
micelles. This method is not suited for dyeing cotton/polyester blends since the
latter fiber is not capable of forming chemical or physical bonds with the polar
dyes. Another disadvantage is that the use of extra chemicals like dispersing agents
complicates the process and therefore it is preferable to use new, non polar reactive
dyes for natural fibers.
Non polar reactive dyes were used on cotton by Schmidt et al. [13] but the results
were not satisfactory, and an unpractical long dyeing time of 4 hours was needed
and during the process, corrosive byproducts were formed. Other researchers have
tried to dye cotton in scCO2 but did not succeed [7]. The successful development of
a process for the dyeing of cotton in scCO2 was carried out by Fernandez [14].
Since this falls beyond the scope of the present work, it will not be discussed
further. Wool was dyed well in scCO2 with non polar reactive dyes by Schmidt et
al. [15], but silk has never been dyed to an acceptable depth [7].
Despite the research on disperse dyeing of non-polar fibers and the reactive dyeing
of polar fibers, there is by far not enough knowledge in both fields, which currently
limits the further development of supercritical fluid dyeing.
1.4.2. Equipment
Pilot plants such as drawn in figure 1.2 have been constructed [16, 17] but, up to
now, no commercial-size machine has been built. This is caused by the high cost
5
associated with large-scale equipment operating around 300 bar. To facilitate
commercialization, the price of large-scale pressure equipment has to be lowered.
To do this, there are some important engineering issues to be faced.
dye
dyeing holder
vessel
DYEING SOLVENT
CYCLE RECYCLE CO2
storage
vessel
The first issue concerns the pressure vessel containing the textile. Since industrial
dye baths have volumes in the order of 1000 liters, the thick-walled supercritical
dyeing vessels will consist of a large amount of steel compared to aqueous dyeing
vessels. This does not only increase the investment costs, but also the operating
costs because the vessel has to be heated to the dyeing temperature in each batch.
Because more energy is needed for the heating, a larger heat exchanger or a longer
process time is required, which increases either the investment or the operating cost
even more.
The second engineering issue is the circulation of scCO2 through the dyeing vessel.
Centrifugal pumps for static pressures of around 300 bar are expensive and weigh
heavily on the equipment cost [16].
6
calculate heat transfer coefficients for the design of a heat exchanger for high-
pressure CO2.
The above three problems have not been solved up to now, which hinders the
implementation of supercritical fluid dyeing in industry.
This work treats the above discussed limiting aspects on both the process (non-
reactive and reactive dyeing) and on the equipment (pressure vessel, circulation
pump and heat transfer).
As was discussed above, the dyeing of natural fibers in scCO2 is a research area
where only a start has been made. Therefore, in chapter 2 experiments are
discussed with new, reactive, non polar dyes on several textile types.
Since it was found in chapter 2 that the presence of a small amount of water in the
scCO2 increases the coloration in reactive dyeing, a model was developed which is
described in chapter 3, to predict how much water is to be added to a supercritical
dyeing vessel to obtain optimum coloration.
7
In chapter 6, the process and equipment design of a 100-liter beam dyeing machine
for supercritical polyester dyeing is treated. A new type of pressure vessel and
centrifugal pump are described, both equipment items are meant to decrease the
process cost significantly.
In the epilogue, recommendations are made for further research and the future of
the technology of textile dyeing in supercritical carbon dioxide is discussed.
References
8
9. S. Sicardi, L. Manna and M. Banchero, Comparison of Dye Diffusion
in Poly(ethylene terephtalate) Films in the Presence of a Supercritical
or Aqueous Solvent, Ind. Eng. Chem. Res. 39, 4707 (2000).
10. W. Saus, D. Knittel and E. Schollmeyer, Dyeing of Textiles in
Supercritical Carbon Dioxide, Text. Res. J. 63, 135 (1993).
11. I. Tabata, J. Lyu, S. Cho, T. Tominaga, and T. Hori, Relationship
Between the Solubility of Disperse Dyes and the Equilibrium Dye
Adsorption in Supercritical Fluid Dyeing, Color. Technol. 117, 346
(2001).
12. J.H. Jun, K. Sawada and M. Ueda, Application of Perfluoropolyether
Reverse Micelles in Supercritical CO2 on Dyeing Process, Dyes and
Pigments 61, 17 (2004).
13. A. Schmidt, E. Bach and E. Schollmeyer, in: E. Reverchon (Ed.), Proc.
6th Conf. Supercrit. Fluids Appl., Maiori, Italy, 557 (2001).
14. M.V. Fernandez Cid, Cotton Dyeing in Supercritical Carbon Dioxide,
Delft University of Technology, Dissertation, Delft (2005).
15. A. Schmidt, E. Bach and E. Schollmeyer, The Dyeing of Natural
Fibers with Reactive Disperse Dyes in Supercritical Carbon dioxide,
Dyes and Pigments 56 (1), 27 (2003).
16. W.A. Hendrix, Progress in Supercritical CO2 Dyeing, J. Ind. Text. 31
(1), 43 (2001).
17. E. Bach, E. Cleve, E. Schollmeyer, M. Bork P. and Koerner,
Experience with the Uhde CO2-Dyeing Plant on Technical Scale. Part
1: Optimization Steps of the Pilot Plant and First Dyeing Results,
Melliand Int. 3, 192 (1998).
9
10
Chapter 2
Abstract
Polyester, nylon, silk and wool were dyed with disperse reactive dyes in
supercritical carbon dioxide (scCO2). The dyes were substituted with either
vinylsulphone or dichlorotriazine reactive groups. Since earlier research showed
that water, distributed over the scCO2 and the textile, increased the coloration,
experiments were done with the vinylsulphone dye with varying amounts of water
in the dyeing vessel, to investigate if there is an optimum water concentration. The
amounts were such, that no liquid water was present. The maximum coloration was
obtained when both the scCO2 and the textiles were saturated with water. At the
saturation point, deep colors were obtained with the vinylsulphone dye for
polyester, nylon, silk and wool, with fixation percentages between 70 and 92%
when the dyeing time was 2 hours. The positive effect of water was due to its
ability to swell fibers or due to an effect of water on the reactivity of the dye-fiber
system. Also the dichlorotriazine dye showed more coloration when the scCO2 was
moist. With this dye, experiments were conducted in water-saturated scCO2,
varying the pressure from 225 to 278 bar and the temperature from 100 to 116ºC.
The coloration of polyester increased with pressure, the results for silk and wool
were not sensitive to pressure. Increasing the temperature had no influence on the
dyeing of polyester, silk and wool. The fixations on polyester, silk and wool, being
between 71 and 97%, were also independent of pressure and temperature.
11
2.1. Introduction
Textiles can be classified into synthetics (e.g. polyester and nylon) and natural
textiles. The latter category can be divided into proteins (e.g. silk and wool) and
cellulosics (e.g. cotton). In this work, polyester, nylon, silk and wool are
investigated. Cotton is also under investigation in our laboratory but will not be
discussed in this work.
When dyeing in scCO2, non-polar dyes are used to enable dissolution. Polyester
(polyethylene terephtalate or PET) is also non-polar and during the dyeing process,
the dye molecules can diffuse into the polymer matrix, where they are physically
bonded. Because of its non-polarity, polyester can be dyed in scCO2 with non-
reactive, so-called disperse dyes. Nylon is non-polar as well, although slightly
more polar than PET, and it can also be dyed with disperse dyes. Since the nylon
molecules have amino end-groups, it is also possible to use dyes that are able to
react with these nucleophiles, forming a covalent bond, as was shown by Liao et al.
[4]. Silk and wool are polar and therefore have no affinity for the non-polar dye
molecules. It is only possible to dye these textiles in scCO2 when the dyes are
reactive towards the amino groups in the protein fibers of silk and wool. These
non-polar reactive dyes are generally called disperse reactive.
In the disperse dyeing of synthetic textiles, the CO2 penetrates and swells the
fibers, thereby facilitating the diffusion of dye molecules through the polymer.
Upon depressurization, the CO2 molecules exit the shrinking fiber and the dye is
12
retained in the textile. The mechanism of PET dyeing in scCO2 has been discussed
extensively, e.g. by Saus et al. [2] and Tabata et al. [3].
In reactive dyeing, four steps can be distinguished in the process, each of which is
potentially rate determining:
1. Transport of dye to the fibers. The dye powder is dissolved in the scCO2
and transported towards the textile. The solubility of the dye determines
the transport rate of dye towards the textile.
2. Diffusion of dye into the fibers. Natural fibers are porous and their
accessibility needs to be sufficient to allow diffusion of dye molecules into
the pores.
3. Adsorption of dye to the textile. Before a dye molecule can react with a
fiber, it needs to be adsorbed to its surface. The affinity or substantivity of
the dye for the textile determines whether or not this step takes place.
4. Reaction of the dye with the textile. The dye has to form a covalent bond
with the amino groups of the proteins.
It is known from Schmidt et al. [7] that a vinylsulphone dye is able to react with the
functional amino groups of nylon and wool in scCO2 but they found no coloration
of silk. Schmidt et al. [8] also found that a dichlorotriazine dye was unable to dye
silk; they did not report on dyeing of wool with a dichlorotriazine dye. However,
these publications describe dyeing in dry CO2, while the patent by Veugelers et al.
[9] claims that the addition of water enhances coloration when natural fibers are
dyed with reactive dyes in scCO2.
13
SO2 HC CH2
Et2N N N
textile NH2
A. Vinylsulphone reaction
Cl N Cl textile HN N Cl
N N N N
HN textile NH2 HN
O2 N N N NEt2 O2N N N NEt2
- HCl
B. Dichlorotriazine reaction
Fig. 2.1. Reaction of vinylsulphone (A) and dichlorotriazine (B) dyes with textile
amino sites
It is not yet known why water has this positive effect on the process. One possible
explanation is that it acts as a modifier, enhancing the solubility of the dye. It could
also be that water works as a swelling agent for the polar, natural textiles. As
explained by Kraessig [10], the macromolecules of a natural fiber are kept together
by intermolecular hydrogen bonds. Water breaks up these interactions and allows
the chain molecules to increase their distance, thereby swelling the textile fiber.
Dye molecules can then penetrate more easily into the fibrous structure. It is
expected that nylon, silk and wool are susceptible to this effect of water and
polyester is not. The third possible effect of water on the supercritical dyeing
process is that it somehow participates in the reaction between the dye and the
fiber. There are, however, no reports on this last possibility.
14
to obtain maximum coloration. Experiments are done with different dyeing times,
to get an indication of the rate of the process. With the dichlorotriazine dye, the
influence of water addition is also investigated. Furthermore, the effects of pressure
and temperature on the coloration of different textiles in moist scCO2 are
investigated with the dichlorotriazine dye. Finally, by dyeing polyester and nylon
in the same batch as silk and wool it is checked if also blends of synthetics and
natural textiles can be dyed in scCO2 with both dyes.
2.2.1 Materials
The orange vinylsulphone dye and the purple dichlorotriazine dye, shown in
Fig.1A and 1B were designed for use in scCO2 and synthesized by a well-known
dye manufacturer. Both powders were used as received. The textiles were all dyed
as pieces of cloth. The polyester (polyethylene terephtalate; 120 g/m2) and the
nylon (polyamide 6.6; 150 g/m2) were knitted and free of spinning oils. The wool
(62 g/m2) and the silk (120 g/m2) were woven. Polyester was received from Ames
Europe, the other textiles were from the Center for Test Materials (Netherlands).
The carbon dioxide (99.97 %) was purchased from HoekLoos. The acetone that
was used in the extractions was technical grade from Chemproha. Demineralised
water was used, in the extractions and in the dyeing vessel.
15
textile
oil heater
autoclave with
heating jacket
CO2 supply
pump oil pump
circulation pump
dye powder
In the experiments in moist scCO2, the cotton was wetted with the desired amount
of water prior to placing it in the vessel. The system was then pressurized with an
air-driven plunger pump, from Williams Instrument Company. The CO2 was
pumped around for 2, 4 or 6 hours, at a flow rate of 0.10 ± 0.02 m3/hr, with a
centrifugal pump with magnetic coupling from Autoclave Engineers. The flow
direction was upwards through the vessel: first through the dye and then through
the textiles. Temperature and pressure increased slowly in the first hour but were
constant afterwards (± 1ºC and ± 2 bar). The stable values in the second hour are
the reaction conditions that are mentioned in the presentation of the results below.
In the experiments of 2 and 4 hours, there was dye powder left between the filter
plates. The amount was different for each experiment and varied from 0.3 to 0.1 g.
In the 6 hour experiments, no remaining dye was found.
The dyed pieces of polyester, nylon, silk and wool were analyzed by measuring the
reflectance curve between 350 and 750 nm with a portable spectrophotometer from
Avantes. The minimum of the curve (Rmin) was used to determine the ratio of light
absorption (K) and scatter (S) via the Kubelka-Munk function [10]:
(1 − R min ) 2
(K/S )dyed = (1)
2R min
16
The ratio K/S is used as a measure of coloration in textile industry and also in this
paper. After this analysis, each sample was stripped of unfixed dye by Soxhlet
extraction with a 50 weight% solution of acetone in water for 30 min. The K/S-
value of the extracted textile (K/S)extr was determined and used to calculate the
percentage of dye molecules that was fixed to the textile (F):
17
100 100
80 fixation 80
60 60
F (%)
K/S
A. Polyester
40 40
(K/S)dyed
20 20
(K/S)
0 0
0 20 40 60 80 100
RH (%)
100 100
80 fixation 80
(K/S)
60 60
F (%)
B. Nylon K/S
40 (K/S)extr 40
20 20
0 0
0 20 40 60 80 100
RH (%)
100 100
80 80
fixation
60 60
F (%)
K/S
C. Silk
40 (K/S) 40
20 (K/S) 20
0 0
0 20 40 60 80 100
RH (%)
100 100
fixation
80 80
60 60
F (%)
K/S
D. Wool
(K/S)dyed
40 40
20 (K/S) 20
0 0
0 20 40 60 80 100
RH (%)
Fig. 2.3. Influence of relative CO2-humidity (RH) on the coloration (K/S) and
fixation (F) of textiles dyed in scCO2 with a vinulsulphone dye during 2 hours at
230 bar and 112ºC, for polyester (A), nylon (B), silk (C) and wool (D)
18
The 5 experiments in this series were all done with the vinylsulphone dye, in 2
hours dyeing at a pressure of 230 bar and a temperature of 112ºC. The coloration
and fixation of polyester were independent of humidity. A dark orange color was
obtained before ((K/S)dyed = 25.5 ± 1.5) and after Soxhlet extraction ((K/S)extr =
19.5 ± 1.5), the fixation percentage was 75 ± 5 %. Nylon was also dyed well. Fig.
2.3.B shows an increase of coloration with humidity, but also in dry CO2 nylon
could be colored deeply (K/S = 20). The fixation percentages showed a slight
increase with humidity. The results for silk show that no coloration was obtained
without water addition, as was found by Schmidt et al. [7]. However, Fig. 2.3.C
shows that, as the humidity increased, so did the coloration and the fixation of silk.
Wool gave light colors in dry scCO2 but, as can be seen in Fig. 2.3.D, a higher
coloration and fixation were obtained when water was added.
For these amino containing textiles, the experiments gave maximum colorations
when the textile and the scCO2 were nearly saturated (RH = 97%). To investigate
the influence of dyeing time, the experiments were repeated at the same pressure
and temperature but with 4 and 6 hours dyeing time, with relative scCO2 humidity
RH = 97%. It was found that the polyester coloration before and after extraction
increased between 2 and 4 hours from (K/S)extr = 19.5 to 49 (Fig. 2.4.A). Between 4
and 6 hours, the coloration did not increase any further for polyester. This indicates
that the polyester was saturated with dye molecules after 4 hours. Also the fixation
(84 ± 1.5) does not change significantly between 4 and 6 hours.
Although silk is colored well by both dyes in moist scCO2, it remains white when
dyeing in dry scCO2, despite the fact that dye is dissolved in the scCO2, as is
indicated by the dark coloration of polyester in dry scCO2. It can be stated from
this that the positive effect of water on the coloration of the amino containing
textiles is not due to an action of water as a solubility enhancer. Because water is a
polar component, it can also not enhance the substantivity between the dye and the
fiber. An increase in fixation with water content, as observed with the amino
containing textiles, indicates that water plays a role in the reaction between the dye
and the fiber. Whether or not the accessibility of the fibers is enhanced by the
19
100
80
60
K/S
(K/S)dyed
40
(K/S)extr
20
0
2 4 6
dyeing time (hours)
100
80
60
K/S
40
20
0
2 4 6
dyeing time (hours)
100
80
60
K/S
40
20
0
2 4 6
dyeing time (hours)
100
80
60
K/S
40
20
0
2 4 6
dyeing time (hours)
Fig. 2.4. Coloration of textiles with a vinylsulphone dye in scCO2 at 230 bar,
112ºC and different CO2-humidities (RH) as a function of time, for polyester (A),
nylon (B), silk (C) and wool (D)
20
swelling effect of water, does not follow from the results, but can also not be ruled
out.
For the amino containing textiles, the results given in Fig. 4B-D show that
coloration after extraction increased with time up to 6 hours; after 4 hours nylon,
silk and wool were not yet saturated with dye. Fig. 2.4.B-D also reveal that, as the
dyeing time of the amino containing textiles progressed, the fixation increased.
This reflects the dyeing mechanism described above: First the dye adsorbs to the
textile surface and, after that, it reacts with the amino sites.
Firstly, two experiments were done with the dichlorotriazine dye: one in dry scCO2
(RH = 4%) and one in moist, almost saturated scCO2 (RH = 97%). As can be seen
in table 2.1, the results correspond qualitatively with those of the vinylsulphone
experiments for all textiles: The coloration before and after extraction and the
fixation of polyester were independent of water addition. The coloration after
extraction of silk was negligible in dry scCO2 but increased with relative scCO2-
humidity, the fixation on silk increased with water addition. Wool could be colored
lightly in dry scCO2 and darker in moist scCO2, the fixation on wool increased
when water was added. As was the case for the vinylsulphone dye, for the
dichlorotriazine dye it follows that the reactivity is increased by water addition.
Whether or not the accessibility of the fibers is also increased does not follow from
the results.
With the dichlorotriazine dye, the influence of pressure and temperature on the
coloration and fixation of polyester, silk and wool was investigated. In all of these
21
experiments, the relative humidity of the scCO2 was 97%. For the calculation of the
amounts of water in the textile and the scCO2, i.e. to calculate how much water had
to be added, the same procedure as above was followed, taking only the mass of
the cotton into account. Because the distribution of water between cotton and
scCO2 depends on temperature and pressure, different amounts of water were
added in each experiment, varying from 22 g at (225 bar, 100ºC) to 30 g water at
(250 bar, 116ºC), to reach the same CO2-humidity of 97%.
In table 2.2 the results are given for varying pressures (from 225 to 278 bar) at
100ºC and 2 hours dyeing time. The colorations (K/S)dyed and (K/S)extr of polyester
were good and increased with pressure at constant temperature, as was reported by
Chang et al. [12] for disperse non-reactive dyes. The increase of dye uptake with
pressure is caused by enhanced dye solubility and by increased swelling of the
polyester. Although less than polyester, silk and wool were dyed well and the
(K/S)dyed and (K/S)extr varied within the experimental error range, i.e. the
colorations were independent of pressure. This is clear from Fig. 2.5, which gives a
graphical representation of the influence of pressure on the coloration after
extraction. Table 2.2 shows that the fixations on polyester, silk and wool are
independent of pressure.
Table 2.2. Influence of pressure on the coloration before ((K/S)dyed) and after
((K/S)extr) Soxhlet extraction and the calculated fixations F of textiles dyed with
dichlorotriazine dye in scCO2 at 100ºC for 2 hours
22
Table 2.3 shows that, when temperature was varied from 100 to 116ºC at constant
pressure (250 bar), the colorations of polyester, silk and wool before and after
extraction showed no significant changes. This is also clear from Fig. 2.6, where
the colorations after extraction are presented graphically. The fixation percentages
of polyester, silk and wool showed no dependency on temperature. Although the
K/S-values for polyester were much higher in all the dichlorotriazine experiments,
the K/S-values for silk and wool, being between 3 and 4, also corresponded with a
dark purple color.
12
10
8 polyester
(K/S)extr
silk
6
wool
4
2
0
225 235 245 255 265 275
p (bar)
Fig. 2.5. Influence of pressure on the coloration (K/S)extr of polyester, silk and wool
dyed for 2 hours at 100ºC with dichlorotriazine dye in scCO2 with relative humidity
RH = 97%.
Table 2.3. Influence of temperature on the coloration before ((K/S)dyed) and after
((K/S)extr) Soxhlet extraction and the calculated fixations F of textiles dyed with
dichlorotriazine dye in scCO2 at 250 bar for 2 hours
23
Pressure and temperature are the two parameters that determine the solubility of
dye in scCO2. Therefore, the negligible influence of both parameters on the
coloration of the amino containing textiles suggests that the solubility is not the
rate determining factor in the process. However, since it is not known how strong
the solubility varies with pressure and temperature, this cannot be stated with
certainty. Research is needed on the influence of pressure and temperature on the
solubility of this particular compound in scCO2.
14
12
10 Polyester
(K/S)extr
8 Silk
6 Wool
4
2
0
100 105 110 115
temperature (ºC)
Fig. 2.6. Influence of temperature on the coloration (K/S)extr of polyester, silk and
wool dyed for 2 hours at 250 bar with dichlorotriazine dye in scCO2 with a relative
humidity of 97%
2.4. Conclusions
24
independent of water addition. For the amino-containing textiles, the coloration
increases with the concentration of water in the scCO2 and the textiles.
Experiments with the vinylsulphone dye show that maximum coloration of nylon,
silk and wool is obtained when both the scCO2 and the textiles are saturated with
water. The fixation percentage of vinylsulphone dye on polyester was 75% and
independent of pressure. Fixations on nylon, silk and wool increased with relative
humidity of the scCO2. In saturated scCO2, values are reached in 2 hours of 94, 85
and 90%, respectively. When dyeing with the dichlorotriazine dye in scCO2
saturated with water, the fixations of polyester, silk and wool are 93, 88 and 79%.
When the dyeing time is varied for the vinylsulphone dye, the coloration of the
amino containing textiles increases as expected. Also the fixation grows in time,
indicating that more and more of the dye molecules present in the textile are
covalently bonded to the amino sites.
The positive effect of water on the dyeing process is caused either by water
facilitating the chemical reaction between the dye and the fiber, or by water acting
as a swelling agent for the textiles. Whatever the reason for this phenomenon, it is
concluded that water should be added in supercritical dyeing of nylon, silk and
wool with disperse dyes containing vinylsulphone or dichlorotriazine reactive
groups.
The experimental results on the dichlorotriazine dye suggest that the solubility and
the reactivity can be ruled out as the rate-determining step of the dyeing process.
However, for this conclusion to be made with certainty, more research is needed on
solubility and reaction kinetics.
References
25
3. I. Tabata, J. Lyu, S. Cho, T. Tominaga and T. Hori, Relationship between
the Solubility of Disperse Dyes and the Equilibrium Dye Adsorption in
Supercritical Fluid Dyeing, Color. Technol. 117, 346 (2001).
4. S.K. Liao, Y.C. Ho, P.S. Chang, Dyeing of Nylon 66 with a Disperse-
Reactive Dye using Supercritical Carbon Dioxide as the Dyeing Medium,
J. Soc. Dyers and Color. 116, 403 (2000).
5. J. Heyna, W. Schumacher and Hoe, German P 925,902 (1949).
6. I.D. Rattee, W.E. Stephen and ICI, BP 772,030 774,925 871,930 (1954).
7. A. Schmidt, E. Bach and E. Schollmeyer, Use of Fiber Reactive Groups in
Supercritical Carbon Dioxide, Melliand Textilberichte 83 (9), 648 (2002).
8. A. Schmidt, E. Bach and E. Schollmeyer, The Dyeing of Natural Fibers
with Reactive Disperse Dyes in Supercritical Carbon Dioxide, Dyes and
Pigments 56 (1), 27 (2003).
9. W.J.T. Veugelers, H. Gooijer, J.W. Gerritsen and G.F. Woerlee, European
Patent EP 1 126 072 A2, 2001.
10. R. McDonald (Ed.), Colour Physics for Industry, 2nd ed., Society of Dyers
and Colourists, Bradford (1995).
11. M. van der Kraan, M.V. Fernandez Cid, G.F. Woerlee, W.J.T. Veugelers,
C.J. Peters and G.J. Witkamp, Equilibrium Distribution of Water in the
Two-Phase System Supercritical Carbon Dioxide-Textile, J. Supercrit.
Fluids, submitted for publication.
12. K.-H. Chang, H.-K. Bae and J.-J. Shim, Dyeing of PET Textile Fibers and
Films in Supercritical Carbon Dioxide, Korean J. Chem. Eng. 13 (3), 310
(1996).
26
Chapter 3
Abstract
When natural fibers are dyed in supercritical carbon dioxide, the addition of a small
amount of water increases coloration. For a process design it is important to know
how much water has to be added to obtain a desired humidity of both textile and
carbon dioxide. In this work a thermodynamic model is proposed to calculate the
distribution of water over the textile phase and the supercritical phase as a function
of pressure and temperature. The phase equilibrium is described with Raoult’s law
for non-ideal fluids. The absorbed water in the textile is a condensed phase and is
modeled here as a non-ideal liquid, using the NRTL-equation. The non-ideality of
the supercritical phase is described by a solubility enhancement factor, a new
equation derived from statistical thermodynamics. Although the model is applied to
cotton, viscose, silk and wool, it can be used for all water absorbing textiles.
27
Symbols
28
Y mole fraction of water in scCO2 [-]
ysat water mass fraction in saturated scCO2 [-]
Z partition function [-]
3.1. Introduction
Using supercritical carbon dioxide (scCO2) as a solvent for dyeing textiles has been
the subject of several researchers in the last few years, as was reviewed by Bach et
al. [1]. Typical dyeing conditions are 373 K and 30 MPa. Veugelers et al. [2] and
Van der Kraan [3] reported that in reactive dyeing of natural textiles the addition of
a small amount of water to the dyeing vessel increased the uptake of dye.
The positive effect of water can be attributed to one or more of the following three
causes. Firstly, water can act as a modifier, increasing the solubility of dye in
29
scCO2. Secondly, water can participate in the reaction of dye with the textile
reactive sites. This option has not yet been investigated. Thirdly, water can have an
influence on the structure of the textile. Natural textiles consist of proteins (e.g. silk
and wool) or cellulose (e.g. cotton). These molecules contain amine or hydroxyl
groups that are capable of forming intermolecular hydrogen bonds that keep the
protein or cellulose chains and, therefore, the whole textile structure together.
When water is added to the textile, it breaks up the hydrogen bonds between the
chains, forms its own hydrogen bonds with the amine or hydroxyl groups and takes
position between the chains. This means that the chains are driven apart and that
the textile volume increases, i.e. water acts as a swelling agent (Nevell [4]). The
swollen structure of the textile allows dye molecules to penetrate the textile and,
therefore, has a positive effect on a dyeing process.
In a dyeing vessel, the water is distributed over the textile and the supercritical
phase. In this work a thermodynamic equilibrium distribution is modeled, taking
into account the non-ideal behavior of water in textile and in scCO2. To quantify
the interactions water-textile and water-scCO2, equations are derived that are fitted
to experimental data from literature. The resulting thermodynamic model enables
calculation of how much water has to be added to a supercritical dyeing process to
obtain a desired water content in the textile and in the scCO2. An example is given
showing how to apply the model in the case that the humidity of the textile is the
factor to be calculated. The case that the humidity of the scCO2 is to be calculated
is an analogous procedure. The model is developed for cotton, silk, wool and
viscose but can also be used for other water absorbing textiles.
30
3.2. Thermodynamic model
A dyeing vessel is considered containing textile, scCO2 and water. All water is
dissolved in either the textile or the scCO2; no liquid water is present. The textile is
modeled here as a homogeneous phase with ntot sites available for water adsorption,
of which nH2O are occupied by adsorbed water molecules. Since the water
molecules that are adsorbed to the textile sites are in fact a condensed phase, they
are modeled here as a non-ideal liquid. When the mole fraction of water in the
textile is defined as θH2O = nH2O/ntot, the equilibrium distribution of water over the
textile and the scCO2 can be described by Raoult’s law for non-ideal fluids (Smith
and Van Ness [5]):
H 2 O = YΦp
θ H 2 O γ H 2 O p sat (1)
pressure of pure water, Y is the mole fraction of water in the scCO2, p is the
pressure inside the dyeing vessel and Φ is the factor describing the non-ideality of
the supercritical phase. Φ is a function of the fugacity coefficient of water in scCO2
and of the Poynting factor and therefore a complex function of pressure and
temperature [5].
In this study, it is convenient to write Eq. (1) in terms of mass and mass density
instead of moles and pressure:
H 2 O = yρ
θ H 2 O γ H 2 O Γρ sat (2)
where ρ sat
H 2 O is the mass density of pure water at the vapor pressure, ρ is the mass
density of the supercritical phase, y is the mass fraction of water in the supercritical
phase and Γ is the factor describing the non-ideality of the supercritical phase. The
advantage of using mass densities instead of pressures is that Γ is a simple function
of density and temperature (see Eq. (19)). Eq. (2) is the basis upon which the
thermodynamic model is constructed.
31
The first term in Eq. (2), θH2O, is defined in Eq. (1) as a mole fraction but is equal
to the following mass fraction:
x
θ H 2O = (3)
x max
where x is the regain, a term used in textile industry to indicate textile humidity. Its
maximum value xmax is the regain in saturated scCO2. The regain is defined as the
mass of water in the wet textile m Htext2 O relative to mass of dry textile m text :
m Htext2 O
x= (4)
m text
Inspection of Eq. (3) and Eq. (4) reveals that θH2O is the same in terms of mass as in
terms of moles.
If the total mass of water distributed over the textile and the scCO2 is mH2O and the
mass of water in the scCO2 is m CO 2
H 2 O , then the water mass balance is:
m H 2 O = m Htext2 O + m CO 2
H 2O (5)
m CO 2 CO 2
(
H 2 O = y m H 2 O + m CO 2 ) (6)
where y is the mass fraction of water in the supercritical phase and mCO2 denotes
the mass of scCO2 in the dyeing vessel, calculated from the volume of the vessel,
the density’s of water and textile (ρCO2 and ρtext) and the mass of (dry) textile mtext:
m text
m CO 2 = ρ CO 2 (V − ) (7)
ρ text
32
The values for the textile density ρtext, taken from Morton and Hearle [6], are given
in table 3.1. The density ρCO2 needed in Eq. (7) is taken from the IUPAC tables
(Angus et al. [7]). The density ρCO2 is also used for ρ in Eq. (2) because the
solubility of water in scCO2 is small and therefore it is assumed that it has
negligible influence on the CO2-density: according to Evelein et al. [8], the value of
y at typical dyeing conditions of 30 MPa and 373 K is y sat = 10-2.
Table 3.1. Density ρtext of dry cotton, silk, wool and viscose rayon from Morton and
Hearle [6]
Textile material ρtext (kg/m3)
cotton 1550
silk 1340
wool 1300
viscose rayon 1520
The terms in Eq. (2) and Eq. (3) that remain to be discussed represent the
thermodynamics of the water-textile mixture (activity coefficient γH2O and
maximum regain xmax) and the thermodynamics of the water-scCO2 mixture (non-
ideality factor Γ).
33
Wiegerink [9] observed that in humid air at atmospheric pressure the logarithm of
the regain x in silk, wool, cotton and viscose is inversely proportional to
temperature for a given relative air humidity. In the case of saturated air the water
regain is at its maximum xmax :
λ
x max = α max exp (8)
T
where αmax and λ are constants and T is temperature in K. Below the saturation
point of the textile, Eq. (8) is valid for calculating x, with a different pre-
exponential factor (α) but with the same λ [9]. Therefore, the ratio x/xmax (= θH2O)
is independent of temperature.
Eq. (8) can be used for supercritical dyeing with parameters determined by fitting
the equation to data from Wiegerink who measured the regains of cotton, silk, wool
and viscose as a function of relative air humidity for several temperatures. Taking
from Wiegerink [9] data points (T, xmax) allows calculation of the parameters αmax
and λ for a textile material (table 3.2). For a given temperature, the value of xmax
can now be calculated from Eq. (8).
activity coefficient
The activity coefficient γH2O in Eq. (2) follows from the NRTL equation (Prausnitz
et al., [10]):
G 21 G 12
ln γ H 2 O = θ 2text τ 21 ( ) 2 + τ12 (9)
θ H 2 O + θ text G 21 (θ text + θ H 2 O G 12 ) 2
34
where θtext ≡ 1-θH2O. The four parameters τ12, τ21, G12 and G21 are determined by
three independent parameters c12, c21 and g:
c12 c 21
τ12 = τ 21 = (10)
RT RT
It has been shown earlier that the chemical potential of water in textile is the same
in air as in scCO2. Therefore, also the activity coefficient is independent of the
surrounding medium. The NRTL-equation can be used for scCO2 when it is fitted
to Wiegerink’s data [9] measured in air. This is done by defining a relative scCO2-
humidity L as:
yρ CO 2
L= (12)
Γρ sat
H 2O
x
θ H 2O γ H 2O = γ H 2O = L (13)
x max
Taking points (L,x) from Wiegerink [9] for a certain temperature, calculating xmax
from Eq. (8) and using Eq. (13), allows calculation of points (θH2O,γH2O ) that are
used to fit the NRTL-equation, i.e. to determine the parameters G12, G21, τ12 and
τ21. Fig. 3.1 shows that temperature has a negligible influence on the activity
coefficient of water in cotton, silk, wool and viscose, from 347 to 377 K.
35
each textile type one set of NRTL-parameters can be used for all dyeing
temperatures between 347 and 377 K.
1 1
B
H 2 O = A exp
p sat (14)
T −C
36
with temperature T expressed in K and the constants taken from Poling et al. [11]:
A = 1.32⋅104 MPa, B = -3884 K and C = 43.0 K. Eq. (14) is combined with the
ideal gas law:
ρ sat
H 2O =
p sat H 2O
RT (15)
1000M H 2 O
where MH2O is the molar mass of water, R is the universal gas constant and the
factor 1000 is introduced to match the units. This allows elimination of vapor
pressure to obtain ρ sat
H 2 O , the density of pure gaseous water at the vapor pressure:
1000AM H 2 O B
H 2O =
ρ sat exp (16)
RT T −C
Eqs. (14) to (16) describe the behavior of ideal gases. For a supercritical phase, the
factor Γ in Eq. (2) corrects for the non-ideality in the fluid. Eq. (2) can be rewritten
as:
yρ CO 2
Γ= (17)
θ H 2 O γ H 2 O ρ sat
H 2O
y sat ρ CO 2
Γ= (18)
ρ sat
H 2O
Inspection of Eq. (18) reveals that Γ is the water concentration in scCO2 (= yρCO2)
relative to the water concentration in the absence of scCO2 (= ρ sat
H 2 O ) , i.e. Γ
describes the extra solubility of water in the gas (or supercritical) phase that is
37
caused by non ideal behavior, or by the interaction of water molecules with CO2-
molecules. Therefore, Γ is called here the solubility enhancement factor.
y sat ρ CO 2 ρ CO 2 κρ CO 2 b1 ρ CO
max
2 ρ CO 2
Γ= =
1 −
max
exp
− (19)
ρ H 2O
sat
ρ CO 2 T M CO 2 ρ CO 2 − ρ CO 2
max
where ρ CO
max 3
2 = 1560 kg/m is the maximum (solid state) density of CO2 (Perry and
Green [12]), MCO2 = 44.0 g/mole is the molar mass of CO2. As is shown in
appendix A, b1 represents the molar excluded volume and can be taken as the Van
der Waals volume b1 = 0.0305 liter/mole (Lide [13]). In Eq. (19), κ is the only
parameter that is to be determined by fitting the equation to literature data. Evelein
et al. [8] give water solubilities in scCO2 as a function of pressure for several
temperatures. These data are used to determine κ = 2.08 K⋅m3/kg and to construct
the fitted curves shown in Fig. 3.2. The solubilities predicted by Eq. (19) are in
good agreement with the data from literature, especially at typical dyeing
conditions of 373 K and 30 MPa. This proves that the approximations introduced in
the derivation of Eq. (19), in appendix A, are justified, at least for the system H2O-
CO2.
38
100
323 K, fit
323 K, data
348 K, fit
enhancement factor
348 K, data
373 K, fit
10 373 K, data
423K, fit
423K, data
1
0 100 200 300 400 500 600 700 800 900
carbon dioxide density (kg/m3)
The aim of the model is to predict how much water has to be added to a
supercritical dyeing process (mH2O) to obtain a desired textile humidity (x), at given
process conditions (p, T, vessel volume, mtext and textile type). The procedure is as
follows:
1. From temperature, ρ sat
H 2 O and xmax are calculated with Eq. (16) and (8)
respectively.
2. From p, T and vessel volume V follow: ρCO2 (IUPAC tables) and Γ
(Eq. (19)).
3. From ρCO2, mtext, ρtext and V, the carbon dioxide mass mCO2 is
calculated (Eq. (7)).
4. The model is now reduced to a system of 6 equations: Eq. (2) to Eq.
(6) and Eq. (9). From experience it has to be known at which value of
x the coloration of a certain type of textile is maximal. Substituting this
x in Eq. (3) and Eq. (4) leaves 6 unknown variables: θH2O, γH2O, y,
39
m Htext2 O , mH2O and m CO 2
H 2 O . The system can be solved and the output
mH2O gives the amount of water that has to be added to a dyeing vessel
to obtain maximum textile coloration.
40
3.3. Conclusions
The solubility and the activity coefficient of water in textile were modeled. The
resulting equations were fitted between 347 and 377 K with experimental data for
air but are also valid under supercritical dyeing conditions.
Appendix A
1 ∂ ln Z
µi = − (a1)
β ∂n i
According to Vera and Prausnitz [16] and Reif [17] the partition function for non-
ideal fluids can be approximated by the generalized Van der Waals form:
3
n
1 2πm 2
Z = 2 [(V − Vex ) exp(−βU)] q nr , v
n
(a2)
n! h β
41
where m is particle mass, h is Planck’s constant, Vex is the volume excluded by
the particles and U is the average attractive potential energy between one particle
and all others in the fluid. The final term, q nr , v , accounts for rotational and
vibrational energies of the particles and according to Gasser [18] it is a function of
temperature only for small molecules like water or carbon dioxide. To allow the
derivation of an analytical solution like Eq. (18), a simplification is needed here: It
is assumed that the rotational and vibrational term q nr , v is negligible compared to
the other contributions in Eq.(a2). The validity of this assumption is checked in a
later stage by fitting the final result of the derivation Eq. (18) to experimental data
from literature.
The total potential U is the sum of ½ n pair potentials (u) where the factor ½ is
introduced because each potential is shared by 2 particles. For u an average value
can be estimated from the attractive term of the Lennard-Jones potential:
∞ 6 ∞ 4
1 1 n σ 6 n 1 8πσ 3 εn
2 V ∫σ V ∫σ r
U = nu = − 4 ε 4 πr 2
dr = −8 πσ ε dr = − (a3)
2 r 3V
where ε and σ are the Lennard-Jones parameters and r is the distance between a
pair of particles.
Substituting the right-hand term of Eq. (a3) in Eq. (a2) and taking the natural
logarithm gives:
3 2 πm 8βπσ 3 εn 2
ln Z = −n ln n + n + n ln 2 + n ln(V − Vex ) + (a4)
2 h β 3V
42
1 ∂ 3 2 πm 8βπσ 3 εn 2
µ=− − n ln n + n + n ln 2 + n ln(V − Vex ) +
β ∂n h β 3V
2
1 3 2 πm Vex 16βπσ 3 εn
= − − ln n + ln 2 + ln(V − Vex ) − + (a5)
β 2 h β V − Vex 3V
Eq. (a5) can be applied for gases, liquids and supercritical fluids. When Eq. (a5) is
substituted in the liquid (L) - gas (G) equilibrium condition µL =µG, the term
3/2ln(2πm/h2β) cancels because it is the same for a gas and a liquid. The
equilibrium condition becomes:
VexL 16βπσ 3 εn L
− ln n L + ln(V L − VexL ) − + =
V L − VexL 3V L
(a6)
VG 16βπσ 3 εn G
− ln n + ln(V − V ) − G ex G +
G G G
V − Vex
ex
3V G
43
b 8βbε 8βbε
ln(VmL − b) − + = ln( V G
− b ) + (a8)
VmL − b VmL
m
VmG
VmL − b 1 1 b 8βbε b
= exp 8βbε ( G − L )( L ) = exp(− L + L ) (a9)
Vm − b
G
Vm Vm Vm − b Vm Vm − b
because the reciprocal of the molar gas volume is negligible compared to that of
the liquid. As discussed above, the pressure is low, so that VmG − b is
approximately VmG and Eq. (a9) becomes:
1 1 8βbε b
= L exp − L + L (a10)
Vm Vm − b
ig
Vm Vm − b
where ig denotes “ideal gas”. This equation is used in combination with Eq. (a18)
that will now be derived.
where ε12 = ε21 represents the binary interaction energy between a water and a
carbon dioxide molecule. The corresponding Lennard-Jones distance is [(σ1
+σ2)/2]3. Substitution of Z in Eq. (a1) and carrying out the differentiation to the
44
particles 1 gives the chemical potential of water µ1 dissolved in supercritical carbon
dioxide. In the differentiation , it is taken into account that the excluded volume is
occupied by both components: Vex = Vex,1 + Vex,2, or, if vi is the volume excluded
by a single particle of component i: Vex = n1v1 + n2v2.
3 2 πm nv 16βπσ13 ε 11 n 1
− ln n 1 + ln 2 1 + ln(V − Vex ) − 1 1 + +
2 h β V − Vex 3V
1 (a12)
µ1 = −
β 16βπ σ1 + σ 2 ε 12 n 2
3
2 n v
− 2 1 + ln q 1,r , v
3V V − Vex
Using the molar excluded volume b1 = 2/3 πNAσ13, introducing the combined
excluded volume b12 = 2/3 πNA [(σ1+σ2)/2]3 and setting µ 1L = µ 1G :
n 1L v1 8βb1ε 11 n 1L 8βb12 ε 12 n L2 n L2 v1
− ln n + ln(V − V ) − L
L L L
+ + − L =
V − VexL V − VexL
1 ex
VLNA VLNA
(a13)
n 1G v1 8βb1ε 11 n 1G 8βb12 ε 12 n G2 n G2 v1
− ln n + ln(V − V ) − G
G G G
+ + − G
V − VexG V − VexG
1 ex
VG NA VG NA
Or, taking one mole of water and neglecting the solubility of CO2 in the liquid
phase:
b1 8βb1ε 11 n 1L
− ln N A + ln(VmL − b1 ) − + =
VmL − b1 VLNA
(a14)
nGv 8βb1ε 11 n 1G 8βb12 ε 12 n G2 n G2 v1
− ln n + ln(V − V ) − G 1 1 G +
G G G
+ −
V − Vex V G − VexG
1 ex
VG NA VG NA
8βb1ε 11 n 1L 8βb1ε 11 n 1G n G2 v1 n 1G v1
Furthermore: >> and >> because the
VLNA VG NA V G − VexG V G − VexG
solubility of water in the gas phase is low. Neglecting both small terms and
substituting b1/NA for v1:
45
b1 8βb1ε 11 n 1L
− ln N A + ln(VmL − b1 ) − + =
VmL − b1 VLNA
(a15)
8βb12 ε 12 n G2 n G2 v1
− ln n + ln(V − V ) +
G G G
−
V G − VexG
1 ex
VG NA
n iα
Introducing the molar concentration of component i in phase α as N iα =
Vα NA
gives:
b1 8βb1ε 11 n 1L
ln(VmL − b1 ) − + =
VmL − b1 VLNA
(a16)
1 VexG N A n G2 N A V G b1
ln( G − ⋅ G ) + 8βb12 ε 12 N G2 −
N1 n1 G
n2 NAV G
1 / N 2 − 1 / N 2,max
G
b1 8βb1ε 11 N A
ln(VmL − b1 ) − + =
V − b1
L
m VLNA
==> (a17)
1 N G2 b NG N
ln G − + 8βb12 ε 12 N G2 − 1 2 2,max
N G N 2,max − N G2
1 N 2,max N 1
N 1G
=
1 − N G2 / N 2,max
==> (a18)
1 b1 8βb1ε 11 b1 ρ CO
max
2 ρ CO 2
exp − + 8β b ε N G
−
Vm − b1 M CO 2 ρ CO 2 − ρ CO 2
Vm − b1
L L L 12 12 2 max
Vm
N 1G 1 8b12 ε 12 N G2 b1 ρ CO
max
2 ρ CO 2
= exp
− (a19)
1 − N 2 / N 2,max Vm
G ig
kT M CO 2 ρ CO 2 − ρ CO 2
max
N 1G N G2 8b ε 1000ρ CO 2 b1 ρ CO
max
2 ρ CO 2
= 1 − exp 12 12 − (a20)
N 1ig N 2,max kTM M ρ max
− ρ
CO 2 CO 2 CO 2 CO 2
46
Setting N1G/N1ig = yρCO2/ρH2Osat = Γ and N2G/N2,max = ρCO2/ ρ CO
max
2 , where ρ CO 2
max
denotes the maximum density of carbon dioxide and introducing the constant κ:
y sat ρ CO 2 ρ CO 2 κρ CO 2 b1 ρ CO
max
2 ρ CO 2
=
1 −
max
exp
− (a21)
ρ H 2O
sat
ρ CO 2 T M CO 2 ρ CO 2 − ρ CO 2
max
References
47
10. J.M. Prausnitz, R.N. Lichtenthaler and E. Gomes de Azevedo, Molecular
Thermodynamics of Fluid-Phase Equilibria, 3rd ed., Prentice Hall, New
Jersey, 1999, 262.
11. B.E. Poling, J.M. Prausnitz and J.P. O’Connel, The Properties of Gases
and Liquids, 5th ed., McGraw-Hill, New York, 2001, A-59.
12. R.H. Perry and D.W. Green, Perry’s Chemical Engineers’ Handbook, 7th
ed., McGraw-Hill, New York, 1997, 2-12.
13. D.R. Lide, Handbook of Chemistry and Physics, 75th ed., CRC Press, Boca
Raton, 1994, 6-48.
14. W. Bobeth, W. Berger, H. Faulstich, P. Fischer, A. Heger, H.-J. Jacobasch,
A Mally and I. Mikut, Textile Faserstoffe, Springer-Verlag, Berlin, 1993,
237.
15. T.L. Hill, An Introduction to Statistical Thermodynamics, Addison-Wesley
Publishing Company, Massachusetts, 1960, 19.
16. J.H. Vera and J.M. Prausnitz, Generalized van der Waals Theory for Dense
Fluids, Chem. Eng. J. 3, 1 (1972).
17. F. Reif, Fundamentals of Statistical and Thermal Physics, McGraw-Hill,
New York, 1965, 426.
18. R.P.H. Gasser and W.G. Richards, An Introduction to Statistical
Thermodynamics, World Scientific Publishing Co., Singapore, 1995, 40.
48
Chapter 4
Abstract
49
4.1. Introduction
The best suited SCF for this application is carbon dioxide, because it is
inexpensive, non-toxic and non-flammable. Due to its green and safe character,
supercritical carbon dioxide (scCO2) is already in use on an industrial scale in the
extraction of caffeine from coffee and tea. Typical conditions for textile dyeing in
scCO2 are 30 MPa and 120ºC, resulting in a liquid-like CO2-density of 585 kg/m3.
The most important advantage of dyeing in scCO2 instead of water is the easier
separation of solvent and residual dye: depressurization after the dyeing leads to
precipitation of the dye and delivers clean, gaseous CO2, so that both compounds
can be recycled.
The review by Bach et al. [2] mentions that natural and synthetic fibers have been
dyed successfully in scCO2. Schmidt et al. [3] have shown that scCO2-dyeing does
not damage the fibers, as long as the temperature is kept below 160ºC.
Most publications in the field of supercritical dyeing treat the measurement and
modeling of dye solubility in scCO2 [4-21]. The present chapter aims to contribute
to a less researched [22-26] subject in polyester dyeing: it describes a quantitative
50
equilibrium study of saturation colorations, adsorption isotherms and distribution
coefficients.
The aim of this work is to investigate the influence of the density and temperature
of the scCO2 on the polyester saturation coloration and the distribution coefficient.
Also the thermodynamic parameters governing the supercritical dyeing process are
determined for two dyes. The coloration and the thermodynamic parameters are
compared to those reported for aqueous dyeing. Finally, the adsorption isotherms
of two dyes on polyester are determined to assess if Nernst adsorption is the dyeing
mechanism, as is the case for dyeing in water.
51
where y is the solubility in mole fraction, ρ is the density of pure scCO2 (kg/m3) at
the dyeing pressure p (MPa) and dyeing temperature T (K) and M0 - M5 are fit
parameters. The value of CCO2 (g/g) is now:
M dye (2)
C CO 2 = y
M CO 2
where Mdye and MCO2 are the molar masses (g/mol) of dye and carbon dioxide.
The thermodynamic affinity (–∆µ0, J/mole) of dye for textile is defined as [29]:
a PET (3)
− ∆µ 0 = RT ln
a CO 2
where R is the universal gas constant (J/mole/K), T is the dyeing temperature (K)
and aPET and aCO2 are the activities of the dye in the polyester and the scCO2 phase.
The activities are defined as the ratio’s of the dye fugacities in the corresponding
phase (fPET and fCO2) relative to the standard-state fugacity f0 , the standard state
being solid dye at the dyeing temperature and 1 bar:
f PET f CO 2 (4)
a PET = a PET =
f0 f0
The fugacities in the PET and the scCO2 follow from the definitions of the activity
coefficients γPET and γCO2:
where CPET and CCO2 are the dye concentrations (g/g) in both phases and f is the
fugacity of pure solid dye at the dyeing temperature and at the dyeing pressure.
Eliminating fPET and fCO2:
52
C PET γ PET f C CO 2 γ CO 2 f (6)
a PET = a CO 2 =
f0 f0
In aqueous dyeing it is common practice [30] to put f = f0, i.e. to assume that the
fugacity of the dye is independent of pressure. This is also assumed for the
supercritical process. The system regarded here is saturated with dye, so that γPET =
γCO2 = 1. The activities are now equal to the concentrations and equation 3 can be
simplified to:
C PET (7)
− ∆µ 0 = RT ln
C CO 2
The ratio of CPET (g/g) and CCO2 (g/g) is the distribution coefficient K:
C PET (8)
K ≡ CO 2
C
In this work, the values for K are determined at several temperatures and scCO2-
densities, using equation 8.
The thermodynamic affinity is equal to the standard free energy change of dyeing
∆G0 (J/mole), and therefore:
where ∆H0 and ∆S0 are the changes in standard enthalpy (J/mole) and entropy
(J/mole/K) during the dyeing process. Combining equations 7, 8 and 9 gives:
∆H 0 ∆S 0 (10)
ln K = − +
RT R
for a saturated dye bath. According to equation 10, constructing a Van ‘t Hoff plot,
i.e. plotting ln K as a function of 1/T leads to a straight line and gives the values of
the standard enthalpy and entropy of dyeing. Equation 10 is known to hold for
53
aqueous dyeing processes [29] and it is investigated here if it is also valid for
supercritical dyeing.
The distribution coefficients are determined here at maximum coloration, i.e. for
polyester saturated with dye. The values for CPET are measured. The values of CCO2
that are to be substituted in equation 8, are the dye solubilities in scCO2, from
equation 2.
In the aqueous dyeing process, polyester is heated with steam of 150 to 200ºC,
prior to dyeing. This is done to relax the polymer molecules in the fiber, to remove
the intermolecular stresses that are formed in the fiber spinning process. Without
this so-called heat-setting, excessive shrinking of the cloth occurs whenever it is
exposed to high temperatures, e.g. in a dye bath. It is investigated in this work what
the shrinking behavior of heat-set and non-heat-set polyester is when it is dyed in
scCO2. Also the tensile strength before and after dyeing are determined, to check
for possible fiber damage.
54
4.3. Experimental
4.3.1. Materials
The properties of the dyes used in this study are given in table 4.1. Their structures
are given in figure 4.1. All dyes were purchased from Sigma Aldrich (The
Netherlands) and used without further purification.
Table 4.1. Properties of the dyes used in this study, with molecular weight (M),
wavelength of maximum absorption (λmax) and the purity according to the supplier
Dye M (g/mole) λmax (nm) purity (%)
Disperse Orange 13 352 427 90
Disperse Orange 3 242 443 90
Disperse Red 1 314 502 95
CH2 CH3
O2N N N NH2 O 2N N N N
CH2 CH2 OH
Disperse Orange 13
Figure 4.1. Molecular structures of the dyes used in this study
55
4.3.2. Dyeing equipment and procedure
The experiments were conducted in a 40-liter stainless steel pressure vessel with a
perforated drum inside that contained 100 g textile cloth. The drum rotated at 60
rpm. The dye was put in a 50 ml porous stainless steel cylinder (pore size 10µm)
that was fixed inside the drum. A simplified scheme of the equipment is given in
figure 4.2. The system was pressurized in 10 minutes, with a air-driven
reciprocating pump (Resato International, The Netherlands). The dyeing vessel was
heated with a steam jacket. During the heating, the scCO2 reached the desired
pressure and temperature in 30 minutes. After that, pressure and temperature
remained constant. During the dyeing, the scCO2 was circulated through the vessel
by a centrifugal pump (Autoclave Engineers, USA) at a flow rate of 30 liter/min, to
intensify mixing inside the vessel.
steam
rotating drum, with
textile and dye
motor 40-liter
pressure vessel
In the experiments in scCO2 where the influence of temperature and density on the
saturation coloration was investigated, the PET was dyed for 6 and 8 hours. No
difference in coloration was measured. Because of this observation and because,
after these experiments, dye powder was left in the porous steel cylinder, it was
concluded that the PET had reached saturation. The distribution coefficient was
calculated with equation 8, taking CPET from equation 12 (below) and CCO2 from
equation 2.
56
Also the experiments in scCO2 for the determination of the sorption isotherms were
done for 6 and 8 hours. Again, no color differences were observed but now all dye
had disappeared from the steel cylinder, leading to the conclusion that after these
dyeing times, the dye distribution between the scCO2 and the PET was at
equilibrium but below saturation. The amount of dye in the PET-cloth (CPET) was
taken from equation 12, the amount of dye remaining in the solvent (CCO2) was
calculated from the dye mass balance:
with mTOTAL denoting the mass of pure dye (g) introduced into the vessel prior to
the experiment, mPET the mass of dyed polyester (100g) and mCO2 the mass (g) of
scCO2 in the dyeing vessel. The latter was calculated as the product of density at
120ºC and 300 bar (585 kg/m3) and vessel volume (40 liter): mCO2 = 23000 g.
To check the shrinkage of the polyester during the dyeing, the weight per unit area
cloth was measured before and after the dyeing. This was done for both heat-set
and non-heat-set polyester.
4.3.3. Analyses
After each dyeing experiment, the dyed PET-cloth was washed with acetone at
room temperature to remove any unfixed dye. The PET was then subjected to a
Soxhlet-extraction with boiling N,N-dimethylformamide (DMF) in an oil bath of
180ºC. After all dye had been removed, the dye concentration in the DMF was
determined with a UV/VIS spectrophotometer, at the wavelength of maximum
absorption given in table 1. The dye concentration in the PET was then calculated
with:
where CPET is the concentration (g/g) of dye in the PET, CDMF is the dye
concentration (g/g) in the DMF, mDMF is the mass of DMF after extraction (g) and
mPET is the mass of PET (g).
57
4.4. Results and Discussion
The solubility fits are constructed for three dyes: Disperse Orange 3, Disperse
Orange 13 and Disperse Red 1. The fit parameters of equation 1 are given in table
4.2. The fits are presented graphically in figures 4.3, 4.4 and 4.5. It can be seen that
the solubilities of Disperse Orange 3 and Disperse Red 1 are described well, that of
Disperse Orange 13 less accurately, but acceptable for our purpose. For Disperse
Orange 3 and Disperse Red1, the fits cover almost the entire dyeing temperature
range, for Disperse Orange 13, the fit range does not cover the higher dyeing
temperatures.
Table 4.2. Fit parameters for equation 1, describing the solubility of Disperse
Orange 3, Disperse Orange 13 and Disperse Red 1 in supercritical carbon dioxide
Dye Disperse Orange 3 Disperse Orange 13 Disperse Red 1
Reference [6] [8] [14]
M0 -0.006745 -48.69 0.07694
M1 -0.2452 0.7175 -0.6607
M2 -0.003408 -0.01720 -5.833·10-4
M3 1.102·10-3 9.821·10-4 1.776·10-3
M4 -0.2199 0.3863 -0.2913
M5 0.6311 4.747 1.448
58
20
y/10 mol/mol
15
10
-6
5
0
300 400 500 600 700
density / (kg/m3)
30
25
y /10 mol/mol
20
15
-6
10
5
0
300 400 500 600 700
density / (kg/m3)
Figure 4.4. Solubility (y) of Disperse Orange 13 as a function of carbon dioxide
density, with drawn lines representing the fitted curves of equation 1 and points
representing literature data [8]: 330 K, 370 K
59
350
300
y / 10 mol/mol
250
200
-6
150
100
50
0
100 300 500 700 900
3
density / (kg/m )
Figure 4.5. Solubility (y) of Disperse Red 1, with drawn lines as the fitted curves
and points representing literature data [14]: 323 K, 353 K, 383 K
For two dyes, Disperse Orange 3 and 13, the saturation coloration and the
distribution coefficient was investigated. The measured values of the saturation
coloration CPET are shown for Disperse Orange 3 in table 4.3, for two scCO2-
60
Table 4.4. Saturation concentrations of Disperse Orange 13 in polyester (CPET) and
in carbon dioxide (CCO2) and the distribution coefficients K
density T pressure CPET CCO2 K
kg/m3 (K) MPa / 10-2 g/g / 10-6 g/g
400 368 165 1.50 9.88 4230
378 178 1.69 16.6 3190
388 192 1.85 27.3 2280
393 199 1.90 34.8 1880
550 368 218 1.83 46.6 998
378 241 2.08 79.8 730
388 264 2.22 134 532
densities, together with the calculated values for CCO2 and the resulting distribution
coefficients K. The corresponding values for Disperse Orange 13 are given in table
4.4. To get a clear impression for the dyeing conditions, also the dyeing pressures
are given in the tables. The saturation coloration of Disperse Orange 13 in scCO2 at
378 K (105ºC) and 550 kg/m3 is 2.1 mass percent, close to the value reported in
literature for aqueous dyeing at 100ºC (2.0 mass percent) [31].
It can be seen from the tables that coloration increases with the temperature and the
density (or pressure) of the scCO2. This was reported earlier for other dyes in
scCO2, e.g. by Chang e.a. [22].
The Van ‘t Hoff plots of the distribution coefficient are given in figure 4.6 for
Disperse Orange 3. The R-squared values for 400 and 550 kg/m3 are 0.976 and
0.998, indicating good agreement of the experimental results with equation 10.
Also figure 4.7, showing the plots for Disperse Orange 13, shows agreement with
equation 10, the R-squared values for 400 and 550 kg/m3 being 0.994 and 0.999
respectively.
61
8.5
7.5
ln K
6.5
5.5
4.5
0.0025 0.0026 0.0027 0.0028
1/T(K)
Figure 4.6. Distribution coefficient K of Disperse Orange 3 as a function of dyeing
temperature T, for different scCO2-densities: 400 kg/m3, 550 kg/m3
8.5
7.5
ln K
6.5
5.5
4.5
0.0025 0.0026 0.0027 0.0028
1/T(K)
Figure 4.7. Distribution coefficient K of Disperse Orange 13 as a function of
dyeing temperature T, for different scCO2-densities: 400 kg/m3, 550 kg/m3.
The fact that equation 10 is valid here, confirms the assumption made between
equations 6 and 7, about the negligible influence of pressure on the fugacity f of the
pure dye was allowed, even though the pressures in scCO2 are far higher than in
aqueous dyeing.
It can be seen in figures 4.6 and 4.7, that the distribution coefficient decreases with
temperature. Regarding equation 10 it can then be stated that the enthalpies of
62
dyeing are negative, i.e. the process is exothermic. This is an apparent
contradiction with the observation made earlier in tables 3 and 4, that the coloration
CPET increases with temperature. The explanation is that also CCO2 increases with
temperature and relatively more than CPET, so that the ratio of CPET and CCO2, the
distribution coefficient, decreases. In other words, the transfer processes of dye
from the standard (solid) state to either the adsorbed state in the PET or to the
dissolved state in the scCO2 are both endothermic processes. The transfer of dye
from the dissolved to the adsorbed state, the actual dyeing step, is exothermic.
There is also a negative entropy of dyeing, caused by the dye molecules taking on a
state of higher order when going from the supercritical to the adsorbed state on the
polyester molecules. As was discussed above, negative enthalpy and entropy
changes are also reported for aqueous dyeing.
Table 4.5 shows values of ∆µ0 that are in or near the range that is characteristic of
physisorption, which is -20 to 0 kJ/mol [32]. The affinity of Disperse Orange 13 for
polyester in water at 100ºC is ∆µ0 = -13 kJ/mol.[31], not equal to but in the same
order as is calculated here for scCO2. This value was determined for polyester that
was heat-set at 195ºC, just as the polyester used in this work. The enthalpies,
entropies and affinities mentioned in the table are all similar to those measured in
water for other disperse dyes [33].
Table 4.5. Standard enthalpy (∆H0), entropy (∆S0) and affinity (∆µ0) of dyeing in
scCO2 at 400 and 550 kg/m3, for two disperse dyes
Dye Density ∆H0 ∆S0 ∆µ0 (100ºC)
kg/m kJ/mole Jmole/K kJ/mol
Disperse 400 -43 -57 -22
Orange 3 550 -41 -61 -18
Disperse 400 -39 -36 -25
Orange 13 550 -37 -44 -21
63
4.4.3. Adsorption isotherm
In these experiments, Disperse Orange 3 and Disperse red 1 were used to dye
polyester. To investigate whether or not the adsorption follows a linear Nernst
isotherm, dyeing experiments were done with both the scCO2 and the polyester
below saturation. The results are given for two dyes in figure 4.8. The R-squared
coefficients for Disperse Orange 3 and Disperse red 1 are both 0.996, showing a
linear dependence of CPET on CCO2.
2.5
Disperse Orange 3 2.5
Disperse Red 1
2 2
CPET / 10 -2 g/g
CPET / 10-2 g/g
1.5 1.5
1 1
0.5 0.5
0 0
0 5 10 15 0 5 10 15 20
CCO2 / 10-5 g/g CCO2 / 10-05 g/g
Figure 4.8. Adsorption isotherms for two disperse dyes, in scCO2 at 550 kg/m3 and
115ºC
To investigate the effect of the dyeing process on the polyester cloth, the shrinkage
and the tensile strength were determined, at the laboratory of Ames Europe. The
results are given in table 4.6. The increase of the mass per unit cloth area of the
non-heat-set polyester increases by 70% during supercritical dyeing. For the heat-
set polyester, this is only 20%, similar to the percentages normally observed in
aqueous dyeing [34]. This indicates that heat-set polyester has to be used for
dyeing in scCO2, to avoid excessive shrinkage. In table 4.6 it can be seen that the
tensile strength of the polyester cloth is not significantly affected by the exposure
to the supercritical dye bath. In the dyeing experiment on which table 4.6 is based,
Disperse Red 1 was used as dye, the temperature was 115ºC and the scCO2-density
was 550 kg/m3.
64
Table 4.6. Specific mass and tensile strength of heat-set and non-heat-set polyester,
before and after dyeing in scCO2 of 115ºC and 550 kg/m3
specific mass tensile strength
(g/m2) (N)
length width
heat-set before dyeing 123 483 457
after dyeing 147 494 476
non-heat-set before dyeing 123 335 479
after dyeing 206 634 529
4.5. Conclusions
When polyester is dyed with disperse dyes in supercritical carbon dioxide, the
saturation coloration increases with temperature and density of the scCO2. The
same coloration can be achieved in scCO2 as in water, as long as the density is
sufficiently high.
The transfer of dye from the dissolved state to the adsorbed state, i.e. the actual
dyeing step of the process, is an exothermic process, although the saturation dye
concentration in the polymer increases with temperature. This apparent
contradiction is caused by the increase of dye concentration in the CO2 with
temperature being relatively larger than the increase of the dye concentration in the
polyester. The dyeing step is accompanied by a negative entropy change. The
values of the enthalpies, entropies and thermodynamic affinities in supercritical
dyeing are in the same order of magnitude as in aqueous dyeing.
65
The linearity of the adsorption isotherms that is observed in water, is also found in
supercritical carbon dioxide, indicating a Nernst adsorption mechanism.
References
66
8. U. Haarhaus, P. Swidersky and G.M. Schneider, High-Pressure
Investigations of the Solubility of Dispersion Dyestuffs in Supercritical
Gases by VIS/NIR Spectroscopy. Part I – 1,4-Bis(octadecylamino)-9,10-
anthraquinone and Disperse Orange 13 in CO2 and N2O up to 180 MPa, J.
Supercrit. Fluids 8, 100 (1995).
9. S.N. Joung and K.-P. Yoo, Solubility of Disperse Anthraquinone and Azo
Dyes in Supercritical Carbon Dioxide at 313.15 to 393.15 K and from 10
to 25 MPa, J. Chem. Eng. Data 43, 9 (1998).
10. S.N. Joung, H.Y. Shin, Y.H. Park and K.-P. Yoo, Measurement and
Correlation of Solubility of Disperse Anthraquinone and Azo Dyes in
Supercritical Carbon Dioxide, Korean J. Chem. Eng. 15 (1), 78 (1998).
11. T. Kraska, K.O. Leonhard, D. Tuma and G.M. Schneider, Correlation of
the Solubility of Low-volatile Organic Compounds in Near- and
Supercritical Fluids Part II. Applications to Disperse Red 60 and Two
Disubstituted Anthraquinones, Fluid Phase Equilibria 194-197, 469
(2002).
12. J.W. Lee, J.M. Min and H.K. Bae, Solubility Measurement of Disperse
Dyes in Supercritical Carbon Dioxide, J. Chem. Eng. Data 44, 684 (1999).
13. H.-M. Lin, C.-Y. Liu, C.-H. Cheng, Y.-T. Chen and M.-J. Lee, Solubilities
of Disperse Dyes of Blue 79, Red 153 and Yellow 119 in Supercritical
Carbon Dioxide, J. Supercrit. Fluids 21 (1), 1 (2001).
14. T. Shinoda and K. Tamura, Solubilities of C.I. Disperse Red 1 and C.I.
Disperse Red 13 in Supercritical Carbon Dioxide, Fluid Phase Equilibria
213, 115 (2003).
15. B. Wagner, C.B. Kautz and G.M. Schneider, Investigations on the
Solubility of Anthraquinone Dyes in Supercritical Carbon Dioxide by a
Flow Method, Fluid Phase Equilibria 158-160, 707 (1999).
16. D. Tuma, B. Wagner and G.M. Schneider, Comparative Solubility
Investigation of Anthraquinone Disperse Dyes in Near- and Supercritical
Fluids, Fluid Phase Equilibria 182, 133 (2001).
17. T. Shinoda and K. Tamura, Solubilities of C.I. Disperse Orange 25 and C.I.
Disperse Blue 354 in Supercritical Carbon Dioxide, J. Chem. Eng. Data
48, 869 (2003).
18. K. Mishima, K. Matsuyama, H. Ishikawa, K.-I. Hayahi and S. Maeda,
Measurement and Correlation of Solubilities of Azo Dyes and
67
Anthraquinone in Supercritical Carbon Dioxide, Fluid Phase Equilibria
194-197, 895 (2002).
19. A.S. Ozcan, A.A. Clifford and K.D. Bartle, Solubility of Disperse Dyes in
Supercritical Carbon Dioxide, J. Chem. Eng. Data 42, 590 (1997).
20. D. Tuma and G.M. Schneider, Determination of the Solubilities of
Dyestuffs in Near- and Supercritical Fluids by a Static Method up to 180
MPa, Fluid Phase Equilibria 158-160, 743 (1999).
21. H.-D. Sung and J.-J. Shim, Solubility of C.I. Disperse Red 60 and C.I.
Disperse Blue 60 in Supercritical Carbon Dioxide, J. Chem. Eng. Data 44,
985 (1999).
22. K.-H. Chang, H.-K. Bae and J.-J. Shim, Dyeing of PET Textile Fibers and
Films in Supercritical Fluids, Korean J. Chem. Eng. 13 (3), 310 (1996).
23. M.R. De Giorgi, E. Cadoni, D. Maricca and A. Piras, Dyeing Polyester
Fibres with Disperse Dyes in Supercritical CO2, Dyes and Pigments 45, 75
(2000).
24. A.S. Özcan and A. Özcan, Adsorption Behavior of a Disperse Dye on
Polyester in Supercritical Carbon Dioxide, J. Supercrit. Fluids, in press.
25. I. Tabata, J. Lyu, S. Cho, T. Tominaga and T. Hori, Relationship between
the Solubility of Disperse Dyes and the Equilibrium Dye Adsorption in
Supercritical Fluid Dyeing, Color. Technol. 117, 346 (2001).
26. M.-W. Park and H.-K. Bae, Dye Distribution in Supercritical Dyeing with
Carbon Dioxide, J. Supercrit. Fluids 22, 65 (2002).
27. I. Kikic, F. Vecchione, P. Alessii, A. Cortesi, F. Eva and N. Elvassore,
Polymer Plasticization Using Supercritical Carbon Dioxide: Experiment
and Modeling, Ind. Eng. Chem. Res. 42, 3033 (2003).
28. A. Jouyban, H.-K. Chan and N.R. Foster, Mathematical Representation of
Solute Solubility in Supercritical Carbon Dioxide Using Empirical
Expressions, J. Supercrit. Fluids 24, 19 (2002).
29. H.H. Sumner, in: A. Johnson (Ed.), The Theory of Coloration of Textiles,
2nd ed., Society of Dyers and Colourists, Bradford (U.K.), 255 (1989).
30. S.M. Burkinshaw, Chemical Principles of Synthetic Fiber Dyeing,
Chapman and Hall, London (1995).
31. E. Merian, J. Carbonell, U. Lerch and V. Sanahuja, The Saturation Values,
Rates of Dyeing, Rates of Diffusion and Migration of Disperse Dyes on
Heat-Set Polyester Fibres, Am. Dyestuff Rep. 53(26), 11 (1964).
68
32. Y. Yu, Y.-Y. Zhuang and Z.-H. Wang, Adsorption of Water-soluble Dye
onto Activated Functionalized Resin, J. Colloid Interf. Sci. 242, 288
(2001).
33. R.H. Peters, Textile Chemistry, Volume III, Elsevier, Amsterdam, 105
(1975).
34. S. Fransman, Ames Europe B.V., Personal communication.
69
70
Chapter 5
Abstract
71
Symbols
a constant -
b constant -
c constant -
cp specific heat at constant pressure Jkg-1K-1
d pipe diameter m
e constant -
g gravitational acceleration ms-2
Gr Grashof number, Gr = gd3ρ(ρb-ρw) /µ2 -
H enthalpy Jkg-1
h heat transfer coefficient Wm-2K-1
k thermal conductivity Wm-1K-1
Nu Nusselt number, Nu = hd/k -
p pressure bar
Pr Prandtl number, Pr = µcp/k -
q heat flow Js-1
Re Reynolds number, Re = ωd/µ -
T temperature K and °C
Uo overall heat transfer coefficient Wm-2K-1
Greek
ρ density kgm-3
µ dynamic viscosity Pa⋅s
ω mass flux kgm-2s-1
Φm mass flow kgs-1
subscripts
b at bulk conditions
cp at constant properties
i inside of pipe
o outside of pipe
pc at pseudo-critical temperature
vp at variable properties
w at wall conditions
c critical
72
5. 1. Introduction
This problem has led to a great deal of research and many publications are to be
found in which Nusselt relations for the calculation of heat transfer coefficients
have been constructed and fitted to experimental work, as is reviewed by Pitla et al.
[1]. The present work is an investigation into the phenomena around the critical
point and into the buoyancy phenomenon, i.e. free convection caused by
temperature-induced density differences. Computational fluid dynamics
simulations of heated supercritical carbon dioxide flows through vertical pipes are
carried out.
The objective is to study when and how the variable properties and the buoyancy
must be taken into account in the Nusselt-relation for heating. For buoyancy
conditions in a heated flow, it is investigated what the position of a pipe and the
direction of the flow should be.
1 1 d o ln(d o d i ) d o
= + + (1)
Uo ho 2k steel dihi
where Uo is the overall heat transfer coefficient corresponding with the outside area
of the tube, ho and hi are the outside and inside film coefficients, do and di are the
outside and inside tube diameters and ksteel is the thermal conductivity of the steel
tube wall. As an example is taken here supercritical carbon dioxide (scCO2) being
heated with steam in a pipe with do=16 mm, di=13 mm and ksteel=16 Wm-1K-1 so
that 2ksteel/(doln(do/di))=10000 Wm-2K-1. The heat transfer coefficient for
73
condensing steam is ho=8000 Wm-2K-1. It can be seen in several publications [1-7]
that, without the influence of variable properties or buoyancy, the tube-side heat
transfer coefficient for scCO2 is in the order of hi=6000 Wm-2K-1. All three
contributions of Eq. (1) have the same order of magnitude, so it is important to take
into account enhancement or impairment of tube-side heat transfer caused by
variations in physical properties or buoyancy.
5.2. Nusselt-relations
h cp d i
Nu cp = = 0.0183 Re 0b.82 Prb0.5 (2)
kb
or:
h cp = 0.0183 ⋅ d i−0.17 ω 0.82 k 0b.5 µ −b0.33 c 0p.,5b (3)
where Nucp is the Nusselt number in case of constant properties, evaluated at bulk
conditions, hcp is the corresponding tube-side heat transfer coefficient, kb is the
thermal conductivity of the fluid at bulk conditions, Reb and Prb are the Reynolds
and Prandtl numbers at bulk conditions, ω is the mass flux and µb and cp,b are the
dynamic viscosity and the specific heat at constant pressure at bulk conditions. In
Eq. (2), the physical properties are assumed constant in radial direction. In the
longitudinal direction of a pipe, the bulk temperature and therefore the physical
properties do change. Eq. (2) delivers the local heat transfer coefficient.
Around the pseudo-critical point, thermal conductivity, viscosity and specific heat
vary strongly. It should be noted here that this temperature is higher than the
critical temperature (31°C) when the pressure is above its critical value (74 bar). As
is shown in Fig. 5.1 for 80 bar (pseudo-critical temperature Tpc = 34°C) , the
variation in the factor c 0p.,5b is much larger than the change in the factor k 0b.5 µ −b0.33 ,
so that it can be expected that the heat transfer coefficient from Eq. (3) will follow
74
the trend of the specific heat, i.e. showing peaks at the pseudo-critical temperature.
Indeed, these peaks were observed in the experiments of Swenson et al. [4] for
water.
Eq. (2) can be used to predict the heat transfer coefficient at supercritical
conditions only at low temperature differences between wall and bulk because then
the physical properties can be regarded as constant in radial direction.
10
6 0 .5
k b Cpµb −0.33
[( W /( mK )) 0.5 (Pa.s) −0.33 )]
wortel
4 0.5
c p,b [(kJ /(kgK))
product
0.5
]
0
25 30 35 40 45
fluid temperature (°C)
Fig. 5.1. Influence of the variation of carbon dioxide physical properties around
the pseudo-critical temperature at 80 bar on the constant-property heat transfer
coefficient of Eq. (3)
a
cp ρw
b
µw
c
kw
e
Nu vp = Nu cp (4)
c p,b ρb µb kb
75
where Nuvp is the variable-property Nusselt number, a, b, c and e are constants, ρ is
the fluid density and c p is the average specific heat that can be calculated from the
specific enthalpy H and temperature T according to Pitla et al. [1]:
Hw − Hb
cp = (5)
Tw − Tb
In practice, not all corrections in Eq. (4) are necessary. Krasnoshchekov and
Protopopov [5] only corrected for specific heat and density and still obtained a
relation that could be fitted to experimental data for a heated carbon dioxide flow
in a vertical pipe:
a
h vp d i cp ρw
b
Nu vp = = Nu cp (6)
kb c p ,b ρb
with a=0.4 at Tb<Tw<Tpc and 1.2Tpc<Tb<Tw,
a=0.4+0.2[(Tw/Tpc)-1] at Tb<Tpc<Tw,
a=0.4+0.2[(Tw/Tpc)-1][1-5{(Tb/Tpc)-1}] at Tpc<Tb<1.2Tpc and Tb<Tw,
and b=0.35-0.05p/pc.
where hvp is the variable-property tube-side heat transfer coefficient, Tb, Tw and Tpc
are temperatures at bulk, wall and pseudo-critical conditions and p and pc are
pressure and critical pressure respectively.
Other forced convection correlations for heating CO2 in a vertical pipe have been
proposed but according to Jackson and Hall [2], Eq. (6) is the most accurate and is
supported by the most experimental data. Although Krasnoshchekov and
Protopopov used a more complex constant-property Nusselt number, Jackson and
Hall [2] found that substituting Eq. (2) in Eq. (6) gave equally good results so this
will also be done in this work.
76
to a sharpened velocity profile relative to normal turbulent flow (Fig. 5.2a, b). This
leads to a steeper shear stress profile and therefore to more turbulence; the heat
transfer is enhanced.
velocity profiles
Fig. 5.2. Schematic drawing of buoyancy effect, in downward heated pipe flow,
leading to a sharpened velocity profile and a steeper shear stress profile, relative
to normal turbulent pipe flow
77
velocity profiles
Fig. 5.3. Schematic drawing of buoyancy effect, in upward pipe flow, leading to
either a flattened velocity profile and decreased shear stress or to an M-shaped
velocity profile and increased shear stress
It was observed by Walisch et al. [6] that heat transfer coefficients for heated
horizontal pipes are below those of vertical pipes with down flow. Since Walisch et
al. conducted their experiments both far away and close to the pseudo-critical
temperature and for different pressures, it is assumed here that the heat transfer
coefficients from the simulations for down flow will be larger than for horizontal
flow for the range of conditions treated here. Therefore, simulations for horizontal
flow are not carried out in this work.
Hall and Jackson [7] found that, in a vertical pipe with up- or down flow, buoyancy
has to be taken into consideration when the bulk Grashof and Reynolds numbers
are such that:
78
a 0.91 1 / 3
cp ρw
b
Gr
Nu mix = 0.0183 Re 0.82
Pr
0.5 1 + 2750 −b2.7 (8)
b b
c p,b ρb Re b
The tube-side heat transfer coefficients predicted by Eq. (8) are used here to
compare with the results from computational fluid dynamics simulations.
In the simulations, the enthalpy Hin of the CO2 flowing into the simulated pipe and
the mass flow rate Φm were known. The program delivered the enthalpy Hout of the
CO2 at the end of the pipe. The heat q that was transported to the CO2 was
calculated with q = Φm(Hout-Hin). The tube-side heat transfer coefficient hi was then
calculated with q = hiA(Tw-Tb). The length of the simulated pipe was short so that
the calculated hi can be regarded as a local heat transfer coefficient. Only tube-side
coefficients were calculated.
Simulations were carried out for constant bulk pressure and for constant bulk
density. In industrial practice these two cases are encountered in continuous and
batch processes respectively. For both cases, the wall temperature was set constant
in the simulations. In a continuous process this is realistic, in a batch heating
process it is not. In the latter case, therefore, the heat transfer coefficients resulting
from the simulations should be regarded as momentary values. It should be noted
79
here that the conditions “constant pressure” and “constant density” refer to the bulk
fluid. On a microscopic level, gradients in pressure and density do exist due to the
transport of thermal energy.
The constant bulk-pressure process was simulated for heated down flow in a pipe
of internal diameter 13 mm, mass flow 0.3 kg/s, bulk temperature Tb = 30 °C. The
tube-side heat transfer coefficients hi were calculated from the simulation results
and plotted as a function of wall temperature in Fig. 5.4. According to Eq. (7),
buoyancy was negligible so that only the variation of physical properties in radial
direction is expected to influence heat transfer.
heat transfer coefficient h i (kWm K )
16
-1
-2
14
12
10
8
6
P=80 bar
4
P=100 bar
2 P=120 bar
0
20 40 60 80 100 120
wall temperature Tw (°C)
Fig. 5.4. Heat transfer coefficient as a function of wall temperature and pressure,
for heated down flow at bulk temperature 30°C. Simulation results showing the
effect of the temperature-induced variation of specific heat
It can be seen in Fig. 5.4 that the heat transfer coefficient follows the same trend as
the specific heat does at the pseudo-critical temperature: showing peaks that
decrease in height and shift to higher temperatures when the pressure increases.
This can be understood by realizing that a peak in the specific heat in the thermal
boundary layer will give a peak in the heat transfer coefficient. However, this
would mean that, at a low wall-to-bulk temperature difference, the peak in transfer
coefficients should be observed when the wall has the pseudo-critical temperature.
From Liao and Zhao [9] it is known that at 80 and 100 bar, the pseudo-critical
80
temperatures are 35 and 45°C respectively. The peaks in Fig. 5.4 lie at higher
temperatures because even at low wall-to-bulk temperature difference, in reality
the wall will have to be somewhat above pseudo-critical to realize a pseudo-critical
temperature in the thermal boundary layer.
To illustrate the influence of the variable properties, the values of hi from Fig. 5.4
were compared to the constant-property heat transfer coefficient hcp calculated from
Eq. (2). In Fig. 5.5 the ratio hi/hcp is plotted as a function of wall temperature. The
horizontal dashed line at hi/hcp = 1 represents the constant-property situation. Going
from low to high wall temperature, it can be seen that first the simulations agree
with the constant-property relation Eq. (2) within 10%, i.e. the ratio hi/hcp is near
unity.
1.2
1.0
0.8
hi / hcp
0.6
P=80 bar
0.4
P=100 bar
0.2 P=120 bar
0.0
20 40 60 80 100 120
wall temperature Tw (°C)
81
On the right-hand side of the peaks in Fig. 5.5, impairment of heat transfer occurs,
i.e. the ratio hi/hcp falls below unity, especially at low pressures. This cannot be
caused by buoyancy, according to Eq. (7). It is caused by the variation of density
and specific heat in radial direction.
density-effect
At high wall temperatures, the density at the wall is significantly lower than in the
bulk and a low-density, gas-like fluid transports heat less effectively. The density
variation in radial direction is less prominent at higher pressure and so is the
resulting impairment of heat transfer.
A comparison between heat transfer coefficients from the simulations and from the
variable-property relation Eq. (6) is made in Fig. 5.6 for the same conditions as in
the 80 bar case of Fig. 5.5. It can be seen that Eq. (6) has an accuracy of ± 30 %.
82
an overestimation of heat transfer by Eq. (2) because of specific heat and density
variation in radial direction.
simulations
1.2
Eq. (6)
1.0
hi/hcp 0.8
0.6
0.4
0.2
0.0
20 40 60 80 100 120
Fig. 5.6. Normalized heat transfer coefficient as a function of wall temperature, for
80 bar and a bulk temperature of 30°C. Comparison of simulation results with Eq.
(6)
1.0
0.8
0.6
hi/hcp
300 bar,
0.4 112°C
80 bar,
0.2 33°C
0.0
50 100 150 200 250 300
pressure (bar)
Fig. 5.7. Simulation results showing the change in normalized heat transfer
coefficient during isochoric heating of carbon dioxide of 615 kg/m3, flowing at 0.3
kg/s in a 13 mm pipe
83
It should be noted here that Eq. (2), containing the specific heat at constant
pressure cp, is used for a process occurring at constant bulk density, where the
heating leads to a pressure increase. This is allowed because Eq. (2) is not used to
calculate the whole heating process from 33 to 112 °C, in which case the specific
heat at constant volume would be expected, but only the momentary heat transfer
coefficient. For a short piece of pipe, such as is simulated here, the bulk-pressure
can be regarded as constant and therefore the use of cp and of Eq. (2) is permitted.
Because in Fig. 5.4 to Fig. 5.7 no buoyancy is active, the influence of varying
physical properties could be investigated. Simulations were also carried out at 120
bar and large Grashof numbers so that not only forced but also free convection is
significant. It was observed earlier that the variation of physical properties has a
limited influence at 120 bar, so that in this case any significant deviation from the
constant-property situation will be caused by buoyancy. In the simulations wall and
bulk temperatures were 134 and 104°C respectively. Fig. 5.8 shows the resulting
heat transfer coefficients as a function of Reynolds number for a heated vertical
flow of CO2 in a 13 mm pipe, both for upward and downward flow.
It follows from Fig. 5.8 that, at large Reynolds number, there is no difference
between up- and down flow. At these conditions, buoyancy is negligible and the
constant-property equation is satisfactory for design purposes. At intermediate Re
values (from the dashed line down to Re = 40000), the heat transfer coefficient for
upward flow is lower than predicted by the constant-property relation Eq. (2) (hi/hcp
<1) while the down flow coefficient stays approximately constant. This is the effect
of the flattening of the velocity profile in the upward flow (Fig. 5.3a). Going from
Re = 40000 further down in Reynolds number, the heat transfer coefficients for
both up- and down flow become larger than predicted by Eq. (2), i.e. the ratio hi/hcp
is larger than unity. The enhancement is caused by the steeper shear stress profile
in the down flow (Fig. 5.2b) and by the W-shaped shear stress profile in the up
flow (Fig. 5.3b). In the buoyancy regime, to the left of the dashed line, the heat
84
transfer for the down flow is larger than for the up flow for the whole range of Re.
At very low flow rates (very low Re), as the regime of pure free convection is
approached, the difference between up- and down flow vanishes, which was also
reported by Jackson and Hall [3]. The limit for the onset of buoyancy from Eq. (7)
(GrRe-2.7 = 5⋅10-6) corresponds with the dashed line; the simulation results confirm
the validity of the buoyancy criterion.
4
heating, upward flow
0
1E+03 1E+04 1E+05 1E+06 1E+07
Re
Fig. 5.8. Influence of buoyancy on the normalized heat transfer coefficient for
carbon dioxide flowing in a vertical pipe at 120 bar. Comparison of simulation
results for up- and downward flow. The dashed line indicates the onset of buoyancy
(Eq. (7)).
In Fig. 5.9, the above discussed simulation results for the heated down flow are
plotted together with the heat transfer coefficients that are predicted by Eq. (8). The
figure shows that the agreement is good, confirming the validity of Eq. (8).
The results from the mixed convection simulations (Figs. 5.8 and 5.9) show that
buoyancy can increase heat transfer significantly relative to the constant-property
situation. For example, if a value of hi/hcp = 2 can be realized, Eq. (1) then predicts
an increase in the overall heat transfer coefficient Uo of 25%.
All simulations and model calculations were carried out for CO2 in 13 mm pipes.
However, the results presented here are also applicable for other systems (other
85
fluids and pipe diameters), as long as the relevant dimensionless numbers (Re, Pr
and Gr) of a regarded system are the same as the ones in this work.
4
Eq. (8)
simulations
3
hi / hcp
2
0
1E+03 1E+04 1E+05 1E+06 1E+07
Re
Fig. 5.9. Influence of buoyancy on the normalized heat transfer coefficient for
carbon dioxide flowing down in a heated vertical pipe at 120 bar. Comparison of
simulations with Eq. (8)
5.5. Conclusions
For a heated vertical flow of CO2 without buoyancy, the constant-property Nusselt
relation is not suitable for pressures below 120 bar as it can introduce an
overestimation of the overall heat transfer coefficient by 30 %. For such conditions,
the variable-property relation of Krasnoshchekov and Protopopov is acceptable for
design purposes.
The criterion that Jackson and Hall developed, to determine when buoyancy is
important, was confirmed by the CFD-results.
In the mixed convection regime in downward heated flow, the CFD simulations
gave results that agreed with the Jackson and Hall buoyancy correction factor
combined with the Krasnoshchekov-Protopopov relation. The action of buoyancy
can lead to an increase in heat transfer coefficient as large as 25 %.
86
A heat exchanger for heating should be placed in a vertical position with the fluid
flowing downwards when buoyancy is active. A vertical up flow or a horizontal
flow will always result in a smaller heat transfer coefficient.
References
13. S.S. Pitla, D.M. Robinson, E.A. Groll, S. Ramadhyani, Heat Transfer from
Supercritical Carbon Dioxide in Tube Flow: A Critical Review, HVAC&R
Research 4 (3), 281 (1998).
14. J.D. Jackson, W.B. Hall, Forced Convection Heat Transfer, in: S. Kakaç,
D.B. Spalding (Eds.), Turbulent Forced Convection in Channels and
Bundles, vol.2, 1979, 563.
15. J.D. Jackson, W.B. Hall, Influences of Buoyancy on Heat Transfer to
Fluids Flowing in Vertical Tubes under Turbulent Conditions, in: S.
Kakaç, D.B. Spalding (Eds.), Turbulent Forced Convection in Channels
and Bundles, vol.2, 1979, 640.
16. H.S. Swenson, J.R. Carver, C.R. Kakarala, Heat Transfer to Supercritical
Water in Smooth-Bore Tubes, ASME Journal of Heat Transfer nov. 1965.
17. A. Krasnoshchekov, V.S. Protopopov, Heat Exchange in the Supercritical
Region During the Flow of Carbon Dioxide and Water, Teploenergetika 6
(12), 26130 (1972).
18. T. Walisch, W. Dörfler, Ch. Trepp, Heat Transfer to Supercritical Carbon
Dioxide in Tubes with Mixed Convection, HTD-Vol. 350, vol. 12, ASME
1997, National Heat Transfer Conference, 79.
19. W.B. Hall, J.D. Jackson, Laminarization of Turbulent Pipe Flow by
Buoyancy Forces, ASME Paper no. 69-HT-55 (1969).
20. National Institute for Standards and Technology (NIST) Chemistry
Webbook, http://webbook.nist.gov/chemistry.
21. S.M. Liao and T.S. Zhao, Measurements of Heat Transfer Coefficients
from Supercritical Carbon Dioxide Flowing in Horizontal Mini/Micro
Channels, J. Heat Transfer 124, 413(2002).
87
)
88
Chapter 6
Abstract
A technical-scale, 100-litre machine was designed and built to test polyester beam
dyeing in scCO2, with respect to the performance of the equipment and the
process.
The design aims at introducing 1 mass percent of dye into 13 kg polyester cloth in
a 2-hour process. Per batch, 95% of the CO2 is recovered and stored for reuse. The
dyeing conditions are 300 bar and 120ºC.
A new pressure vessel was designed, consisting of a steel liner with carbon fibers
wound around it to take up the radial forces. The axial pressure forces are taken up
by an external steel yoke keeping both lids of the vessel in place.
To circulate the CO2 with the dissolved dye through the textile roll, a low-pressure
centrifugal pump was placed inside the dyeing vessel.
Although the equipment designed in this chapter is smaller, the process is the same
as in the industrial-scale 1000-litre machine discussed in chapter 7. This gives the
opportunity to test and optimize the performance of the process. The two most
costly items of an industrial-scale machine, the dyeing vessel and the circulation
pump, were both especially designed for the technical-scale machine in a cheaper
form, thereby facilitating the scale-up of the supercritical dyeing process.
89
6.1. Introduction
The dyeing machine developed in this chapter is large compared to the dyeing
machines described in literature [1] and in chapters 2 and 4 but is still an order of
magnitude smaller, in volume, than the industrial-size (1000-liter) machine
discussed in chapter 7. The aim of the currently discussed 100-liter dyeing machine
is testing the process with respect to:
1. Effect of heating and cooling rate. It is known from aqueous dyeing that
too fast a temperature rise or drop results in uneven dyeings of the PET. It
is to be tested if this phenomenon also puts a restriction on the heating and
cooling rate in scCO2.
2. Evenness of dyeing as a function of dyeing time, CO2-flow rate, beam
diameter and the diameter of the holes in the beam.
3. Dye dosing. An appropriate way of introducing and dissolving the dye
powder into the scCO2 is to be developed.
4. The leveling time. After all dye has diffused into the PET, the inside of the
textile roll is darker than the outside. When the CO2-flow is now
continued, the color will be equilibrated. It should be determined how long
this leveling takes.
90
5. The required time and temperature of rinsing. After dyeing, the CO2
contains residual dye that is to be flushed out of the vessel to avoid
contamination of the machinery and the cloth. The optimal rinsing time
and temperature must be determined.
6. Problems with deposition of cyclic trimers. It has been observed by
Montero et al. [2], that PET-dyeing in scCO2 can lead to deposition of
ethylene terephtalate trimers inside the dyeing machinery. This problem is
to be investigated with the 100-liter machine.
It should be noted that the machinery was designed and built, but no experimental
results are available at this moment. Both the process and the equipment were
designed in close cooperation with FeyeCon D&I B.V. (The Netherlands).
The machine is designed for one set of boundary conditions (table 6.1). The aim of
the process is to introduce 1 mass% of dye into 13 kg of polyester. The textile, as
rolled up on the beam, has a density of 500 kg/m3. Since chapter 4 showed
sufficient coloration at 120ºC and 300 bar, these values are taken as the working
temperature and pressure in the design.
91
6.3. Process steps
The consecutive steps of the process are given in table 6.2, together with the times
for which the corresponding equipment items are designed. The aim is a total
process time of 2 hours. Figure 6.2 gives a simplified process flow diagram.
dye condenser
dyeing holder
vessel
DYEING SOLVENT
CYCLE RECYCLE CO2
storage
vessel
Figure 6.2. Simplified process flow diagram for the supercritical beam dyeing
process
92
After the textile and the dye powder are loaded, the dyeing vessel is pressurized
with CO2. In the last part of the pressurization step, the circulation pump and the
heat exchanger are started. The CO2 is heated at the same time as new CO2 is
pumped into the machine until the desired temperature (120ºC) and pressure (300
bar) are reached.
During the dyeing step, the CO2 is circulated from the dyeing vessel, over the dye
holder, the heat exchanger and back into the dyeing vessel. In this way, dye
dissolution and PET dyeing take place simultaneously.
After 45 minutes, the dye dosing stops. In the leveling time that follows, the flow
rate, temperature and pressure are kept constant. After this step, the dye is
distributed over the PET and the scCO2. Depressurization would lead to
precipitation of dye on both the textile and the inner parts of the dyeing machine
and therefore the vessel is rinsed with fresh CO2. During the rinsing, clean CO2
from the storage vessel is pumped into the machine while CO2 containing residual
dye is removed through the reducing valve. The circulation pump and the heater
are still running at this time.
The reducing valve keeps the pressure in the dyeing vessel at 300 bar during the
rinsing step. The CO2 exiting the valve enters the separator, where it is gasified to
precipitate the dye in order to obtain a flow of clean CO2 in the direction of the
condenser.
The gaseous CO2 is condensed and stored in liquid form. During the rinsing, the
pressure in the CO2-recycling part is kept constant by keeping the temperature in
the storage vessel at the saturation temperature at 60 bar (22ºC), using the coolant
flowing through the condenser.
The aim of the rinsing step is to lower the dye concentration in the vessel, so that
depressurization does not lead to precipitation of dye onto the cloth and the
machinery. In the rinsing step, dye will be extracted from the fibers. This negative
effect is minimized by lowering the temperature from 120 to 80ºC while the
pressure is kept constant by the pressurization pump. However, as is illustrated for
5 dyes in appendix 6.1, this leads to a decrease in dye solubility of maximally a
factor 2. When the coloration of the polyester is to be above one half of the
93
maximum coloration, the CO2 at the beginning of the rinsing step contains dye
concentrations above one half of the solubility, following from the linearity of the
absorption isotherms in chapter 4. Cooling then results in precipitation of dye in the
machine and on the cloth. To avoid this, the first part of the rinsing time can be
carried out at 120ºC and only then the temperature is lowered to 80ºC.
When the dyeing machine is regarded as ideally mixed, the time t (s) required to
decrease the concentration from c0 to c in the vessel volume V (0.10 m3) by rinsing
with volume flow rate ΦV (pump flow 1.2·10-4 m3/s) is given by integrating the
mass balance:
dc
V = −Φ V c (6.1)
dt
from t = 0; c = c0 to t = t; c = c, leading to:
tΦ
c = c 0 exp − V (6.2)
V
From equation 6.2 it follows that the first 5 minutes of the rinsing have to be
carried out at 120ºC, the concentration is then c = 0.7 c0. The next 5 minutes,
rinsing and cooling occur simultaneously. In this way, the temperature of 80ºC is
reached at the moment that the dye concentration c = 1/2 c0, i.e. after 10 minutes of
rinsing. It is also calculated that a total of 20 minutes of rinsing leads to a decrease
in dye concentration of the scCO2 by 75%, i.e. c = 1/4 c0 . It is assumed that the
remaining 25% does not significantly affect the coloration of the next batch.
The process and equipment design allows changes in the times of heating, cooling,
dyeing, leveling and rinsing as well as changes in CO2 flow rate, beam dimensions
and in the dye dosing device, so that all six issues mentioned in section 6.1 can be
investigated.
94
pump. From B to C, the CO2 is pumped into the dyeing vessel. This is assumed to
be an isentropic process. Energy dissipation in the pump will bend the top of line
BC to the right. On the other hand, the CO2 in the dyeing vessel will lose part of its
heat to the steel and the polyester, bending the top of line BC to the left.
From point C on, the circulation and heating of the CO2 are started while the
pressurization pump still runs. The effect is a simultaneous pressure, temperature
and density increase, leading to point D that represents the dyeing conditions
(120ºC, 300 bar). From D to E, the CO2 is cooled isobaric from 120 to 80ºC; the
pressurization pump is used to simultaneously increase the density in the dyeing
vessel from 585 to 745 kg/m3.
During the rinsing period, the conditions inside the dyeing vessel remain those of
point E (80ºC, 300 bar). The CO2 that is used for the rinsing, follows the route
EFGABE (line BE is not drawn in the diagram). From E to F, the CO2
T (ºC)
D
supercritical
120
phase region
E
80
C critical point
31
liquid phase gas phase
22 G
region B A F region
gas-liquid two-phase region
Figure 6.3. Temperature-entropy diagram for CO2, showing the solvent cycle
during the supercritical dyeing process
passes through the regulating valve, undergoing an isenthalpic expansion into the
two-phase region, ending at the 60 bar isobar (line AG). The liquid mass fraction in
point F, being 0.4, represents the amount of CO2 that is to be evaporated in the
separator. During evaporation, the CO2 goes to the saturated gas line (point G). In
95
the condenser it crosses from G to A through the two-phase region. As liquid CO2
reenters the storage vessel, the cycle is closed.
After the rinsing, the pressure in the dyeing vessel is lowered through the
regulating valve until it is equal to the pressure in the storage vessel (60 bar). The
pressure in the dyeing vessel is then further reduced to 10 bar with the gas booster.
The final temperature in the dyeing vessel is estimated to be between –20 and
20ºC. This corresponds with a density of (21±2) kg/m3, giving a total amount of
95% of CO2 that is recycled per batch.
The vessel was designed by Solico B.V. (The Netherlands). The internal length and
diameter are 1350 and 350 mm. The volume available in the vessel for the scCO2
is 100 liters. The working and design pressures are 300 and 350 bar, respectively,
the working and design temperatures are 120 and 150ºC.
96
flanged CO2 exit connection
yoke plates
spindle
positioner
To load and unload polyester rolls, the yoke is lowered so that one of the lids can
be moved away from the vessel by means of a pneumatic linear drive underneath
the vessel. Attached to the lid, a cart on wheels, supporting the beam with textile, is
then also pushed out. Inside the vessel, the cartwheels are guided by rails, welded
onto a stainless steel shell that covers the lower 115º of the liner. Figures 6.5 A, B
give a schematic impression of the opened vessel and a cross-section of the cart
loaded with textile, respectively.
97
1 2 3 4 5 6 7
4
3
8
2
A B
Figure 6.5. Schematic representation of the dyeing vessel in open position, with a
longitudinal section (A) and a cross-section of the loaded cart (B)
When the yoke is lowered or raised, contact with the lids is to be avoided. A
distance of 3 mm is realized by pressing the lids against the vessel with:
• the pneumatic linear drive, on the beam-side and
• springs attached to the lid on the pump-side.
When the vessel is pressurized, the lids move 3 mm outwards until there is contact
with the yoke. On the lids, lip seals are placed, designed to slide along the internal
surface of the liner. Regarding the internals of the dyeing vessel, the movement of
98
connection for
electrical wiring
of pump
pump casing beam carbon fibre layer
motor casing
seal bearings polyester liner
magnetic spools
99
the lids is compensated by allowing the O-rings in figure 6.7 to slide 6 mm.
Because of the 3 mm movement of the pump-side lid, the CO2 entrance and exit of
the vessel are connected to the rest of the machine by flexible stainless steel hoses.
Materials of construction.
All parts of the vessel and the internals are made of stainless steel (AISI 316), with
the exception of:
• the lip seals, made from Rulon®, with PEEK back-up rings,
• the 4 O-rings in figure 6.7, made of a Viton® core with Teflon® jacket and
• the axes of the cartwheels are made of brass because this material is dry-
running. The use of steel is avoided here because the required grease on the
bearings would dissolve into the scCO2.
In the choice of the material for the liner and the composite, care is taken that the
strain of the liner is higher so that the pressure-induced deformation of the liner
will follow that of the composite. For the liner, duplex (stainless steel 1.4462) is
chosen since it has a large tensile strength, relative to other stainless steel types.
For the composite material, the standard carbon fiber type T 700 is taken because it
is cheaper and has a higher stiffness and tensile strength than other composite
materials, like aramid or glass fiber [4]. The matrix to keep the carbon fibers
together is a high-temperature epoxy resin, because it binds well with carbon fibers
and it can be applied at the design temperature of the dyeing vessel (150ºC).
The physical properties of the liner and the fiber material are given in table 6.3.
Calculations [4] show that at the working pressure of 300 bar, the stress in the liner
in tangential direction is 430 MPa, below the creep limit. At this pressure, the liner
is pressed outwards against the composite layer. At a pressure of 2070 bar inside
the vessel, the composite layer is at its tensile strength of 2450 MPa.
100
Table 6.3. Mechanical properties of the liner and the composite layer for the 100-
litre dyeing vessel [4]
Liner Material Duplex stainless steel
Thickness 10 mm
Creep limit 460 MPa
Fatigue limit (2·106 cycles) 565 MPa
Composite layer Material Carbon fiber in epoxy resin
Thickness 40 mm
Tensile strength 2450 MPa
Fatigue limit (2·106 cycles) 510 MPa
Pump capacity
At 120ºC and 300 bar, dye solubilities lie between 10-4 and 10-5 g dye / g CO2 [5-
8]. It is assumed here that:
• the CO2 is saturated, containing 10-5 g dye / g CO2, when leaving the dye
holder and
• the CO2 contains a negligible amount of dye when exiting the textile roll.
To introduce 1 mass% of dye into 13 kg PET, a total amount of 13000 kg CO2
needs to be circulated in the 45 minutes mentioned in table 6.2. Since the density of
scCO2 at 120ºC and 300 bar is 585 kg/m3, this means that a pump capacity of 30
m3/hour is needed. The resulting CO2 velocity in the 2” SCH 160 pipes is 5.7 m/s.
Pressure drop
The required head of the circulation pump is determined by several pressure drop
contributions. The calculations are given in appendix 6.2. It is concluded that the
total pressure drop over the machine, to be delivered by the circulation pump, is 2
bar.
The combination of capacity and head have led to the choice of a 3000 rpm
centrifugal pump with a maximum capacity of 30 m3/hour at 3 bar head, designed
in cooperation with and supplied by Packo Inox N.V. (The Netherlands). The
electric motor for the pump has a capacity of 7.5 kW, is controlled with a
frequency drive and was delivered by Electromotorenfabriek Nijmegen (The
Netherlands).
101
Materials of construction
All parts of the motor and pump are made of stainless steel (AISI 316 and 304),
with the exception of:
• The electrical copper spools of the motor.
• The resin in which the spools are fixed is a product from
Electromotorenfabriek Nijmegen (The Netherlands). This product was
tested at the Delft University of Technology and found to be not soluble in
scCO2 and not damaged by repeated depressurizations.
• The bearings. Since grease dissolves in scCO2, dry-running bearing are
used. The balls are made of silicium nitride and run in steel cages with a
dry-lubrication coating. The bearings are a product from SKF Nederland.
• The electrical wiring to the motor is copper, coated with Teflon®.
• The rotor is made from soft iron.
Because an air-driven piston-cylinder pump was found to be suitable for liquid CO2
in the 40-litre machine discussed in chapter 4, such a device was also chosen here.
The pump has to fill the vessel in 15 minutes (table 6.2) and rinse the vessel
sufficiently in 20 minutes. The latter duty determines the required flow rate,
following from equation 6.2 as 1.2·10-4 m3/s at 300 bar. This leads to the choice of
a type P200-65-2 from Resato Int. B.V. (The Netherlands). The pump has two
heads; both suction pipes are chilled in 1.5 meter double-pipe exchangers by water
from the cooling machine.
To prevent dye powder from reaching the dyeing vessel, which would lead to
staining of the textile, the dye has to be molecularly dissolved in the scCO2. A
device is used consisting of a combination of a filter housing and a powder dosing
device, as shown in figure 6.8.
102
After the dye has been dosed, it is kept dye powder falling
from entrainment by a stainless steel into the dye holder
candle filter, with pore size 10 µm.
The scCO2 flows past the dissolving candle filter
particles and through the filter
material towards the dyeing vessel.
The volume of the filter housing is 3
liters. The dosing device has yet to be
developed; the possibility of dosing CO2
powder into a pressurized machine has entering
been proven by Eggers et al. [9].
CO2 and dissolved
dye exiting
6.5.5. Heat exchanger
Figure 6.8. Dye dissolution unit with
As discussed above, each dyeing cycle dosing and candle filter
involves heating and cooling of the
CO2, polyester and steel. To save on
equipment cost, this is both done in one heat exchanger. Because in a dyeing
process the ease of cleaning of the tubes is important, a single-pass exchanger is
used. Two configurations are possible for this equipment item. Firstly, the steam
and cooling water compartments on the shell-side can be separated, as shown in
figure 6.8. Secondly, it is possible to send both the steam and the cooling water
through the same compartment, as is shown in figure 6.9.
In the case of figure 6.9, care should be taken to avoid flow of steam into the
cooling water system and vice versa, because cooling water systems are designed
for lower pressures than steam systems and because the presence of water in steam
pipes leads to water hammer. Therefore, figure 6.9 contains six more valves than
figure 6.8. Four valves separate the cooling water from the steam system and two
valves are needed to squeeze water out of the shell and into the drain, using
pressurized air, when the duty from the exchanger switches from cooling to
heating. The advantage of the configuration in figure 6.8 is an easier control; it is
therefore used in current industrial aqueous dyeing processes. The advantage of the
configuration shown in figure 6.9 is a smaller, cheaper heat exchanger and it is
therefore chosen for the 100-litre dyeing machine.
103
regulating valve
steam supply
partition
cooling water return
CO2 in
baffles
condensate return
regulating valve
cooling water supply
Figure 6.8. Possible configuration for alternate heating and cooling, with separate
compartments for steam and cooling water
pressurised air
condensate return
drain
Figure 6.9. Heat exchanger configuration for the 100-litre dyeing machine, with a
single shell-side compartment for both steam and cooling water
104
The temperature of the scCO2 in the dyeing vessel is measured and determines the
flow rate of the steam or cooling water into the heat exchanger via the
corresponding regulating valves. Since the heating and cooling times are yet to be
investigated, the heat exchanger is designed for relatively short, minimum heating
and cooling periods, in both cases 5 minutes.
Heating
The circulation pump, and thus the heating, starts when a pressure of 150-200 bar
is reached. It was concluded in chapter 5 that, for these conditions, the influence of
temperature-induced variations of physical properties in radial direction of a heat
exchanger pipe can be neglected, i.e. a constant-property relation for calculation of
the heat transfer can be used. Also buoyancy can be neglected according to chapter
5.
Walish et al. [10] measured that for Reynolds number larger than 105, upward,
downward and horizontal flows of scCO2 all have the same heat transfer
coefficients (±20%). Since the Reynolds number in the dyeing process is 6·105, a
horizontal heat exchanger for the 100-litre dyeing machine is designed, using the
constant-property equation 5.2. A heating time of 5 minutes is calculated when 23
pipes of 1/4” SCH 40 (external diameter 13.72 mm, wall thickness 2.24 mm, length
1950 mm) are used. A steam temperature of 150ºC is taken, the tube-side heat
transfer coefficient is 8000 Wm-2K-1, making the overall coefficient 2000 Wm-2K-1,
according to equation 5.1.
Cooling
The tube-side coefficient is 8000 Wm-2K-1. The water, from a cooling tower, has a
temperature of 20ºC. The shell-side coefficient is calculated with Kern’s method
[11a] to be 9000 Wm-2K-1. This value is an order of magnitude larger than what is
normal for a shell-side liquid. The reason is the small flow area between the pipes,
leading to large Reynolds numbers and fast heat transfer. The overall heat transfer
coefficient from equation 5.1 is now 2000 Wm-2K-1. The time needed to cool the
scCO2, polyester and steel from 120 to 80ºC is 5 minutes.
The resulting heat exchanger (figure 6.9) was designed in cooperation with and
supplied by GTI Siersema Procestechnologie (The Netherlands). It is built at an
105
incline of 1º to allow the condensate and cooling water to flow out. For the same
purpose, a space of 5 mm is left open between the baffles and the shell wall. To
allow thermal expansion of the exchanger, a bellow is implemented in the shell
wall. All parts of the machine are made of stainless steel (AISI 304), with the
exception of the Teflon® O-rings in the heads.
In the regulating valve, the pressure of the CO2 changes from 300 bar to 60 bar.
Since this process is fast and the piping is insulated, the expansion is isenthalpic, so
that the temperature-entropy diagram of carbon dioxide gives the temperature
(22ºC) and the mass fraction of liquid (0.4) in the CO2 downstream of the reducing
valve, flowing into the separator vessel.
The valve selection is based on the required flow coefficient KV, generally defined
as:
K V = Q G ∆p (6.3)
where Q is the volume flow (m3/hour), G is the specific gravity or the CO2-density
relative to the density of water at 20ºC and ∆p is the pressure drop over the valve
(bar).
106
6.5.7. Separator
The maximum amount of liquid that is to be boiled off by the steam jacket of the
separator is 0.05 kg/s. A conservative estimation
of the inside film coefficient of 800 W/m2/K [11b]
gives an overall heat transfer coefficient of 400
W/m2/K. With a steam temperature of 150ºC and
internal and external vessel diameters of 0.183 and
0.220 m, it follows that the height of the steam
jacket should be 0.2 m. To compensate for fouling
by dye powder, a height of 0.3 m is taken. The
content of the separator is 20 liters.
107
6.5.9. Cooling machine
As cooling medium, water with 10% ethylene glycol is used. Since the required
cooling capacity for the suction pipes of the pressurization pump is small, the
capacity of the cooling machine is determined by the heat flow through the
condenser, amounting 12 kW. A CH380 chiller from Piovan (Italy) is chosen, with
a capacity of 50 kW.
6.5.10. Condenser
The flow through the condenser during the rinsing is 0.084 kg/s, requiring a
condenser of minimally 12 kW. The outside film coefficient, following from the
Nusselt relation for condensation outside horizontal tubes [11c], is 2200 W/m2/K.
The overall coefficient becomes 1000 W/m2/K, so that the required pipe length is
20 m, when the cooling medium is 10ºC and flowing through a pipe of 16 mm
outside diameter and 2 mm wall thickness. Inside the condenser, two pipe spirals of
each 16 m are placed concentrically, as shown in figure 6.11.
To save on equipment cost, the condenser and the storage vessel are combined into
one structure. The storage vessel has a volume of 250 liters, and working and
design pressures of 60 and 100 bar, respectively. The liquid level in the storage
vessel is detected with a differential pressure transmitter.
All parts of the condenser and the storage vessel are made of stainless steel (AISI
316), with the exception of the Teflon®/steel flange seal.
108
CO2-gas entrance
2 exits for
cooling liquid
inner cooling spiral
outer cooling spiral
2 entrances for
cooling liquid
flange connection between
gas condenser and storage vessel
liquid
CO2-liquid exit
Figure 6.11. Integrated condenser-storage vessel unit for the 100-litre dyeing
machine
Table 6.4. Solubilities (g/m3) of several dyes in supercritical carbon dioxide at 300
bar, for different temperatures (T)
Dye T (ºC) Reference
80 120
Solvent Brown 1 62 73 [12]
Disperse Blue 79 260 425 [13]
Disperse Red 324 36 45 [13]
Disperse Red 153 21 21 [14]
Disperse Yellow 119 12 26 [14]
109
Appendix 6.2 Pressure drop calculation for the 100-litre beam-dyeing
machine.
All pressure drop contributions are calculated for rinsing conditions: 300 bar, 80ºC.
dp µΦ V
= (6.4)
dr 2παL
Integrating equation 6.4 from the inside (radius r) to the outside (radius R) of the
textile roll gives the pressure drop ∆p:
µΦ V R
∆p = ln (6.5)
2παL r
110
With r = 0.03 m and R = 0.14 m, it follows that the pressure drop contribution from
the textile roll is 0.29 bar.
Φ V sµ
∆p = (6.6)
Aα
with s = filter layer thickness (0.0008 m), A = filter surface (0.04 m) and α is the
permeability coefficient of the filter material (9·10-12 m2). The contribution to the
pressure drop is relatively small: 0.04 bar. The joint pressure loss in the entrance
and exit of the dye holder (total KW = 0.5 [17b]) and in the sharp 90º bend inside
(KW = 1.5) is 0.2 bar. The total pressure drop over the dye holder is 0.24 bar.
111
entrance and exit effects is 0.05 bar, giving a total pressure drop over the heat
exchanger of 0.35 bar.
The total pressure drop in the 100-litre dyeing machine is concluded to be 2 bar.
References
112
10. T. Walish, W. Doerfler and Ch. Trepp, Heat Transfer to Supercritical
Carbon Dioxide in Tubes With Mixed Convection, HTD volume 350,
ASME National Heat Transfer Conference 12 (1997) 79.
11. J.M. Coulson, J.F. Richardson and R.K. Sinnot, Chemical Engineering
vol.6: Design, 2nd ed., Pergamon Press, Oxford, 1993, a600, b569, c636.
12. A. Ferri, M. Banchero, L. Manna and S. Sicardi, A new Correlation for the
Solubilities of Azoic and Anthraquinone Derivatives in Supercritical
Carbon Dioxide, J. Supercrit. Fluids 32(1-3), 27 (2004).
13. A. Ferri, M. Banchero, L. Manna and S. Sicardi, An Experimental
Technique for Measuring High Solubilities of Dyes in Supercritical Carbon
Dioxide, J. Supercrit. Fluids 30 (1), 41 (2003).
14. H.-M. Lin, C.-Y Liu, C.-H Cheng, Y.-T. Chen and M.-J. Lee, Solubilities
of Disperse Dyes of Blue 79, Red 153 and Yellow 119 in Supercritical
Carbon Dioxide, J. Supercrit. Fluids 21(1), 1 (2001).
15. W.A. Rapp, Y. Yang, G. O’Neal, H. Boyter and G. Hughey, Evaluating
Liquor Flow Profiles in Fabric Beam Dyeing, AATCC Review 2, 42 (2002).
16. J.M. Smith, E. Stammers and L.P.B.M. Janssen, Fysische
Transportverschijnselen I, Delftse Uitgeversmaatschappij, 6th ed., Delft,
1989, a42, b47.
113
114
Chapter 7
7.1. Introduction
In the previous chapters, it has been shown that textile dyeing in supercritical
carbon dioxide is technically possible. Whether or not it is also economically
attractive is investigated in this chapter by comparing the cost of a polyester beam
dyeing process in supercritical CO2 with the cost of the same process in water.
Dyeing methods other than beam dyeing are possible (e.g. jet-dyeing, drum dyeing)
but with beam dyeing, the ratio of textile volume and bath volume is the largest. A
small bath volume per unit textile volume means smaller equipment and less CO2
to recycle over the storage vessel. In other words, for a supercritical process to
have low investment- and operating-costs, a beam dyeing process is the best
option.
The raw material of a polyester finishing plant is usually PET fibers contaminated
with spinning oil, an additive that functions as a lubricant during spinning, weaving
and knitting. After the PET is woven or knitted, the spinning oil has to be removed
from the cloth by passing it through a washing machine. This step is the same for
aqueous and supercritical dyeing and therefore these washing costs are left out of
consideration in the comparison of both processes. Before supercritical or aqueous
dyeing, the polyester has to be heat-set to avoid excessive shrinking of the textile in
the dye bath, due to thermal effects. Also this step is identical for the aqueous and
the supercritical process and is left out of consideration.
115
It is assumed that pressurised air, steam and cooling water from a tower are present
at the site. The use of these utilities is incorporated in the evaluation, but the
required compressor, steam boiler and cooling tower are left out of consideration.
The comparison between both processes is done for dyeing batches of 300 kg PET,
an amount that is in the same order as in current industrial practice. The density of
solid polyester is 1380 kg/m3; rolled upon a beam in woven or knitted form it is
250-750 kg/m3; in this work it is taken to be 500 kg/m3 in the roll. The perforated
pipe (beam) onto which the PET is rolled is assumed to be 0.1 m in diameter for
the supercritical and 0.3 m for the aqueous process. Table 7.1 gives an overview of
the boundary conditions for both processes.
Table 7.1. Boundary conditions for the supercritical and aqueous dyeing process
scCO2-dyeing aqueous dyeing
For the industrial supercritical dyeing process, the same process steps and times are
used as in chapter 5, leading to a total batch time of 2 hours. Table 7.2 gives an
overview. The same process flow diagram applies as was used in chapter 6.
116
Table 7.2. Process steps and times for polyester dyeing in supercritical carbon
dioxide
Process step time (min)
Loading textile and dye 3
Pressurisation 15
Dyeing 45
Levelling 15
Rinsing 20
Depressurisation 20
Unloading textile 2
Total batch time 120
The pressure drop calculations are done in the same manner as in chapter 6, leading
to a drop of maximally 0.3 bar over the PET roll and 0.6 bar over the piping. The
pressure drop over the heat exchanger is 0.2 bar and also over the dye holder 0.2
bar. The total pressure drop, to be overcome by the circulation pump, is 1.3 bar.
Dyeing vessel
A dyeing vessel is used with a stainless steel liner reinforced by carbon fibres
wound around the circumference, thereby taking up the radial pressure forces. The
two lids are kept in place by a steel yoke that takes up the axial pressure forces.
117
The length of the vessel is 3 m: 2 for the PET roll and 1 m for the pump and motor
discussed below. As the textile roll is 0.63 m, the inner diameter of the dyeing
vessel is taken as 0.7 m.
The advantage of using a carbon fibre vessel instead of an entirely steel vessel is
twofold. Firstly, a steel vessel of inner diameter 0.7 m and a design pressure of 350
bar would have a wall thickness of 0.11 m, making the vessel mass 7000 kg,
without yoke or closure. For heating and cooling the CO2, PET and steel, there is
now twice the amount of steam and cooling water per batch needed. The second
disadvantage of a steel pressure vessel is that it requires a heat exchange area twice
as large as the carbon fibre vessel, if the heating and cooling times are to be equal
to the carbon fibre case. The extra costs for cooling water and steam and for the
larger heat exchanger lead to an increase of 6 % in the process cost relative to the
carbon fibre case.
Circulation pump
The choice and price of the circulation pump is determined by the combination of
flow (145 m3/hour) and pressure drop (1.3 bar) and by the working pressure of 300
bar. The high flow means that a centrifugal pump is required. The high working
pressure demands that a magnetic coupling be used between the pump and its
motor. Such a pump would cost approximately 100 k€ [2] and weighs heavily on
the investment cost. An alternative is to use a standard low-pressure centrifugal
pump and place it inside the pressure vessel, together with the motor. The cost of
pump and motor is now reduced to 20 k€ [3]. It can be calculated from the tables
below that the total process cost is then reduced with 5%. Dry-running bearings are
used and a special resin for the electrical spools of the motor, as was discussed in
chapter 6.
118
Dye holder
The pressure vessel where the dye powder is dissolved in the scCO2 is a stainless
steel filter housing of 200 mm internal diameter and height 600 mm. The dye is put
into the vessel, upstream of the filter, so that the scCO2 flows past the dye particles
and then through the filter.
Heat exchanger
As was discussed in chapter 6, there are 2 possible configurations for the heat
exchanger: one in which the steam and cooling water flow through the same shell-
side space and one in which the steam- and cooling water parts are segregated. The
latter configuration requires less valves,is easier to use in an industrial environment
and is therefore chosen here.
The heat exchange area for heating and cooling the scCO2, steel and PET is
calculated in the same manner as in chapter 6. As is shown in chapter 5, the heat
transfer coefficients for heating and cooling can be calculated with constant-
property Nusselt relations. Both the steam- and the cooling water part of the shell
have a heat exchange area of 3.4 m2. Heating and cooling can be accomplished in
10 minutes if 50 stainless steel pipes of type ½” shedule 40 are used. Both the
heating and the cooling part should be 1 m in length, the shell diameter is 250 mm.
Separator
The precipitation and collection of residual dye is done in a separator vessel. By
means of a steam jacket, the liquid fraction of the CO2 that results from the
expansion from 300 to 60 bar is boiled off and flows to the storage vessel. The
separator is designed in the same way as in chapter 6, resulting in a vessel with an
internal diameter of 0.3 m and a height of 1 m.
119
Condenser
The condenser is designed to liquify the flow of CO2 leaving the separator during
the rinsing step: 0.83 kg/s. The same condenser configuration is taken as in chapter
6: CO2 flowing on the shell-side and the cooling water on the tube-side. With a
cooling water temperature of 0ºC, the required area of heat exchange is 6.4 m2. The
calculation procedure is similar to the one in chapter 6.
Cooling machine
The cooling water needed for the condenser is brought to 0ºC with an electrical
cooling machine. The required capacity is estimated to be 140 kW.
Storage vessel
The CO2 is stored at 22ºC and 60 bar, in liquid form. The required pressure vessel
has a volume of 1200 liters.
The capital cost of the main plant items discussed above are listed in table 7.3.
Table 7.3. Investment cost for the main plant items of an supercritical dyeing
machine
Main plant item Capacity Purchase cost (k€)
Dyeing vessel 1150 liter 90a
3
Circulation pump 150 m /hr 20b,c
Pressurisation pump unit 70 liter/min 20d
Dye holder 20 liter 20e
Heat exchanger 2 x 3.4 m2 30e
Separator 70 liter 20e
Gas booster unit 50 liter/min 14f
Condenser 6.4 m2 20e
Cooling machine 140 kW 55g
Storage vessel 1200 liter 20h
Total main plant items: 309 k€
120
Price estimations from: aSolico B.V., bPacko Inox N.V., cDACE Price booklet,
d
Resato Int. B.V., eEstimations based on costs of the 100 liter pilot machine from
chapter 6, fMaximator GmbH, gPiovan S.p.A., hVan Steen Apparatenbouw.
Table 7.4. Overview of the secondary equipment costs for a supercritical dyeing
machine
Cost (k€)
Instrumentation and control 25a
Piping 10b
Valves 20a
Frame and isolation 20a
Working hours 14c
Total secondary costs: 89 k€
a
Estimations based on costs of the 100 liter pilot machine from chapter 6, bNoxon Stainless B.V., c10 weeks à
€35/hr.
From tables 7.3 and 7.4 it follows that the cost of producing the supercritical
dyeing machine is 398 k€. If an overhead of 25 % is assumed, the purchase cost for
a textile dyer would be 500 k€ for a beam dyeing machine for 300 kg polyester
batches. This is to be compared with the purchase cost of a machine for aqueous
beam dyeing: 100 k€ [4].
For both the supercritical and the aqueous process, the annual capital charge a is
calculated with the annuity equation:
i
a= (7.1)
1 − (1 + i) − N
121
Table. 7.5. Capital costs for polyester beam dyeing in scCO2 and water
CO2 Water
Investment k€ 500 100
Annual capital charge k€ / year 70 14
Batch time hour 2 3.5a
Production capacity ton / year 525 300
Capital charge € / batch 40 14
€ / (kg PET) 0.13 0.047
a
Data from Ames Europe B.V. and Stork Prints B.V.
The prices of the different utilities and compounds that are needed to operate a
dyeing machine are listed in table 7.6.
The use of utilities and compounds is calculated from the specifications of the
equipment and the demands of the process. The heat balance is made with the data
from table 7.7.
122
Table 7.7. Data for the heat balances of the dyeing processes
Material Specific heat Heat of Amount in dyeing machine (kg)
(J/kg/ºC) vaporisation CO2-process Aqueous
(kJ/kg) process
Steel 460 1700 1000b
PET 1300 300 300
CO2 2200a 140 560 0
c
Water 4200 2200 0 1000b
a
Average value, taken between 20 and 130ºC. bData from Stork Prints B.V. c At 150ºC
Supercritical process
In table 7.8 the calculated operational costs are shown for the supercritical process.
Table 7.8. Operational costs per batch for polyester beam dyeing in scCO2
Item Amount per batch Price (€)
3
Pressurised air 1110 Nm 7.55
Steam 137 kg 3.55
Electricity 105 kWh 10.50
3
Cooling water 16 m 0.80
CO2 20 kg 2.20
Dyes 6 kg 60
a
Maintenance 1.2
Labour costb 0.67 hour 18.67
Total operating cost: 104.47 € / batch
0.35 € / (kg PET)
a
Maintenance is 3 % of the capital cost. b 1 operator for 3 machines
The steam consumption in the table is the sum of the amount of steam needed to
heat all steel, PET and CO2 and the amount that is needed to vaporise the liquid
fraction of CO2 that enters the separator during the rinsing step of the process.
The electricity use of the process is calculated by adding the use of the motor of the
centrifugal pump and the use of the cooling machine.
The cost of the dye powder in table 7.8 is based on the current price of dye for
aqueous dyeing: 15 €/kg. This includes dispersing agents and since these are not
123
needed in supercritical dyeing, the dye will be cheaper. It is therefore roughly
estimated to be 10 €/kg. It follows from the distribution coefficients given in
chapter 4 that the amount of dye remaining in the scCO2 after the dyeing process is
small compared to the amount in the PET; it is therefore neglected.
Aqueous process
The dye for the aqueous process contains around 50 % actual dye molecules, the
rest is mainly dispersing agents. Therefore, to obtain 2 mass% of dye into the 300
kg of PET, 12 kg dye is needed. Also in the aqueous process, the dye concentration
in the machine after the process is neglected.
After the dyeing step, the PET is washed in 1000 liter water with surfactants at
90ºC and subsequently rinsed 3 times with 100 liter water at 60ºC. The time
required for dyeing, washing and rinsing together is 3.5 hours [4]. Finally, the PET
is dried, using steam. The resulting operating costs are given in table 7.9.
Table 7.9. Operating costs per batch for aqueous polyester beam dyeing
Process step Item Amount per batch Price (€)
Maintenance 3
Dyeing Steam 253 kg 6.58
Electricity 50 kWha 5
Water 1000 litera 2.27
Dyes 12 kg 180
b a
Chemicals 3 kg 9
Labour cost 32.67c
Washing and rinsing Steam 319 kg 8.30
a
Water 4000 liter 9.08
Chemicals 12.3a
a
Electricity 50 kWh 5
Drying Steam 910 kg 23.66
Total operating costs: 297 €/batch
0.99 € / (kg PET)
a
Information from Stork Prints B.V. bLevelling agents and pH modifiers. c Labour cost for dyeing, washing and
rinsing together, 1 operator for 3 machines.
124
7.7. Total process costs
The costs for supercritical and aqueous dyeing are summarised in table 7.10. It is
clear that the capital costs are higher and the operating costs are lower for the
supercritical process. Because the process costs are largely determined by the
operating costs, the process costs are lower when polyester is dyed in scCO2.
Table 7.10. Comparison of costs (€ / (kg PET) for supercritical and aqueous
polyester dyeing
CO2 Water
Capital cost 0.13 0.047
Operating cost 0.35 0.99
Process cost 0.48 1.04
For the process time, 2 hours was taken in the above calculations. It is possible
that, in industrial practice, the time will be longer. It will, however, always be
shorter than for the aqueous process. To illustrate the influence of the time on the
costs of the supercritical process, the calculations were performed also for a time of
3 hours. The capital and operating costs now become 0.20 and 0.42 €/(kg PET)
respectively. The total process cost increases from 0.48 to 0.64 when the time
changes from 2 to 3 hours.
In the calculation of the operating costs for supercritical dyeing, the price of dye
was estimated roughly to be 10 €/kg, since there is no dye commercially available
for dyeing in scCO2. The value must lie between the price of the press cake, the
unpurified product of dye synthesis (4 €/kg) and the price of dye for aqueous
dyeing, containing dispersing agents (15 €/kg). This means that the process costs
for the supercritical process lie between 0.36 and 0.58 € per kg polyester.
In the supercritical process evaluated above, the dye remaining in the CO2 after
dyeing, is separated from the CO2 by vaporisation of the latter. This separation
technique requires continuous vaporisation, condensation and pressurisation of the
CO2 during the rinsing step. An energetically prefferable method is to remove the
125
dye from the CO2 by adsorption onto a medium like activated carbon. This
separation technique was investigated for dye and scCO2 on a laboratory scale by
Von Schnitzler [5], with bleaching earth as adsorptive material. It is unknown how
this would affect the process time and it is also unknown what the price of
regeneration or disposal of the adsorptive medium is. Therefore, at this moment,
the costs of a supercritical process with this alternative separation cannot be
estimated.
It will be clear that the main environmental advantage of supercritical dyeing is the
reuse of 95 % of the solvent. Because the CO2 is obtained as a waste product from
industrial processes like combustion or ammonia synthesis, the venting of the 5 %
CO2 is not a pollution on the account of textile dyeing. As discussed in chapter 1,
in polyester dyeing 5-10 % of the dye is not used and ends up in the wastewater. In
scCO2-dyeing, all dye can be reused, which forms a second environmental
advantage of the supercritical process. In short, the supercritical process is
environmentally superior to the aqueous process when the use of compounds is
regarded. Table 7.11 gives an overview of the use of compounds and energy in
both the supercritical and the aqueous process. The disperse dyes that are used in
126
supercritical dyeing are the same as in aqueous dyeing, so that the human and
ecological toxicity of these compounds does not play a role in the comparison of
the processes.
It can be seen in table 7.11 that different types of energy are used in both dyeing
processes. The use of cooling water is neglected here. For a comparison with
respect to energy consumption, it is calculated how much carbon dioxide is formed
in the production of the pressurized air, steam and electricity. The generation of 1
Nm3 pressurised air of 7 bar requires 0.125 kWh [6], the CO2-production for 1
kWh electricity is 0.61 kg [7] and the CO2-production for steam formation is 118
kg CO2 / ton steam [7]. The resulting amounts of carbon dioxide associated with
the energy use in the dyeing processes are given in table 7.12. Supercritical dyeing
requires less energy and is therefore associated with a 26% lower CO2-emission.
Table 7.11. Environmental comparison between the supercritical and the aqueous
polyester dyeing process, regarding the uses of compound and utilities
compound/ unit CO2 water
utility
water kg 0 5000a
CO2 kg 20 0
dye pure dye kg 6 6
dispering agent kg 0 6
chemicals kg 0 3
3
pressurized air Nm 1110 0
steam kg 137 1380c
electricity kWh 105 100
3
cooling water m 16 0
a b c
For dyeing, washing and rinsing. 5% of the amount of CO2 that is used per batch (560 kg). For dyeing,
washing, rinsing and drying.
127
Table 7.12. Carbon dioxide emission associated with the energy used in
supercritical and aqueous dyeing, per 300 kg polyester batch
energy equivalent CO2-emission (kg CO2)
supercritical dyeing aqueous dyeing
steam 16 163
electricity 64 61
pressurised air 85 0
total 165 224
7.9. Conclusions
From the process and equipment design for beam dyeing of 300 kg batches of
poyester in scCO2, it follows that the purchase cost for the machine is 500 k€. The
equivalent machine for dyeing in water costs 100 k€. The large difference in
investment cost is caused by the high pressure of the supercritical process and by
the fact that 96 % of the CO2 is recycled.
The operating costs for the supercritical process are lower than for the aqueous
process: 0.35 and 0.99 € per kg of polyester, respectively. This is mainly caused by
the higher rate of dyeing and the lower price of dye for the supercritical process.
The overall effect is that it is more than twice as cheap to dye polyester in scCO2
than it is in water. The process costs are 0.48 and 1.04 € per kg of polyester,
respectively.
The economic evaluation was based on a supercritical process where the residual
dye is separated from the CO2 by gasifying the latter. When, instead, the dye is
removed by adsorption, the supercritical process could be made cheaper. At this
moment, no estimation can be made as to how much cheaper.
128
References
129
130
Epilogue
As was concluded in this thesis, one of the main reasons that the supercritical
process is economically superior to the aqueous process, is the higher rate of
dyeing. It is therefore recommendable to investigate the kinetics of both disperse
and reactive dyeing. Knowledge on the kinetics and the identification of the rate-
determining step in non-reactive and reactive dyeing open the possibility of further
reducing the process times.
Regarding the equipment, both the rotating drum machine (chapter 4) and the beam
dyeing machine (chapter 6) are to be tested for their applicability in reactive
dyeing. The process time and the evenness of the beam have to be measured and
optimized in the 100-liter dyeing machine described in chapter 6, both for disperse
and reactive dyeing. Finally, an equipment item should be designed, built and
tested to introduce the dye powder into the pressurized dyeing machine.
It is clear that a large driving force exists for the development and implementation
of textile dyeing in scCO2. Water pollution is becoming an ever greater
environmental concern and it is likely that the regulations on the quality of treated
wastewater will become more stringent in the future. Therefore, the end-of-pipe
treatments will require more effort and cost, so that the economical advantage of
scCO2 will become even greater. Both the environmental and the economical
incentive will increase the demand for a clean dyeing process.
For the implementation of a new dyeing process in industry, the environmental and
economical implications, the product quality (color depth, fastness) and the
131
robustness of the process in practice have to be ascertained. Regarding polyester
dyeing, the environmental advantage is clear. The second two issues were
investigated satisfactorily in this thesis. The development of the 100-liter polyester
beam dyeing machine, described in chapter 6, creates the possibility of carrying out
large-scale experiments to test the robustness of the process. Only then can a full-
scale 1000-liter pilot machine be designed, built and tested in an industrial
environment.
132
Dankwoord
Tijdens de vier jaren van mijn promotieonderzoek op API heb ik met veel mensen
samengewerkt. De mede- “API-bewoners” zorden ervoor dat het niet alleen een
leerzame en productieve tijd was, maar ook een vrolijke.
Met Wim Veugelers heb ik het genoegen gehad veel samen te werken. Een
genoegen vanwege je efficiënte manier van werken en je praktijkervaring, Wim,
maar ook omdat je een bijzonder prettige collega bent, altijd bereid tijd vrij te
maken om anderen te helpen.
Jan, Martijn, Carel, Stefan, André, Jos en Gerard: Hartelijk dank voor de inzet bij
het tekenen en construeren van de opstellingen en jullie praktische adviezen.
Bovendien wil ik jullie bedanken voor de vele keren dat er reparaties en
veranderingen aan de apparatuur moesten worden uitgevoerd.
Een grote hulp bij het onderzoek waren Marjan, Özcan, Frans, Elmehdi en Zinaida.
Hun werk, in de vorm van experimenten en CFD-simulaties, is overal in dit
proefschrift terug te vinden.
Bij de analyses van de geverfde lappen is meermalen de kennis van Michel en Paul
van grote waarde gebleken. Hen wil ik bedanken voor de hulp bij zaken als
chemicaliën en analytische apparatuur.
Tijdens mijn gehele promotie heb ik veel geleerd van de mensen die vanuit Stork
Prints B.V. op het project werkten. Vooral Jan-Willem Gerritsen en Peter
Leerkamp bedank ik hiervoor. Ook de medewerkers van Ames Europe en Triade
wil ik hierbij bedanken voor hun bijdragen. Ludo en Hans van Solico zijn we
dankbaar voor het prachtige ontwerp van het koolstofvezelversterkte verfvat.
133
Vanaf het begin van mijn API-tijd heb ik veel steun en ideën gekregen van de
medewerkers van FeyeCon D&I B.V. Behalve Geert en Wim zijn Ernst, Frank,
Gerard en Hubert met hun ervaring op superkritisch gebied behulpzaam geweest.
Bedankt!
Uschi, jouw steun is heel belangrijk geweest de afgelopen jaren. Ook voor je
geduld wil ik je bedanken, er ging nogal wat tijd en energie in deze klus zitten. Nu
hebben we gelukkig weer meer tijd voor elkaar.
Tot slot wil ik opmerken dat ik zonder de goede zorgen van mijn ouders nooit tot
een promotie gekomen zou zijn. Bedankt voor alles!
Martijn.
134
Curriculum Vitae
135
136
Publications
M. van der Kraan, M.M.W. Peeters, M.V. Fernandez Cid, G.F. Woerlee, W.J.T.
Veugelers and G.J. Witkamp The Influence of Variable Physical Properties and
Buoyancy on Heat Exchanger Design for Near- and Supercritical Fluids, J.
Supercrit. Fluids 34, 99 (2005).
M. van der Kraan, M.V. Fernandez Cid, G.F. Woerlee, W.J.T. Veugelers and G.J.
Witkamp, Equilibrium Study on the Disperse Dyeing of Polyester in Supercritical
Carbon dioxide, submitted to The Textile Research Journal.
M. van der Kraan, M.V. Fernandez Cid, G.F. Woerlee, W.J.T. Veugelers and G.J.
Witkamp, Dyeing of Natural and Synthetic Textiles in Supercritical Carbon
Dioxide with Disperse Reactive Dyes, submitted to The Journal of Supercritical
Fluids.
M. van der Kraan, M.V. Fernandez Cid, G.F. Woerlee, W.J.T. Veugelers, C.J.
Peters and G.J. Witkamp, Equilibrium Distribution of Water in the Two-phase
System Supercritical Carbon Dioxide – Textile, submitted to the Journal of
Supercritical Fluids.
M. van der Kraan, O. Bayrak, M.V. Fernandez Cid, G.F. Woerlee, W.J.T.
Veugelers and G.J. Witkamp, Textile Dyeing in Supercritical Carbon Dioxide,
Proceedings of the 6th International Symposium on Supercritical Fluids, Versailles,
France, 2003, 2119.
M. van der Kraan, M.M.W. Peeters, M.V. Fernandez Cid, G.F. Woerlee, W.J.T.
Veugelers and G.J. Witkamp, Heat Transfer in High-pressure Carbon Dioxide,
Proceedings of the 6th International Symposium on Supercritical Fluids, Versailles,
France, 2003, 517.
M. van der Kraan, M.V. Fernandez Cid, G.F. Woerlee, W.J.T. Veugelers and G.J.
Witkamp, Dyeing Natural and Synthetic Textiles in Supercritical Carbon Dioxide
137
with Reactive Dyes, Proceedings of the 4th International Symposium on High
Pressure Technology and Chemical engineering, Venice, Italy, 2002, 199.
M. van der Kraan, E. Abdaou, M.V. Fernandez Cid, W.J.T. Veugelers, G.F.
Woerlee, G.J. Witkamp, Maximum attainable Colouration of Polyester in
Supercritical Carbon Dioxide, Proceedings of the 7th Italian Conference on
Supercritical Fluids and their Applications, Trieste, Italy, 2004.
138