You are on page 1of 12

Textile Research Journal

http://trj.sagepub.com Application of Contact Angle Measurement to the Manufacture of Textiles Containing Microcapsules
Fabien Salan, Eric Devaux, Serge Bourbigot and Pascal Rumeau Textile Research Journal 2009; 79; 1202 DOI: 10.1177/0040517508100724 The online version of this article can be found at: http://trj.sagepub.com/cgi/content/abstract/79/13/1202

Published by:
http://www.sagepublications.com

Additional services and information for Textile Research Journal can be found at: Email Alerts: http://trj.sagepub.com/cgi/alerts Subscriptions: http://trj.sagepub.com/subscriptions Reprints: http://www.sagepub.com/journalsReprints.nav Permissions: http://www.sagepub.co.uk/journalsPermissions.nav Citations http://trj.sagepub.com/cgi/content/refs/79/13/1202

Downloaded from http://trj.sagepub.com by Ngoc Nguyen Thi Thuy on October 22, 2009

Textile Research Journal

Article

Application of Contact Angle Measurement to the Manufacture of Textiles Containing Microcapsules


Abstract The efficiency of a binder to link microcapsules on a textile surface depends on the compatibility of the different interfaces of the products involved in the coating process. The choice of a binder adapted to the microcapsules was determined in this study by the comparison of the surface energy components induced by the contact angle measurement method and washing tests. It was found that a polyurethane-based binder was the most suitable to link melamine formaldehyde microcapsules. Furthermore, the adhesion of microcapsules was closely dependent on the chemical nature and structure of the textile support.

Fabien Salan1 and Eric Devaux


Laboratoire de Gnie et Matriaux Textiles (GEMTEX), UPRES EA2461, Ecole Nationale Suprieure des Arts et Industries Textiles (ENSAIT), BP 30329, 59056 Roubaix Cedex 01, France

Serge Bourbigot
Laboratoire des Procds dElaboration des Revtements Fonctionnels (PERF), UPRES EA1040, Avenue Dimitri Mendeleev Bt. C7a, BP 108, 59652 Villeneuve dAscq Cedex, France

Pascal Rumeau
Institut Franais du Textile et de lHabillement, Direction Rgionale Rhne-Alpes PACA, Avenue Guy de Collongue, 69134 Ecully Cedex, France

Key words binder, coating textiles, microencapsulation, surface energy, washing test

For the last decade, functional textiles have been developed to enhance textile performances according to the consumers demand and to include a large range of properties with a higher added value. One of the possible ways to manufacture functional textile products is the incorporation of microcapsules or the use of microencapsulation processes for textile finishing. Although microencapsulation has been in use for a long time in many industries, including pharmaceutical, carbonless copy, agricultural, food processing, cosmetics and bulk chemistry, it has only recently been introduced in the textile industry [1]. Many of the substances are encapsulated for potential textile applications. During the 1990s, this technique made progress in the field of phase change materials (PCMs) for thermoregulation [2], in durable fragrances for aromatherapy or controlled release [3], and in cosmetics for cosmetotextiles which can impart skin care benefits and promote a feeling of well-being. Other applications developed include dyes (thermochromic, photochromic, reactive or dispersed) [4], insect repellents [5], antimicrobials [6], and fire retardant or intumescent compounds [7]. Microencapsulation is a technique that allows an active substance to be entrapped by a suitable polymer wall on a very small scale. The functional performance of the microcapsules depends on the morphology, the chemical nature

and the surface characteristics of the polymeric shell influenced by the process parameters [8]. The choice of a particular process is determined by the solubility characteristics of the active compound and the shell material. Thus, the formulation of sophisticated shell and the development of the technologies allow a wide variety of functionalities to be obtained. The step of encapsulation allows the active substance to be integrated in a textile coating, in a textile by impregnation or exhaust bath [9], or directly incorporated in different artificial fibers, e.g. polyacrylonitrile fiber [10], polypropylene fiber [11], or polyacrylonitrile-vinylidene chloride fiber [12]. They will remain effective as long as the coating or the fibers stay intact [1315]. All common coating processes, such as knife over roll, knife over air, or screen printing, can be adapted to apply microcapsules to fabric. The method for manufacturing coating composition has been widely described in the patent literature. Nevertheless, few papers published in the literature give account

1 Corresponding author: Laboratoire de Gnie et Matriaux Textiles (GEMTEX), UPRES EA2461, Ecole Nationale Suprieure des Arts et Industries Textiles (ENSAIT), BP 30329, 59056 Roubaix Cedex 01, France. e-mail: fabien.salaun@ensait.fr

Textile Research Journal Vol 79(13): 12021212 DOI: 10.1177/0040517508100724

The Author(s), 2009. Reprints and permissions: http://www.sagepub.co.uk/journalsPermissions.nav

Downloaded from http://trj.sagepub.com by Ngoc Nguyen Thi Thuy on October 22, 2009

Application of Contact Angle Measurement to the Manufacture of Textiles Containing Microcapsules F. Salan et al. of the formulation of coating, finishing of fabrics, and therefore the evaluation of their characteristics, more specifically thermal and durability properties [16, 17]. However, the choice of the process influences the microencapsulated fabric behavior and the yield [9]. The choice of the appropriate fibrous structure is linked to the capacity of the textile structure to carry a sufficient quantity of the microcapsules and whether it can provide an appropriate heat transfer to microcapsules. The following study developed the different steps taken into consideration in order to link efficiently the microcapsules to textile supports with a binder. It was based on two approaches; the first one was a thermodynamical approach of the wetting that characterizes the interfaces between the textile itself, the binder and the microcapsules, and the second one was the washing fastness of the microcapsules once linked on the textile with the binder. A knowledge of the wettability and surface free energy of microcapsules is very important in the design of textile formulations. This type of information can help in the selection of the components if the interfacial interactions and compatibility of the formulation components are known. Thus, the first step consisting of the comparison between the components of the binders surface energy and those of the resins that form the microcapsule shells helped us to perform our formulation. Then, the second step, validated by the ISO 6330 norm washing fastness picked up the best binder according to the best washing fastness results.

1203

TRJ

Figure 1 Contact angle measurement system

Figure 2 Schematic of a sessile drop, contact angle (), and the three interfacial tensions (lv: liquidvapor, sv: solidvapor, and sl: solidliquid) on substrate.

Background
Contact Angle Measurement: Surface Energy Determination
Surface energy determination by means of contact angle measurement has been frequently used to investigate the surfaces properties of textiles [1820]. These studies showed that surface energy determination is an appropriate method to investigate textile surface characteristics such as adhesion properties. Contact angle measurement (Figure 1) is a well-known technique in many fields such as contamination control, adhesion, surface treatments, and polymer film modification. The measurement of contact angles yields data which reflect the thermodynamics of a liquid/ solid interaction. In this study, surface energies were estimated using the geometric mean approach of Owens-Wendt [21], taking into account the dispersive and polar components of the surface energy and based on Young-Dupr equation (equation (1)) [22]: SV + SL + LV cos = 0 (1)

where is the equilibrium contact angle, measures between the tangents to the liquid-vapor and solid-liquid interfaces within the liquid phase, under the action of the three interfacial forces, i.e. solid-vapor (SV) (mN.m1 or mJ.m2), solidliquid (SL) (mN.m1 or mJ.m2), and liquid-vapor (LV) (mN.m1 or mJ.m2) tensions (Figure 2). The relationship between the contact angle () of the liquid phase deposited onto a solid phase is derived from the general Fowkes expression which considers the polar and dispersive contributions for both solid and liquid designated as S and L, with a superscript d or p for the dispersive and polar contribution, respectively.
1 d d 2 1 p p 2

L ( 1 + cos ) = 2 ( L S ) + 2 ( L S )

(2)

Thus, from equation (2) for two different test liquids it is possible to determine the surface energy of the solid phase (S) as the sum of two components: a dispersive component d ( S ) attributable to London attraction, and a specific (or p polar) component ( S ) owing to all other types of polar interactions such as hydrogen bonding.

Downloaded from http://trj.sagepub.com by Ngoc Nguyen Thi Thuy on October 22, 2009

TRJ

1204

Textile Research Journal 79(13)

Relation between Surface Energy and Adhesion


Adhesion is a manifestation of the attractive forces that exist between all atoms or molecules and which fall into two broad categories, i.e. primary (chemical bond) and secondary (van der Waals force and hydrogen bonds). To assess wetting and adhesion properties, it is necessary to consider the work of adhesion between two phases. The work of adhesion (Wadh) describes the work necessary to separate two solid phases (S1 and S2) so that two new surfaces of unit area are formed. Therefore, it refers to the free energy difference between two defined states, the first of two phases in contact in equilibrium and the second comprising the two phases separate in equilibrium in vacuo [23]. The work of adhesion (Wadh) of the liquid drop on a substrate is given by: W adh = SV SL LV (3)

Preparation of Microcapsules
Prepolymer solution synthesis: 100 ml of a 37 wt% strength aqueous formaldehyde solution was adjusted to pH 9 with triethanolamine in 500 ml three-neck flask equipped with an anchor stirrer and a condenser at 80 C. After adding 21 g of melamine, a clear solution was formed and was stirred until boiling occurred. After cooling down to 62 C, pH was reduced to 1 with sulfuric acid (10 wt%). When the contents had acquired a milky turbid appearance, 76 ml of methanol was added. The content of the flask was cooled to 40 C over the course of an hour. The solution was neutralized with triethanolamine, which terminated the reaction and the excess alcohol was distilled off. This resin may be written as MF5.6Me, where M represents the melamine, F the formaldehyde, Me the methanol, and the subscript is the molar ratio of formaldehyde per melamine. The MF5.6Me-He was synthesized by the process as described above, where the methanol solution was replaced by a methanol/1,6-hexanediol mixture. The reaction scheme for the synthesis is given in Figure 3. Microencapsulation: the microencapsulation of core material was carried out in a 500 ml three-neck round-bottomed flask equipped with a mechanical stirrer via an in situ polymerization. Prior to the encapsulation, 24 g of nhexadecane, 24 g of n-eicosane and 2 g of TEOS were emulsified into a continuous phase containing 125 g of the prepolymer solution (MF5.6Me or MF5.6Me-He), 200 ml of distilled water and 10 g of a binary mixture of Tween 20 and Brij 35 at pH 4 stirring at 13,500 rpm using an ultra turrax high speed homogenizer (Ika T 25 basic, Germany) at room temperature. The stirring speed was decreased to 700 rpm with an anchor stirrer after 15 min, and the emulsion was stirred for 4 h at 55 C. Then, the pH of the solution was adjusted to 9 with 50 wt% triethanolamine solution to complete the reaction.

Combining equations (1) and (3) yields the following equation, relating the contact angle to the work of adhesion: W adh = LV ( 1 + cos ) (4)

In our case, it was important to know whether this interface was stable towards liquids and more especially towards water. Thus, the work of adhesion is: W adh = S1/water + S2/water S1/S2 (5)

Adhesion is not assured when the interface is immersed in a liquid like water and Wadh is negative, since the free energy of the system is reduced and therefore the separation of the two solids is favored.

Experimental
Materials
Melamine (99%, Acros Organics) and formaldehyde (37 wt% aqueous, Acros Organics) were used as monomers; n-hexadecane, n-eicosane and tetra ethyl orthosilicate (TEOS) (Acros Organics) were used as core material. Nonionic surfactants, Tween 20 and Brij 35 (Acros Organics), were used as emulsifiers. Methanol (99.8%) and 1,6hexanediol (99%) were also obtained from Acros Organics. For pH control, triethanolamine, sulfuric acid and citric acid were used (Aldrich).

Sample Preparation
Textile supports: a 100% cotton fabric (566 dtex warp and 564 dtex weft yarns at densities of 26 ends/cm 16 picks/ cm, weighing 270 g/m2, thickness of 0.50 mm) (COTTON) and a 100% polyester fabric (345 dtex warp and 290 dtex weft yarns at densities of 18 ends/cm 17 picks/cm, weighing 120 g/m2, thickness of 0.22 mm) (PES) were chosen as the specimens. Binders: the commercial binders used in this study were: Alcoprint PB-66 (ethyl butyl acrylate from Ciba Specialty Chemicals), Alcoprint PB-HC (polyethyl acrylate from Ciba Specialty Chemicals), Dicrylan AS (polydimethylsiloxane polyacrylate from Ciba Specialty Chemicals), Dicrylan PMC (polyurethane from Ciba Specialty Chemicals), Airflex EP 177 (ethylene vinyl acetate copolymers from Air Products), and Airflex EN 428 (ethylene vinyl acetate copolymers from Air Products).

Downloaded from http://trj.sagepub.com by Ngoc Nguyen Thi Thuy on October 22, 2009

Application of Contact Angle Measurement to the Manufacture of Textiles Containing Microcapsules F. Salan et al.

1205

TRJ

Figure 3 Formation of cross-linked amino resin shell.

Impregnation: the textile impregnation of the two fabrics was made under the different baths containing the binder and the microcapsules at different concentrations. It was carried out by immersion of the fabrics at 2 meters per min in the different formulation baths. Once impregnated, the samples were pressed by a BENZ vat padding device without pressure in order to keep the microcapsules intact. Then, the drying treatment step was performed in a BENZ frame under ventilation at 0.5 meter per min speed during 4 min at 100 C (to evaporate the water) and 4 min at 150 C (to ensure adequate binder crosslinking).

Diiodomethane: l = 50.8 mJ/m2, ld = 48.5 mJ/m, lp = 2.3 mJ/m. The contact angles were measured by two referent liquids, which were water and diiodomethane on two melamine formaldehyde resin films (MF5.6Me or MF5.6Me-He). MF5.6Me film was obtained from the polycondensation of a methanol-etherified methylolmelamines prepolymer, and MF5.6Me-He film from the polycondensation of a hexanediol-etherified methylolmelamines prepolymer and then compared with the measures obtained on films of six crosslinked binders that had been put into an oven at 150 C for 5 min. These films were very uniform spatially, but did exhibit hydration, which meant that their contact angle varied according to the exposure of the tested liquid on the vapor. Therefore, each contact angle was measured twice, as a function of time, with t = 0 corresponding to the first contact of the liquid with the dry surface and at the time the angle became steady-stable, when the two surfaces were equilibrated generally after a few seconds. During the rest of this study, we took only the second angle measures into account. Each surface was measured with five independent drops of the two referent liquids, allowing us to distinguish independently the dispersed and the non-dispersed components of each material. No significant variations were found within any single surface. Drops were always placed at least 10 mm apart and never closer than 10 mm to an edge.

Characterization Techniques
Characterization of the Solid Surface: Contact Angle Method
The thermodynamic theory on surface properties allowed the relation describing the surface adhesion properties to be obtained. Thus, the interface study of the binder/microcapsule was based on a wetting approach by the contact angle measure that characterizes the interaction between a liquid and a solid. Contact angles were estimated with a goniometer equipped with a special optical system and a camera (Figure 1). A drop of liquid was placed on a polymer film and the image was immediately sent via the camera to the computer for analysis. The volume of the drop was about 6 l. Constant values for the test liquids used for contact angle measurements were as follows [24]: Water: l = 72.8 mJ/m2, ld = 21.8 mJ/m, lp = 51.0 mJ/m;

Downloaded from http://trj.sagepub.com by Ngoc Nguyen Thi Thuy on October 22, 2009

TRJ

1206

Textile Research Journal 79(13)

Washing Machine ISO Test


The washing fastness tests of the interface binder/microcapsules/textile were made in the Wascator washing machine, half charged, at 40 C, wool program, under the norm NF EN ISO 6330. After each washing cycle, the samples were conditioned under normal values: 20 C for the temperature and 65% humidity for the humidity rate.

Scanning Electron Microscope (SEM) Images


The conditions of the attached microcapsules and the characterizations of the interface binder/microcapsules and binder/textile were observed by electronic scanning microscopy (Philips XL 30 ESEM), with an accelerating voltage of 20.0 or 25.0 kV and 2,000 or 5,000 magnifications.

Fourier Transform Infrared (FTIR) Spectroscopy


The structure of the shell polymer was analyzed by FTIR spectra. Samples were ground and mixed with KBr to make pellets. FTIR spectra in the transmission mode were recorded using a Nicolet Nexus, connected to a PC, in which the number of scan was 32 and the resolution was 4 cm1.

Results and Discussion


Microcapsule Characteristics
Two types of microcapsules containing PCMs were obtained by an in situ polymerization. The surface morphology of the microcapsules and chemical structure of the polymeric shells were investigated using SEM and FTIR spectroscopy, respectively (Figures 4 and 5). As shown in Figure 6, the resulting microcapsules had relatively uniform sizes, spherical shape and smooth surface. No destruction of the capsule walls due to mechanical agitation was perceivable. The main difference between the two samples was their mean diameter. The particles from MF5.6Me resin had a mean diameter of about 1.5 m, whereas those of MF5.6Me-He had one of about 5 m. The FTIR spectra of the microcapsule shells (MF5.6Me and MF5.6Me-He) are shown in Figure 5. Characteristic broad band responsible for hydroxyl, imino and amino stretching was observed around 3350 cm1. Alkyl C-H stretching vibration was found around 2950 cm1. The C-N multiple stretching in the triazine ring was observed around 1551 cm1. C-H bending vibration in CH2 was found at 1490 cm1 and 1360 cm1 due to methylene bridges. The characteristic absorption bands of aliphatic CN vibration appeared between 1200 and 1170 cm1. The band at 1650 cm1 corresponded to the stretching mode of C=N. Characteristic triazine ring bending could also be observed at 810 cm1. C-O-C stretching due to ether bridge at 1110 cm1 was also present in Figure 5. Figure 4 SEM images of microcapsules from MF5.6Me (a) and MF5.6Me-He (b) resins.

As shown in Table 1, the surface free energy and its specific components decreased with the incorporation of hexanediol in the polymer shell, while the London dispersive component increased. As expected, the work of adhesion for water on microcapsules with MF5.6Me shell (127.4 mJ/ m2) was higher than MF5.6Me-He microcapsules (99.2 mJ/ m2) (Table 1). These results indicated that the surface of microcapsules was modified with increased hydrophobic groups during the polymerization. The chemical difference between the two amino polymers was based on the triazine substitution lengths that varied from 2 to 6 atoms of carbon. This modification involved a higher non-dispersive behavior for MF5.6Me microcapsules than MF5.6Me-He microcapsules.

Surface Energy of Binders


The results of the contact angles and the surface energy calculations are shown in Table 1. The test results indicated that the polyacrylate binders all had much higher surface energies than the other polymers. These higher surface energies were mainly due to the contribution of the polar

Downloaded from http://trj.sagepub.com by Ngoc Nguyen Thi Thuy on October 22, 2009

Application of Contact Angle Measurement to the Manufacture of Textiles Containing Microcapsules F. Salan et al.

1207

TRJ

Figure 5 FT-IR spectra of MF5.6Me (a) and MF5.6Me-He (b) resins.

Figure 6 Washing fastness of coated samples.

component. The presence of ester groups in the polyacrylate backbone may explain this high polarity. Furthermore, the presence of butyl groups to polyacrylate resulted in an increase of the contact angle with water correlated to a decreased diiodomethane contact angle value. Thus, the length of the lateral chain led to a decrease in the polar component and governed the surface energy of the polyacrylate binders. The vinyl acetate copolymers and the polyester urethane presented low surface energy. It seemed that the values were closely linked to the dispersive component of the materials. The contact angle values with water

and diiodomethane decreased in the presence of acrylic acid in vinyl acetate copolymer (Airflex EN 428). Therefore, this binder was more hydrophobic than the Airflex EP 177. From these results, the component of surface energy and therefore their wetting behavior were closely dependent on the chemical structure of the binders.

Downloaded from http://trj.sagepub.com by Ngoc Nguyen Thi Thuy on October 22, 2009

TRJ

1208

Textile Research Journal 79(13)

Table 1 Contact angle measures and surface energy components.


Polymer Contact angle Component of surface energya (mJ/m) Work of adhesion (mJ/m)

water diiodomethan
() MF5.6Me methanol-etherified methylolmelamine hexanediol-etherified methylolmelamine 27.4
e ()

sd
26.0

sp
39.6

s
65.6 Wadh MF5.6Me -waterb 137.4

39.4

MF5.6Me-He

68.7

42.0

32.0

10.6

42.6

Wadh MF5.6Me-He -waterb 99.2 Wadh MF5.6Me -Bic Wadh MF5.6Me-He -Bic 12.1 22.8 20.5 54.1 64.3 58.2

Alcoprint PB-66 Alcoprint PB-HC Dicrylan AS Dicrylan PMC Airflex EP 177 Airflex EN 428
a

ethyl butyl acrylate polyethyl acrylate polydimethylsiloxane polyacrylate polyurethane ethylene vinyl acetate copolymers ethylene vinyl acetate copolymers

28.7 22.6 20.5 79.5 91.1 83.8

38.2 70.6 65.9 49.5 54.2 51.8

26.7 10.1 12.1 30.1 29.9 29.7

38.3 59.7 57.7 16.0 12.0 14.3

65.0 69.8 69.8 36.1 31.9 34.0

14.1 14.3 13.4 13.6 14.1 13.8

is the surface energy, and sd and sp are the dispersive and polar components, respectively (equation (2)); b Bi: binder; c calculated from equation (4).
s

Interface Binder/Microcapsules
Binder Results
The results of the Wadh calculations with water showed that the adhesion was assured for the binders Alcoprint PB-66, Dicrylan PMC, Airflex EP 177 and Airflex EN 428 with either MF5.6Me or MF5.6Me-He. The measures of the contact angles allowed us to find out the more suitable binders for the link binder/microcapsules (made of MF5.6Me or MF5.6Me-He). They were Alcoprint PB-66, Dicrylan PMC, Airflex EP 177 and Airflex EN 428, because their dispersive and non-dispersive components were close to the ones of the microcapsule wall and because their Wadh was positive when wetted by water. The values of the components of the surface energy that we measured were parallel to the values given in the literature, even if they were slightly higher [25].

Interface Binder/Microcapsules/Textile
The wetting of textile fabrics can be explained by a physical phenomenon basis of the system solid-liquid-air. The measurement of the fabric surface wetting is a very sophisticated and complicated process because of the fabric roughness, heterogeneity and because of the diffusion of the liquid into the fiber and the capillarity of the fiber

assembly. Contact angles can range from 0 (complete wetting) to 180 (non-wetting). Therefore, on a fiber it is more convenient to visualize its wet ability by observing the waters contact angle on its surface; when it is near 0, then it means that the fiber is hydrophilic and almost completely wettable. However, on the textile fabric itself, the pores created by the yarn twilling involves the ability of the liquid to withdraw or to spread. Therefore, the experimentally measured contact angles on a textile surface can vary considerably from the contact angle for an ideal system. Direct contact angle measurements on textile surfaces were not reliable for our study. Thus, the adhesion of the adequate binders determined with the interface binder/microcapsule method with the textile surfaces was characterized by the washing fastness (NF EN ISO 6330) of the coated samples. Furthermore, as Airflex EP 177 and Airflex EN 428 had the same component surface energies, we chose to work with Airflex EN 428 only. The binders did not react the same way depending on the chemical structure and the contexture of the textiles fabrics; this was due to their internal physical characteristics, but also to their different surface energies. On all SEM micrographs (Figures 7 and 8), we observed a total wrapping of the binder around the textile fiber. However, the behavior of the binder on the cotton fabric and on the PES fabric varied. On the cotton, the binder tended to coat

Downloaded from http://trj.sagepub.com by Ngoc Nguyen Thi Thuy on October 22, 2009

Application of Contact Angle Measurement to the Manufacture of Textiles Containing Microcapsules F. Salan et al.

1209

TRJ

Figure 7 SEM micrographs of PES coating with Alcoprint PB-66 (a), Dicrylan PMC (b), Airflex EN 428 (c); with Alcoprint PB-66 + microcapsules (d), Dicrylan PMC + microcapsules (e), Airflex EN 428 + microcapsules (f); with Alcoprint PB-66 + microcapsules after washing test (g), Dicrylan PMC + microcapsules after washing test (h), Airflex EN 428 + microcapsules after washing test (i).

the twilling parts of the yarns, which involved a surface coating by creating crosslinked areas as the binder could not penetrate the pores of the fabric. On the PES, the binder wrapped the fibers homogeneously inside the core of the fabric. The incorporation of the microcapsules in the binder formulation did not modify the cotton surface coating. It seemed that the microcapsules created crosslinked areas of coating between the PES fibers attributed to an increase in viscosity of the formulation. The microcapsules were linked to the surface of the fibers either by being completely wrapped by the binder, or by creating a linkage with the binder between the fibers. The different binders did not have the same behavior; Alcoprint PB-66 seemed to have a better adhesion to the fiber than to the microcapsules, as it created agglomerates of microcapsules instead of fixing them individually to the fiber. The washing test (Figure 6) confirmed this trend,

since after 20 washing cycles, the loss of weight was higher for fabrics treated with Alcoprint PB-66 than with the other binders for both cotton and PES. The coatings on PES fabric were more resistant than on cotton, as more impregnated in the core of the fabric, so more protected by the friction during washing. On the SEM micrographs (Figures 7 and 8) of the coated textiles after 20 washing cycles, we observed holes at the fiber surface, which was more subject to frictions. These holes represented the place where a microcapsule was linked with the fiber, so it meant that the adhesion was better for the binder and the microcapsule than for the binder and the fiber especially in the case of Alcoprint PB-66. This involved the problem of the adhesion of the binder to the fiber that appeared more brittle than the adhesion of the microcapsules and the binder. The behavior differences between the binders with the washing fastness tests were in accordance with the differ-

Downloaded from http://trj.sagepub.com by Ngoc Nguyen Thi Thuy on October 22, 2009

TRJ

1210

Textile Research Journal 79(13)

Figure 8 SEM micrographs of COTTON coating with Alcoprint PB66 (j), Dicrylan PMC (k), Airflex EN 428 (l); with Alcoprint PB-66 + microcapsules (m), Dicrylan PMC + microcapsules (n), Airflex EN 428 + microcapsules (o); with Alcoprint PB-66 + microcapsules after washing test (p), Dicrylan PMC + microcapsules after washing test (q), Airflex EN 428 + microcapsules after washing test (r).

ences already noticed on the adhesion of the binder and the microcapsules, as for Alcoprint PB-66, the values of work adhesion were already worse than the other ones. According to the different tests, surface energy calculations and washing tests, it appeared that Dicrylan PMC was the more efficient binder to confine the melamine formaldehyde microcapsules to the textiles, both cotton and PES.

Binder Concentration Influence


The efficiency of adhesion was linked to the nature of the binder and to its proportion used for the coating formulation (Table 2). When the weight of the binder represented 33% of the coating formulation, less weight loss was observed with a formulation of 45 g/m on the fabrics. Thus, the impregnation mechanism could be divided into two steps. First, the binder and the microcapsules filled the

core of the textile, and second, once this area was full, the rest of the formulation settled at the surface of the coated structure. Therefore, the more binder/microcapsules on the textile surface (g/m2), the higher the loss of weight after 20 washing cycles, because the coating settled on the surface was more sensitive to frictions than the coating penetrated through the textile. On the PES fabric, it was clear that a high concentration of microcapsules reduced the adhesion of the coating to the textile fabric. On the cotton fabrics, adhesion was not affected by the microcapsule concentration. This was probably due to the cotton fiber surface that was not smooth enough, so the presence of the microcapsules increased the specific contact surface, and thus a better adhesion. This adhesion could be improved by the presence of hydroxyl groups on the cotton, which could create a covalent bond with the functional groups attached to the polymer binder.

Downloaded from http://trj.sagepub.com by Ngoc Nguyen Thi Thuy on October 22, 2009

Application of Contact Angle Measurement to the Manufacture of Textiles Containing Microcapsules F. Salan et al.

1211

TRJ

Table 2 Washing fastness of coated samples (c: microcapsules).


Textile supports Formulation (g/m) (c + Dicrylan PMC) 59 70 53 44 PES 26 90 66 32 23 57 68 47 29 COTTON 28 63 16 26 17
a

Weight of binder in the formulation (wt%) 15 25 33 33 33 45 100 30/(3)a (5)a 15 25 33 33 33 45 100 30/(3) (5)
a a

Weight loss of coating after 20 washing cycles 43 26 20 10 17 12 10 17 35 38 20 19 27 28 12 43 22 85

Weight of blocked isocyanate in the formulation (wt%).

The presence of blocked isocyanate in the formulation reduced the weight lost after washing. The unblocking during the binders reticulation allowed the formation of covalent bonds the binder/fiber and microcapsule/binder. We noticed that we could fix up to 15% of microcapsules on cotton and 65% on PES without requiring the use of the binder. In fact, the microcapsules were locked between the twilling of the fibers and the isocyanate helped to increase their adhesion. Indeed, the loss of weight of the coating on the textile surfaces that were impregnated by microcapsules alone was of 100% after several washing cycles.

Conclusion
The adhesion of microcapsules on a textile support is a complex phenomenon linked to the physical and chemical characteristics of the different involved compounds. The adhesion study was divided into two steps; the analysis of the binder/microcapsules based on the thermodynamic wetting approach and the analysis of the binder/textile interface based on the NF EN ISO 6330 test norm. We noticed a good result correlation of these two approaches.

By the wetting and washing approaches, binder based on polyurethane polymer was the most suitable for melamine formaldehyde microcapsules. The use of a blocked isocyanate in the coating formulation allowd the washing fastness to improve. Better results were obtained with PES fabric than cotton fabric, mainly influenced by the structure of the textile substrate. A dense structure led to a brittle coating on the surface, whereas with a less dense structure, the binder wrapped the fibers homogeneously inside the core of the fabric which ensured a better adhesion. Regarding the concentration of binder/microcapsules that should be used for the best washing fastness, according to the final desired effect of the microcapsules, it has been proved that it was not necessary to have high concentrations of both binder and microcapsules. Indeed, it was more judicious to determine a good proportion of microcapsules/binder (66/33 wt%) to limit the loss of coating weight after washing.

Downloaded from http://trj.sagepub.com by Ngoc Nguyen Thi Thuy on October 22, 2009

TRJ

1212

Textile Research Journal 79(13)


13. Cox, R., Synopsis of the New Thermal Regulating Fiber Outlast, Chem. Fibers Int. 48, 475479 (1998). 14. Pause, B. H., Development of Heat and Cold Insulating Membrane Structures with Phase Change Material, J. Coated Fabrics 25(7), 5968 (1995). 15. Shim, H., McCullough, E. A., and Jones, B. W., Using Phase Change Materials in Clothing, Textile Res. J. 71(6), 495502 (2001). 16. Pause, B., Interactive Thermal Insulating System having a Layer Treated with a Coating of Energy Absorbing Phase Change Material Adjacent to a Layer of Fibers Containing Energy Absorbing Phase Change Material, US Patent 6 217 993 (2001). 17. Zuckerman, J. L., Pushaw, R. J., Perry, B. T., and Wyner Daniel, M., Fabric Coating Composition Containing Energy Absorbing Phase Change Material, US Patent 6 207 738 (2001). 18. Di, J., Perwuelz, A., and Gueguen, V., The Surface Energy of PET Fabrics Coated with Silicone, J. Textile Inst. 92, 184192 (2001). 19. Zhu, L., Perwuelz, A., Lewandowski, M., and Campagne, C., Wetting Behavior of Thermally Bonded Polyester Nonwoven Fabrics: the Importance of Porosity, J. Appl. Polym. Sci. 102(1), 387394 (2006). 20. Dumitrascu, N., and Borcia, C., Adhesion Properties of Polyamide-6 Fibres Treated by Dielectric Barrier Discharge, Surf. Coating Technol. 201(34), 1171123 (2006). 21. Owens, D. K., and Wendt, R. C., Estimation of the Surface Free Energy of Polymers, J. Appl. Polym. Sci. 13, 17411747 (1969). 22. Dupr, A., Thorie Mechanique de la Chaleur, Gauthier-Villars, Paris, 1869, In Particle-particle adhesion in pharmaceutical powder handling, (Podczeck, F., Ed.), Imperial College Press, London, UK, pp. 2930 (1998). 23. Packham, D. E., Work of Adhesion: Contact Angles and Contact Mechanics, Int. J. Adhes. Adhes. 16(2), 121128 (1996). 24. Fowkes, F. M., Attractive Forces at Interfaces, Ind. Eng. Chem. 56(12), 4052 (1964). 25. Brandrup, J., Immergut, E. H., and Grulke, E. A., (Eds.), Polymer Handbook, IVth Edition, John Wiley and Sons, New York, USA (1999).

Literature Cited
1. Nelson, G., Application of Microencapsulation in Textiles, Int. J. Pharm. 242, 5562 (2002). 2. Mondal, S., Phase Change Materials for Smart Textiles an Overview, Appl. Therm. Eng. 28(1112), 15361550 (2008). 3. Wang, C. X., and Chen, S. H. L., Fragrance-release Property of -Cyclodextrin Inclusion Compounds and Their Application in Aromatherapy, J. Ind. Textiles 34(3), 157166 (2005). 4. Aitken, D., Burkinshaw, S. M., Griffiths, J., and Towns, A. D., Textile Applications of Thermochromic Systems, Rev. Prog. Color. 26, 18 (1996). 5. Fei, B., and Xin, J. H., N, N-Diethyl-m-toluamide-containing Microcapsules for Bio-cloth Finishing, Am. J. Trop. Med. Hyg. 77(1), 5257 (2007). 6. Ghosh, S. K., Functional Coatings and Microencapsulation: a General Perspective (Chapter 1), In Functional Coatings by Polymer Microencapsulation Wiley-VCH, Weinheim (2006) (ISBN 352731296X). 7. Giraud, S., Bourbigot, S., Rochery, M., Vroman, I., Tighzert, L., Delobel, R., and Poutch, F., Flame Retarded Polyurea with Microencapsulated Ammonium Phosphate for Textile Coating, Polym. Degrad. Stabil. 88(1), 106113 (2005). 8. Yadav, S. K., Suresh, A. K., and Khilar, K. C., Microencapsulation in Polyurea Shell by Interfacial Polycondensation, AIChE J. 36, 431438 (1990). 9. Monllor, P., Bonet, M. A., and Cases, F., Characterization of the Behaviour of Flavour Microcapsules in Cotton Fabrics, Eur. Polym. J. 43(6), 24812490 (2007). 10. Lennox-Kerr, P., Outlast Technologies Adapts Space-age Technology to Keep Us Comfortable, TTI J. 17(7), 2526 (1998). 11. Zhang, X. X., Wang, X. C., Tao, X. M., and Yick, K. L., Energy Storage Polymer/MicroPCMS Blended Chips and Thermo-regulated Fibres, J. Mater. Sci. 40, 37293734 (2005). 12. Zhang, X. X., Tao, X. M., Yick, K. L., and Wang, X. C., Structures and Properties of Wet Spun Thermo-regulated Polyacrylonitrile-vinylidene Chloride Fibers, Textile Res. J. 76(5), 351 359 (2006).

Downloaded from http://trj.sagepub.com by Ngoc Nguyen Thi Thuy on October 22, 2009

You might also like