Download as pdf or txt
Download as pdf or txt
You are on page 1of 107

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/236107775

Wetting and Wicking of Textiles

Book · January 2006

CITATION READS

1 1,403

4 authors, including:

Asis Patnaik R. S. Rengasamy


Cape Peninsula University of Technology Indian Institute of Technology Delhi
62 PUBLICATIONS 418 CITATIONS 80 PUBLICATIONS 595 CITATIONS

SEE PROFILE SEE PROFILE

Anindya Ghosh
Government College of Engineering & Textile …
125 PUBLICATIONS 266 CITATIONS

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

ISSIP 2017: 2017 First International Symposium on Signal and Image Processing View project

aerogel and sound insulation View project

All content following this page was uploaded by Anindya Ghosh on 13 July 2017.

The user has requested enhancement of the downloaded file. All in-text references underlined in blue are added to the original document
and are linked to publications on ResearchGate, letting you access and read them immediately.
This article was downloaded by: [Majumdar, Abhijit][Indian Institute of Technology]
On: 21 April 2009
Access details: Access Details: [subscription number 906467071]
Publisher Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House,
37-41 Mortimer Street, London W1T 3JH, UK

Textile Progress
Publication details, including instructions for authors and subscription information:
http://www.informaworld.com/smpp/title~content=t778164492

Wetting and Wicking in Fibrous Materials


A. Patnaik; R. S. Rengasamy; V. K. Kothari; A. Ghosh

Online Publication Date: 01 January 2006

To cite this Article Patnaik, A., Rengasamy, R. S., Kothari, V. K. and Ghosh, A.(2006)'Wetting and Wicking in Fibrous Materials',Textile
Progress,38:1,1 — 105
To link to this Article: DOI: 10.1533/jotp.2006.38.1.1
URL: http://dx.doi.org/10.1533/jotp.2006.38.1.1

PLEASE SCROLL DOWN FOR ARTICLE

Full terms and conditions of use: http://www.informaworld.com/terms-and-conditions-of-access.pdf

This article may be used for research, teaching and private study purposes. Any substantial or
systematic reproduction, re-distribution, re-selling, loan or sub-licensing, systematic supply or
distribution in any form to anyone is expressly forbidden.

The publisher does not give any warranty express or implied or make any representation that the contents
will be complete or accurate or up to date. The accuracy of any instructions, formulae and drug doses
should be independently verified with primary sources. The publisher shall not be liable for any loss,
actions, claims, proceedings, demand or costs or damages whatsoever or howsoever caused arising directly
or indirectly in connection with or arising out of the use of this material.
WETTING AND WICKING IN FIBROUS
MATERIALS
A. Patnaik1, R. S. Rengasamy, V. K. Kothari and A. Ghosh
doi:10.1533/tepr.2006.0001
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

Abstract: This issue reviews developments in wetting and wicking of fibrous materials,
covering characterization of wetting; wetting and wicking of fibers, filaments, yarns,
and fabrics; factors affecting wetting and wicking of fibrous assemblies; mathematical
models of wetting and wicking; and application areas of the same.
Key words: Wetting, wicking, fibrous materials.

1. INTRODUCTION
Wetting and wicking are important phenomena in the processing and applications
of fibrous materials. Various aspects of liquid–fiber interactions such as wetting,
transport, and retention have received much attention both in terms of fundamental
research and for product and process development [1]. In fiber composites, the
performance of the composites is governed by the adhesion between the fibers and
resin binders. The adhesion between the fibers and resin is influenced by the initial
wetting of the fibers by the resin as this decides the subsequent resin penetration
between the fibers and voids content. On the other hand, surgical fabrics should not
let liquid and solid particles pass through easily. The mechanics of liquid coating are
greatly influenced by the combined abilities of the liquid and solid to achieve complete
wetting at the given speed of solid motion [2]. Wetting and wicking behavior of
fibers is important to describe fluid transport in fibrous media such as filters, coalescers,
and sorbents.
A spontaneous transport of a liquid driven into a porous system by capillary forces
is termed wicking because capillary forces are caused by wetting; wicking is a result
of spontaneous wetting in a capillary system [3]. Wetting is a prerequisite for wicking.
A liquid that does not wet fibers cannot wick into a fabric. The wetting and wicking
behavior of the fibrous structure is a critical aspect of performance of products such
as sports clothes, hygiene disposable materials, and medical products. Wetting and
wicking processes occurring during wearing of clothes have a practical significance
in clothing comfort.

2. WETTING
Wetting of a fibrous assembly affects many manufacturing processes, as well as the
end use performance of materials. Wetting is a complex process complicated further
by structure of the fibrous assembly e.g. yarns, woven/nonwoven/knitted fabrics, and
pre-forms for composites.
There is clear distinction between two terms that are sometimes used interchangeably,

© The Textile Institute


2 Textile Progress doi:10.1533/tepr.2006.0001

‘wetting’ and ‘wettability’. The wetting of a solid surface is understood to be the


condition resulting from its contact with a specified liquid under specific conditions.
Wettability is the potential of a surface to interact with liquids with specified
characteristics [4]. According to Harnett and Mehta [5], ‘wettability’ is the initial
behavior of a fabric, yarn, or fiber when brought into contact with a liquid. It also
describes the interaction between the liquid and the substrate prior to the wicking
process.
For a liquid to move in a fibrous medium, it must wet the fiber surfaces before
being transported through the interfiber pores by means of capillary action. While the
interactions of molecules in the bulk of a liquid are balanced by an equal attractive
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

force in all directions, the molecules on the surface of a liquid experience an imbalance
of forces. Hence, there is presence of free energy at the surface of the liquid. The
excess energy is called ‘surface free energy’, which tends to keep the surface area of
the liquid to a minimum and restricts the advancement the liquid over the solid
surface. For a liquid to wet a solid completely, or for the solid to be submerged in a
liquid, the solid surfaces must have sufficient surface energy to overcome the free
surface energy of the liquid. The surface free energy can be quantified as a measurement
of energy per area. It is usually termed ‘surface tension’ and is quantified as force per
length, with units mN/m or dynes/cm.
The forces involved in the equilibrium that exists when a liquid is in contact with
a solid and a vapor at the same time are given by the following (Young–Dupré)
equation [4, 6–8]:
γSV – γSL = γLV cos θ (1)
where γ represents the interfacial tension that exists between the various combinations
of solid, liquid, and vapor; the subscripts S, L, and V standing for solid, liquid, and
vapor, and θ is the equilibrium contact angle (Fig. 1). The term γLV is denoted as the
surface tension of the liquid for the liquid/vapor interface. The term γLV cos θ, has
been called the ‘adhesion tension’ or ‘specific wettability’. Equation (1) is valid only
for a drop resting at equilibrium on a smooth, homogeneous, impermeable, and non-
deformable surface. The equation has been widely used to explain wetting and wicking
phenomena. The contact angle θ is the angle between the tangent to the liquid–vapor
(air) interface and the solid–liquid interface [4, 8–9].

γ LV

Vapor

Liquid θ

γ SL γ SV

Solid

Fig. 1 Equilibrium state of a liquid drop on a solid surface [33 Miller]

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 3

2.1 Wetting Mechanism


Wetting of textiles involves several primary processes: immersion, capillary sorption,
adhesion, and spreading [10]. During immersion (Fig. 2a) or capillary sorption (Fig.
2b) a solid–vapor interface disappears and solid–liquid interface appears. The work
of immersion, WI, or the work of penetration, WP, performed during capillary sorption,
is defined as the free energy change when the contacting solid and liquid are separated.
For spontaneous penetration, e.g. a positive capillary rise, the work of penetration
has to be positive. This is the case when the interfacial energy of the solid in contact
with vapor exceeds the interfacial tension between the solid and the liquid. For
interfaces of unit area [10]:
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

WI = WP = γSV – γSL (2)


Adhesion is the attraction between two surfaces in contact (Fig. 2c). When the
contacting surfaces are those of a solid and a liquid, the work of adhesion, WA or WSL,
is equal to the change of surface free energy of the system when the contacting liquid
and the solid are separated [10].
The separation results in the loss of their interface with interfacial tension, γSL, and
formation of two new surfaces with surface tensions γSV and γLV. The work of adhesion
per unit area of interfaces is given by the Young–Dupré equation (1). This is the total
attraction per unit area between the two phases and it can be expressed in general
terms as:
WA = γLV + γSV – γLS (3)
Assuming that equation (1) is valid, and combining it with equation (3), leads to an
expression for the work of adhesion in terms of two measurable properties:
WA = γLV + γLV cos θ (4)

γ SV

γ SL
γ SL
wI = γ SV – γ SL wP = γ SV – γ SL
(a) (b)

γ LV
γ SV
γ LV
γ SL
γ SL
wS = γ SV – γ SL – γ LV
wA = γ LV + γ SV – γ SL (d)
(c)

Fig. 2 Wetting mechanism: (a) Immersion of a solid in a liquid; (b) capillary sorption; (c)
adhesion between a liquid and a solid; (d) spreading of a liquid on a solid [10 Kissa]

© The Textile Institute


4 Textile Progress doi:10.1533/tepr.2006.0001

This can be transformed to


WA = γLV(1 + cos θ)
Application of equation (3) to a liquid yields the work of cohesion, WC, which is the
reversible work to pull apart a liquid column creating two liquid surfaces, each
having an interfacial tension γLV:
WC = 2γLV (5)
Spreading is the flow of liquid at least two molecular layers thick over a solid (Fig.
2d). During spreading, the solid–liquid and liquid–vapor interfaces increase, whereas
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

the solid–vapor interface decreases. The work of spreading, WS, is the reversible
work equal to the free energy change that occurs when the solid and liquid are
separated (reversal of spreading), and it is given by:
WS = γSV – γLV – γSL (6)
For spreading to be spontaneous, the work of spreading, WS, has to be positive.
Equations (2), (3), and (5) are valid only for ideal, smooth, homogeneous, impermeable,
and non-deformable surfaces. Since textile fibers do not have such ideal surfaces,
their wetting phenomena are more complicated. In addition, the prediction of wetting
phenomena, e.g. spreading, from wetting energy is difficult because a direct method
for determining γSV, a term found in equations (2), (3), and (6), is not available. It is
more convenient to use the forces in balance at a three-phase (solid, liquid, and
vapor) boundary as an indication of wettability.
When a liquid drop is placed on fabric, it will spread under capillary forces [11].
As suggested by Gillespie [12], the spreading process may be split conveniently into
two phases, I and II, when some of the liquid remains on the surface and when the
liquid is completely contained within the substrate, respectively, as shown in Fig. 3.

Phase I

Phase II

Fig. 3 Schematic illustration of the two phases in drop spreading of liquid in fabrics [12
Gillespie]

For two-dimensional circular spreading in textiles during Phase II, Kissa [13]
developed Gillespie’s equation to propose the following exponential sorption:
A = K(γLV/η)uVmtn
where A is the area covered by the spreading liquid, K is the capillary sorption
coefficient, η is the viscosity of the liquid, V is the volume of the liquid, t is the
spreading time, and the values of the exponents u, m, and n are 0.33, 0.67, and 0.33
respectively.
© The Textile Institute
doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 5

Kissa’s exponents can be applied only for fibers that are impermeable to liquids.
When the liquid diffuses into the fibers (e.g. water into cotton fiber), the exponents
depend on the nature of the liquids, fibers, and drop volume [13].
The wetting process, as applied to textiles, can be resolved into two independent
processes, often competitive with one another. One, the escape of occluded gases
from the substrate, is essentially mechanical, and the other, the rate of advance of the
liquid phase into the mass to be wetted, is a physiochemical phenonenon [14].

2.2 Characterization of Wetting


The surface wetting characteristics of fibers affect their processability into finished
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

products and the performance of the latter in applications in which fibers contact
fluids. The characteristics are the cosine of the contact angle, the work of adhesion
(WA), and the surface energy of the solid, all related but describing different behaviors
[15].
The cosine of the contact angle is a parameter often used to represent wettability
of a material, and to predict capillary force using the Laplace model, and imbibition
rate of a fluid in fibrous networks using the Washburn-Lucas model [15–17].
To get the best information about wetting of a solid, the useful parameters are
work of adhesion and solid surface energy. The work of adhesion characterizes
cohesion (net attraction) between solid and liquid and is a parameter considered
when selecting materials for applications such as coatings and composites. The work
of adhesion depends, in part, on the nature of the solid surface. In many instances
where a property or process is influenced by wetting behavior, it is the work of
adhesion that is the critical factor [4]. Wetting hysteresis is the ratio of the work of
adhesion in the receding mode to that in the advancing. Differences in hysteresis
among fibers shed light on differences that exist in their chemical and physical
structures [8]. The surface energy of a solid ( γ SV) determines the inherent ability of
that solid to interact with a contacting liquid, and the contact angle is the result of
such interaction [15]. The chemical nature and physical structure of the solid influence
the solid surface energy.
Various methods for quantifying liquid–solid surface interaction are outlined below.

2.2.1 Contact Angle


The contact angle serves as a convenient means for visualizing and describing the
geometry of solid–liquid contact. It depends on three interfacial tensions γSV, γSL, and
γLV. If γSV is larger than γSL, then cos θ is positive, and the contact angle θ must be
between 0 and 90°. If γSV is smaller than γSL, then the contact angle must be between
90° and 180° [3, 9–10]. With increasing wettability, the contact angle decreases and
cos θ increases. Complete wetting implies a zero contact angle, but equating θ = 0
may lead to incorrect conclusions [3, 10]. It is better to visualize that, when the
contact angle approaches zero, wettability has its maximum limit [10]. A high contact
angle for water with the surface means that the water will run off it; a low contact
means that water will wet the material [9]. According to Adam [18], equilibrium
conditions cannot exist when the contact angle is zero, and the equation (1) does not
apply. The distinction between zero and non-zero contact angles is really the difference
between spreading and non-spreading of a liquid across a solid. When a liquid comes

© The Textile Institute


6 Textile Progress doi:10.1533/tepr.2006.0001

into contact with a solid, there is some wetting whether it spreads or not [19]. For
convenience, most workers use contact angle or its cosine function as a direct measure
of wettability or degree of wetting. According to Miller [4], contact angle is not the
cause of wetting but its consequence, and is determined by the net effect of three
essentially independent attractive forces, as seen in equation (1). If a few solids are
compared for their wettability with a given liquid, the lowest contact angle observed
among the solid–liquid combination implies that the solid has greatest attraction for
the said liquid. If two liquids are compared with respect to the same solid, the
difference in contact angles has no clear cut significance by itself, unless the
corresponding liquid–liquid interactions are adjusted by some manner. In both cases,
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

the changes in contact angle can indicate only relative differences [4].
The term ‘contact angle’ has several meanings. The intrinsic contact angle or the
true contact angle is the angle at very short (molecular) distance from the contact line
on the solid [20]. The equilibrium contact angle is the single valued intrinsic contact
angle described by the Young-Dupré equation [4, 6–8] for an ideal system. However,
a real solid–liquid system may exhibit several stable contact angles. An experimentally
observed contact angle is an apparent contact angle, measured on a macroscopic
scale – for example, through a low power microscope [20]. Although measuring the
true intrinsic contact angle at the contact line is very difficult [21], on rough surfaces
the difference between the apparent and intrinsic contact angles can be considerable
[20]. Shuttleworth and Bailey [22] have suggested that the apparent contact angle is
the sum of the intrinsic contact angle and the slope angle of the surface at the point
of contact. The slope angle can be positive or negative. Since the Young-Dupré
equation applies to the interfacial tensions and the intrinsic contact angle at the
contact line, the results obtained by substituting an apparent contact angle into that
equation can serve only as an approximate guide [20]. This obviously raises the
question of which contact angle to be used in describing a wetting process. According
to Kissa [3], the contact angle concept is useful, but the problem associated with it is
the confusion of equilibrium contact angle with an apparent contact angle in real non-
equilibrium systems.
The magnitude of the cosine of the angle is of fundamental importance rather than
the angle itself. The lack of precision in the measurement of a contact angle will
influence the accuracy of the cosine values. As the contact angle increases above 15°,
the percentage error in the cosine function increases markedly and reaches infinity at
90°. On the other hand, very low contact angles represent an almost constant value of
cosine of contact angle. In summary, Miller and Young [8] concluded that direct
measurement of contact angles as a means for determining the wetting phenomenon
for filament–liquid systems is an unsatisfactory approach both from the experimental
and statistical points of view.
Several methods are available to measure the contact angles which are based on (i)
Goniometry and (ii) Tensiometry.
In Goniometry, the shape of a drop of test liquid placed on a solid is analyzed.
Contact angle is directly measured by measuring the angle formed between the solid
and the tangent to the drop surface, utilising image capturing. Using goniometry,
analysis of very small quantities of liquid on surfaces of relatively flat or regular
curvatures is possible. The assignment of the tangent line that defines the contact

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 7

angle is difficult and relies on the consistency of the operator. The conditions that
produce advanced and receded contact angles are sometimes difficult to reproduce;
as such, goniometry is not suitable for analysis of the effects of wetting on changes
in contact angle. The velocity of a liquid on solid is difficult to control. Further, the
amount of sample for each measurement is limited and multiple measurements should
be used to characterize a surface. Hence, fibers are not easily studied by goniometry
[23].
In Tensiometry, contact angles are calculated indirectly from the wetting force
when a solid is brought into contact with a test liquid, using the Wilhelmy principle
(discussed in subsequent sections), and the perimeter of the solid and surface tension
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

of the wetting liquid. In this technique, the force measured as the solid is immersed
into liquid represents an average force representing all points along the perimeter of
the solid at that depth. Thus the contact angle is an already averaged value. Wetting
force can be objectively measured over an entire range of velocities, from static to
rapid wetting. Variations of contact angles, both advancing and receding, can be
obtained for the entire length of the sample. This is useful in estimating the wetting
hysteresis, which is the ratio of the work of adhesion in the receding mode to that in
the advancing [8]. Analysis of fibers using tensiometry is much easier than with
goniometry [23].
Contact angle measurement by placing a drop on a horizontal surface and using
direct or indirect optical scanning of the drop profile has some uncertainties; the
resulting contact angle depends on drop size and drop deposition energy [24], as well
as on the possibly atypical physical or chemical characteristics at the arbitrarily
selected three-phase intersection [25]. Since textile fibers do not have ideal surfaces,
their wetting phenomena are complicated by surface roughness, heterogeneity, and
adsorption of liquids or surfactants with a consequent change of surface energy. On
such non-ideal surfaces, the measured (apparent) contact angle θ exhibits hysteresis
[26–29]. The contact angle displayed after the liquid front has advanced is usually
higher than the contact angle formed by the receding liquid, or the receding contact
angle, and so the advancing contact angle is usually employed in discussions of
wicking.
According to some authors, for many liquid–solid combinations there is a noticeable
difference between the contact angles observed for advancing and receding wetting
[28, 30–32]. The receding angle is always smaller and the corresponding work of
adhesion is larger. The possible reasons proposed are: (i) the advancing, or receding,
or both, angles are not in equilibrium; (ii) the receding angle is smaller because the
interface then is in contact with a pre-wet surface. A Wilhelmy scanning experiment
negates this idea. If a receding scan is interrupted and is changed to the advancing
mode before the sample breaks free of the liquid, the contact angle is observed to
increase, usually back to the value found for the first immersion. Thus, pre-wetting
of the surface is not a factor. At this time, a good explanation of this wetting hysteresis
is not available, though there is a fairly widespread belief that its fundamental cause
comes from the surface heterogeneity of nearly all solids, especially organic polymers
[31, 33].
Liquid droplets placed on a surface may produce an advancing angle if the drop is
placed gently enough on the surface, or a receding angle if the deposition energy

© The Textile Institute


8 Textile Progress doi:10.1533/tepr.2006.0001

forces the drop to spread further than it would in the advancing case. Wilhelmy
scanning usually guarantees that the wetting direction is known, and is the best way
to provide reproducible hysteresis data. Whatever method one uses to measure wetting,
precautions should be taken that evaporation does not change a supposedly advancing
wetting condition into a receding one [30]. In fact, the best indication that one has
produced true advancing wetting for a static situation (when the interface is apparently
not moving) is the observation that the contact angle decreases with time (i.e. with
evaporation). If, on the other hand, the angle does not change, it very likely was a
receding angle to begin with.
Capillary flow is not determined by a constant advancing contact angle θA, as is
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

frequently assumed, but depends on a dynamic contact angle corresponding to the


instantaneous velocity of the moving meniscus. The contact angle of a moving liquid
front (dynamic contact angle) can be quite different from the contact angle formed by
a static liquid [34, 35]. Dynamic contact angle can depends on the velocity of the
moving contact line, on time, or on both.

2.2.2 Work of Adhesion


A more useful and direct measure of interaction between solid–liquid is the work of
adhesion, WSL or WA, which has already been discussed in the preceding sections and
is given by equations (3) and (4). Work of adhesion is equal to the sum of two
quantities, one of which is a property of the liquid and the other the result of interaction
between the two phases. Work of adhesion is determined from the known values of
the surface tension of the liquid, and measuring the contact angle through goniometry
or tensiometry.
The two ways of using wetting data (work of adhesion and contact angle) are
compared in Table 1, for three polymers tested with four liquids [33]. Toluene and
ethylene glycol form the same contact angle (57°) on nylon. The work of adhesion of
nylon with ethylene glycol is 73.5 erg/cm2, whereas with toluene is 43.9 erg/cm2,
indicating that nylon is more strongly attracted to the latter liquid. As another example,
the attraction of 95% ethanol to polypropylene is much less than that of water to
nylon, despite the lower contact angle for the former solid–liquid system than the
later. These illustrate quite clearly the limitations of the contact angle as a quantitative
index of solid–liquid interaction; in other words the usefulness of the work of
adhesion.

Table 1 Work of Adhesion Compared with Contact Angle*

EtOH (95%) Toluene Ethylene glycol Water



Nylon WSL 43.6 43.9 73.5 95.7
(θ ) (18°) (57°) (57°) (71°)
Polyester WSL† 42.6 44.3 70.4 90.9
(θ ) (26°) (56°) (61°) (75°)
Polypropylene WSL† 37.7 – 61.0 77.4
(θ ) (47°) – (74°) (86°)

*From Miller[33].

erg/cm2.

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 9

2.2.3 Evaluation of Wetting Force Using the Wilhelmy Dynamic Method


Wetting of fibers or filaments is extremely difficult to measure by drop deposition
and accurate drop profile observations [36–38], and reproducibility of such results is
also poor. For such geometry there is a much better experimental approach, based on
the Wilhelmy wetting force technique originally developed to measure surface tensions
of liquids [39]. Wilhelmy dynamic procedure is considered the most effective and
useful way of assessing the contact angle [15]. Further, the major advantage of the
Wilhelmy method is that it provides an estimate of the dynamic contact angle when
the results are extrapolated to a macroscopic scale, as in wetting processes in textiles
and composites. According to Miller et al. [40], Wilhelmy-type wetting force
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

measurements utilizing sensitive instrumentation are being used more and more to
quantify the surface energy of solids generally, and of filaments in particular. Advancing
and receding contact angles can be obtained from wetting force measurements alone,
without the need for independent determinations of specimen perimeter or liquid
surface tension.
When a solid is partially immersed in a liquid, the liquid exerts an attractive force
on the solid in addition to buoyancy effects [33, 39]. The vertical component, Fw, of
this force can be measured by weighing the solid before and after immersion, usually
by suspending it from a hangdown weighing device as indicated in Fig. 4. The source
of this attractive force is the surface tension of the liquid, and Wilhelmy [39] derived
the relationship:
Fw = γLVP cos θ (7)
where Fw = measured wetting force in mN; P = perimeter of the solid in m.

Filament Filament partially


suspended freely immersed
F1 F2

θ
γ LVP
Fb
Mg
Mg

Fig. 4 Forces on a suspended filament [33 Miller, 39 Wilhelmy]: F1 = Mg = weight;


Fb = buoyancy; F2 = F1 – Fb + FW

Figure 5 shows the elements involved for a cylindrical solid partially immersed in
a liquid with vertical walls normal to the liquid surface. The dashed line around the
solid represents the system boundary. In the experiment, the liquid reservoir is moved
slowly up or down at a constant speed so that viscous and inertial forces can be
neglected, or is in a static condition. The forces involved are as follows [33]:

© The Textile Institute


10 Textile Progress doi:10.1533/tepr.2006.0001

PAAS

γ SV

θ γ LV
γ SL y

PLAS
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

mg

Fig. 5 Diagram of the external and internal forces on a partially immersed suspended cylinder
[33 Miller]

mg is the weight of the solid plus the weight of liquid contained within the control
volume; as the dashed line is moved closer to the solid surface, mg approaches a
limiting value ms g, the weight of the solid; F is the upward pull on the solid,
transmitted by the suspension wire attaching it to the balance, which is equal to the
force measured by the balance.

FA = PA AS = Air pressure × cross-sectional area of the cylinder, a downward force on


the solid.
FB = PLAS = Upward force (buoyancy) on the solid produced by the liquid pressure
at depth y. This is equal to PAAS + ρLgyAs, y being the depth of immersion measured
from the horizontal surface of the liquid, and ρL the density of the liquid.
Wetting force (Fw) is the vertical component of the liquid–vapor surface tension,
a downward or upward pull. This is an external force because it originates in the
liquid–vapor interface outside the control volume.
Under equilibrium, the measured force, F is given by:
F = γLVP cos θ + msg – ρL gy As (8)
Equation (8) is the general form of the Wilhelmy relationship, which includes the
weight in air of the solid, usually nulled out in the experiment, and the buoyancy
force, which must be considered if there is anything other than zero immersions.
Once these forces are accounted for, equation (8) becomes Fw = γLV P cos θ, the
classic Wilhelmy expression for the corrected measured force. Although the Wilhelmy
wetting force technique is an effective tool for studying fiber–liquid interactions, the
Wilhelmy equation is based on an assumption that may not be absolutely correct. The
magnitude of the attracting force vector is considered to be equal to γLV. As previously
mentioned, there is no reason why the total attraction between solid and liquid could
not be greater than 2 γ LV. To account for this, the Wilhelmy equation might be changed
to [33]:
Fw = P cos θ (aγLV) (9)
where ‘a’ is a parameter of liquid–solid attraction. This would acknowledge that the
interphase attraction is related to γLV but is not necessarily equal to it.
© The Textile Institute
doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 11

2.3 Wetting of Fibers and Filaments


For fibers as well as for other solids, wetting phenomenon can be divided into two
general classes: equilibrium wetting, where liquid and solid phases once placed in
contact are no longer externally perturbed; and dynamic wetting, where the liquid or
solid (or both) is kept in motion relative to the other phases throughout the wetting
process [4, 8].

2.3.1 Measurement of Wettability of Fibers and Filaments by Tensiometry


Methods reported for measurement of wetting force are discussed below.
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

2.3.1.1 Wilhelmy Method


Several workers have studied wettability using this technique [8, 15, 30, 33, 39–48].
A general arrangement for a Wilhelmy wetting force experiment is shown in Fig. 6
[33]. To carry out measurements on textile fibers or filaments, it is necessary to use
a recording microbalance with sensitivity of the order of 0.1-10 µg, depending on the
size of the filament; a high-sensitivity balance for finer filaments. A liquid reservoir
kept over an elevator is moved up and down, so that the surface of the liquid scans the
surface of the solid [40]. Such scanning must be done slowly (e.g. 0.2 mm/min) and
at constant velocity to avoid filament/fiber bending, and disturbances in the recorded
force coming from viscous friction between the liquid and fiber [47].

Electronic microbalance

Chart recorder output


Fiber hook
Force Monofilament
Wetting chamber
Time

Fiber preparation cut


Reversible elevator

Fig. 6 Apparatus for Wilhelmy wetting force measurements on single filaments [33 Miller]

To mount a filament sample, a hook is glued to each end. It is then cut with a sharp
blade to produce two specimens about 10 mm in length, as illustrated in Fig. 6 [33].
In this manner, data can be obtained from adjacent surface regions. The weight of the
sample suspended from the microbalance and hook in air is zeroed out before the
liquid contacts the sample. The liquid is raised until a force change is registered as a
result of initial contact with the sample [8, 33]. The liquid is raised further to produce
a specific depth of immersion of the sample, the drive is reversed and the liquid
moves down until it breaks free of the specimen [40].

© The Textile Institute


12 Textile Progress doi:10.1533/tepr.2006.0001

A typical recorder force plot is shown in Fig. 7. When the liquid first contacts the
bottom of the specimen is easily identified as Fw(a). The upward scan shows a
negative slope because of increasing buoyancy [40]. At time ts, the drive is reversed
to produce a receding wetting scan. The interval A-B represents the transition of the
meniscus from a larger advancing wetting angle, θa, to a smaller receding one, θr.
During this transition period the position of the three-phase interface does not change,
i.e. scanning does not take place although the liquid surface is moving downward
[30]. The events that occur are illustrated in Fig. 8 [41]. When the solid surface is
being scanned, there is always some ‘noise’ in the force trace due to localized surface
variations which perturb the wetting force to some minor degree [41]; the lack of
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

‘noise’ in region A-B indicates that no scanning is taking place during A-B.

F
M
FM

P
Fw(r) FB
g
B Recedin

Fw(a) Advancing
A
0
tS 2tS Time

Fig. 7 Typical wetting force scan [40 Miller]

To balance

θa θr < θ < θa θr θr
(a) (b) (c) (d)

Fig. 8 Sequence of events when an advancing scan is changed to receding [41Bayramli]: (a)
the advancing angle θa before reversal; (b) contact line remains fixed while the liquid level
drops and θ begins to decrease; (c) the receding angle θr is established; (d) scanning
recommences, the contact line moving downward

As long as the liquid surface remains in contact with the vertical surface of the
filament, the receding force (corrected for buoyancy) remains the same. Eventually,
the contact line reaches the bottom edge of the solid and must change its orientation.
It moves around the corner, which, no matter how sharp the cut, is not exactly square
on a microscopic level [42, 43]. However, on a macroscopic scale the perimeter can
be considered as unchanged during the transition, which is described in Fig. 9. The
liquid surface must maintain the same receding angle of contact as it moves around
the edge, but the effective angle with the vertical wall, which determines the measured
force, changes [42, 43]. At some point in this transition, the direction of the pulling
vector is straight down and the maximum contribution of the surface tension force is

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 13

θr

θr θr θr
(a) (b) (c) (d)

Fig. 9 Transformation of the interfacial contact line from the vertical to the bottom surface
during pull-out for θ r < 90° [42 Lane, 43 Padday]; (b) and (c) show the edge greatly magnified
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

recorded. This is represented in Fig. 7 by the rise in the force reading from P to M
[40]. At M the effective angle with the vertical wall is essentially zero, and the
Wilhelmy relation reduces to:
Fw(M) = PγLV
The coincidence of maximum force reading and zero effective angle may not be
exact if buoyancy end effects are significant [42, 43]; however, the error in making
this assumption is minor when small filaments are used.
Precise measurement of the wetting force is possible only when the sample filament
is straight, rigid, and vertically oriented during immersion [33]. Hsieh et al. [45] used
short length fibers. Barraza et al. [47] cut fine glass fibers to 5 mm to avoid bending
of fibers while submerging into liquids. This may allow the sample to be rigid; it does
not guarantee that a crimped fiber will remain extended or vertically oriented. Crimped
fibers or very fine and long filaments cannot be expected to comply with this requirement
without some loading on their bottom end (sinkers). The added load introduces a
complication because of its considerable buoyancy effect, which has to be corrected
for to determine wetting force. A method for handling such materials [33] is shown
in Fig. 10. Two ends of a looped fiber are glued to a triangular tab and the closed end

To balance

Mounting tab

Glue

Fiber

Platinum sinkers
1
2

Fig. 10 Method for handling crimped or very fine filaments [33 Miller]

© The Textile Institute


14 Textile Progress doi:10.1533/tepr.2006.0001

of the loop is linked to a platinum wire sinker to produce two vertical sections in
parallel. A wetting force scan is carried out and then an additional sinker load is
added and a second scan is performed.
The measured forces in each of these cases would be [33]:
F1 = Fw – Fb – B1
and
F 2 = F w – F b – B 2,
where Fb = buoyancy of immersed fibers, and B1, B2 = buoyancy of submerged
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

sinker(s).
The above can be rearranged to give a ratio:

B1 F – Fb – F1
R= = w
B2 Fw – Fb – F2
or
RF2 – F1 + Fb ( R – 1)
Fw =
R–1
If Fb can be eliminated (e.g. by extrapolation), then:
Fw = (RF2 – F1)/R – 1
The ratio R can be determined with considerable accuracy by weighing each of the
platinum sinkers. Therefore, Fw can be determined from the two measured forces, F1
and F2, and the actual buoyancy of the sinkers need not be determined. The magnitude
of the measured wetting forces is twice the amount for a single fiber, which is
advantageous when very fine fibers are used [33].
According to Minor et al. [36], the pull on a single fiber during immersion is too
small for convenient measurement; the pull on several fibers must be measured
simultaneously. A frame or zither having 50 evenly spaced fibers, each in U shaped
form was therefore prepared. This allowed the pull on 100 fibers to be measured
simultaneously. Spacing devices (a fine-tooth comb with the teeth sawn off nearly to
the root), one at the top and the other at the bottom of the fibers, supported by pillars
at the sides) were used to space each fiber far enough apart to avoid capillary action
in the inter-fiber spaces. Miller and Young [8] also described a methodology to hang
many filaments/yarns side-by-side for studying the wettability. A linear plot of observed
force of all filaments Fw as a function of depth of immersion (L) is obtained, with
intercept I and slope S. The wettability of a filament (W) is then calculated as
W = (1/NP) [(I – (xS/N) + Bs)]
where N = number of vertical filaments, P = perimeter of a filament, x = horizontal
filament length, and Bs = buoyancy force of the submerged holder and sinker.
The ratio of the derived cross-sectional areas of the filaments (from the slope of
plot of the Fw versus depth of immersion) to the ones measured from the scanning
electron microscope is found to be somewhat greater than unity. A significantly high
ratio indicates that the liquid does not completely wet the fiber [8].

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 15

According Carroll [25], the Wilhelmy method, when applied to fibers, suffers
from two disadvantages: (1) difficulty in ascertaining the perimeter of the fiber at the
three-phase contact line; and (2) it is necessary to know independently the interfacial
tension of the two fluid phases. The first one arises when the liquid does not wet the
fiber completely, as, for example, indicated by the high ratios of derived to measured
cross-sectional areas of small-lobe polypropylene filaments in the experiments conducted
by Miller and Young [8]. This suggests that independent measurement of fiber perimeter
through scanning electron micrographs has to be carried out and this value may be
substituted in arriving at the wettability values from the wetting force values.
Whang and Gupta [15] characterized the wetting of cellulosic fibers. The fibers
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

were attached to sinkers to keep the specimens straight during immersion. The wettability
of the system, W (‘specific wettability’, ‘adhesion tension’, ‘wetting index’, as termed
by different workers) was defined earlier as:
W = Fw /P = γLV cos θ
The force, F, measured by the microbalance as the result of the suspended sample
with sinker immersed in liquid is [15]:
F = mg + Fw – Fbs – Fbf ± Fv (10)
where mg is the combined weight of the sample and sinker in air; Fw is the actual
wetting force; Fbs is the buoyancy force of the sinker; Fbf is the buoyancy force of the
sample; and Fv is the viscous frictional force of the fluid on the filament. The sign of
Fv is negative for the advancing and positive for the receding modes. The first term
in the right side of equation (10) is usually tared out at the beginning of the experiment.
The fourth term is omitted as it is negligible compared to the wetting force. The fifth
term that is related to the viscosity and speed of immersion was found to be negligible
in the experiment [15]. The buoyancy force of the sinker was calculated from the
dimensions of the sinker and density of the fluid and this value was substituted in
equation (10). Substituting the measured values of P and known values of surface
tension of the liquid γLV, the contact angles (advancing and receding) were calculated.
Work of adhesion was calculated from equation (4), WA = γLV (1 + cos θ).
Whang and Gupta [15] have determined polar (γ sp ) , dispersive (γ sd ) , and total
surface energies (γS), of fibers by immersing a fiber separately in two different liquids
(deionized ultra filtered water and methylene iodide) whose polar and dispersive
contributions to the surface tension are known. From the Young–Dupré equation for
liquids, i = 1, 2. (water, methylene iodide)
–γS + γSLi = – γLi cos θi (11)
Using Kaelble’s [49] model for solid–liquid interfacial energies as
γSLi = γS + γLi – 2 (γ Sd γ Li ) – 2(γ Sp γ Li
d 0.5 p 0.5
) (12)
The contact angles were measured on these two liquids, θ1 and θ2. The polar and
dispersive values of the surface energy of the fibers were determined by substituting
the known values of components of surface tension, γ Ld 1 , γ Ld 2 , γ Lp1 , and γ Lp 2 in
equations (11) and (12).
The wetting force at the three-phase line was measured by the Wilhelmy technique

© The Textile Institute


16 Textile Progress doi:10.1533/tepr.2006.0001

using fibrous solids/liquid/liquid systems. Advancing and receding contact angles


were calculated from the wetting forces during fiber immersion and emergence. The
dispersive and polar components of surface free energies of the fibers were determined
from the advancing and receding contact angles, respectively [29].
The circumferential dimension of Kevlar 49 single fibers, their wetting force and
work of adhesion with an epoxy resin were evaluated by Hsieh et al. [45]. Fiber was
cut to a length of few mm and attached to a sinker. The wetting characteristics of the
fibers were studied in water and epoxy resin in stepwise steady-state wetting, dynamic
wetting, and multiple immersions. The steady-state wetting force measurement was
made by stopping the liquid level at a location along the fiber.
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

Advancing and receding contact angles on single elastomer-modified glass


fibers as small as 6.2 µm in diameter, were measured at slow meniscus velocities
(2.0 µm/s) with probe liquids of various hydrophobic/hydrophilic characteristics
[47].
Several experimental configurations of the Wilhelmy method have been reviewed
for measuring surface tension and contact angles in wetting and nonwetting systems
where the sensing element is moved through a liquid/fluid interface. Adhesion tension
measurements of several lipid types at the oil/water interface provide examples where
contact angles are assumed in order to calculate interfacial tensions. One of the lipids
studied was the natural secretion of human skin, which is highly surface active as a
result of the presence of 1% of polar material. The method was also applied to single-
fiber adhesion tensions, where the interfacial tension is assumed in order to calculate
the contact angles of fibers exposed to surface treatments [44].
Contact angles, work of adhesion, fiber surface energy, and interfacial energy have
been calculated for glass fibers, utilised as reinforcement for thermoplastic and thermoset
composites, using the Wilhelmy method with various liquids [46]. Measurements
were made with a Cahn DCA 322 microbalance based system. The wetting liquid
reservoir, mounted on an elevator, was raised or lowered at a slow, steady rate (40
µm/s). The liquid was then raised until a force change was registered as a result of
first contact with the suspended sample. The liquid was then raised further to produce
a specific depth of immersion of the sample. The apparent weight at this depth was
recorded on a chart recorder and was fed to a computer for storage and analysis. The
final depth used for the experiments was 5 or 10 mm. Thereafter, the liquid in the dish
was lowered at a carefully controlled rate while the electrobalance measured the
force. In most cases, five or six specimens of each kind of sample were run and a plot
of force as a function of depth of immersion was constructed. A typical curve is
shown in Fig. 11. As the liquid is first bought into contact with the fiber, the force
jumps to the advancing interline condition. The slope of the line results from the
buoyancy correction. When the liquid level movement is stopped and reversed, and
the meniscus is reoriented, the force trace yields the receding contact angle. The
perimeters of the fibers were calculated from the measurements in the liquid hexadecane,
in which a total wetting out of the fiber by the liquid took place (θ = 0). The fibers
surfaces were analyzed by scanning electron microscopy [46].
Li et al. [48] used the Wilhelmy balance method and the solidification front
method to determine the surface tension of fibers for predicting the adhesion bond
strength between reinforcing fibers (carbon and an aromatic polyamide) and resin. It

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 17

Buoy
ancy
slope
Out Rece
ding
angle
A
2

h 1
D

1 2
Force

Adva
ncing
angle
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

Zero
depth
Immersion depth

Fig. 11 Force vs immersion depth for the Wilhelmy technique [46 Van de Velde]

is suggested that the reproducibility and accuracy of these techniques enable the
surface tension of small-diameter fibers to be determined.
The fact that the dynamic contact angle does not change when the filament is
advancing at a slow rate leads to a useful application of the Wilhelmy wetting-force
measurement on a moving filament. By arranging to record the wetting force
continuously as the liquid–air interface moves up or down the filament, one can
monitor random or systematic surface irregularities with considerable precision and
sensitivity. Bendure [44] was the first to demonstrate this technique, using a polyester
filament that most likely had titanium dioxide particles embedded on its surface. An
example of the reproducibility of such a scan is shown in Fig. 12.

Receding wetting (r)


(3)

Force (1) (2)


(weight)

Liquid movement

Fig. 12 The results of three continuous wetting-force scans over the same length of fiber
surface [44 Bendure]

The processing behavior and the mechanical properties of polymer composites are
influenced by the properties of both the bulk materials and the interface. The mechanical
properties of fiber-reinforced polymers are strongly influenced by the surface
characteristics of the fibers and the polymer matrix. Wetting measurements have been
extensively used to investigate fiber surfaces [50]. Dynamic wetting measurements
and micromechanical and macromechanical testing have been used to characterize
the fiber/matrix interaction. Carbon and glass fibers with various resins were studied.
© The Textile Institute
18 Textile Progress doi:10.1533/tepr.2006.0001

The experimental results of dynamic wetting show that the shape of the tensiogram
depends both on fiber surface and on resin chemistry. The slope of the force versus
stage position plots represents the fiber/matrix interaction as swelling or acid-base
interaction and depends on stage speed and fiber roughness. Wilhelmy wetting tests
can be used to study the fiber/matrix interaction but conclusions about the resulting
interphase can be drawn only in conjunction with micromechanical and
macromechanical testing.
The wetting behavior of polypropylene fibers has been studied extensively using
the Wilhelmy technique [51, 52].
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

2.3.1.2 Liquid Membrane Method


Kamath et al. [53] described a technique for characterizing the wettability of relatively
long filaments based on scanning the filaments with a liquid membrane. Liquid
membrane scanning involves contact of the fiber with two liquid surfaces; the force
recorded by the electrobalance is thus a function of both the advancing and receding
wetting forces.
The total force is given by:
F = P γLV (cos θa – cos θr) (upward)
or
F = P γLV (cos θr – cos θa) (downward) (13)
where F is the force recorded by the balance, P is the wetted perimeter (measured
independently), and θa and θr are the advancing and receding contact angles.
The liquid membrane technique is particularly useful for fibers of uniform perimeter.
If significant perimeter changes are encountered along the length of the specimen,
they have to be determined independently and used in the calculation of the contact
angle. The choice of liquids for the membrane is guided not only by the requirement
that interactions between fiber and liquid provide an optimum differentiation between
treated and untreated domains on the specimen surface, but also by the necessity that
the membrane be stable. Aqueous media are preferable because of their relatively
low rate of evaporation, but very dilute aqueous solutions of surfactants pose problems
because membranes of these liquids tend to be unstable due to concentration fluctuations
and associated surface flows. Organic liquids of low volatility can also be used. This
method can be used for studying surface distribution of finishes on relatively long
filaments, even with crimp, and wetting hysteresis. The technique overcomes the
limitations of specimen size and crimp which are inherent in the bulk immersion
method for evaluating wettability changes along a filament [53].

2.3.2 Methods based on Goniometry


Various methods based on goniometry are discussed below.

2.3.2.1 The Shape of a Drop Surrounding the Fibers/The Sessile Drop Method
According to Carroll [25] and Jones and Porter [54], measurement of contact angles
by methods that work well with solids of flat plate shape does not give very accurate
or reproducible results when applied to fibers. The principal reason for this lies in the

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 19

fact that when a thin cylindrical solid supports a liquid meniscus, the part of the
meniscus close to the solid has principal curvatures, both of which are very high (of
the order of the reciprocal of the cylinder radius) and which are of opposite sign. This
high curvature of the meniscus jeopardizes the methods, which depend for success
upon accurate visual observation of the meniscus. It is not possible to draw a tangent
to the meniscus at the three-phase contact line with much confidence. A more viable
method, based on the reflection of a point light-source by the meniscus, has been
described [54], but this again suffers in reliability when the fiber size becomes so low
(100 µm or less) as to make the meniscus very highly curved.
Carroll [25] described a method for determining the contact angle on fibers which
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

is analogous to the sessile drop method for flat plates. It is shown that when the
Goucher Number (the ratio of fiber radius to capillary length) is small, then capillary
rather than gravity forces mould the drop shape. In such cases, the profile of the drop
as described by the length, surface area and volume of the drop, is a sensitive function
of the contact angle. Filaments of about 200 µm diameter are used. The filament is
driven into an aqueous phase via an oil phase and then stopped. The entrained oil
breaks up into droplets and then is either photographed (using transmitted or reflected
light depending on the opacity of the fiber) through a low power microscope or
measurements are made directly by means of a micrometer. Analytical expressions
have been derived relating the length of filament, surface area and volume of the
drop, and Laplace excess pressure inside a liquid drop adhering to a cylindrical fiber
to the linear drop dimensions and the contact angle. The method has the great advantage
that it is applicable to drops on fibers of any size, provided only that gravity effects
are negligible. A further advantage is that measurements are possible without in any
way disturbing the drop–fiber system. The method is restricted to contact angles
below about 60°, but as in principle it is possible, in two liquid systems especially, to
invert the system (water drops in oil instead of the converse), the limitation will in
many cases apply only to the measurement of intermediate values of the contact
angle (60° < θ <120°) [25].
Yamaki and Katayama [24] derived the shape of a drop (length, maximum radius,
and apparent contact angle) attached to a monofilament using Laplace’s excess pressure
inside the liquid drop. The relations between contact angle and the dimensions of
drops are theoretically obtained. Thus, it is possible to calculate the contact angle if
the drop shape is measured from photomicrographs of the drops. They verified the
drop shape which they derived, with the actual drop shape observed. Contact angle
between epoxy resin and various kinds of monofilaments have been measured by this
method. According to them, this method has practical utility for measurements of the
contact angles between liquids and monofilaments.
Carroll [55] discussed the effect of system parameters on the deposition of drops
from a dispersion onto the surface of a collecting fiber. The relative predominance of
a particular deposition mechanism depends on the stage of the overall process, and
stages are identified where the contact area changes are insensitive to the contact
angle.
Minor et al. [36] examined the physical behavior of small droplets (weighed in the
range of 1 µg) of organic liquids on a variety of textile fibers. On smooth-surfaced
fibers of approximately cylindrical cross-section, the droplets formed either unduloids

© The Textile Institute


20 Textile Progress doi:10.1533/tepr.2006.0001

or clamshell-shaped beads clinging to one side of the fibers. Unduloid droplets always
have zero contact angles while clamshell droplets form contact angles above 60°.
Contact angles of various liquids on wool fibers are measured using the liquid droplet
method of Le et al. [56]. A computational technique greatly simplifies contact angle
calculations. Measurements of contact angles of liquids on wool fibers, however, are
limited, in that the unduloid shape of the liquid droplet can only form when the
contact angles are less than 60° [56].

2.3.2.2 Contact Angle Measurement by a Reflected Light Beam


A modified procedure for measuring the contact angle of a liquid drop on a solid has
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

been described [57]. The method is based on the observation of the angle at which
light from a point source is reflected from a liquid drop surface at its contact point
with a solid. It gives results that are the same, within experimental error, as those
obtained by the usual ‘drop profile’ method. However, it is limited to angles <90°.
A simple apparatus and technique are described for measuring contact angles of
liquids on small-denier fibers by Grindstaff [37]. This technique is based on the level-
surface method as described by Adam and Morrell [58]. This can be used to obtain
either advancing or receding contact angles. The apparatus is arranged so that the
profile of the region where the liquid drop forms the contact angle on the fiber is
viewed through a microscope and the reservoir which contains the liquid is rotated
until the liquid surface is level right up to the point of contact with the fiber. The
angle between the fiber and the liquid surface is read on a protractor. The fiber
sample is mounted on one side of the axis of rotation of the reservoir so that the liquid
advances up the fiber when the reservoir is rotated and the advancing contact angles
are measured. Receding contact angles are measured by rotating the reservoir in the
opposite direction. Some of the limitations of the instrument outlined by Grindstaff
[37] are:
• Contact angles between highly volatile liquids and fibers cannot be measured;
• The geometry of the fiber samples must be such that they can be oriented with an
edge which is straight, perpendicular to the liquid surface; this makes measurements
on some natural fibers, for example cotton, difficult; and
• Very low contact angles (15° or less) are difficult to measure.
The first limitation has not proved to be a problem since most liquids of interest are
not very volatile; however, this limitation could be overcome by enclosing the apparatus
so that the liquid can be equilibrated with the vapor. The second and third limitations
are inherent in the level-surface method. Common errors in contact angle measurements
on fibers using the level-surface method are discussed by Grindstaff and Patterson
[38].

2.3.2.3 Critical Surface Tension of Fibers Using a Homologous Series of Liquids


Zisman [59] measured advancing contact angles of a homologous series of liquids on
a solid and plotted cos θ values against the surface tension of the liquids. The lines
were extrapolated to the same γLV value at cos θ =1. Zisman named this γLV value the
‘critical surface tension’ of the solid, and proposed that only liquids having surface
tensions below this value spread on the solid. When the plot of cos θ against γLV

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 21

resulted in a rectilinear band, Zisman chose the intercept of the lower side of the band
at cos θ = 1 as the critical surface tension. Although critical surface tension is an
empirical concept with several deficiencies and limitations, critical surface tension
data have been very useful in evaluating the wettability of fibers [60] and developing
water- and oil-repellent finishes for textiles and paper [10]. The values of surface
energy obtained by this method are largely characterized by the dispersive fraction of
the total energy. The best method available for determining surface energy on a
material that has significant polar interactions is ascribed to Kaelble [49]. This method
involves measuring the contact angles of a solid with two dissimilar liquids whose
dispersive and polar contributions to surface tensions are known.
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

A method, based on that of Grindstaff [37], that can be used to rapidly measure the
critical surface tension, γc, of individual wool fibers at all positions along their length
except the root and the tip is described by Byrne et al. [61]. Contact angles were
measured with various liquids whose surface tensions (γLV) and dispersive components
(γ LV
d
) are known. By extrapolating the plot of cos θ versus γLV to cos θ = 1, the
critical surface tension of the fiber (γc) was estimated. Liquids having a high ratio of
d to γ
γ LV LV yielded a slightly higher γc than with those liquids having a lower ratio.
This indicates that γc depends on the test liquid used. Chemical treatment of wool
fibers increases γc slightly but also modifies the proportion of dispersion, polar, and
H-bonding components of the wool surface [61].

2.3.2.4 Dynamic Methods


The efficiency of wet processing of textiles can depend on the dynamic wetting
behavior of the system because it decides the minimum exposure time [62]. The
static arrangement of a liquid in equilibrium contact with a solid surface can be
altered by either the liquid or solid being caused to move by some external force.
Most studies of this effect have been performed by forcing liquids through capillary
tubes [63]. The shape of the advancing liquid–vapor interface is observed to be
noticeably sensitive to its velocity. More pertinent to filament wetting is the situation
where a moving solid penetrates the surface of a liquid, passing either from the dry
side into the liquid or vice versa.
Under static conditions, contact angles are known to depend on the balance of
surface energies and gravity at the line. In a coating process, the solid surface is
steadily moving and the contact angle departs from its static value nonlinearly, increasing
within the range of complete wetting. At higher substrate speeds, the contact line
loses it dynamic stability and visible air bubbles may be entrained into the coating
layer. Dynamic contact lines and air entrainment are important factors in fiber coating.
Parameters of wetting dry fibers with non-newtonian liquids have been studied
experimentally. Dynamic contact angles are strong functions of the apparent viscosity
of the liquid at the particular wetting speed, regardless of whether the liquid is
newtonian or non-newtonian.
Ellison and Tejada [64] used direct photographic observations of the interface
when a continuous length of nylon filament was drawn into a liquid, and they reported
the increase in contact angle. Inverarity [65] carried out a more extensive study in a
similar manner with a variety of filaments. His results revealed a general trend
wherein the contact angle, while practically constant at low velocities, increased

© The Textile Institute


22 Textile Progress doi:10.1533/tepr.2006.0001

systematically at higher velocities, and in many cases reached nearly 180° (Fig. 13).
Inverarity’s [65] data showed that contact angle was dependent on the logarithm of
the product of liquid viscosity and filament velocity. These results were interpreted
to mean that viscous drag could be a dominant factor in the formation of a dynamic
contact angle. Schwartz and Tejada [66] measured dynamic contact angles, up to 90°,
for a series of organic liquids on moving filaments of stainless steel and polymers,
500 µm in diameter. They interpreted their results in terms of three different mechanisms
governing the dependence of contact angle on filament velocity:
(i) Initially the dynamic contact angle stays equal to the equilibrium value as the
velocity is increased (the Elliott and Riddiford [63] or Hansen and Miotto [67]
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

mode);
(ii) In the next velocity range, the relation between contact angle and velocity is
controlled by molecular motion along the filament surface, the retarding force
is dependent on an interfacial viscosity (the Blake and Haynes [68] mode); and
(iii) Finally, a surging mode of flow predominates, with the retarding force dependent
primarily on the bulk viscosity (when θ <90°).

1
cos θ

–1
0 2 4
log ηV

Fig. 13 General nature of the dependence of the dynamic contact angle on filament velocity
and liquid viscosity [65 Inverarity]

The movement of dry filaments into a liquid was discussed by Dyba and Miller
[62]. When the filament is stationary (if there is positive wetting), there is a raised
meniscus and the contact angle is less than 90°. When downward motion is imparted
to the fiber, the meniscus recedes and the contact angle becomes larger. In general,
it is possible to impose a velocity on the filament sufficient to eliminate the meniscus
rise, or, in other words, to bring the contact angle to 90°. A further increase in
velocity may produce an actual depression of the interface with a contact angle
exceeding 90°. Even if the initial static condition does not show positive wetting, the
same general trend of increasing contact angle with increasing velocity can occur.
A simpler experimental approach was used by Dyba and Miller [62] to obtain a
single-valued property for a given liquid–filament combination – the velocity required
to produce a 90° contact angle (designated the ‘rise-canceling velocity’, VC). A
mechanical drive system advances the dry filament at a variable speed into a liquid
held at a fixed temperature. Through an optical system, the image of the intersection
of the filament with the surface of the liquid is observed. This arrangement makes use
of the fact that undispersed reflection from the liquid surface will occur only if the
surface is flat; any curvature will cause a portion of the light to be removed from the
final image on the screen. The above procedure is best suited for studying the effect

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 23

of the wetting liquid properties, i.e. viscosity or the presence of dissolved solutes or
surfactants [69]. VC does not seem to be very sensitive to the chemical nature of the
filament, even such extremely different materials as nylon 66 and Teflon (PTFE)
show just about the same response in a given liquid. On the other hand, both the
physical and chemical aspects of the wetting liquid significantly affect this property.
Rise-canceling velocity shows a dependence on surface tension, surface adhesion,
and viscosity, which control the effectiveness of wetting under processing conditions
[62].
A major limitation of all high-speed composite processing is the wetting of fibers.
In molding processes, resins are forced to penetrate through fiber networks, driving
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

out the air that is initially in the space between individual fibers. Incomplete wetting
or impregnation often results in the formation of voids or air entrapment. In one
study, wetting of glass fibers was visualized using oil with the same refractive index
as glass [70]. Using a video-enhanced microscope and a high magnification video
camera, the flow front and a qualitative assessment of entrapped air bubbles by image
analysis was carried out. This method is useful to form guidelines of optimum processing
parameters of fiber wetting, namely viscosity, vacuum, mold temperature, and pressure,
for resin transfer molding and structural reaction injection molding.
Direct dynamic observation of wetting on a micron level can be done on the ESEM
(environmental scanning electron microscope). This instrument permits high
magnification and observation of wetting on a microstructure. Wetting phenomena
can be examined continuously, and images can be taken during the process in the
ESEM [71, 72]. The presence of liquid water in the ESEM specimen chamber allows
observations of in situ hydration without the need for coating or drying the sample
[71].
Microscopic observations of dynamic wetting of untreated and oxygen plasma
treated polypropylene fibers on the ESEM have been investigated [71]. The relative
humidity in the ESEM chamber was raised to 100% by controlling the specimen’s
temperature and the chamber pressure to produce water condensation on the fiber
surfaces. Dynamic wetting, i.e. the formation, spreading, and coalescence of water
droplets on fiber surfaces, was observed and recorded in real time. Contact angles
were measured from the ESEM images, using image analysis software which revealed
the dynamic wetting of textile fibers on a micron level. Adjustment of the viewing
angle in the ESEM placed the image in a good position for contact angle measurement.
Further, in ESEM gravity effects on droplets was negligible due to the micron range
droplets (less than 100 µm in diameter) [71].
Polypropylene fibers have been increasingly used in oil-related areas, such as oil
sorbents, oil filters, and oil separators. The wetting behavior of the fibers by oil is of
particular importance in these applications. The ESEM technique was used in this
study to investigate the oil wetting of polypropylene fibers [72]. The wettability was
investigated by contact angle measurements with the oil. Oil drops were added onto
fibers using a micro-injector. The contact angles were measured from ESEM micrographs
using image analysis and a theoretical approach. ESEM observation reveals the wetting
of single fibers on a microscopic level [72]. The axisymmetrical barrel-shaped droplets
on a single fiber observed in the ESEM can be further analyzed using Carroll’s
approach [25].

© The Textile Institute


24 Textile Progress doi:10.1533/tepr.2006.0001

A hydrodynamic study on the wettability of fiber assemblies [73] was performed


using an apparatus in which the dip-solution could be introduced to the sample
chamber under constant hydrostatic water pressure. The estimation of effluence and
pressure loss produced by the mass of fibers and air in the fiber assembly gave a flow-
resistance value and permitted a comparison of wettability. Various effects on the
flow resistance were determined on wool fibers having poor wetting properties and
the effectiveness of forced flow of the dip-solutions on level wetting was observed.

2.3.2.5 Wettability Test by the Sink–Float Method


This method is based on the principle that a solid placed on a liquid floats when the
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

surface tension of the former is less than that of the latter. In this case, the surface
tension of the liquid is in excess of the gravitational force of the solid and makes it
difficult for the solid to penetrate into liquid. When the surface tension of the solid
equals or just exceeds that of the liquid, free energy transfer from solid to liquid
attracts the liquid towards the solid surface, wets the solid completely (contact angle
becomes zero), and the solid submerges into the liquid. In one study, snippets of fiber
about 1-cm long were placed on the surface of a number of test liquids whose surface
tension values were known [74]. When the fibers were observed to sink in a liquid,
the contact angle was assumed to be zero and the surface tension of that liquid was
taken as the critical surface tension of the solid. Good agreement was found between
the results of this method and the sessile drop technique of Zisman [59] on a number
of polymeric materials.
This method essentially determines the surface energy of the solid. The critical
surface tension of a solid is an empirical quantity [75] which depends on the polar
and dispersive components of the surface tension of the test liquid [76]. Because of
hydrogen bonding between liquid and solid, H-bonding liquids generally do not give
the same critical surface tension as non-H-bonding liquids [77]. The test is time
consuming as it requires preparation of a series of liquids with a range of surface
tension values. However, the sink–float method provides an adequate measure of
relative wettability of samples [76]. Pittman [78] investigated the wetting behavior of
wool and mohair fibers using the technique developed by Mutchler et al. [74]. Ward
and Benerito [76] concluded that the sink–float technique, using either of two liquids,
namely a mixture of 20% dioxane in water or 1% acetone in water, could be used to
sort washed and unwashed cotton fibers. Cotton fibers from four varieties were
washed with water using two procedures that included several combinations of
temperatures and water volumes. All washed cottons sank in either of these two
solutions, while unwashed samples floated.
Bateup et al. [79] have measured the critical surface tension (γc) of wool fibers
with various classes of liquids. Nonpolar organic liquids and aqueous surfactants
gave similar values of critical surface tension. Lower values observed in butanol/
water mixture were attributed to butanol adsorption on the fiber surfaces. Removal of
lipid material from the fiber surfaces by extraction increased the γc from 30 mN/m to
37 mN/m. Exposure of wool fibers to corona discharge in air at atmospheric pressure
also increased γc [79]. Mutchler et al. [74] presented a theory of flotation of fiber-
shaped solids on a liquid of lower density. Within the usual range of fiber dimensions
and densities, provided the cross-section shows no protruding cusps, a very small

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 25

contact angle is sufficient to float the fiber. If the contact angle is zero, the fiber will
sink. The critical surface tension of a fiber surface can therefore be established by
placing samples of the fiber on a series of liquids of progressively increasing surface
tension. The critical surface tension lies between the surface tensions of the liquid in
which the fiber just sinks and the liquid in which it just floats.

2.4 Wetting of Yarns and Fabrics


This section discusses the wetting of yarns and fabrics.

2.4.1 Wetting of Yarns


Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

The wettability of yarns during air-jet texturing, of yarns and sewing threads during
finishing, and of high-performance yarns used in composites is critical, as it influences
the processability of the yarns and performance of the final products.
The reason for composites made from high-modulus carbon yarns and epoxy
resins exhibiting very low shear strength is widely accepted as being due to poor
wetting of the yarn by epoxy resins.
According to Acar et al. [80], wetting of the feed yarn in the air-jet texturing
process produces a lubrication effect to reduce the friction between the filaments and
the contacting parts. This causes a decrease in frictional forces and consequently an
increase in the resultant forces acting on the filaments, which in turn results in
textured yarns with improved properties.

2.4.2 Wetting of Fabrics


Wetting of fibrous assemblies including fabrics are of paramount importance in the
process of dyeing and finishing and in composites. Fabric–liquid interactions are
involved in so many aspects of day-to-day human life. Fibrous materials are designed
to absorb, transport, or reject liquids for applications in the home, filtrations, and
medical applications. In the case of surgical textiles, the barrier effect of fabric
against particle-loaded liquid is determined by the resistance to flow of liquid and
sorption of liquid on the fabric surface [81].

2.4.3 Methods of Measuring Fabric Wettability


Although the wettability of a single fiber/filament can be derived directly and consistently
from wetting force measurements using the Wilhelmy technique, wetting force
measurements on fabrics are difficult to perform because the dynamic liquid–fiber
interaction processes in fabrics involve not only wetting but also liquid uptake in the
pore structures and, for some fibers, liquid absorption within the fibers [1, 82]. For
instance, the readiness of a fibrous assembly to be wetted can be measured in terms
of the length of time for the material to sink below the surface of the liquid [83] or
the time for a drop of liquid to disappear from the material surface [84]. These
methods provide the collective measurements of simultaneous wetting, transport, and
absorption processes, and their results can be highly influenced by fabric geometry.
Another way to measure fabric wettability is the sessile drop approach of obtaining
the solid–liquid contact angle, assuming smooth surfaces. However, fabric wettability
measured by such an approach can be erroneous because of the inherent roughness of
fabric surfaces. If the intrinsic contact angle for a liquid on a solid is less than 90º,

© The Textile Institute


26 Textile Progress doi:10.1533/tepr.2006.0001

the apparent contact angle decreases with increasing surface roughness [85]. Therefore,
fabrics with identical fiber content may yield different apparent wetting contact
angles depending on fabric geometries or the roughness of the surfaces.
Wetting of liquids through fabrics can be classified under two general categories
of phenomena: flow in the plane of the fabric, and flow through the fabric in a
direction normal to the plane. Due to structural anisotropy of fabrics, these two
modes of liquid movement would not be affected similarly by changes in either fabric
or liquid properties [86]. Studies of forced flow through the plane of a fabric can be
performed in one of two ways: (i) the liquid is forced through at a constant velocity
and the resistance to passage is measured in some manner, or (ii) the liquid is forced
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

by a controlled pressure gradient and the resultant flow rate is monitored. For the
second case there are also two alternatives: either the driving pressure remains constant
or else it is systematically increased with time. The latter arrangement is convenient
for establishing the critical breakthrough pressure of a fabric barrier [33]. The
measurement of volume flow rate of a liquid through a fabric under a given pressure
is difficult to perform for fabrics which are very thin or highly wettable or very thick.
In the first two cases, significant pressure would not be encountered to be measured
reliably. In the last case, build-up of a significant opposing hydrostatic head defeats
the constant-pressure requirement [86].

2.4.3.1 Forced Flow at Constant Velocity


Miller and Clark [86] developed an experimental method wherein a liquid is brought
into contact with a fabric and passed through it at a constant flow rate while measuring
the resulting changes in pressure. A similar experimental method was also reported
by Miller [33] in another study. The initial wetting response as the liquid contacts the
fabric, and the steady-state wet-through resistance as the liquid breaks through the
fabric can be estimated. A piston moving at a constant velocity inside a cylinder
pushes a column of liquid into and through the fabric sample, while the pressure at
a fixed position in the liquid is measured and recorded by a sensitive transducer (Fig.
14) [33]. The inside diameters of the cylinder and the sample holder are the same, so
that the volume rate of movement of the liquid column remains constant during the
experiment. The pressure transducer is rigidly linked to the piston so that the hydrostatic
head remain constant; the pressure reading, therefore, does not change until some
interaction occurs between liquid and fabric. Experiments are conducted at very low
velocity to avoid viscous drag effects.
The general pressure-response as a function of time is illustrated in Fig. 15. O is
the initial static pressure when the liquid surface is below the fabric. With the movement
of the piston, pressure increase slightly to O′, a dynamic pressure as a result of
cylinder-wall drag. When the liquid first contacts the fabric, the pressure decreases to
A′ if there is any spontaneous uptake of liquid. If there is no uptake the pressure
remains unchanged, but in no case is there any resistance to flow at the moment of
first contact. As the liquid front is forced through the fabric, the pressure can increase
(not always), reaching a maximum (A) at the moment when the liquid first appears on
the top surface of the fabric. After breaking through the fabric, the pressure drops
abruptly to a steady-state value (B). The final dynamic pressure reading (B) represents
the viscous drag exerted by the fabric on the moving liquid. This drag is quite small,

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 27

O-rings

Fabric

Liquid
Pressure
transducer
Piston
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

Constant
speed drive

Fig. 14 General concept of the TRI wet-through resistance monitor [33 Miller]

Pressure

O′
C
O

Ar
Time

Fig. 15 Generalized pressure–time profile obtained from the constant flow rate flow-through
experiment [33 Miller]: O = initial static pressure; O’ = pre-contact dynamic pressure

unless the liquid is very viscous. C represents the hydrostatic pressure when the
piston is stopped. The difference between A′ and O or O′ is the measure of the initial
wetting response of the fabric. The ‘wet-through resistance’ is measured by the
difference between A and B. The viscous drag is quantified by the difference between
B and C [33, 86].

2.4.3.2 Controlled Pressure Wet-through


An arrangement [33] for controlled pressure wet-through studies on fabrics is shown
schematically in Fig. 16. The driving pressure is produced by the movement of the
piston of a syringe pump, which causes compression of the air behind the wetting
© The Textile Institute
28 Textile Progress doi:10.1533/tepr.2006.0001

Burette Manometer
Sealing
Support ring
screen Fabric

Contact
chamber Ph

A
B C Glass capillary
tube
Pressure
reservoir
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

Syringe pump

Fig. 16 Instrumentation for controlled pressure wet-through experiments [33 Miller]

liquid, forcing it up through the underside of the fabric. The large pressure ballast
chamber maintains the driving air pressure effectively, independent of the movement
of the liquid. The amount of liquid used is sufficient that the trailing meniscus
remains in the horizontal capillary when the leading edge of liquid encounters the
fabric. To perform a constant pressure wet-through experiment, the liquid is brought
up to the fabric at a very small driving pressure. If the fabric has any significant
resistance, the movement of liquid in the capillary stops at first contact. Stopcock C
is then closed and the pressure is built up to the desired level. The experiment is
started by opening C, and the movement of the meniscus through the capillary is
monitored as a function of time. The sensitivity of the measurement depends on the
bore of the capillary. When breakthrough occurs, the meniscus abruptly moves more
rapidly and thus identifies that event. In this way, volume uptake before breakthrough
and breakthrough time under a specific pressure gradient can be determined in a
single experiment. The effective gradient is determined by correcting the manometer
reading for the hydrostatic head (Ph) between the sample and the capillary tube.
Alternatively, the syringe pump is allowed to compress the air continuously to
systematically build up the driving pressure. In this way, a wide range of linear rates
of increase can be produced. Under such conditions, the trailing meniscus enters the
capillary and moves at a fairly rapid rate until the resistance of the fabric is encountered,
at which point the meniscus slows down (or stops) and then moves slowly until
breakthrough occurs. By following the movement in the capillary and noting the
pressure reading when the speed-up at breakthrough occurs, one can conveniently
find the critical breakthrough pressure. This pressure depends on the rate of pressure
increase, and such dependence can be ascertained by performing the experiment with
various rates of piston movement [33].
Both constant flow and controlled pressure wet-through methods have their strong
points and limitations. The former can be used to observe both negative and positive
resistances during wetting-through, but does not tell us much about the effect of
pores other than the largest. The latter can be performed under modest driving pressures
© The Textile Institute
doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 29

so that the behavior of the system between the time of first contact and breakthrough
can be observed. However, for conducting a controlled-pressure wet-through experiment,
the fabric needs to have significant resistance for the liquid to pass through, otherwise
there will be no way to identify the contact or breakthrough [33].

2.4.3.3 Sessile Drop Method


Contact angles on surgical fabrics can be measured by the sessile drop method [81]
using the drop shape analysis system G10/DSA10, produced by Krüss GmbH [87].
The basic instrument of the system-contact angle meter G10 is presented in Fig. 17.
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

Video-microscope
Syringe

Light
source

Sample stage

Fig. 17 Contact angle meter G10 [87 www.Kruss.de]

The sample stage of the instrument can be precisely adjusted along all three axes
using precision drives. Drop image contrast is optimized using the variable intensity
illumination. A zoom objective of the microscope allows the drop image to be shown
at the maximum size possible. The optics can be tilted by up to 2° to achieve the best
angle of view. This is especially important for such rough and uneven surfaces as
textile fabrics.
The instrument is equipped with a micrometer-driven syringe and a mechanical
stage [87]. A dosing system uses a syringe controlled by a stepper motor to produce
drops at a controlled flow rate. This ensures that the liquid drop is dosed exactly and
positioned reliably for measurement. Windows software is used to process images
acquired by the video camera and digitizing card of the DSA10 system. Software uses
a special algorithm to define the drop boundary reliably. Images of sessile drops are
fitted to yield contact angle data. Contact angles are measured at the tangent lines to
the surface and the mean value θ = (θ1 + θ2) is calculated (Fig. 18).
Both distilled water and salt solution (NaCl 0.9%) are used as test liquids. The
diameter of the drops is kept small (2–5 mm) so that the contact angle is not influenced
by the drop size. Contact angles are measured a few seconds after the drop is put
down from the syringe [81].
Wetting by water and surfactant solutions was studied by Herlinger et al. [88] on
different fabrics as a function of time. Since wetting with water proceeds relatively
quickly on hydrophilic, well-prepared fibers, two rapid methods of measurement
were developed. The first method was to follow the physical contact angle between

© The Textile Institute


30 Textile Progress doi:10.1533/tepr.2006.0001

θ1 θ2

Fig. 18 Tangent lines at the borders of the drop [87 www.Kruss.de]

a drop of liquid and a flat textile material as a function of time. In the second method,
the increase in weight of a sample placed horizontally, in contact with a liquid
surface, was determined as a function of time. The fabrics tested were cotton, nylon,
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

polyester, acrylic, and cotton/polynosic [88].

3. WICKING
The transport of a liquid into a fibrous assembly, such as a yarn or fabric, may be
caused by external forces or by capillary forces only. An understanding of capillarity
is important in wet processing of textiles, especially dyeing. The transport of liquid
has been a subject of study for fabrics of many kinds but usually in connection with
the water repellency of fabrics [89]. In most of the wet processing of fibrous materials,
uniform spreading and penetration of liquids into pores are essential for the better
performance of resulting products [90]. Wicking can only occur when a liquid wets
fibers assembled with capillary spaces between them. The resulting capillary forces
drive the liquid into the capillary spaces. Wicking can be visualized as a spontaneous
displacement of a solid–air interface with a solid–liquid interface in a capillary
system. In a simple case, such as wicking in a capillary tube, the area of the liquid–
air interface across the capillary is very small relative to the area of the wetted
capillary wall, and does not change during wicking. Hence, the only considerable
change is the increase of the solid–liquid interface and decrease of the solid–air
interface. For the process to be spontaneous, free energy has to be gained and the
work of penetration has to be positive [3]. This is the case when the interfacial energy
of the fiber surface in contact with vapor (air) γSV exceeds the interfacial energy
between the liquid and the fiber surface γSL:
WP = γSV – γSL (2)
The work of penetration WP is a measure of the energy required for capillary penetration.
Since γSV and γSL are exceedingly difficult to measure independently, workers have
attempted to estimate the surface energy indirectly by interaction with liquids. The
Young-Dupré equation (1) [4, 6–8] permits expression of the differences between γSV
and γSL with the measurable quantities γLV and cos θ
Relating capillary penetration to capillary pressure can draw a similar conclusion.
When a liquid in a capillary wets the walls of the capillary, a meniscus is formed. The
surface tension of the liquid causing a pressure differences ∆P across the curved
liquid–air (vapor) interface, related to the curvature of this interface according to the
Young-Laplace equation, is [21]:
∆P = γLV (1/R1 + 1/R2) (14)
For a capillary with a circular cross-section, the radii of the curved interfaces R1 and
R2 are equal:
© The Textile Institute
doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 31

∆P = 2γLV/R (15)
If the capillary is circular with radius r, the meniscus will be approximately hemispherical
with a constant radius of curvature,
R = r/cos θ (16)
The capillary pressure is therefore
∆P = 2γLV cos θ/r (17)
When the capillary wall is completely wettable by the liquid, then cos θ = 1. For a
positive capillary pressure, the values of θ have to be between 0° and 90°. Capillary
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

pressure is inversely related to the capillary radius.


The term ‘wicking’ has been used in a broader practical sense to describe two
kinetically different processes: a spontaneous flow of a liquid within the capillary
spaces accompanied by a simultaneous diffusion of the liquid into the interior of the
fibers, or a film on the fibers [91]. If the penetration of liquid is limited to the
capillary spaces and the fibers do not imbibe the liquid, the wicking process is termed
‘capillary penetration’ or ‘capillary sorption’. The sorption of the liquid into fibers
can cause swelling [92–95], reduce capillary spaces between fibers, and complicate
kinetics [96–97]. The interpretation of wetting results can be misleading if the effects
of sorption in the fibers or finishes on fibers are overlooked [98–99].
According to Schwartz [100], wicking is one example of the more general set of
phenomena termed ‘capillarity’. Capillarity encompasses all the dynamic and kinematic
effects and behavior of phase interfaces. The interfaces are regarded as geometrically
sharp when solving practical wicking problems. However, at the molecular level, the
boundary is in the range of a few molecular diameters. For wicking to be significant,
the interfacial areas in the system should be large in comparison with the bulk
volumes. In porous systems, the area of the pore walls must be large as compared
with the volume of the solid. In particulate systems, the particles must be small in at
least one dimension, and must be closely spaced. This latter condition ensures that
the ratio of solid–liquid (SL) interfacial area to liquid volume will be large. These
requirements follow from the fact that interfacial free energy, like other forms of
energy, is an extensive quantity [100].
The capillary spaces in yarns and fabrics are not uniform, and an indirectly determined
effective capillary radius has to be used instead of the radius r [3]. Further, fibrous
materials encounter roughness on the surfaces and walls of the pores. A liquid may
spread along grooves or rugosities on a surface, even if it does not spread on a smooth
surface of the same solid. The driving force for such surface wicking depends on the
geometry of the grooves, surface tension of the liquid, and free energies of the solid–
gas and solid–liquid interfaces [101]. Schwartz [102] has proposed a general
thermodynamic approach to all wicking and spreading phenomena for cases when
the spreading coefficient, S ≤ 0. The driving force for the movement of a liquid
interline normal to itself, a distance ds along a solid surface, is given by:

dF = γ  d ( A ) – cos θ a d ( A )  (18)
LV 
ds  ds
LV
ds SL 

© The Textile Institute


32 Textile Progress doi:10.1533/tepr.2006.0001

where F is the total free energy of the system, and ALV and ASL are the liquid–vapor
and solid–liquid interfaces. If dF/ds is negative, the liquid will advance over the
unwetted solid surface. Equation (18) provides the basis, for a given wicking geometry,
of sorting out the relative importance of the surface tension and the contact angle to
the wicking process. For a liquid thread advancing from a circular pore where the
radius is large into a region where it is small, equation (18) may be written to account
for the difference in contact angle between that at the advancing interline and that at
the receding interline:

dF = γ  d ALV – γ cos θ dASL – γ cos θ dASL 


A R
LV  (19)
d s 
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

LV a LV r
ds  ds ds
A R
dASL dASL
At the advancing meniscus, = 2πrrA, and for receding meniscus, =
ds ds
2πrrR, where rA and rR are the capillary radii in advancing and receding modes,
respectively. Substitution into the equation (19), noting that dALV/ds = 0 leads to:
dF = 2πγ (r cos θ – r cos θ ) (20)
LV R r A a
ds
The net capillary pressure propelling the thread forward is obtained by dividing each
term on the right-hand side of equation (20) by the appropriate cross-sectional area
of the capillary, i.e. π rA2 , respectively [102].
Equation (18) can be used to compute the conditions for a simple type of surface
wicking, as shown in Fig. 19, where a liquid from a large reservoir is moving in a
straight-edged, horizontal trough at a depth R0 measured along each side [101]. The
angle of the trough bottom is φ (in radians). When R0 and φ are sufficiently small, the
liquid surface may be regarded as a portion of a right cylinder. Under such conditions,
advance of the liquid along the channel gives d ALV/ds = R0 φ and d ASL/ds = 2R0.
Equation (18) then gives:
dF = γ R (φ – 2 cos θ ) (21)
LV 0 a
ds
Surface wicking will thus occur in this case only when 2 cos θA > φ, and once again
the process is favored by low contact angle and high surface tension. Similar
computations based on equation (18) may be used to obtain surface wicking conditions
for other geometries, such as those pertaining to the grooves between parallel cylindrical
textile fibers, etc [101].

φ R0

Fig. 19 Surface wicking in an angular groove [101Berg]

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 33

For a theoretical treatment of capillary flow in fabrics, the fibrous assemblies are
usually considered to consist of a number of parallel capillaries. The advancement of
the liquid front in a capillary can be visualized as occurring in small jumps. The
advancing wetting line in a single capillary stretches the meniscus of the liquid until
the elasticity of the meniscus and the inertia of flow are exceeded. The meniscus
contracts, pulling more liquid into the capillary to restore the equilibrium state of the
meniscus. The movement of the liquid in a non-homogeneous capillary system, such
as a fibrous assembly, is discontinuous for another reason as well. The wetting front
advances into the capillary system in small jumps because the irregular capillary
spaces have various dimensions [3].
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

Most textile processes are time limited, and the rate of wicking is therefore important.
However, the wicking rate is not solely governed by interfacial tensions and the
wettability of the fibers, but by other factors as well. The wicking rate depends on the
capillary dimensions of the substrate and the viscosity of the liquid [103, 104].
The mass rate (M) at which a liquid moves through a porous channel is related to
the pressure difference (p or ∆P) across the channel, in the absence of gravitational
force and neglecting inertial forces (as acceleration is small), by Poiseuille’s law [33,
105].
M = πp ρLlr4/8ηh
If the pressure differences ∆P is due to capillary forces, then
M = π ρL r3 γLV cos θ/4η h
The volume rate (V) is
V = π r3 γLV cos θ/4η h
Linear rate of flow (u) is
u = dh/dt = rγLV cos θ/4η h (22)
where h is the height of liquid rise in the capillary channel.
The Lucas–Washburn [16, 17] equation for the rate of fluid flow against gravitational
forces, through a porous solid as a bundle of round capillary tubes, each being very
small with radius r, is given by [106, 107]:
γ LV r cos θ
dh /dt = – r 2 ρL g /8η (23)
4ηh
where g is the acceleration due to gravity. The second term is negligible when either
the flow is horizontal or r is very small (r2 = 0).
When the capillary forces are balanced by the gravitational forces, liquid rise stops
and equilibrium is reached, as given by
γLV cos θ 2πr = πr2ρLgh
Hence, equilibrium wicking height heq = 2γLV cos θ /rρL g
According to the Lucas-Washburn equation [16, 17] neglecting gravitational forces,
the wicking height, h, is directly proportional to the square root of time, t [108]:
h = (rγLV cos θ/2τ2η)1/2t1/2 = kt1/2 (24)
where hτ is the actual distance traveled, τ being the tortuosity factor.
© The Textile Institute
34 Textile Progress doi:10.1533/tepr.2006.0001

3.1 Wicking in Yarns


There are some theories regarding wicking of yarns [89, 109]. The surface area of a
liquid–air interface is unique and may always be identified with its plane geometrical
area, but the surface of a solid–liquid interface depends very much on the condition
of the solid surface. Wenzel [85] proposed that if a solid surface is roughened so that
unit plane geometrical area has an actual surface area R times that of the smooth
surface, the energy gained in forming the solid–liquid interface will be R(γSV – γSL)
and the apparent contact angle θ of the roughened surface will be given by:
R(γ SV – γ SL )
cos θ ′ = = R cos θ (25)
γ LV
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

Alternatively, Bartell and Shepherd [110] have proposed that where the microscopic
cavities of the rough surface make an angle φ with the general direction of the solid
surface, the apparent advancing and receding contact angles are:
θ a′ = θ + φ (26)
θ r′ = θ – φ (27)
It is also permissible to include in θ a′ effects due to sorbed material on the fiber
surface.
The contact angle varies, however, with the tendency of the liquid to move along
the surface, being at a maximum when the liquid is advancing over a dry surface.
Then for a meniscus moving slowly into a rough capillary, the driving capillary
pressure is:
2γ LV cos θ a′ (28)
p=
r
For a liquid penetrating a horizontal capillary, no gravitational force is involved. In
addition, because the acceleration is small, inertia forces may be neglected and the
surface and viscous forces at the surface simply equated [111]. Accordingly, equation
(22) is transformed to:
γ LV cos θ a′ r
u= (29)
4η l
Considering that fibers in textile assemblies form capillaries of effective radius ref,
then the horizontal water transport rate becomes:
γ LV cos θ a′ re t
h2 = = ks t (30)

where ks is the capillary liquid transport constant for the penetration of a given liquid
into a definite assembly of fibers [89]. Equation (30) is similar to equation (24). In a
twisted yarn, fiber follow a helical path due to twist. The corrected liquid transport
rate constant, Kc, is related to twist T as [89]:
Kc = Ks (1 + 0.87D.T)2 (31)
where D is yarn diameter and T is the twist per inch.

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 35

The factors that determine k can be separated by considering the well-known


Lucas-Washburn equation [16, 17] for the rate of flow in a cylindrical capillary tube,
from which equation (30) is deduced [112].
Since a yarn is not geometrically equivalent to a capillary tube, a suitably derived
equivalent radius of the open cross-sectional area with in the yarn must be substituted
for r. It may be assumed that yarns are composed of oriented cylindrical fibers and
that r will be proportional to the square root of the open correctional area. This
assumption is compatible with the concept of the hydraulic depth widely used in
considering flow through porous media. It is also noteworthy that equation (30)
applies only to a system where the free surface of the liquid reservoir feeding the
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

capillary tube is substantially flat, i.e. the capillary pressure on the reservoir surface
is zero [112].
Lord [109] discussed a theory for yarn wicking that can also be applied to fabrics.
He concluded that wicking height is likely to be a function of twist multiple, type of
fiber, type of yarn structure, and packing, as well as fiber migration.
Wicking rate in a vertically hung yarn or fabric is practically not constant and
in fact decreases because of the increasing mass of the ascending liquid. Pan et al.
[113] transformed the Lucas-Washburn [16, 17] equation for ascension of a liquid
wicking in a hung fibrous sample dipped into the liquid reservoir and extended it into
the case of radial expansion of wicking liquid originating at the centre of a flat
sample as:
dmA/dt = (K/mA) – L
where mA is the liquid mass uptake, K is a constant, and L is the distance traveled by
the liquid.

3.1.1 Measurement of Wicking Behavior in Yarns


Several methods are available for the measurement of wicking in yarns and they are
described below.

3.1.1.1 Microscopic Observation of Vertical Wicking Height


[89, 109, 112, 114–118]
Vertically hung yarn (usually under tension to make the yarn straight) is partially
immersed in test liquid or water. In some cases, many yarns are mounted side-by-side
over frames, and the bottom frame is dipped into the test liquid such that a few
lengths of yarns are submerged in the liquid [112]. A suitable dye solution of very
low concentration is added to facilitate easy observation of the liquid front in the
yarn. The time required for the liquid to rise through for fixed intervals of distance
on the yarn, or vice-versa, is noted. Water transport curves (i.e. plots of the square of
the distance of water travel against time) for cotton, viscose, wool, and wool–Dacron
spun yarns were reported by Hollies et al. [89]. The linearity of the curves gave a
good indication that equation (30) applies to water flow in these yarns. Water transport
rate constants were determined from the slopes of the plots. They conducted experiments
on vertical wicking of nylon filament yarns with different numbers of filaments and
twist levels. Yarn diameters were measured before and after wetting. From the average
wet diameter, mean capillary radii were calculated. The corrected water transport

© The Textile Institute


36 Textile Progress doi:10.1533/tepr.2006.0001

rates for all the nylon yarns were indeed almost constant, which supported the form
of the capillary theory given in equation (30).
Sengupta and Murthy [117] described a method for measuring the vertical wicking
height. A test length of 30 cm of the yarn was randomly taken. One end of the yarn
was tied onto a gem clip, and the other end, to a hook of a small weight (3g). The yarn
was hung on the notch of a nail fixed to a wooden plank held by an iron stand. Yarn
was slowly lowered into the measuring jar containing 3% Procion Blue HGR (a
reactive dye) till the weight just touched the dye solution. The yarn was further
lowered down into the measuring jar in such a way that the lower end, of about 5 cm,
was below the liquid surface. Simultaneously, a stopwatch was pressed and the rise
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

of dye solution from the fluid surface in the yarn was continuously watched through
the eyepiece of a cathetometer microscope. The time required to cross every 1 cm
mark was noted. Fifty tests were carried out for ring and open end yarns, spun with
varying twist factors. Further, the vertical wicking height after 12 hours was also
noted for all the yarn samples used; 25 readings were taken for each sample.
Subramaniam et al. [118] used the above test procedure for measuring vertical wicking
height in Siro spun yarns.
Ring and compact yarns were wound at constant tension around a frame on which
two raised platforms were mounted (Fig. 20). The ends on the platform were fixed
with Sellotape and the remaining portions of the yarns were cut and removed. The
raised portions ensured that there existed a gap between the yarns and the frame so
that when the frame was dipped into water, the movement of the water through the
yarns was not hindered. A scale was attached next to the yarns on the frame. Four
yarns were mounted at a time [116]. A beaker containing distilled water with 1% dye
solution (Reactive M8B) was prepared. The frame was slowly lowered into the solution
so that 2 cm lengths of yarn remained submerged in the solution. Simultaneously, a
stop watch was pressed and the rise of dye solution through the yarn was continuously
monitored by means of a cathetometer. The distance traveled by the dye solution was
noted as a function of time until it became virtually constant. For each yarn, 12
samples were studied. The average wicking height of the 12 yarn samples at regular
intervals of time was determined [116].
Sengupta et al. [114] studied the equilibrium and dynamic wicking behavior of
air-jet textured yarns. Methylene blue solution, with a concentration of 1%, was used

Raised platform
Clamp
Frame

Attached Yarns
scale

Beaker

Fig. 20 Schematic representation of wicking test [116 Chattopadhyay]

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 37

for the above study. Air-jet textured yarn with a tension of 0.017 cN/dtex was hung
vertically from a clamp into the bath containing the prepared dye solution. The
tension in the yarn was maintained by attaching the required weight to the bottom
end of the yarn by means of a hook. A scale was attached very near to the yarn to
measure the equilibrium wicking height. This height was measured after 12 hours
from the time of dipping the yarn into the dye bath. In the case of wicking rate, the
time required for the dye solution to travel upwards along the yarn for every 1cm
length of yarn, up to a maximum of 5 cm, was noted with the help of a stopwatch and
a traveling microscope.
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

3.1.1.2 Measurement of Wicking Rate in a Predefined Yarn Bundle Using


Tensiometry
The wetting of powders and porous solids (yarns) is complicated by the presence of
porous architecture. According to the Washburn equation [17], when a porous solid
is brought into contact with a liquid, the rise of the liquid into the pores of the solid
is given by [119]:
T = (η / CρL2 γ cos θ ) Μ 2
where T is the time after contact, C is a constant characteristic of the solid sample,
and M is the mass of liquid absorbed on the solid.
Using tensiometry, M is measured and a straight-line plot of M 2 versus T gives the
slope (η/C ρL2 γ cos θ). From the slope obtained using a low surface tension liquid
(n-hexane, where cos θ is 1) as the wetting liquid, the material constant C is estimated.
Then using the test liquid, the experiment is performed. From the known values of T,
M, η, C, ρL, and γ, the values of cos θ and the contact angle are computed. Dynamic
methods using the above principle for measuring the wettability of a bundle of
multifilament yarns based on the rate of wicking are reported by Chwastiak [103],

and as a powder coating method by Perssin et al. [120]. In one method, 40–50 loops
of carbon filament yarns (Thornel) are pulled parallel through a Teflon tube so that
the porosity of the fiber bundle inside the tube is kept at 0.5 [103]. This is just to
define the geometry of the fibers in the yarn bundle for the wettability measurement.
The excess yarn is cut away flush with one end of the tube and trimmed to leave about
2 mm of yarn protruding from the other end. The yarn loading procedure is illustrated
in Fig. 21. The loaded tube is hung vertically from one arm of a Cahn RH electrobalance,
the end with the protruding yarn pointing downward. A pan containing the wicking
liquid is raised under the tube until the yarn ends are immersed in the liquid. The
liquid wicks up into the yarn bundle and the weight gain of the tube is recorded as a
function of time. The measured weights must be corrected for the buoyancy and
surface tension acting on the end of the yarn that is immersed in the liquid. The
magnitude of this correction is found at the end of the test when the weight of the
tube is constant because all the pores in the yarn bundle are full of liquid. The
difference in weight between the tube with the yarn ends immersed in the liquid and
the tube with the yarn bundle ends withdrawn out of the liquid into air is the amount
of the correction for buoyancy and surface tension. This correction is then applied to
all weights from the start to the end of the test. The test is usually continued for at
least 10 minutes after the weight gain has stopped to ensure that constant weight has

© The Textile Institute


38 Textile Progress doi:10.1533/tepr.2006.0001

String

Tube
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

Flare

40 to 50 strands of carbon yarn

Fig. 21 Loading yarn into a tube [103 Chwastiak]

been achieved. When the yarn does not wet, i.e. the contact angle is equal to or
greater than 90°, no liquid at all is wicked by the yarn bundle. The wettability of
carbon yarn is determined in various liquids from wicking rates and contact angles.
It is reported that the precision of the method is ± 7%. The accuracy depends on the
accuracy of the constant, C, characterizing the pore spaces in the yarn bundle. It is
claimed that this method has the advantages of being fast, independent of operator
judgment, and only requires a sample size of about 1 g [103].
The above method is highly useful for studying the wettability of reinforcing
filaments with resins for fiber-composite applications. In composites, the adhesion
between the reinforcing fiber and the matrix influences the mechanical properties of
the material. Measurements of surface energies can predict the compatibility of these
materials. Often, the interface can be engineered by modifying the fiber and matrix.
As the nature of physical interaction between liquids and solids is important for
practical interest, the measurement of contact angles by this method and the calculation
of surface energies can lead to a better understanding and even an improvement of
these interactions by modification [46].

3.1.1.3 Other Methods


Ansari and Kish [121] developed a method based on the electrical resistance of yarns
and a liquid having reasonable conductivity. A yarn sample, held vertically by a small
weight, is in contact with a movable contacting sensor (A) in the measuring unit (Fig.
22). The liquid container (B), with a submerged stripped wire, can be raised to dip the
yarn end in the liquid. When liquid wicking through the yarn reaches the movable
contacting sensor, the yarn resistance decreases and an electrical current passes through
the liquid, the wetted section of the sample, and the contacting sensor. A sudden
decrease of yarn resistance to a certain predefined value, defined in the computer
software, causes an electrical current to actuate the stepper motor and to rotate the
spindle (C). Stepwise rotation of the spindle wraps the connection thread and pulls
the sensor to the dry section of the test yarn. For each step of the spindle rotation, the

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 39

Step meter

C
Holder

PC
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

A IO
dig
ita
l
Yarn sample AD

B
1.1
A
V

Fig. 22 Diagram of the apparatus used for determination of the time required for vertical
wicking of water in given length of yarn [121 Ansari]

computer, using suitable software, records the height of the sensor and the time. A
plot of time versus height is constructed. This method works with yarns having high
electrical resistance and liquids of reasonable conductivity.
Kamath et al. [115] developed an electronic method to determine the time required
for spin finishes to wick horizontally into a certain length of yarn. Since the yarns
studied (polyester and polypropylene) had high electrical impedance, this method
will work only with finish liquids of reasonable conductivity. With non-conducting
organic liquids, wicking can be measured manually by adding a fluorescent dye to
the liquid and using light of appropriate wavelength to identify the liquid front.
Chen et al. [122] developed a PC-based imaging system to measure the wicking
behavior of yarns. The set-up allows recording the droplet shape changes and thus the
kinetics of droplet absorption. The input parameters of this method are limited to the
initial droplet volume and the time for the droplet to disappear.
Perwuelz et al. [123] studied the capillary flow in polyester and polyamide yarns
and glass fibers using a technique based on the analysis of CCD images taken during
the capillary rise of a colored liquid in yarns.
An experimental apparatus has been developed for studying water transport behavior
along a bundle consisting of a few textile fibers measured by an electrical capacitance
technique [124, 125]. As shown in Fig. 23, sample fibers are placed between two
metal plates (electrodes); this assembly can be regarded as a condenser (Fig. 23a).

© The Textile Institute


40 Textile Progress doi:10.1533/tepr.2006.0001

(a) (b) (c)

Fig. 23 A model for water transport through fiber assemblies placed between two electrodes
[125 Tagaya]

When the end of a fiber bundle comes into contact with liquid water, the water starts
to travel along the fiber bundles by displacing air (Fig. 23b). The inner part of the
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

condenser consists of fibers, air, and water. The capacitance of the condenser is
changed by introducing water into the condenser because the dielectric constant of
water is about eighty times greater than that of air. Therefore, water transport can be
sensitively detected by measuring the change in electrical capacitance of the condenser.
Ito and Muraoka [124] measured electrical capacitance on an apparatus similar to
that constructed previously by Tagaya et al. [125], but much more sensitive. The
apparatus of Tagaya et al. [125] was developed to measure a relatively large amount
of water transported through fabric samples. In the outline of the circuit reported by
Tagaya et al. [125], the electric current generated by the change in electric capacitance,
which is caused by the transport of water, is converted into a voltage (Vo) by a
current-to-voltage (I-V) converter circuit:
Vo = Rf Cx Vp–p f
where Rf is the feed-back resistance of the I-V converter, Vp–p is the output potential
of the wave generator, f is the frequency, and Cx is the electric capacitance, which is
influenced by water transport. The drawback of this method is that the sensitivity of
the apparatus is not as high, due to the low signal-to-noise (S/N) ratio, coming from
the fact that one of the electrodes is connected not to the true ground but to the virtual
ground.
The principle of this method is that the change in electric capacitance is detected
as the difference in pulse widths, instead of electric current, by using two kinds of
monostaple vibrators [124].
Due to yarn heterogeneity, wicking measurements have to be undertaken on a
large number of samples in order to have a good statistical representation [123].
Sample numbers are generally 30–60 [114].

3.2 Wicking in Fabrics


High and uniform absorbency of cotton fabric is a desirable quality in nearly every
wet finishing operation and many finished fabrics. Absorbency of fabrics is influenced
by their wicking ability. Wicking occurs when a fabric is completely or partially
immersed in a liquid or in contact with a limited amount of liquid, such as a drop
placed on the fabric. Capillary penetration of a liquid can therefore occur from an
infinite (unlimited) or limited (finite) reservoir. Wicking processes from an infinite
reservoir are immersion, transplanar wicking, and longitudinal wicking. Wicking
from a limited reservoir is exemplified by a drop placed onto the fabric surface
[3].

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 41

Based on the extent of interactions with fibers, wicking processes can be divided
to four categories [3].
(i) Wicking of a liquid, no significant diffusion into the fiber surface, e.g. hydrocarbon
oil wicking into a polyester fabric at ambient temperature; capillary penetration
is the only process operating;
(ii) Wicking accompanied by diffusion of the liquid into fibers or into a finish on the
fiber, e.g. water wicking in a cotton fabric and diffusing into fibers, water
wicking into a soil-release treated polyester fabric, and diffusing into the finish.
Two simultaneous processes are operating – capillary penetration and diffusion
of the liquid into the fibers;
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

(iii) Wicking accompanied by adsorption on fibers, e.g. an aqueous surfactant solution


wicking into a polyester fabric. Several processes are operating simultaneously
– capillary penetration of the liquid, diffusion of the surfactant in the liquid, and
adsorption of the surfactant on fibers; and
(iv) Wicking involving adsorption and diffusion into fibers, e.g. an aqueous surfactant
solution wicking into a cotton fabric. Several processes are operating
simultaneously – capillary penetration, diffusion of the liquid into the fibers,
diffusion of the surfactant in the liquid, and adsorption of the surfactant on
fibers.

3.2.1 Wicking From an Infinite Reservoir


Capillary penetration of a liquid from an infinite reservoir has been the subject of
numerous studies [1, 96, 98, 99, 126–131]. The wicking processes and methods of
measurement of wicking from an infinite reservoir are outlined below:

3.2.1.1 Immersion
Trounson et al. [132] reported that wicking during immersion occurs when the fabric
(or fibrous assembly in general) is completely immersed in a liquid and the liquid
enters the fabric from all directions. A liquid wicking into an immersed yarn or fabric
displaces most of the air in the fibrous assembly and causes it to sink.
The canvas disk-wetting test measures the sinking time of a disk submerged in a
liquid. A canvas disk, 1 inch in diameter, is placed into a Gooch funnel. The funnel
is inverted in the solution in order to immerse the canvas disk about 0.5 inches below
the surface of the liquid. The time required for the disk to begin to sink is recorded.
However, the introduction of the disk into the liquid inadvertently causes hydrodynamic
forces that may affect the measure of sinking time. The test is no longer used.
The Draves test measures the wetting time of an unboiled, naturally waxed cotton
skein submerged in a surfactant solution. The skein is attached to a hook that is tied
to an anchor and placed into a graduated cylinder containing the surfactant solution
[133–135]. The sinking time of the skein is related to the advancing contact angle and
the dynamic surface tension of the wetting front [134–136]. When the adsorption of
the surfactant is strong, the surfactant in the advancing wetting front may be depleted.
In this case, the wetting time depends on surfactant diffusion into the wetting front.
A compact surfactant structure with a centrally-located hydrophilic and branched
hydrophobe facilitates diffusion, but extensive hydration of the surfactant decreases
the diffusion rate and increases the wetting time. The Draves test is routinely used to
© The Textile Institute
42 Textile Progress doi:10.1533/tepr.2006.0001

evaluate wetting agents [134-135]. McLaughlin et al. [137] have stated that immersion
sorption is so common in textile processing that its complexity is usually overlooked.
In sink–float tests, the force of the impact when the fabric specimen is simply
dropped onto the liquid can lead to erratic results, as in the ‘drop-square’ test [138].
A mechanical device has been designed to minimize the impact of the fabric–liquid
contact. A fabric disk, resting on a platinum wire ring, is lowered gently onto the
surface of the liquid, and the sinking time is recorded [96, 97]. The hydrophilicity
and wettability of fabrics treated with various finishes are determined by comparing
the sinking time in a hydrocarbon solvent (decane) and in a surfactant solution [96].
When the finish on a fabric is the only variable, the test correctly predicts soil release
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

performance and repellency. The sinking of a disk impregnated with a hydrocarbon


oil, e.g. decane or hexadecane, in an aqueous detergent simulates intrinsic soil release
under static conditions [10].

3.2.1.2 Longitudinal or In-plane Wicking


It is common practice to use in-plane wicking measurements to evaluate the absorbing
power or liquid transport capabilities of fibrous sheet materials. Most versions of the
test methods used for this purpose start out by dipping one end of a sheet into a liquid
and monitoring its subsequent upward movement into the sheet, either by following
the position of the liquid front (rate of liquid travel) or by gravimetric or volumetric
changes. If the upward distance traveled by the liquid becomes long enough, there
will be a noticeable effect of gravity on the flow rate [139]. Equilibrium will happen
when the capillary action is balanced by gravity, i.e. by the weight of the raised liquid
[3, 140]. Moreover, water uptake in a vertically hung cotton fabric shows a gradient
distribution pattern, with a saturated zone near the water-fabric interface followed by
diffused and steadily distributed zones [128]. During vertical upward wicking, the
flow of liquid is unsteady due to gravity effects. At the onset of absorption in a
vertical capillary system, the absorbed liquid is relatively close to the liquid source,
and the effect of gravity can be neglected in this situation. However, at a longer
period of time (or upward wicking distance), gravity plays an increasingly important
role.
Hsieh and Yu [1] have described a gravimetric measurement for wetting and transport
in woven fabrics. Hsieh [82] discussed capillary theory for this gravimetric method.
Initially, when the liquid contacts the fabric, the capillary pressure is greater than the
weight of the liquid in the fabric and the liquid continues to rise. The liquid stops
rising at the equilibrium height (heq), when capillary forces are balanced by the
weight of the liquid. Upon reaching equilibrium, the force detected by the balance
reading (∆Beq) becomes constant, indicating stabilization of the wetting and wicking
processes. The balance reading after separating the fabric from the liquid (∆Bsp)
indicates the amount of liquid retained by the fabric (Wt). The fabric wetting force
(Fw) is decoupled as:
Fw = (∆Beq – Wt)g
The fabric wetting contact angle (θ) is calculated as:
θ = cos–1 (Fw/Pγ) (32)

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 43

where P is the fabric–liquid perimeter, which can be estimated from the force measured
in hexadecane as:
P = Fw /γhexadecane
The water contact angles on materials, namely acetate, nylon, and polyester in
film form [141], the calculated contact angles on fabrics using equation (32), and
contact angles measured using tensiometry on constituent fibers [1], were similar,
though the films were obtained from a different source. According to Hsieh and Yu
[1], this shows that the wetting contact angles obtained from equation (32) provide
intrinsic wetting characteristics of fabrics.
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

If the liquid retention (Wt) for hexadecane and water are termed as Wth and Wtw,
then liquid retention capacity (Cv) and vertical water retention values (Cw) are calculated
by dividing Wth and Wtw values, respectively, by the fabric weight. The liquid retention
capacity (Cl ), calculated from the overall porosity of the fabric (φ), fiber density (ρf)
and liquid density (ρL) can be obtained as [82]:
Cl = (ρL /ρf)(φ)/(1 – φ)
At equilibrium, the sum of the external (Fwo) and the internal (Fwi) surface attraction
forces [82] are:
∆Beqg = Fwo + Fwi = Fw0 + Pπ ri2 = 2πγ (ro cos θo + ri cos θi) (33)
where P is the capillary pressure, ro and ri are the external and internal capillary radii,
and θo and θi are the external and internal liquid–solid contact angles, respectively.
The contribution of Fwo or Fwi to the force detection (∆Beq.g) varies depending on ri
and ro and the wetting properties of the internal and external surfaces. If these wetting
properties are the same, i.e. θi = θo, the contributions of Fwo and Fwi will be proportional
to the magnitudes of ro and ri.
Using hexadecane, the values of ri and ro can be calculated from the Fwi and Fwo
values determined in hexadecane.
For the test liquid (water), θi and θo can be calculated:

 Fwi 
θ i = cos –1   (34)
 2 π ri γ 
and

θ o = cos –1  wo 
F
(35)
 2 π ro γ 
The equilibrium height of the liquid rise (heq) is:
2γ cos θ i
heq = (36)
gρL ri
The calculated height of the liquid column (hcalc) using the measured Wt and ri values
is given by:
Wt
hcalc = (37)
π ri2 ρL
© The Textile Institute
44 Textile Progress doi:10.1533/tepr.2006.0001

In the simultaneous measurement of liquid wetting and transport in a vertically hung


fabric described above, the initial force detection represents the collective response
of both wetting and wicking phenomena [1]. On scoured cotton fabrics, the initial
force detection is very steep. Due to the liquid–fabric contact, the initial sections of
curves are similar for cotton fabrics of different lengths. As wicking becomes
predominant, profiles of the balance readings of these fabrics are further separated in
direct proportion to the fabric lengths [82].
A syphon test method has been reported by Lennox-Kerr [142] and Phukon [143].
A rectangular strip of the test fabric is used as a syphon, by immersing one end in a
reservoir of water or saline solution and allowing the liquid to drain from the other
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

end, placed at lower level, into a collecting beaker (Fig. 24). The amount of liquid
transferred at successive time intervals can be determined by weighing the collecting
beaker [5]. Lennox-Kerr [142] has taken the rate of mass transfer of liquid when a
constant flow through the syphon has been attained, as an indicator of wickability.

Reservoir

Test fabric

Collecting beaker

Fig. 24 Syphon test [143 Phukon]

Test methods based on monitoring the position of the liquid front in a vertically-
held fabric have been described [9, 129, 139, 140, 144–147]. A test method has been
used in which a strip of fabric is suspended vertically with its lower edge in a
reservoir of distilled water, as shown in Fig. 25 [9]. The rate of rise of the leading
edge of the water is monitored. To detect the position of the water line, a dye can be
added to the water or, in the case of dark colored fabrics, the conductivity of the water
may be used to complete an electrical circuit. The height of rise of water in a given
time is taken as a direct indication of the wickability of the test fabric. The simple
form of the test does not take into account the mass of the water that is taken up. This
depends on the height the water has risen to, the thickness of the fabric, and the
water-holding power of the fabric structure. One way of allowing for this is to weigh
the fabric at the end of the test, and hence obtain the mass of the water taken up by
the fabric. The mass can then be expressed as a percentage of the mass of the length
of dry fabric, which is equivalent to the measured height of the water rise.
The experimental set-up for the measurement of vertical wicking of fabrics is
described [145, 146]. Water-soluble ink is used to mark dots to indicate distance on
the fabric. To observe the movement of the liquid front, the time is recorded when the

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 45

Test fabric

Height of
water
Reservoir
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

Fig. 25 Wicking test [9 Saville]

marking dot begins to blur. The 1.5 cm × 7 cm strips were lowered to touch the water
reservoir, at which moment the timer was started. The liquid flows were monitored
until either a distance of 5 cm or a time of 30 minutes was reached [145]. In one
study, a dye and water were used to inspect the liquid front. However, the dye and
water became separated at the surface of fabrics treated with some cationic surfactants
[5].
Ravichandran et al. [147] described a method for measuring vertical wicking.
Samples were marked 1.27 cm from the top and bottom of the fabric strip and hung
vertically from a cross-bar into a dilute aqueous solution of Procion Red HE-3B.
Wicking distances corresponding to five minutes were measured for all samples
except those treated with anthranilic acid. Samples treated with anthranilic acid were
measured for wicking after three minutes because they wicked so rapidly and were
not long enough for a five-minute test.
Liquid penetration into a piece of polyester nonwoven fabric hung vertically above
the liquid was studied theoretically and experimentally by Kurematsu and Koishi
[129]. The phenomena of liquid ascent were analyzed using Washburn’s equation
[17], which rarely indicated the effect on the fiber bed structure such as porosity and
specific surface. Heights of liquids ascending in polyester nonwoven fabric can be
calculated with the improved Kozeny–Carman’s equation, for fluid flow through a
porous media, including a term for the capillary force between two parallel surfaces
of the nearest fibers in three fiber orientating models of polyester nonwoven fabric.
Theoretical ascending heights for carbon tetrachloride, ethyl alcohol, n-butyl alcohol,
n-decyl alcohol, toluene, cyclohexane, and diglycidylether bisphenol A (epoxy resin)
appear to be in fair agreement with the experimental heights determined by taking
into account liquid evaporation.
Miller et al. [140] presented theoretical and experimental cases of downward and
upward wicking. The equations describing these two processes have somewhat different
forms. The results indicate that downward wicking can produce an essentially constant
feed rate over a very long time period, whereas upward wicking rates decrease
continuously with time but never actually reache zero. According to them, if it is
necessary to bring about complete filling of all the available free volume within a
porous sheet, the best way is to test under downward wicking.
© The Textile Institute
46 Textile Progress doi:10.1533/tepr.2006.0001

The wetting and wicking abilities of plasma–treated linen fabric were investigated
by Wong et al. [139], using contact angles and upward and downward wicking
methods. In both the experiments, the distance water travels along the fabric is
determined as a function of time.
Hollies et al. [144] also discussed a horizontal water transport apparatus for fabrics.

3.2.1.3 Transplanar or Transverse Wicking


Transverse wicking is the transmission of water through the thickness of a fabric, i.e.
perpendicular to the plane of the fabric [3, 9]. It is also termed ‘demand wetting’
[148], or ‘spontaneous transplanar liquid uptake’ [105]. It is perhaps of more importance
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

than longitudinal wicking because the mechanism of removal of liquid perspiration


from the skin involves its movement through the fabric thickness. Transverse wicking
is more difficult to measure than longitudinal wicking as the distances involved are
very small and hence the time taken by a liquid to transverse the thickness of fabric
is short [9]. Sorption of water in a towel is one practical example [3].
Miller and Tyomkin [105] have shown that liquid entry into a pore need not be
restricted to contact angles less than 90° when the pore walls are curved. Similar
research work was also reported by Miller in another study [33]. This is essentially
the case with fibrous materials where almost all pores have a splayed opening to
some degree. In pores with straight walls, formation of a concave meniscus exerts a
driving force for the liquid to move in and the contact angle in this case is less than
90°. Miller and Tyomkin illustrated the sequence of events that might be expected
when a liquid front comes in contact with the entrance of a pore as shown in Fig. 26.
The configuration c is not possible as the convex meniscus shape would produce a
pressure gradient opposing liquid rise. The only way the liquid surface can form its
proper contact angle is for the contact line to move into the pore to some extent. This
produces a concave meniscus (configuration d), and in this case the contact angle is
more than 90°.
Various test methods are reported to measure transplanar wicking [5, 9, 33, 105,
149–151]. A porous plate test is used to study transverse wicking. The apparatus for
the plate test is shown in Fig. 27 [5, 9, 150]; it consists of a horizontal sintered glass
plate, kept moist by a water supply whose height can be adjusted so as to keep the
water level precisely at the upper surface of the plate. A fabric is placed on top of this
glass plate and draws water from it at a rate which depends on its wicking power. It
is important that the water level is adjusted to touch the underside of the fabric but not
to flood it. The rate of uptake of water is measured by timing the movement of the
meniscus along the long horizontal capillary tube. The equipment is arranged so that
the head of water supplying the glass plate does not change during the experiment.
Given the diameter of the capillary tube, the mass of water taken up by the fabric in
a given time can be calculated. A problem encountered with the method is that a load
has to be placed on top of the fabric to ensure contact with the sintered glass plate.
This compresses the fabric and, as a consequence, can potentially change it absorption
characteristics. A contact pressure of 0.098 KPa has been used by Harnett and Mehta
[5].
A schematic view of the instrument developed by the Department of Textile

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 47

Air
a Solid

Liquid

Air
b Solid

Liquid
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

Air

c Solid

Liquid

Air
d Solid
E
D
θa
Liquid

Fig. 26 Entry of a liquid into a fabric pore [33 Miller]: (a) liquid surface just before contact
with the pore; (b) not unless θa = 180º; (c) not possible with convex meniscus; (d) allowable
mode of liquid entry

Weight

Capillary tube
Sintered glass
plate

Test fabric

Fig. 27 The porous plate test [9 Saville]

Technology, Indian Institute of Technology, Delhi, and used for conducting transverse
wicking tests, is shown in Fig. 28. A fabric sample is placed on a horizontal base plate
which is connected to a liquid reservoir by means of a syphon tube. The fabric sample
is covered by a cover plate so as to ensure intimate contact between the base plate and
the fabric [116]. The spatial relationship between the bottom surface of the test
specimen in contact with the liquid in the syphon tube and the liquid level in the
reservoir is adjustable. At the beginning of the test, the base plate position is brought

© The Textile Institute


48 Textile Progress doi:10.1533/tepr.2006.0001

Fabric Syphon tube


Cover plate

Base plate Liquid reservoir

0123456 Electronic balance

Height
adjusting knob
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

Fig. 28 Schematic view of transverse fabric wicking tester [116 Chattopadhyay]

to a height which is slightly above the liquid level in the reservoir. After placing the
sample on the base plate, the height is slowly reduced so that the liquid meniscus in
the syphon tube comes into contact with it and wicking starts. The liquid reservoir is
placed on a suitable electronic balance. The balance indicates the weight of water that
has left the reservoir and the difference between the two consecutive readings shows
the weight of water absorbed by the fabric [116].
According to Burgeni and Kapur [152], using a porous plate with a considerable
number of small pores enables the liquid level to remain at the surface of the plate
even when the reservoir level has dropped well below it.
The TRI demand wettability apparatus shown in Fig. 29 is a variation of the
porous plate technique [33]. Before the test sample is put in place, the filter paper is
wetted, and a negative pressure gradient is established by first lowering the wetting

Weight
Sample Filter
paper
Vent Perforated
Pressure plate
head
Reservoir ∆P
Flexible Movable
tubing clamp

Rigid tube
Recorder Balance Clamp Flexible
tubing

Fig. 29 Instrumentation for studying demand wettability under negative pressure gradients [33
Miller]

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 49

chamber until the filter paper is slightly below the liquid level in the reservoir. Once
the paper is completely wet, the wetting chamber is raised to establish the negative
hydrostatic head. The fabric sample is attached to a weighted compression plate by
means of double-sided tape and placed on the wet filter paper. The balance records
the progress of liquid uptake. Data obtained at different negative pressure gradients
can be extrapolated to zero-gradient, usually using semi log plots that produce straight
lines.
A demand wettability apparatus for needle punched nonwoven fabrics is shown in
Fig. 30. The apparatus consists of a burette which is connected by fluid transport
tubing to an opening in the specimen cell. An aluminum cylinder is fitted over the
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

opening. The samples are die-cut to fit the cylinder exactly. A stainless steel piston
fitting closely inside traverses downward and exert a uniform pressure of 2.16 kN/m2.
Additional cylindrical weights with holes drilled in the centre and fitting on the
piston rod allow provision for systematically varying the environmental pressure to
three different levels, namely 2.16, 6.87, and 16.69 kN/m2. From the amount of liquid
absorbed and the time of absorption, the absorbent capacity and average rate of
absorption are determined [153].
Wicking test equipment has also been described by other workers [151].
In the case of demand wetting, the hydrostatic head h produced by the rise of

A – Air bleed B
B – Burette
C – Cylinder
D – Wicking initiating mechanism
E – Leveling knobs
F – Spirit level
D
C
F

Fig. 30 Demand wettability apparatus [153 Gupta]

© The Textile Institute


50 Textile Progress doi:10.1533/tepr.2006.0001

liquid in the fabric pore is trivial and as such the term gravitational pressure can be
replaced by ‘applied opposing’ or ‘negative pressure’ PH or ∆P (shown in Fig. 29), in
equation (23). Then the appropriate version of the Washburn [17] relationship is
[105]:
dh/dt = (2γ cos θ/r – PH)(Πr4/8η h)
Plate tests measure absorbency of fabrics, especially towels [109, 150]. Some of
the tests apply pressure to enhance the flow of liquid, but therefore deviate from a
true wicking process. Alternatively, the fabric is contacted with aqueous surfactant
solution by raising the liquid level from below the fabric until it contacts the fabric
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

[88]. A test developed by Lennox-Kerr [142] uses a capillary tube or a wick to deliver
a liquid continuously to the fabric and the area of the wetted region is measured.

3.2.2 Wicking from a Limited (Finite) Reservoir


The capillary penetration of a drop indicates several important properties of a textile
fabric, including repellency [10], absorbency, sorption of stain, and stain resistance.
Either the drop absorbency time [5, 90] or the spreading rate of the liquid within the
fabric is measured [11, 12, 96, 97]. Wicking of minute drops of liquid has become
very important for printing technology [154]. Wicking of a limited amount of liquid,
such as a drop placed onto the surface of a textile fabric, is a more complicated
process than wicking from an infinite reservoir [3].
In the standard spot tests [5, 155], a drop of liquid (either distilled water or, for
highly wettable fabrics, 50% sugar solution) is delivered from a height of approximately
6 mm onto a horizontal specimen of the test fabric (preconditioned at 20°C, 65%
RH). The region of the fabric on which the drop falls is illuminated by a beam of light
to create a bright reflection from the liquid surface, and the elapsed time between the
drop reaching the fabric surface and the disappearance of the reflection from the
liquid surface is measured. The disappearance of the reflection is assumed to indicate
that the liquid has spread over and wetted the fabric surface. The elapsed time t
recorded is taken as a direct measure of the fabric wettability; the shorter the time,
the more wettable the fabric [5].
While some correlation between wettability and wickability has been observed,
the published literature indicates certain additions or modifications to the standard
test in order to assess wicking [5, 91, 142]. DeBoer’s [91] approach is to measure the
area of the wetted region of the fabric at the moment reflection ceases, or the mass
of liquid absorbed as a percentage of the dry weight of area of fabric equal to the area
of the wetted region, the latter one being defined as ‘saturation value’ (V). Lennox-
Kerr’s [142] approach is to replace the drop by a continuous supply of liquid (e.g. a
capillary tube or a saturated fabric wick in contact with the test specimen) and to
measure the rate of increase in diameter of the wetted region. In this case, precautions
should be taken to minimize the head of pressure in the supply (e.g. by feeding from
a reservoir at the same height as the test fabric) and to ensure that the supply tube or
wick is readily capable of delivering liquid at the rate demanded by the test specimen.
The effect of contact pressure between supply and test specimen should also be
considered.
A radial wicking or ring test is described, consisting of an assembly of horizontal

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 51

micro graduated pipettes feeding into a vertical length of 3 mm bore PTFE tubing
which just touches the sample of fabric stretched on an embroidery ring. The volume
of water desorbed as well as the area spread is measured by observing with a TV
camera both the fabric as well as the pipettes with time [156, 157].
In another method using the above principle, a fabric piece (10 cm × 10 cm) is
placed between two glass plates (15 cm × 15 cm). The top plate has a hole of 5 mm
diameter at the centre (Fig. 31). 0.1 ml of dye solution is added through the hole by
a micropipette. The solution spreads over the fabric and forms a ring, the area and
shape of which corresponds to the ease of water transport in the fabric, as well as the
uniformity in different direction [157].
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

Ring – Test
0.1 ml

Microburette

A
C
B

Fig. 31 Assembly for ring test [157 Achwal]

Gillespie [12] measured the radius of the spot covered by a drop wicking into
paper and developed an equation based on Darcy’s law. Unlike paper, most textile
fabrics are not isotropic, and the spreading liquid does not usually form a circle with
a well-defined radius. Kissa [13, 96] therefore measured the area covered by the
spreading liquid by a photometric technique. Kawase et al. [11, 158, 159] used the
image analysis studies for spreading of liquids in textile assemblies.
According to Kawase et al. [11] the spreading of water in cotton fabric is a typical
case of diffusion of liquid into fibers. They plotted area of spreading versus time for
each volume of the droplet. This showed two straight portions with a high correlation
before and after the time at which the droplet disappeared (about 0.2 minutes). The
first linear spreading rate curve represents the spreading during Phase I, and the
second strength line that during Phase II. In another work, Kawase et al. [159] used
an image analyzer technique for objective measurement of spreading areas. In this
apparatus a fabric is mounted on a wooden frame. A measured amount (0.05 to 0.20
ml) of n-decane is introduced onto the fabric through a sealable orifice by a micropipette.
The area of the spreading liquid and time are recorded simultaneously on videotape
with a CCD-camera as a black/white image.
The wicking process is kinetically quite different when capillary penetration is
accompanied by diffusion of the liquid into the fibers or into a finish on the fibers.
Sorption within the fibers decreases the volume of the liquid flowing in the capillary
spaces, and reduces the interfiber spaces available for capillary penetration because
of swelling of the fibers [3]. The area covered by the liquid spreading within the
© The Textile Institute
52 Textile Progress doi:10.1533/tepr.2006.0001

fabric does not correlate with the drop absorbency time. When the drop absorbency
time is used to evaluate fabric absorbency, an inadequate drop volume can produce
misleading results if capillary penetration is accompanied by absorption of the liquid
within a finish or into the interior of fibers [98, 99]. When the capillary spaces in a
fabric are not uniform, the liquid may not spread as a continuous front, but may
instead penetrate some capillaries before others, forming a ‘fingering’ pattern. A non-
uniform wicking pattern complicates measurements of wetted area, even when using
imaging techniques. However, the pattern itself provides useful information about the
characteristics of the fabric [3].
When the liquid is a surfactant solution, drop absorbency time and rate of capillary
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

penetration depend on the nature of the surfactant and its concentration. The
interpretation of wicking data is complicated by the small size of the liquid reservoir,
which has a pronounced effect on the effective surfactant concentration. Drop absorbency
time decreases with increasing surfactant concentration and becomes constant well
above the critical micelle concentration (CMC). The need to increase the surfactant
concentration above CMC is due to the depletion of surfactant as it is adsorbed on
fibers. Since the liquid reservoir is of limited size, the amount of surfactant adsorbed
cannot be replenished, and its concentration drops. Hence, the drop absorbency time
and capillary penetration of surfactant solution do not depend only on the wettability
of the fabric, but on surfactant–fiber interactions in a complicated manner [88].
Herlinger et al. [88] employed a photographic technique to measure the contact angle
of a drop placed on a fabric as a function of time. The contact angle changed from
initial values ranging from between 40° and 130°, to 0°, as the drop moved into the
fabric. A surfactant decreased the drop absorbency time, but a hydrophilic treatment
of the fabric was found even more effective.
In order to obtain much broader information on the moisture management properties
of the fabric, there is a need to develop a test method that will quantitatively determine
both absorption and wicking properties [160, 161]. A method is described which
simulates the absorption of sweat by a fabric from a profusely sweating body and its
transport across the fabric surface [161]. The test sample, of dimensions 140 mm x
150 mm, is placed on the support base so that the center of the fabric matches as near
as possible the center of the porous plate (Fig. 32). The area of absorption is limited
to the surface area of the porous plate. Wicking then follows as soon as the area under
the porous plate becomes saturated. Normally, wicking is allowed to continue up to
the prescribed distance and a near-circular wicking pattern is obtained with most of
the fabrics. With some knitted fabrics, however, the wicking pattern is elliptical, and
allowance has been made in the software to take this into consideration in area
calculation. Connection to the computer allows for accurate data acquisition, with
measurement of water loss from the reservoir being taken from the balance every
second, or as many as three readings per second with certain fabrics. From the loss
of mass of water from the reservoir, the fabric absorption and wicking characteristics
are computed using software. The initial portion of the plot of total water uptake
versus time reveals that the absorption rate corresponds to absorption of water on the
fabric up to the area equal to that of the porous plate. Water uptake beyond this point
is related to wicking rate.
While wetting and wicking refer to what may appear to be separate phenomena,

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 53

Mass, 592 g (pressure,


25.3 g cm–2)

Perforated plate with attached foam

Fabric sample, 140 mm × 150 mm


Porous plate area, 23.75 cm2

15 mm
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

Water

Support base
Computer

Electronic balance

Fig. 32 Absorption-wicking apparatus [160 D’Silva, 161 Anand]

they can be described by a single process – liquid flow in response to capillary


pressure. This process is described mathematically by Darcy’s equation [162]:
kdpc
V=
ηdx
where V (cm/s) is the apparent velocity of the liquid (volume flow rate divided by
cross-sectional area), k (cm2) is the permeability that describes the ease with which
liquid water flows through porous media, and pc (g/cm s2) is the capillary pressure
that is the driving force for liquid movement. In order to use Darcy’s equation, the
values of permeability and capillary pressure must be known. These parameters vary
greatly as a function of the liquid content of a fabric. If Darcy’s equation is to be used
to describe liquid transport in fabrics for the wide range of conditions associated with
what is traditionally referred to as wetting and also wicking, then the capillary pressure
and permeability values must be known for conditions ranging from a fabric with no
liquid in it to a fabric where all void spaces are filled with liquid. Once capillary
pressure and permeability are determined for saturations ranging from near zero to
100%, liquid transport related to both wicking and wetting behavior can be described
by Darcy’s equation [162].
Meeren et al. [163] have described a method for testing rate of wicking for cotton
terry fabrics.

3.2.3 Other Methods


Some other test methods are reported for measuring the wicking behavior of fabrics
[5, 164–166]. According to Zhuang et al. [166], at higher ambient temperatures or
during strenuous bodily activity, the wearer perspires profusely so that clothing worn
next to the skin becomes saturated with perspiration. Ideally, the liquid on the body
© The Textile Institute
54 Textile Progress doi:10.1533/tepr.2006.0001

surface or in the inner layer of clothing should be transferred to the outer layer so as
to keep the skin dry and allow the liquid to evaporate. Moisture vapor may form in
clothing under certain circumstances, which may result in liquid transfer back from
the outer layer to the body, even if perspiration is not significantly present on the
skin. Therefore, it is important to understand transfer wicking mechanism in terms of
wearer comfort for functional sportswear.
There is no standard method for measuring transfer wicking. Based on Spencer-
Smith’s approach [167], Zhuang et al. [166] studied transfer wicking.
An investigation of liquid interaction within clothing systems showed that the
amount of liquid transferred largely depends on the performance of individual fabrics
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

as well as on the way in which they contact each other [166]. Transfer wicking in
woven and knitted fabrics has been investigated by other workers [168, 169].
A Laser-Droplet Anemometry (LDA) technique to measure fluid velocity through
a web is described by Howaldt and Yoganathan [170]. LDA measurements are usually
taken in systems where the velocities are in the range of cm/s to m/s, whereas in the
case of fluids wicking in fibrous materials, the steady state velocities are of the order
of mm/s or less. A test cell was developed and described to ensure that LDA was
capable of detecting such low velocities.

4. FACTORS AFFECTING WETTING AND WICKING IN FIBERS


AND FIBROUS ASSEMBLIES
Various factors affecting wetting and wicking in fibers and fibrous assemblies are
discussed below.

4.1 Fibers
Factors such as type of fibers, chemical purity, orientation of molecules, surface
contamination, surface finish, cross-sectional shape, surface roughness, pre-wetting,
annealing, argon glow discharge and corona treatments, presence of surfactants,
alkaline hydrolysis, washing, bleaching, and mercerization are found to influence the
wetting behavior of fibers.
Wetting force increases linearly with diameter of filaments, indicating that wettability
of filaments is the same, irrespective of some filament diameters [8]. Whang and
Gupta [15] tested wetting characteristics of some finish-extracted cellulosic fibers,
namely cotton, regular rayon (roughly round but crenulated shape), and trilobal
shaped rayon, using the Wilhelmy technique. The magnitude of the wetting force of
fibers increases with perimeter but the relation is linear only for the receding and not
for the advancing liquid front. The wettability index of all the three fibers was nearly
the same in receding but widely different in advancing. The wetting index while
receding is governed mostly by the chemical make-up of the fiber, the index during
advancing being additionally affected by physical and morphological structures that
include molecular orientation, crystallinity, roughness, and surface texture. The wetting
parameters of cellulose acetate fibers of different sizes and morphologies were quite
different in the advancing mode but were very nearly the same in the receding mode.
Infrared spectroscopy of these three cellulosic fibers confirms that they are chemically
similar [51].
The cosines of advancing contact angles were 0.93, 0.83, and 0.57 for trilobal

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 55

rayon, cotton, and regular rayon fibers respectively. This reflects the relative abilities
of these fibers to attract and imbibe fluid by capillary action in fibrous assemblies.
The trilobal viscose fiber had a high wetting index, followed by cotton and regular
rayon fibers. A separate study [171] indicated that webs of similar constructions
made from trilobal rayon and cotton had much higher rates of water absorption than
did a web containing regular rayon. The receding contact angles for these fibers are
similar due to their similar chemical structures and they are much lower than the
advancing contact angles [15].
Surface contamination on fibers, roughness, and molecular structure of fibers are
the factors responsible for wetting hysteresis [172]. The wetting hysteresis for cotton,
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

trilobal rayon, and regular rayon were 1.06, 1.01, and 1.25 respectively. Very little or
no hysteresis for the trilobal rayon fiber and high for regular rayon fiber may be
explained on the basis of chemical purity, cross-sectional morphologies, and orientation
of molecules in the fibers. The birefringence values of both the rayons were the same
and lower than that of cotton. The trilobal rayon fibers had high purity, were smoother
and had more homogeneous surfaces than regular rayon fibers. These differences are
partly responsible for the difference in the hysteresis values of the two rayon fibers.
The surface structural differences were evident in the wetting force fluctuations. A
very high fluctuation was observed in the wetting force trace for regular rayon
followed by cotton and trilobal rayon fibers. Further, the longitudinal ridges on the
surface of the trilobal fibers would allow them to imbibe fluid ahead of the fluid–
fiber interface in the advancing mode, to a state comparable to that in the receding
mode, and led to relatively lower hysteresis in trilobal fibers [15].
Higher work of adhesion during advancing and, as a result, lower hysteresis for
cotton compared to regular rayon, is due its higher chemical purity and molecular
orientation than the regular rayon. It was shown that contact angle hysteresis was due
to the heterogeneity of the fiber surfaces [15].
Several types of T-300 carbon and polytetrafluoroethylene fibers immersed in
hexamethyl disiloxane silicone oil and an epoxy resin were studied in relation to their
dynamic fiber wettability. After progressive immersions, decrease in the wettability
was observed for the carbon fibers, but not for the Teflon (PTFE) fibers, suggesting
that absorption could be occurring on the carbon surface [173]. The wettability of
polypropylene fibers in water is extremely low compared to polyester and nylon.
However, receding wettability for polypropylene is high compared to others. Pre-
wetting of polypropylene fibers in water had a much more significant effect in enhancing
its wettability as compared to the response of nylon or polyester fibers under the
same conditions [8].
Annealing of polyester at 200°C under slack conditions increases water wettability,
whereas annealing under taut conditions shows no increase in wettability. Annealing
at elevated temperatures causes an increase in shrinkage caused by regular chain
folding and such effects may be reflected in the surface characteristics of polyester
fibers or migration of oligomeric materials to the fiber surface. Annealing of
polypropylene fibers both under taut and slack conditions increased their water
wettability, but the shrinkage phenomenon had a much more dramatic effect. Since
the repeat unit for polypropylene is quite short, it is easier for disordered material to

© The Textile Institute


56 Textile Progress doi:10.1533/tepr.2006.0001

adjust to quasicrystalline packing, and thus mild annealing would alter the wetting of
polypropylene [8].
Argon glow discharge significantly improved water wettability of Kevlar 49 fibers
but lowered their wettability in epoxy resin. SEM micrographs of Kevlar 49 fiber
surfaces after glow discharge showed disappearance of surface nodules and indentation.
The observation on wettability in water and resin and the effects of glow discharge on
wettability suggested that the polar nature of the Kevlar fiber surface might be a
dominant factor in determining its wettability in these liquids rather than the effect of
increased surface roughness. Poor wettability of Kevlar 49 in epoxy resin after glow
discharge is reflected by low interfacial shear strength between fiber and resin [45].
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

Velde and Kiekens [46] investigated wettability and surface analysis of glass
fibers used as reinforcement for thermoplastic and thermosets composites. Glass
fiber treated with a finish for thermosets has a higher surface polarity and lower
advancing contact angles in water and ethylene glycol than others. The presence of
a high surface energy component on the glass surface tends to resist dewetting of the
receding fluid front, lowering the receding angle.
Glass fibers immediately after desizing showed a smooth tensiogram during Wilhelmy
scanning in water. The same fibers after 12 hours of exposure to the atmosphere
exhibited a jagged force-trace during the immersion mode. After combined exposure
to air and water, desized glass fibers showed higher advancing angles and hysteresis.
Admicellar-modified fibers due to their high hydrophobicity, showed a higher average
advancing contact angle compared to that of the fully wettable desized fibers. Sized
fibers, especially those with coupling agents, and admicellar-polymerized styrene–
isoprene are more chemically heterogeneous than the desized ones [47].
Polar matrix materials such as nylon 66 are predicted to adhere much better to
graphite and glass filaments than non-polar resins such as polyethylene. These
predictions correlate well with the notch sensitivity of composites, except for silane
systems, where wetting is not the main mechanism for better adhesion and strength
properties [174].
Rough surfaces give rise to fast spreading along troughs offered by the surface
roughness [175]. Alkaline hydrolysis causes pitting of the surface of polyester fibers
and improves their wettability, as indicated by contact angle measurements [176].
The enhanced wettability is due to an increase in either the number of polymer
hydrophilic groups or their accessibility to water, and/or an increase in the roughness
of the sample surfaces.
In order to improve the sorption characteristics of a cellulose fabric during textile
finishing, different pre-treatment processes such as washing, bleaching, and
mercerization are applied. Pre-treatment increases the sorption ability and makes the
material more accessible to chemicals used in the finishing processes. Fibers with the
highest moisture sorption have the smallest contact angle [120]. Contact angles of
raw and pre-treated regenerated fibers are shown in Fig. 33. Alkaline purification has
the biggest influence on viscose fibers. In the case of lyocell and modal fibers,
influence of alkaline purification is smaller in comparison with viscose fibers and no
essential reduction in contact angles can be seen. In the case of viscose fibers, after
washing, the alkaline solution of the washing agent easily penetrates into less orientated
amorphous regions and breaks down the interactions between the cellulose

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 57

80

75

70
ϕ/°

65
Raw
60
Washed
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

55 Bleached
Viscose Modal Lyocell

Fig. 33 Contact angle, φ, of raw and pre-treated regenerated cellulose fibers [120 Perssin ]

macromolecules. The diameter of the fibers increases and the structure becomes
loose leading to better accessibility of fiber interfaces to liquid. The result is a
smaller contact angle, and wettability and sorptivity improvement of the viscose
fibers. In comparison with viscose fibers, modal and lyocell fibers have a higher
degree of crystallinity and higher molecular orientation, which means that only a
small quantity of washing agent can penetrate into less-ordered amorphous regions
of the fibers. This result also implies a smaller pre-treatment effect on the hydrophilic
character of the fibers. The improvement of sorption characteristics due to washing
in an alkaline medium can be explained by the increase of voids [120].
In emulsion systems, the presence of surfactant will in general modify the rate of
attachment of drops to the fiber by Type 1 and Type 2 mechanisms [55]. The surfactant
will adsorb at both the fiber–water and oil–water interfaces and will thus alter the
interaction between the incoming drop and either the bare fiber or a previously
deposited drop. These two interactions will usually differ because the adsorbed surfactant
will be present at the respective interfaces at different concentrations and in different
molecular conformations. Also, the system geometry for the two types of deposition
is different. A situation is thus conceivable where surfactant adsorption on the fiber
surface is low, relative to that at the oil–water interface [55]. A key element in the
wetting behavior appears to be the boundary condition arising from a surfactant
balance at the contact line, which gives rise to surfactant accumulation and hence to
surface immobilization. Once established, immobilization tends to be self-maintaining
(hysteresis) and surfactant concentrations many orders of magnitude greater than the
equilibrium level can be maintained near the contact line, disproportionably reducing
static contact angle [177].
Perwuelz et al. [178] investigated wetting at a silicone oil/ water/polyester fiber
interface. The contact angle at these interfaces was measured using tensiometry, as
shown in Fig. 34. If the fiber is first covered with oil, it is then difficult for water to
displace the oil phase and the advancing oil–water contact angle measured is greater
than 90° in most of the cases studied. On the contrary, when the fiber emerges from
the water phase, the receding contact angle is low.
Hsieh et al. [179] studied the moisture sorption on the wetting behavior of Kevlar
fibers in water and epoxy resin. When dried by preconditioning at 0% relative humidity

© The Textile Institute


58 Textile Progress doi:10.1533/tepr.2006.0001

Balance Balance

Fibre

Fibre
θo θo
Oil Oil
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

θow θow
Water Water

Receding mode Advancing mode

Fig. 34 Schematic organization at the air: silicone: PET fiber interface and at the silicone:
water: PET fiber interface during tensiometric experiments [178 Perwuelz]

(R.H.), water wettability of Kevlar 49 fibers was superior to that of Kevlar 149 fibers.
Resin wettability of the dried Kevlar 49 fibers, on the other hand, was lower than that
dried of Kevlar 149 fibers. Wettability in water and resin of these two fibers was
affected differently by moisture. Exposure to 97% R.H. moisture level significantly
lowered water wettability of Kevlar 49 fibers but did not affect the wettability of
Kevlar 149 fibers in water. Resin wettability of Kevlar 49 fibers was improved upon
exposure to moisture, but the opposite was observed on Kevlar 149 fibers.
Surface wetting measurements and several analytical techniques including FTIR/
ATR, DSC, TGA, and pyrolysis-GC/MS were used to characterize the noncellulosic
components of developing cotton fibers as intact components of the fiber structure.
Water contact angle measurements are most sensitive to the presence of hydrophobic
compounds on the surfaces of cotton fibers of all ages and to their removal by
alkaline scouring [180].
Hangey [133] reported that nylon fibers used in carpets have a built-in fluorochemical
compound that reduces the surface energy of the fiber and hence increases resistance
to soiling and staining in end-uses.
The wetting behavior of several flax and polypropylene fibers can be characterized
by measuring the wetting rates (penetration velocities) of a series of liquids using the
capillary rise technique [181]. The fiber surface tensions are established from plots of
the normalized wetting rate as a function of the surface tensions of the liquids,
assuming, in analogy to Zisman’s method, that the maximum of the normalized
wetting rate corresponds to the solid surface tension. The method employed to separate
flax fibers from the rest of the plant has a large influence on the fiber surface tension.
Grafting small amounts of maleic acid anhydride (MAH) onto the polypropylene
fibers’ surfaces does not affect the wetting behavior and therefore the surface tension,
whereas grafting larger amounts (10 wt%) of MAH causes the polymer surface
tension to increase significantly.
Corona treatment was given to polypropylene fibers after extrusion but prior to

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 59

drawing and crimping. Treatment resulted in an average decrease of 5–10° in the


advancing contact angle and 10–25° in the receding contact angle for water on the
fibers. In addition, there was a sharp change in wetting and friction properties of
fibers with corona treatment when the amount of spin finish was between 0.12 and
0.13 wt%. These effects are attributed to improved wetting of the treated fibers by
spin finishes, leading to a more uniform spreading of other finishing agents on the
fiber surface [182]. Wakida and Tokino [183] investigated the effect of discharge
treatments such as low-temperature plasma treatment on the surface characteristics
of different textile fibers. The plasma treatment causes mainly chemical modification
and increases wetting and adhesion.
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

Mullins et al. [184] described the effect of fiber orientation on the fiber wetting
process. Glass filter fibers in various combinations were oriented at various angles
within a plane defined by the airflow direction and were supplied with distilled water
in aerosol form. The behavior and flow of the liquid collected by the fibers were
observed and measured using a specially developed microscope cell, detailed in the
paper. A mathematical model was developed and sensitivity analysis of the model
was conducted to determine the most important parameters. This will aid design of
wet filtration systems such that maximal self-cleaning can be accomplished with
minimal water use.

4.2 Yarns
Various parameters, such as yarn structure, yarn tension, twist, fiber shape, number
of fibers in yarns, fiber configuration, finish, and surfactants influencing wicking of
yarns, are discussed below.
Lord [109] reported that open-end yarn wicks faster and more evenly than ring
yarn, but elevates about the same volume of water for a given yarn count. According
to Sengupta and Murthy [117], for any given vertical wicking height, the wicking
time of open-end spun yarn is less than that of ring-spun yarn. Microscopical
examinations of yarns showed that dye had wicked to a greater height in the core of
the open-end yarn than in the surrounding sheath fibers. No such differentiation was
noticed with ring yarns. Open-end yarns have a relatively denser core and less dense
skin when compared to ring yarns.
Chattopadhyay and Chauhan [116] studied the wicking behavior of ring and compact
spun yarns (Fig. 35). The rate of water rise was very fast at the beginning and slowed
down gradually, as observed by various workers. In the first minute it was very
difficult to distinguish the differences between the behaviors of different yarns. The
equilibrium wicking heights observed for ring yarns were more than those of compact
yarns. Ring yarns wicked faster than compact yarns. Coarser yarns wicked faster than
finer ones.
As the packing coefficient of compact spun yarns is greater than that of corresponding
ring yarns, the average capillary size would be less in compact yarns than ring yarns.
It has been stated by Staples and Shaffer [185] that smaller capillaries may create
sufficient drag to slow down the rise in liquid height. According to Chattopadhyay
and Chauhan [116] there must be an optimum capillary size that will cause fastest
entry of water into the yarn pores. Larger than optimum pores will also slow down
entry due to low capillary pressure. Hence, both too small and too large pores are

© The Textile Institute


60 Textile Progress doi:10.1533/tepr.2006.0001

10

8
Wicking height (cm)

5
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

50 Ne ring yarn
40 Ne ring yarn
4 Comparative chart for different count yarn 50 Ne compact yarn
40 Ne compact yarn
3
0 5 10 15 20 25
Time (min)

Fig. 35 Wicking of ring and compact yarns [116 Chattopadhyay]

detrimental to quick wicking. The slowing down of height rise with time for any yarn
can be ascribed to the gravity action of the water column within the capillary, which
acts against the capillary pressure.
Sengupta et al. [114] investigated the wicking behavior of air-jet textured yarns.
The structure of air-jet textured yarns is unique in the sense that, in addition to the
core, the surface of the yarns shows different types of loops of varying shapes and
sizes. The wicking behavior can be affected by both core and surface structure.
Figure 36 shows the relationship between equilibrium wicking height and percentage

80
Equilibrium wicking height, mm

With spin (r = 0.58)


finish
60

(r = 0.72)

Circular filament yarn


40 Without spin finish
Without spin With spin finish
finish Trilobal filament yarn
Without spin finish
With spin finish

20
20 40 60 80
Floats and arcs %

Fig. 36 Relationship between equilibrium wicking height and percentage of floats and arcs for
wet- and dry-textured yarns with and without spin finish [114 Sengupta]

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 61

of floats and arcs for both dry- and wet-textured yarns. For the same percentage of
floats and arcs, the trilobal filament yarns show better wicking properties. The
equilibrium wicking height tends to increase with an increase in the percentage of
floats and arcs. As the air-jet textured yarns have bipartite structure with core and
surface loops, the configuration of loops and their frequency may influence the
wicking behavior, since part of the liquid travels through the periphery of the yarn.
The presence of long drawn-out loops such as floats and arcs offers a less tortuous
path for the liquid to travel; as a result, a higher percentage of floats and arcs leads
to a higher wicking height.
The equilibrium wicking height and wicking rate are higher for air-jet textured
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

yarn than for the corresponding feeder yarn. Equilibrium wicking height initially
increases and then decreases with increasing tension on the yarns during wicking.
The initial increase in height is due to partial alignment of filaments. Further increase
in tension may bring the filaments closer to each other, reducing the capillary radii or
possibly introducing discontinuity in the capillaries [114].
With open-end yarns, the highest wicking is found with a twist multiple of about
4.0, but wicking height is not found to be greatly sensitive to changes in twist
multiple [109]. Over the twist range of 4 to 12 tpi, the change of water transport rate
follows the capillary transport laws fairly closely, high twist yarns exhibiting low
rates. At lower twists, the decrease in transport rate is apparently due to a substantial
reduction in the number and continuity of the interfiber capillaries [89].
Sengupta and Murthy [117] found that wicking was highly sensitive to the twist
and structure of ring and open-end spun yarns. For the ring spun yarns, the wicking
time increases steeply as the twist increases, whereas for the open-end spun yarn, the
increase is gradual. The results of overnight vertical wicking (Fig. 37) also show a
similar trend, i.e. for any given twist, open-end yarn wicks more than ring-spun yarn.
Ansari and Kish [121] investigated the wicking behavior of polyester spun yarns
produced with varying twist levels. It was observed that the wicking rate decreases

Overnight wicking
8
Vertical wicking height, cm

2
Ring spun
Open-end spun

0
23.95 28.75 38.32 47.90 57.48
Tex twist factor

Fig. 37 Effect of twist factor on wicking height [117 Sengupta]

© The Textile Institute


62 Textile Progress doi:10.1533/tepr.2006.0001

with increase of twist factor from 22 to 49 tex0.5 x turns/cm, due to reduction of


capillary size. Twisted filament yarn shows a lower wicking rate than a yarn without
twist. In capillary penetration of liquids, tortuosity affects wicking. Twists in the
yarns influence the size of inter-fiber capillaries as a result of the helical path of the
fibers in the yarns [89]. Minor et al. [112] observed similar findings on nylon filament
yarns of different twists.
Minor et al. [112] reported the role of circular shaped nylon and crenulated viscose
filaments on wicking rate, using loose (low twist) and tight (high twist) yarns. They
selected the twist levels for the loose and tight yarns such that the ratio of the
equivalent capillary radii (rloose/rtight) was 1.3. From the rate of rise of various liquids,
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

they deduced the ratio of equivalent capillary radii, which were all much higher than
the value 1.3 calculated from the yarn geometry. Further, from the yarn geometry, for
the loose yarns the ratio of equivalent radii rviscose/rnylon was 1.45, and for the tighter
yarns it was 2.23. As the rate of wicking (ds/dt), i.e. the slope of plot of rate curves,
is proportional to equivalent capillary radius, the ratios of wicking rates for viscose
and nylon yarns under loose and tight constructions were expected to be 1.45 and
2.23. However, the wicking rates in loose viscose yarns were abnormally high. It is
evident from the crenulated cross-section of the viscose filament that the open cross-
sectional areas in a loose yarn, calculated from the microscopically observed diameter,
will be greater than the real cross-sectional area. This would account, at least in part,
for the liquid migration in this yarn being faster than predicted by the theoretical
calculations. When the viscose yarn is tightened by tension and twisting, the irregular
cross-section affords the possibility of very tight packing. If the crenulations mesh
like gear teeth, the open space can be reduced greatly without any corresponding
reduction in the yarn diameter. Thus the wicking rate in the tight viscose yarns could
readily be diminished to the level observed.
Ito and Muraoka [124] reported that water transport (measured by the capacitance
method) is suppressed as the number of fibers in the yarn decreases. From the results
of rayon, nylon, and PET fibers, they found that, when the number of fibers is greater
than ten, liquid water moves along even untwisted fibers. But when the number of
fibers is less than five, wicking occurs only for twisted fibers. When the number of
fibers was reduced to three, wicking did not always occur. When the number of fibers
was reduced to two, there was scarcely any wicking (only once in ten times), and
there was none with a single fiber. This result indicates that the mechanisms of water
transport for an isolated single fiber differs from water sorption in a fiber bundle or
assembled fibers where capillary spaces exist [4].
The dynamic wicking behavior of both circular and trilobal filament air-jet textured
yarns is shown in Fig. 38. Yarns of trilobal filaments have faster wicking rates than
circular filament yarns [114]. The trilobal filament yarn has a larger core diameter
(0.388 mm) compared to that of the circular filament yarn (0.326 mm). The inherent
bulkiness of trilobal filament may be the reason for the better wicking rate in trilobal
filament yarns.
Hollies et al. [89] reported that differences in yarn surface roughness give rise to
differences in wicking of yarns and fabrics made from the yarns. Wool fibers form
rough yarns of high apparent contact angle because of the natural crimp and more
random distribution of fibers in the yarns, whereas the yarns of synthetic fibers are

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 63

8
Circular
Trilobal

6
Wicking height (cm)

With spin finish

4
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

2
Without spin
finish

0
0 10 20 30 40
Square root of time (s)

Fig. 38 Wicking height as function of square root of wicking time for both circular and
trilobal filament textured yarns before and after spin finish removal [114 Sengupta]

compact, well aligned, and hence smooth in terms of contact angle. This type of
behavior is in agreement with evidence from the wicking and structural properties of
different blended fabrics discussed in a previous study [186]. Increase in yarn roughness
due to random arrangement of its fibers gives rise to a decrease in the rate of water
transport, and this is seen to depend on two factors directly related to water transfer
by a capillary process: (i) the effective advancing contact angle of water on the yarn
is increased as yarn roughness is increased; (ii) the continuity of capillaries formed
by the fibers of the yarn is seen to decrease as the fiber arrangement becomes more
random. The measurement of water transport rates in yarns is thus seen to be a
sensitive measure of the properties of fiber arrangement and yarn roughness in textile
assemblies [89].
Liquids containing surfactant on fiber surfaces affect the wicking flow [88, 108].
Kamath et al. [115] reported on the effect adding a few drops of fluorocarbon surfactants
(FC-129 and F) to the spin finish liquid (P-65) on wicking rates in polyester yarns.
Surfactants are often used to promote spreading of liquids on the surface of solids.
Fluorosurfactants are very efficient in this regard because of the very low surface
tensions of their solutions. It is found that fluorosurfactant FC-129 has essentially no
effect on wicking where as fluorosurfactant F exerts a significant negative effect on
the wicking rate. The experimental results shows that, although both FC-129 and F
are surface active and reduce the surface tension of the P-65 solution, this does not
translate into any improvement in wicking rate. On the contrary, the wicking rates are
reduced, although the reduction is not significant in the case of FC-129, because the
product σ cos θ decreases as the surface tension of the liquid goes below the critical
surface tension σc of the solid surface.
In two component systems such as surfactant solutions, absorption of surfactant
molecules on the solid surface at the wicking front results in an increase in the

© The Textile Institute


64 Textile Progress doi:10.1533/tepr.2006.0001

surface tension of the liquid, with a concomitant decrease in cos θa [149]. This is
shown schematically in Fig. 39. Equilibrium conditions are reestablished when the
molecules diffuse from the more concentrated regions into the leading edge of the
meniscus [115]. These effects, often termed transient effects, arise due to depletion
and replenishment of surfactants at the liquid surface [149]. The overall results of
adsorption and surface tension gradient-related contractions at the surface (Marangoni
effect) are erratic wicking behavior and a lower wicking rate. These effects are more
pronounced in dilute solutions and decrease as concentration increases. The wicking
rate increases with concentration, approaching Lucas-Washburn behavior [16, 17].
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

Marangoni contraction
σ
σ′
σ′ > σ

Fig. 39 Diagram showing Marangoni contraction of a liquid meniscus within a pore when
depletion of a surface active component increases surface tension [149 Hirt]

Wolfrom and Nuessle [187] investigated the influence of temperature on the observed
wetting time and the relationship between wet pick-up and observed wetting time for
cotton skeins. Most wetting agents follow a typical curve characterized by decreasing
wetting time up to about 60°C, an increase to a peak near 85°C, followed by a sharp
drop as the solution approaches the boil. With soap and a few other surfactants, the
curves level off after the initial decrease so that there is no peak, with a marked drop
beyond 85°C. The peak occurs at approximately the melting point of the cotton wax
and it seems that once the wax is completely melted, the skein will sink rapidly in any
surfactant solution.
Wool yarns may be chemically treated with finishes of low surface energy (usually
fluorocarbon polymers) to retard the wicking of aqueous stains into the yarn. Such
finishes are widely used on carpet piles [188]. Treatment of woolen yarns with
several typical fluorochemical finishes, showed a considerably reduced rate of liquid
uptake, the most effective by a factor of 300 relative to scoured, propanol/water-
rinsed yarn. The fluorochemicals that gave no change in oil repellency between the
0.5% and 3% levels also showed the smallest change in wicking rate. In the other
cases, an increase in oil repellency led to a decrease in wicking rate, in line with
previous suggestions [132, 137] that wicking involves the adsorption of the surfactant
hydrophobe on the hydrophobic fiber surface.
Air jet textured yarns with spin finish have faster wicking and more liquid rise
than finish-free yarns, due to higher hydrophilicity of filament surfaces. The wicking

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 65

rate is more sensitive to the spin finish rather than the structure of the air-jet textured
yarns and constituent filaments, indicating the importance of initial wettability as a
prerequisite for better wicking [114].
Chen et al. [122] showed a linear relation between the wicking time Tw and the
initial droplet volume squared V02 for wicking of droplets of a spin finish on a
polypropylene yarn. The slopes of the plot were very close to those predicted by the
Lucas-Washburn equation [16–17].

4.3 Fabrics
Factors such as size, shape, alignment and distribution of fibers, fiber combinations,
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

yarn structure, fabric construction parameters, fabric position in multilayer system,


desizing, scouring, bleaching, alkaline hydrolysis, enzymatic treatments, plasma, UV
and ozone treatments, property of liquid, surfactants, type of finishes, use of electrolytes
in disperse dyeing, and laundering of cotton are here discussed.
Hsieh et al. [189, 190] reported that, in the case of fibrous structures, woven,
nonwoven, or knitted fabrics, a distribution of pore sizes along any planar direction
is expected. Wicking rate and liquid transported in a fabric depend on these pore sizes
and their size distribution. The capillary principle dictates that smaller pores are
completely filled first and are responsible for the liquid front movement. As the
smaller pores are completely filled, the liquid then moves to the larger pores. The
sizes and shapes of fibers as well as their alignment will influence the geometric
configurations and topology of the interfiber spaces or pores, which are channels
with widely varying shape and size distribution and may or may not be interconnected
[82, 128, 190]. The shape of fibers in an assembly affects the size and geometry of
the capillary spaces between fibers and consequently the wicking rates. The flow in
capillary spaces may stop when geometric irregularities allow the meniscus to reach
an edge and flatten [3]. The distance of liquid advancement is greater in a smaller
pore because of the higher capillary pressure, but the mass of liquid retained in such
a pore is small. A larger amount of liquid mass can be retained in larger pores but the
distance of liquid advancement is limited. Therefore, fast liquid spreading in fibrous
materials is facilitated by small, uniformly distributed and interconnected pores,
whereas high liquid retention can be achieved by having a large number of large
pores or a high total pore volume [82].
Wicking is affected by the morphology of the fiber surface, and may be affected
by the shape of the fibers as well. The common belief that fiber shape does not affect
wetting is valid only for the wetting of single fibers. The shape of fibers in an
assembly such as yarn or fabric affects the size and geometry of the capillary spaces
between the fibers, and consequently the rate of wicking [20].
Comparative studies carried out by Hollies et al. [144] on the rate of movement of
water along fabrics have shown that the penetration of the capillaries formed by the
fibers in the yarns accounts for most aspects of water transport behavior. Both the
amount of water carried by the fabric and the distance that it travels in unit time are
influenced considerably by the randomness of the arrangement of the fibers in the
yarns. The same factor seems to control the ease of wetting of the surface of the
fabrics.
The distribution of a liquid in a vertically hung fabric may exhibit a pattern,

© The Textile Institute


66 Textile Progress doi:10.1533/tepr.2006.0001

especially when wicking is accompanied by sorption in fibers [128]. Longitudinal


wicking in a fabric does not predict soil release when swelling of fibers or a hydrophilic
finish on fibers reduces capillary spaces and retards wicking [3].
A study on polyester surgical woven fabrics using the sessile drop technique
revealed that fabrics constructed from microfilament yarns have higher contact angles
than the others. Aseptic fabrics (sterilized) have mostly higher contact angles than
non-aseptic fabrics. High contact angles of liquid on fabrics are the result of high
hydrophobicity of the fabrics [81]. It was proposed that surfaces that are rough on a
nano-scale, tend to be more hydrophobic than smooth surfaces because of the extremely
reduced contact area between the liquid and solid as analogous to so called lotus-
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

effect (repellency of lotus leaves) [191]. This gives a self-cleaning effect to surgical
fabrics, i.e. particles adhered on the fabric surface are captured by rolling water due
to the very small interfacial area between particle and rough fabric surface [81].
Hsieh and Yu [1] studied the water-wetting properties of woven fabrics and their
constituent single fibers. Overall, contact angles measured on fabrics and on constituent
single fibers are the same. The contact angles of fabrics are not affected by fabric
configurations such as length and fabric–water interface depth. Variations in liquid–
solid interfacial dimensions (perimeter) and contact angle are much larger for single
fibers than for the fabrics. Larger variations are observed on contact angles on single
fibers for cellulosic materials (cotton, rayon, and acetate) than on the fabrics. Greater
variation in size and cross-sectional shape of these cellulosic fibers may explain the
larger standard deviation in their perimeters and perhaps wetting properties. However,
polyester fabrics show greater variation in wetting parameters than constituent fibers.
The low wettability of polyester fibers causes inconsistent wetting along the liquid–
fabric interface, leading to differential liquid uptake among the fabric specimens.
Ring yarn fabrics wick faster than compact yarn fabrics (Fig. 40) [116]. Fabrics
made from coarser yarns (40S Ne) show faster wicking than those made from finer
yarns (50S Ne). Wicking behavior of fabrics follows the same order as that of yarns.

19.5

18.5
Wicking height (cm)

17.5

16.5

15.5
Fabric made from
14.5 Ring spun yarn 40 Ne 50 Ne
Compact spun yarn 40 Ne 50 Ne
13.5

12.5
0 2 4 6 8 10 12 14 16
Time (min)

Fig. 40 Wicking behavior of knitted fabrics [116 Chattopadhyay]

Comfort characteristics of plain-woven fabrics containing viscose staple fiber,


twistless and hollow fibrous assemblies and core-sheath type DREF-III yarn in the
© The Textile Institute
doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 67

weft, have been reported by Das and Ishtiaque [192]. Three different types of fabrics
with different types of DREF III yarn in the weft were produced, viz. fabric A (59 tex
yarn with staple viscose fiber in both core and sheath in weft), fabric B (118 tex yarn
with staple viscose fiber in core and water-soluble staple PVA fiber in sheath in weft)
and fabric C (118 tex yarn with staple PVA in core and staple viscose in sheath in
weft). The PVA was dissolved out completely at 60ºC. They found that fabric B has
the highest wicking value in the weft direction followed by fabric C, and that fabric
A showed the lowest wicking in the weft direction. The warp-way wicking for all the
fabrics was almost the same, which may be due to use of the same warp yarn for all
three fabrics. Fabric B had twist-less and parallel-channeled fibrous assembly in the
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

weft. The twist-less fibrous assembly, due to more parallel fibers, forms small pores
and channels and absorbs more water through capillary pressure. Small, uniformly
distributed and inter-connected pores, and channels facilitate fast liquid transport.
Bhargava [193] has investigated the wettability of terry towel fabrics of various
constructions using a sink time method. Fabrics with higher loop density (repeat with
3-pick) sink faster than those with lower pick density (4, 5, and 6-pick weaves).
Vertical wicking tests show that wicking rate is the fastest in the fabric with 3-pick
followed by fabrics with 4, 5, and 6-pick density. Fabrics having loops only on one
surface, take a lesser time for sinking in water an have a faster vertical wicking rate
and higher water absorbency than those with loops on both surfaces. The percentage
of water absorbed by fabrics with high loop density is more than that of those having
lower loop density.
Yoon and Buckley [194] reported the vertical wicking behavior of various cotton
knit fabrics. They found that wicking rate was higher in the wale direction than in the
course direction. They also observed substantial variation in wicking behavior as the
fiber composition varied in polyester/cotton fabric samples. 100% cotton and 50/50
blend fabrics showed a very rapid wicking behavior but the wicking rate sharply
dropped as the polyester content increased. For 100% polyester fabric, they did not
observe any amount of wicking within the time-scale of the experiment. The liquid
transport properties of a fabric as a whole are essentially determined by the energetics
between the fiber surface and the liquid. Bulk properties of the fiber material, such
as regain, do not have any significant influence on the liquid transport properties.
The effect of pile height, pile density, areal density, thickness, etc. on the surface
water absorption characteristics and wicking behavior of different cotton terry fabrics
have been investigated [195]. Wicking heights both in warp and weft directions were
found to increase with increase in pile height for a given pile density. A combined
increase in pile density and height showed increasing wicking height along warp and
weft. Water absorption and wicking height increase with thickness and gsm of the
fabrics.
In multilayer fabrics, wicking between the fabrics is by means of a vapor diffusion
process. Wicking will not begin when the moisture content of the wet fabric is less
than that of 30% moisture regain. Pore size and pore volume play a major role.
Wicking behavior also depends upon the external pressure applied on the fabric layer
and fabric position. When the back side of a dry fabric contacts the face side or the
back side of the wet fabric, transfer wicking will not takes place (Fig. 41). This is
because of increase in capillary size. The face-to-face contact between dry and wet

© The Textile Institute


68 Textile Progress doi:10.1533/tepr.2006.0001

Face to face Face to back


Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

Back to back Back to face

Fig. 41 The position of the cloth during transfer wicking [196 Ramachandran]

fabric has higher wicking behavior than that of dry fabric contacting the back side of
the wet fabric. This is because the capillary size in the smooth area is less and this
causes the water to be retained. So, the position of the fabric is essential in the
wicking behavior of knitted fabrics [196].
Transfer wicking properties of Aquator and polyester eyelet knit fabrics have been
compared [196]. Liquid transportation in the Aquator fabric is high so that water/
perspiration from the skin that is in contact with the fabric is transferred to the second
layer where it escapes to the atmosphere within a short period. In the polyester eyelet
knit fabric, the liquid transportation is even so it requires more water in the first layer
to transfer the water to the second layer. The Aquator more easily allows the water
to fill the interstices than the polyester eyelet knit fabric. This is shown in the Fig. 42.

Absorbent layer Absorbent layer

Conductive layer aquator Conductive layer polyester


eyelet knit

Fig. 42 Distribution of water in the first layer of aquator and polyester eyelet knit fabrics
[196 Ramachandran]

According to Adler and Walsh [168], vapor diffusion is the major mechanism of
moisture transport between two layers of fabrics at low moisture level. In woven
fabrics, wicking commences only when the regain is more than 30% whereas in
knitted fabrics wicking does not commence at this regain level. Crow and Osczevski
[169] conducted transfer wicking experiments on a range of knitted and woven fabrics
used for active wear. Their results indicated that the amount of water that wicks from
one layer to another depends on the sizes and volumes of the pores.
© The Textile Institute
doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 69

Weiyuan et al. [197] studied moisture transfer in knitted fabrics made from different
fiber combinations of polypropylene and hydrophilic fibers treated with resins. The
polypropylene fiber, which is knitted in the wrong side (inner side), transferred the
moisture easily to the hydrophilic fibers in the right side (outer side). The polypropylene
combined with silk and polyester had low moisture transfer properties whereas the
polypropylene combined with viscose/cotton showed greater moisture transfer.
According to them, knitted fabrics have more pores in their structure because of
lower cover factor and they will have better liquid transmission than woven fabrics.
Plain knitted fabrics with finer polypropylene in the wrong side show good transmission
of liquid, so they are better for sportswear.
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

Kim and Hsieh [198] studied the wetting and absorbency of nonionic surfactant
solutions on cotton fabrics. The wetting properties of hydrophobic cotton fabrics are
greatly improved by addition of nonionic surfactants in the aqueous systems.
There were no significant changes in the pore-size distribution of cotton and
polyester fabrics before and after treatment with a commercial antistatic agent containing
cationic surfactant. Antistatic finishes on the cotton reduced the surface energy of the
fiber as indicated by higher values of contact angles, and reduced both the wicking
and water uptake rates. When the cationic surfactant is absorbed on the fabric, the
hydrophilic end apparently attaches to the surface of the cotton, and the hydrophobic
alkyl chain protrudes out, resulting in a hydrophobic surface. However, the antistatic
treatment significantly increased the wicking rate and rate and amount of water
uptake of the polyester fabric [145].
The aqueous retention (W) in cotton fabrics is positively related to their wetting
behavior (cos θ) and negatively related to the liquid γ [198]. For maximum liquid-
holding capacity, an uncompressed fibrous assembly requires wettable fibers with
high wet modulus. The use of wet cross-linked rayon, which has the requisite
combination of wet and dry properties, leads to significant improvements in absorbency
[199]. The vertical liquid retention capacity (H), water retention (W), W/H ratio, and
cosine of contact angle are higher for desized–scoured fabrics than those only desized.
The desizing process substantially reduced the cosine of contact angle and water
retention from the values observed for the original cotton fabrics, due to the removal
of water-soluble size and exposing the raw cotton fiber surfaces, which consists of
hydrophobic cuticles and noncellulosics [198].
Quantitative methods of measuring water wettability using wetting force and retention
are highly sensitive to differentiating the effectiveness of scouring and bleaching
processes on cotton assemblies. The improved water wettability and retention of
scoured and bleached cotton fabrics are reported by Hsieh et al. [189]. The water
wetting contact angles (CAs) of raw cotton fibers and gray plain weave and satin
weave cotton fabrics are found to be same. The forms of the cotton assemblies that
are scoured and the duration of the scouring affect the degree of improvement in
water wetting properties. The water wetting CAs of single fibers improve to a much
greater extent when scoured for two hours in loose fiber form than in the yarn or
fabric form. Water wetting contact angles of single fibers taken from the scoured
plain weave fabrics are identical to those of fabrics and yarns scoured under the same
conditions. Scouring improves water wettability and water retention, even when the
pore volume in the fabric is reduced. Bleaching improves surface wettability and

© The Textile Institute


70 Textile Progress doi:10.1533/tepr.2006.0001

water retention without affecting the fabric pore structure. The water retention properties
of scoured and bleached cotton fabrics are directly related to their wetting properties
in water, or inversely related to their water wetting CAs [189]. Alkaline and bleaching
treatments of cotton [200] and enzymatic scouring of cotton [201] enhance the wetting
and wicking properties of textile materials [202].
Altering the chemical and physical properties of fibrous assemblies using hydrolysis
by alkaline or enzymes on polyester [203], and argon glow discharge treatments to
several synthetic fiber assemblies improves the wetting and wicking properties of
textile materials [204]. Alkaline hydrolysis is a surface reaction [205] that increases
the surface polarity, enabling polar interaction or hydrogen bonding with water
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

molecules, thus increasing the water wettability of the polyester and water retention.
By varying the conditions during hydrolysis, it is possible to vary the degree of
hydrolysis from mere surface hydrolysis to extensive removal of polymer, leading to
changes in pore size, pore distribution, and connectivity of pores, which are the
factors for improving liquid retention in hydrolyzed polyester [203].
Nandy et al. [206] reported that there is a substantial increase in the wicking
height with the increase in weight loss of polyester fabric due to hydrolytic action.
The wicking height was the maximum for a sample treated with sodium hydroxide
and methanol, followed by that for the sample treated with only sodium hydroxide
[206]. This may be due to the flow of water through the inter-fiber spaces in the yarn
by the interfacial tension between water and fiber surfaces [194].
Hsieh et al. [190] investigated the contributions of surface wetting and pore structure
to liquid retention of fibrous materials using regular and micro denier polyester
fabrics as the porous media. Various surface wetting contact angles and pore structures
were produced by hydrolysis of these PET fabrics using aqueous sodium hydroxide.
Increasing levels of weight loss, porosity, and thickness reduction were observed
with increasing hydrolysis temperatures. The effects of hydrolysis temperature on
wetting and pore structure were different for the two PET fabrics. Lower contact
angles and larger vertical liquid retention, observed for micro denier polyester fabrics
compared to regular denier fabrics, could result from a higher concentration of polar
groups or greater roughness. Varying the lengths of hydrolysis time also imposed
different effects on wetting and the pore structure of the micro denier fabric.
The ability of various enzymes to improve the surface wetting properties of raw
and pre-treated cotton fabrics has been studied by Hartzell and Hsieh [202]. When
applied alone to raw cotton fabrics, pectinase, lipase, and protease treatments provide
very little improvement in water wetting and retention properties. A combination of
pectinase and cellulase significantly improves water wetting and retention properties
similar to those of commercially scoured cotton fabrics. Water pretreatment at 100°C
enhances the effectiveness of subsequent pectinase and cellulase reactions. Pectinase
following the 100ºC water pretreatment produces improved wettability with the least
weight loss and dimensional change. Cellulase treatments on scoured fabrics further
enhance water wettability and whiteness, but with significant reductions in fabric
weight, thickness, and strength [202]. Cotton fabrics treated with cellulase enzyme
were found to have improved wettability [207]. Hydrolyzing enzymes (lipases) improve
the water wetting and absorbent properties of regular polyester fabrics more than
alkaline hydrolysis under optimal conditions, accompanied by full strength retention

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 71

compared to the significantly reduced strength and mass from alkaline hydrolysis.
The wetting and absorbent properties of sulfonated polyester and micro denier polyester
fabrics are also improved by lipase [203].
In recent years, surface modifications of fibers by corona/plasma treatments have
opened up new possibilities in relation to fiber wettability [182]. Plasma surface
treatment causes changes to a limited depth; bulk properties of even the most delicate
materials remain unchanged [208]. Tsai et al. [204] found that the treatment of
polypropylene meltblown webs with one-atmosphere glow discharge plasma using
CO2 gas improved the wettability, as indicated by the increased critical surface tension
of the substrate. The plasma treatment induced a chemical reaction that converted
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

hydrocarbon to carboxyl, carbonyl, and hydroxyl polar groups and which was
irreversible. Other molecules are also polarized. These two factors are responsible for
improved wettability. The decay of critical surface tension of the treated substrate
levels-off after 3 days. The initial decay is due to depolarization of material with
time. It is pertinent to note that plasma-treated fibers have shown surface roughness
in SEM photomicrographs, resulting from chemical reactions and micro etchings on
the fiber surface. However, it was pointed out that surface roughness is not a primary
reason for improved wettability, but may increase it [204]. A plasma treatment on
polypropylene fabrics using saturated CF4 gas decreased water wettability due to
implantation of fluorine atoms into the surface of the PP fibers [209].
Wong et al. [139] have presented wetting and wicking behavior of linen treated
with low-temperature oxygen and argon plasma. Wetting and wicking abilities of
plasma-treated linen were investigated using contact angles and upward water wicking
methods. The results showed that the contact angle between the liquid and the low
temperature plasma treated fabric surface decreased. The oxygen plasma treatment
can increase the downward wicking rate under all treatment conditions. However,
prolonged exposure of linen to 100W argon plasma and 200W oxygen plasma leads
to a significant reduction of the downward wicking rate. The downward wicking
method is more suitable for distinguishing the effects of plasma treatment under
various conditions. They concluded that plasma treatment greatly improves the wicking
properties of linen, making it more absorbent. According to Kwon [209], CF4 plasma
treatment can probably be a simple and effective method to increase resistance of PP
fabrics to water uptake.
Micheal and Zaher [210] evaluated the effect of ultra violet/ozone treatments for
different times on the characteristics of wool fabrics with respect to wettability and
permeability. The results indicated that the improvement in wetting may have been
due to surface modifications; this meant that an increase in the amorphosity of the
treated samples, oxidation of the cystine linkage on the surface of the fabrics, and the
formation of free-radical species encouraged dye penetration and aggregation inside
the fiber pores as well as bond formation.
Yatagai [211] has investigated the correlation between water contact angle, work
of adhesion and water absorbency for rayon and polyester fabrics soiled with various
oily substances. Positive correlations were found between work of adhesion (oil–
water) and relative water absorbency values for oil-soiled fabrics. Soiled rayon fabrics
were less wettable than the unsoiled ones and the same trend was observed mostly
with polyester fabrics except for some oils.

© The Textile Institute


72 Textile Progress doi:10.1533/tepr.2006.0001

D’Silva et al. [160] reported some interesting results for wicking of fabrics. Although
cotton has the highest absorption capacity and rate, the amount wicked, as well as the
wicking rate, was no different from that of untreated polyester fiber fabric. The
reason for this behavior could be due to the swelling of the cotton. The polyester
fabrics do not swell in contact with water. A faster rate of wicking observed across
soil-release-finished polyester fabric was due to the modification of fiber surface
properties, governed by the extent of hydrophilicity imparted by the soil-release
finish. The wicking rate was therefore faster, and the amount of water wicked was
much higher.
Hawkyard et al. [207] found that crease-resistant and flame-retardant finish on
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

bleached and mercerized cotton significantly improved the wettability. A considerable


increase in the wetting time was observed after treatment with softening agents.
Rhee et al. [145] studied the effects of durable-press, stain-repellent, antistatic
finishes, and laundering on vertical wicking and demand wettability of fabrics. An
increment in vertical wicking rate for cotton fabric after treatment with durable
finishes [146] was attributed to the increase of total pore volume, although no noticeable
changes were observed in contact angles [145]. This indicated that there was no
significant difference in fiber wettability after durable-press treatment. After the
durable-press treatments, the water-uptake rate (from demand wettability test) and
amounts were reduced for fabrics treated with DMeDHEU and formaldehyde and
increased for one treated with DMDHEU, but the differences were not significant.
Vertical-wicking tests correlated better with pore-volume changes than the rates obtained
from the demand wettability test.
A study on stain-repellent finishing using fluorine-containing acids on cotton
showed that the shorter chains of CF2 and oxidation improved the vertical wicking
rate, whereas longer CF2 chains decreased the wicking rate due to increase in the
hydrophobicity of the fiber surface. Water uptake from demand wettabilty tests was
found to decrease after treatment but this was not statistically significant. Stain-
repellent finished polyester fabrics showed dramatic improvement in wicking rates
and demand wettability for finishes with longer chains and lower contact angles [145].
Laundering of cotton affects shrinkage and pore size. The wicking rate increases
after consecutive launderings due to reduction in contact angles as well as increased
pore volume. However, as the number of launderings increases from one to five to
ten, the wicking rate gradually decreases and after ten washings there is no significant
difference between unwashed and laundered cotton. In the case of polyester fabric,
wicking rate significantly increases after five and ten washes, due to the reduction in
contact angles. In the case of loose fabric, a significant increase in pore size could
occur during laundering leading to a higher wicking rate, wicking amount and rate of
demand wettability values after laundering [145].
The inclusion of a synthetic polymeric electrolyte in a disperse dye pad bath was
shown to decreases only slightly the rate of capillary wicking of the bath in a 50/50
intimately blended polyester–cotton fabric, but to bring about a significant reduction
of disperse dye migration. The significant reduction of migration is attributed to
flocculation of the dye to produce particles too large to move within the capillaries
of the fabric [212].
Multicomponent liquids such as surfactant solutions obey the Lucas-Washburn

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 73

equation [16, 17], where the wicking rate is proportional to the effective surface
tension at the meniscus (higher than the equilibrium surface tension of the surfactant
solution due to surfactant depletion from the advancing meniscus caused by adsorption
at the liquid–fiber interface [108]). The difference between the effective and equilibrium
surface tensions decreases as the surfactant concentration increases. When the surfactant
concentration is equal to CMC or above, convergence of the effective and equilibrium
surface tension is observed. Differences between wicking performances of various
surfactants can be explained by differences in both the extent of their adsorption at
the fiber surface and their diffusivity [108].
The moisture transport process in clothing under humidity transience is one of the
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

most important factors influencing the dynamic comfort of a wearer in practical wear
situations. Yi [213] studied the moisture diffusion into hygroscopic fabric process by
measuring the moisture take-up and temperature changes of fabrics made from wool,
cotton, porous acrylic, and polypropylene under humidity transience. The mechanisms
of the coupling effect between moisture diffusion and heat transfer depend on a
number of properties which are functions of water content, including moisture sorption
capacity, diameter, water vapor diffusion coefficient, density, and heat of sorption.

5. MATHEMATICAL MODELS OF WETTING AND WICKING IN


FIBERS AND FIBROUS ASSEMBLIES
Various mathematical models of wetting and wicking in fibers, filaments, yarns, and
fabric are discussed in the following sections.

5.1 Modeling of Fiber Wetting


Direct observation of contact angle of a liquid on a very thin filament (5µ) is very
difficult. Yamaki and Katayama [24] and Carroll [25] have developed the theory of
liquid drop shape attached to a monofilament under negligible gravity to compute the
contact angle, using Laplace’s equation:
1 + 1 = ∆P (38)
R⊥ RII γ
where the two radii of curvature RII and R⊥ are defined in Fig. 43. ∆P is the pressure
at the liquid–vapor interface and γ is the surface tension of the liquid.
In their approach [24, 25], the contact angle calculation comprises an analytical
expression relating droplet length, maximum drop radius, and the fiber radius to the
contact angle. The boundary conditions are: a tangent equaling zero at the maximum
drop height defined by the point A in Fig. 44, and tan θ slope at the drop foot defined
where the filament begins to be wetted, point B in Fig. 44. By introducing these
boundary conditions, the differential equation (39) describes the drop profile z (x) as
a function of the fiber radius x1, the drop height x2, and the contact angle θ.
–d z x –2 + na
= (39)
+ d x [( n 2 – x –2 )( x –2 – a 2 )]1/2
where the dimensionless parameters are defined as:
z = z/x1, x = x/x1, n = x2/x1, a = (n cos θ – 1)/(n – cos θ) and the fiber axis is in the
z-direction, as shown in Fig. 44.

© The Textile Institute


74 Textile Progress doi:10.1533/tepr.2006.0001

P⊥ RII
Oll
O⊥
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

R⊥

PII (D)
x z
ile
of
pr
op
Dr

Fig. 43 An axisymmetrical drop on a single fiber: localization of radii of curvature. The radii
of curvature are defined in the two orthogonal plans, one, PII, containing the drop profile along
the z-axis, the second P⊥ orthogonal to the z-axis contains the curve describing the revolution
around the z-axis [214 Rebouillat]

x
A

θ θ1
X1 B
z
o Fibre

X2 Liquid

Fig. 44 Geometrical parameters for the description of a drop on a single fiber. x1 is the fiber
radius, x2, the maximum drop height, θ, the contact angle, θi the inflection angle and L is the
drop length [214 Rebouillat]

The drop length and inflection angle can be calculated from equation (39), which
leads to ‘Length’ and ‘Inflection’ methods for contact angle calculation using the
drop profile measurements [214].
In the length method, equation (39) is integrated with boundaries from the drop
foot at the point B, where x = x1, to the drop height at the point A, where x = x2, and
the expression of the wetted length L can be found. The reduced drop length L =
(L/x1) is given by equation (40) as a function of the contact angle θ, the fiber radius
x1, and the drop height x2.
© The Textile Institute
doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 75

L = aF (ϕ , k ) + nE (ϕ , k ) (40)
1 1
2
where k and ϕ1 are defined as: k2 = 1 – (a2/n2), sin2 ϕ1 = (n2 – 1)/(n2k2). F(ϕ1, k) and
E (ϕ1, k) are elliptic integrals which are defined [214].
The contact angle calculation can be computed with equation (40), which is implicit
in terms of the contact angle θ. Therefore the calculation process must be an iterative
one. Measurements of the geometrical parameters, namely fiber radius x1, the drop
height x2, and the drop length L, are required. From an initial value of the contact
angle θ, the drop length can be calculated with equation (40) and the measurements
of x1 and x2. Then, the calculated value of the drop length can be compared to the
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

measured drop length. The calculation is repeated by adjusting the contact angle
value until the calculated and the measured drop lengths are equal [214].
Carroll [25] plotted the reduced drop length, L/x1, as a function of the reduced
drop height, n = x2 /x1, for increasing values of the contact angle up to 60°. The
reliability of the contact angle determination by the ‘length method’ depends obviously
on the accuracy and the reproducibility of the measurements of the geometrical
parameters. The measurements of the fiber radius x1, the drop height x2, and the drop
length L can be conducted from a photograph or can result from a computer image
analysis via an edge detection algorithm. In every case, the length method requires
fastidious and rather inaccurate localization of the drop extremities [214].
By deriving the differential equation (39), the expression of the inflection angle of
the drop profile can be obtained as a function of the contact angle θ, the fiber radius
x1, and the drop height x2. The inflection point is characterized by the maximum
slope of the drop profile z (x) defined by equation (39). This inflection point is given
for (d2 z /dx–2) = 0, which is equivalent to: x = 0 or x = ± na .
Then, for the non-zero solution, the inflection angle is given as function of n and
θ by:

( )
tan θ i = dx
dz x = na
= ± n–a
2 an
(41)

where

a = n cos θ – 1 and n = 2
x
n – cos θ x1
Yamaki and Katayama [24] provided a graph of the contact angle as a function of
the inflection angle and the drop length. But no attempt was made to further integrate
this approach in the study of the wetting phenomena since, in the author’s judgment,
the drop thickness and the inflection angle did not vary largely with the contact angle
θ. Robouillat et al. [214] proposed a new approach for the calculation of the contact
angle θ via θi, and n. Equation (41) can be squared, then its solutions leads to
equation (42). This equation theoretically yields two values for the parameter a as a
function of the inflection angle, θi, and the reduced drop height, n. Only the value of
a, smaller than unity, can be kept in equation (42) for the calculation of the contact
angle θi.

a = n(1 + 2 tan2 θ i ) ± n (1 + 2 tan 2 θ i ) 2 – 1 (42)

© The Textile Institute


76 Textile Progress doi:10.1533/tepr.2006.0001

The parameter a, defined to formulate equation (39), leads to the simple expression,
equation (43), of the contact angle θ as a function of a and the reduced thickness n.

 (1 – n 2 )( a 2 – 1) 
θ = A tan   (43)
 na + 1 
The inflection angle θ i, is measured via computer-aided image analysis. n is
calculated from the measurements of the fiber radius x1 and the drop height x2. Then,
the value of parameter ‘a’ is calculated with equation (42) and substituted in equation
(43) to compute the contact angle θ. The accuracy given by the inflection method is
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

better than that given by the length method [214].


Three distinct cases of shapes that occur when a droplet contacts a fiber are: (i)
film flow (not very common, usually broken into distinct droplets by Rayleigh
instability); (ii) a series of usually axisymmetric barrel-shaped droplets, commonly
connected by a film (in the order of nm); and (iii) axially asymmetric clamshell-
shaped droplets. Fig. 45 gives examples of the three distinct cases [184].

b θR
a θA

2h

(a) (b) (c)

Fig. 45 Example of the spectrum of fiber wetting: (a) film flow (axisymmetric) along a fiber
with no droplets present; (b) barrel-shaped droplets (generally axisymmetric) with connecting
film; (c) clamshell-shaped droplets of various sizes (radially asymmetric) with no connecting
film. The gray area represents the fiber and the film thickness has been exaggerated for reasons
of visibility [184 Mullins]

The most important forces involved in droplet motion on fibers are the drag force,
the gravitational force, and the change in surface tension induced by the change in
droplet profile as the fiber is angled. To determine the drag force (Fd) acting on a
droplet on a fiber or array of fibers in a parallel-sided channel, Stokes’ drag equation
with Faxen’s correction is used [215]:
–1
 3 4
b 
5
Fd = 6 π bUµ 1 – 1004   + 0.418   + 0.210   – 0.169   
b b b
 l l l l 
(44)
where l = 2h – a, and a is the fiber radius. 2h is the distance between fiber axes or the
distance between the fiber axis and the edge of the channel; 2l is the channel width

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 77

surrounding each fiber; b is the droplet radius (perpendicular to the fiber at the widest
point); U is the face velocity; and µ is the viscosity of the gas (Fig. 45). Faxen’s
correction applies only to the velocity in the channel and it assumes that the droplet
remains spherical in shape.
The gravitational force (Fg) acting on the droplet is:
Fg = mg = 4 πb3 ρLg (45)
3
where m is the mass of the droplet.
The adhesion force (A) of the droplet to the fiber can be given by:
A = (T cos θR – T cos θA) 2πra – mg = 0
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

at the point where the droplet flows down the fiber, where T is the tension per unit
length of contact line, and θR and θA are the contact angles on advance and retreat,
respectively (Fig. 45). The variation in interfacial tension with changing droplet
profiles (thus changing θR and θA) on differently angled fibers is far more difficult to
quantify, as the only suitable equations are in 2D, not 3D, and cannot readily be
converted into spherical coordinates [184].
Mullins et al. [184] considered the droplet on the fiber (Fig. 46), where gravity is
along the z-axis, airflow (hence drag) is along the x-axis, and the fiber is inclined at
θf = 3π/2 to 2π radians. Then the combined gravitational and drag forces along the
fiber (FQ) will be the sum of the component vectors of Fg and Fd:
FQ = cos (θg) Fg + cos (θd) Fd

z
Fd
x θ
f θd
θg
Droplet

Fg
A FQ

Fiber

Fig. 46 Diagram of forces acting on fiber/droplet and angles. Drag force (Fd) is parallel to the
x axis; gravitational force (Fg) is parallel to the z axis. θf, θg, and θd relate to the fiber, gravity,
and drag, respectively [184 Mullins]

This can be then maximized for FQ to determine the fiber angle (θMAX) where droplet
flow will be maximized:

F 
θ MAX = tan –1  d  + 3π
 Fg  2

with Fd and Fg as given in equation (44) and (45).

© The Textile Institute


78 Textile Progress doi:10.1533/tepr.2006.0001

The experimental results were compared to a theoretical model developed to


describe the behavior. The theory and experimental result showed good agreement.
The developed theory allows an optimum angle to be determined for the internal
filter fiber structure in the design of wet filters [184].

5.2 Modeling of Wetting of Fibrous Materials


Lukas et al. [216] described a method which applies the Ising model and Kawasaki
thermodynamics, combined with a Monte Carlo computer simulation technique to
study the liquid–fiber interaction and the wetting behavior of the fiber network. They
discussed the various interactions occurring in a liquid–fiber mixture and the energy
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

components associated with these interactions. Then a criterion of state change according
to Kawasaki dynamics [217] is established as a mechanism describing the liquid–
fiber wetting dynamics. Finally, the Monte Carlo stochastic approach is employed to
simulate the liquid wetting process.
The wetting of a fiber mass is quite different and unique compared to that of a
solid plane [216]. A fiber mass is heterogeneous, anisotropic, and not straight in
general, and ideally has a value of mean curvature equal to 1/2b where b is the fiber’s
radius, while the mean curvature for a plane is 0. The consequences of this lead to
different conditions for the complete wetting in these two cases.
Brochard [218] has dealt with the problem of single fiber wetting where a fiber of
radius b is covered by a layer of liquid film with the equilibrium thickness e0 (Fig.
47). He used the spreading coefficient S to express explicitly the aforementioned
differences.
S = γSV– γSl – γLV (46)

e0

Film
2b
Fiber

Fig. 47 A fiber of radius b covered by liquid film of equilibrium thickness e0 [216 Lukas,
218 Brochard]

Equation (46) is useful for understanding the liquid–fiber interactions during the
wetting process. For the complete wetting of the fiber to occurs, S > 0.
This condition for a fiber was eventually changed into another inequality:
S ≥ e0 lb
According to a two-dimensional Ising model, the mutual interaction energy of a
cell (Hi) or (Hj) are related to the two spin variables σi and σj which can take either
+1 or –1 as:

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 79

Hi = Hj = E (σi , σj)
This interaction energy is negative when both the spins have +1 or –1, the two
spins then attracting each other, but is positive when the spins have opposite signs
with a repulsion situation existing between them. Further, a spin σi may interact with
an external gravitational field with energy of:
Hg(i) = Eg (σi , yi)
where yi is the vertical coordinate of the cell i. The interaction energy is positive or
negative depending on the sign of the spin σi. Lukas et al. [216] considered the
problem of liquid–fiber mass wetting as a two dimensional regular and square lattice
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

of cells. For a two-dimensional case, the issue of fiber curvature becomes irrelevant
and the model will only exhibit the heterogeneous and anisotropic features of a fiber
network. Each cell is denoted by the index i and is occupied by either of the two
fluids (liquid or air). When the cell is filled by the liquid or air, it has a spin σi with
a value of +1 or –1 respectively. When incorporating the fiber network into the
system, the second kind of Ising spin variable Fi is defined. The cells occupied by
fibers have Fi = +1, or alternatively Fi = –1.
The spin variables σ and F can overlap in each lattice cell. During the computer
simulation process, the lattice cells can interchange their positions, reflecting the
movement of liquid during the wetting process whereas the fiber spins remain stationary
during this process. If this system with overlapped cells is placed in the field of
gravitation, then interaction of the cells with the gravitational field, interaction among
the fluid cells reflecting liquid surface tension and liquid–liquid cohesion, and the
adhesive interaction between the cells of the fluid and the fiber interface will occur.
The energy of the fluid spin in the gravitational field is given by:
Hg(i) = Gg σi yi
where Gg is a constant.
It is assumed that spin–spin interaction between the fluid cells occurs only among
cells within a certain proximity to each other. The assumptions for three types of
neighboring cells in terms of closeness, and an equation for the interaction energy for
a liquid cell are given. An equation governing the sum of the energy contribution of
overlapped cells by the fluid and fiber, the nearest fiber cells, and the next-nearest
neighbor fiber cells and the source of system wetting dynamics by the total Hamiltonian
energy system of the lattice model is presented. Considering the Ising model as a
thermodynamic system connected with a thermodynamic reservoir, energy exchange
can occur between them. The probability of spin exchange given by Boltzmann’s law
is presented. The treatment of fiber orientation in the system is described [216].
One result of the liquid wetting simulation in a fiber network obtained by Lukas
et al. [216] is shown Fig. 48. It is seen that at different times t of an artificial unit,
defined as the period for 1000 trials of spin exchange, the liquid is gradually wetting
the fiber network, and the liquid films on and between the fibers are observed in Fig.
48a. The height of the liquid front in the fiber network increases with times as shown
in Fig. 48b, where the numbers are in relative units.
Lukas et al. [216] carried out a computer simulation of the dynamic process of
spreading a liquid drop on a fiber, and the results are provided in Fig. 49, depicting
© The Textile Institute
80 Textile Progress doi:10.1533/tepr.2006.0001

t=2 t = 20 t = 40 t = 60
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

t = 100 t = 200 t = 300 t = 500


(a)
Height of liquid front
200

150

100

50

0 Time step
0 100 200 300 400 500
(b)

Fig. 48 The wetting simulation results in relative units [216 Lukas]: (a) wetting progress with
time in a fiber network; (b) change of the height of the liquid front during wetting

the changing shape of the liquid drop after seven different time periods. The Ising
model can describe a complex physical phenomenon in a very simple binary form,
yet can still be able to account for all the mechanisms involved and yield realistic
results. This makes the model a very attractive, practical and powerful tool to study
wetting phenomenon. Also, the constants associated with the energy terms have very
clear physical meanings, and they in fact represent the properties of the medium
involved.
Zhong et al. [219] also used the two-dimensional Ising model to simulate liquid

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 81

Time step 3 Time step 1000

Time step 102 Time step 3648

Time step 10000


Time step 247
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

Time step 548

Fig. 49 Simulation of the spreading of a liquid drop on a fiber [216 Lukas]

wetting of fibrous assemblies. The adhesive energy between fiber and liquid and the
cohesive energy within the liquid were calculated by applying the Lifshitz theory
under the assumption that the interactions between a fiber cell and a liquid cell and
between two liquid cells are dominated by Van der Waals forces. The work of extrinsic
force, the surface tension, was taken into account by its contribution to the energy of
the system. To test the validity of the model, wicking experiments were conducted
which involved a set of polypropylene filament yarns with the same yarn count but
different fiber finenesses. Computer simulations were also made to show the relationship
between the traveling height of the liquid front in the fibrous assemblies and the time
taken. Experimental and simulation results showed considerable accord.
Chen et al. [220] reported experimental observations on wetting of fiber mats for
composite manufacturing and explained the results with the aid of a simple air-
entrapment model based on the concept of two levels of porosity of fiber mats. A
simple model that includes liquid bypassing with initial air trapping, subsequent
capillary invasion of regular fiber bundles with air compression, and finally mobilization,
is proposed to explain air-entrapment phenomena. The simple model successfully
rationalized the observed air trapping and compression during initial liquid–fiber
contact.

5.3 Modeling of Wicking in Yarns


Chen et al. [122] developed a mathematical model to describe wicking kinetics.
When a droplet of wetting liquid is deposited on a yarn, it spontaneously wicks into
the yarn due to the capillary forces associated with the given structure and geometry
of the void spaces between the filaments. The model describes the droplet disappearance
just after the yarn section underneath the droplet has been saturated with the liquid.
When the droplet size is sufficiently small, the effect of gravity on the droplet shape
is negligible and the droplet may be considered axisymmetric. Quantitatively, the
approximation is valid if the characteristic length of the capillary wave l cap = γ ρL g
is larger than the droplet radius Rd. Typically, parameter lcap is in the order of a
millimeter, so the smaller droplets were the objectives of their study. Pressure difference
between the droplet Pd and the liquid front Pf causes movement of the liquid front

© The Textile Institute


82 Textile Progress doi:10.1533/tepr.2006.0001

along the yarn. It is assumed that the droplet takes an equilibrium shape at each
instant of time, since the process of liquid imbibition is slow. Then the pressure can
be expressed by means of the droplet curvature 2H and the hydraulic radius of the
yarn pores R as:
Pd = Pg + (2H)γ
Pf = Pg – (2γ cos θa)/R
where Pg is the atmospheric pressure. Applying Darcy’s law for the flow rate:

dL = k ( Pd – Pf ) = 2 kγ ( H + cos θ a / R )
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

(47)
dt ηL ηL
where k is the permeability of the yarn. Equation (47) must be complemented by both
a condition of mass conservation and an expression for the droplet curvature [122].
Based on equation (47), the condition of droplet wicking in a yarn is dL/dt > 0.
Expressing the droplet curvature through n and cos θ, the condition for droplet
wicking is:
n – cos θ + α > 0 (48)
n2 – 1
As seen from equation (48), the model predicts that wicking can still occur if the
apparent contact angle θa is greater than 90°, yet n is sufficiently small [122].

5.4 Modeling of Wicking in Fibrous Structures


Zhong et al. [221] used a two-dimensional Ising model to study in-plane fluid flow
through fibrous structures under pressure. For wicking, they considered the Hamiltonian
energy system and the work of surface tension. To construct the energy of the system,
additionally, the work of applied pressure to the internal energy of the system and the
mechanical energy were taken into account. The internal energy of the system, E, is:
E = H + Eγ – pV (49)
where H is the Hamiltonian energy of the system, Eγ is the work done by the interfacial
tension, p is the pressure applied to the system, and V is the fluid volume.
A fluid flowing through a fibrous structure under an applied pressure exhibits a
substantial frictional loss which is expressed in thermodynamic variables such as
pressure. Assuming a quasi-equilibrium process, Kawasaki’s dynamics [216] are
extended to the case of a fluid flowing under driving pressure. For a streamline flow,
the modified Bernoulli equation gives the mechanical energy of the system I as:

p 2 p u2
I=y+ + u = y0 + 0 + 0 – δ (50)
ρL g 2 g ρL g 2 g
where y, p, and u are the height, static pressure, and velocity of the fluid in an
arbitrary position during flowing; y0 , p0 , and u0 are the height, static pressure, and
velocity of the fluid at the inlet, and δ is the frictional loss per unit weight of the fluid.
For an in-plane flow, y = y0. The fluid flow is very slow and, neglecting the
friction due to container wall, the entire loss term can be given by:

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 83

Wδ = ρLgVδ′ = α0(p0 – p)V (51)


The term α0 is a coefficient that denotes the extent of the resistance of the fibrous
structure to the flowing fluid, which ranges between 0 and 1. In an extreme case
when α0 = 0 (i.e. no fibrous structures in the flow path), there is no resistance.
Therefore, the total energy of the system is the summation of the total internal energy
and the total mechanical energy:
ET = H + Eγ + ρLgVy0 + α(p0 – p)V (52)
where α is the structural coefficient, equal to (1 – α0).
The pressure terms in equation (52) are further expounded: If the pressure at the
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

inlet is p0, and that at the flow front is atmospheric pa at all times, for an in-plane
radial flow with a constant influx U, the pressure gradient along the flow direction,
according to Darcy’s law is:
U = – K dp (53)
2 π rd 0
dr
where r is the distance of any point in the flow field to the center of the inlet, d is the
thickness of the fibrous structure, and K0 is the permeability. Assuming

K1 = U and integrating equation (53) gives


2π d K 0

K1 ln =  r  = p 0 – p (54)
 r0 
where r0 is the radius of the inlet. The pressure at the flow front, Pa, which approximates
to a circle with radius ra, is expressed by:

K1 ln  a  = p 0 – p a
r
(55)
 r0 
Then,
α ln (r / r0 )
E T = H + Eγ + ρL gVy 0 + ( p0 – pa )V (56)
ln ( ra / r0 )
and
dET = dH + dEγ + dEP (57)
where EP denotes the last term in equation (56).
The difference of the total energy dET after and before two cells in the lattice
exchange their states, was used in a Monte Carlo simulation to predict the penetration
process of fluids in fibrous structures. The simulation procedure used was slightly
different from the earlier one [219], where each Monte Carlo step represented 1
second. Further, two variables n and t were defined, n to denote the net increase in the
fluid cells in the whole system during simulation when the fluid is driven by the
pressure flowing through the fibrous structure. When n reaches a certain value, which
indicates a certain amount of net increase in the influx volume of the fluid, the value
of the time variable t is increased by 1, representing the elapse of 1 second.

© The Textile Institute


84 Textile Progress doi:10.1533/tepr.2006.0001

The predicted values of the flow front were compared with the experimental
results. Good agreement was found between the experimental and simulation results,
thus indicating the applicability of this model in predicting the behavior of in-plane
fluid flow through porous media under pressure. One of the most attractive advantages
of Ising’s model for the study of in-plane fluid flow through fibrous structures is that
the burden of computation does not become an aggravation, no matter how intricate
the geometry of the interface between fluid and media. This also contributes to the
model’s ability to reveal the behavior of fluids flowing in fibrous structures in greater
detail. And the model can be easily adapted to other systems composed of different
fluids and media by altering relevant parameters [221].
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

Wiener and Dejlová [222] proposed a model based on liquid wicking into fiber
bundles. The equilibrium of the wicking process was not sufficiently elaborated, but
at the same time a series of phenomena was found on the equilibrium, which from the
practical standpoint influences the more interesting dynamics of the process. Most
authors understand the equilibrium of the process only as a physical state without
considering the peculiarities related to such a state [222]. Important factors affecting
wicking in textile structures are frequently neglected, and then the results are affected
by useless error [115]. The relationships accepted for capillaries are commonly used
for analyzing the wicking equilibrium. The essential difference between the classical
capillary and fabric is that the capillary is a ‘closed’ and the textile fabric an ‘open’
capillary system. The liquid does not flow out in all directions from the capillary;
solid walls oppose it. There are no firm walls in the textile fabric, the liquid in the
system is held by its surface tension.
The proposed model is based on a simplified description of the thread structure,
and works with the textile description of its structure [222]. The following textile
parameters are included in the model: fiber fineness, number of fibers at the cross-
section of thread, fiber shape factor, and the filling. The formation originated by
wicking liquid into the longitudinal textile was described in detail; the separate
phenomena have been discussed. Important parameters are used in the wicking model.
Parameters with very small influence on normal threads are neglected. The proposed
model of wicking allows the functional dependence of suction height on the thread
parameters to be expressed in an analytical form. The influence of the number of
fibers in the cross-section of the bundle was tested experimentally. The influence of
the tested parameter was analyzed by well-arranged linearizations. This work could
form the basic study for a future system that would allow prediction of the wetting of
multidimensional textile formations under real conditions.
The importance of fiber orientation characteristics in nonwovens, which directly
influence the in-plane fluid flow in nonwovens, was explored by Kim and Pourdeyhimi
[223]. The framework was developed by Mao and Russell [224, 225]. For completeness
Kim and Pourdeyhimi [223] presented a summary below, and demonstrated how it
might be extended to determine spreading. The drag force per unit length acting on
a single fiber surrounded by similar fibers, all oriented along the direction of flow
can be deduced:
8πη q
fp = (58)
S

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 85

where
S = [2 ln φ – 4φ + 3 + φ2]
φ is the volume fraction of solid materials, and q is the superficial flow velocity of the
fluid stream.
The drag force per unit length acting on a single fiber oriented perpendicular to the
fluid flow is give by:
8 πη q
fP = (59)
T
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

where

 1 – φ2 
T =  ln φ +
 1 + φ 2 
The drag force acting on the fiber in the direction θ can be obtained by considering
the contribution of each of the components of the drag force and flow velocity in the
equations (58) and (59):
8πη q cos 2 (θ – α ) 8πη q sin 2 (θ – α )
f (θ ) = + (60)
S T
where α is the fluid flow direction over the fiber.
The total drag force on the material is equal to the sum of the individual drag
forces acting on each fiber. Hence, for a flow in a particular direction, say θ, the drag
force acting on all the fibers, where the unit volume of the fabric consists of n fibers,
will be:
π
f ′ (θ ) = n
∫0
Ω (α ) fα (θ ) dα (61)

where n is the number of fibers of unit length per unit volume and Ω(α) is fiber
orientation function.
The fiber orientation distribution Ω is a function of the angle θ. The integral of the
function Ω from an angle θ1 to θ2 is equal to the probability that a fiber will have an
orientation between the angles θ1 to θ2. The function Ω must satisfy the following
conditions:
Ω(θ + π) = Ω(θ) (62)
π

∫0
Ω (θ ) dθ = 1

The pressure gradient in the flow direction due to the drag force in a unit volume
of the fabric is equal to the pressure drop per unit length of flow resulting from the
drag force in this direction. Thus,
∆P = f ′(θ ) (63)
L
The established theory of laminar fluid flow through homogenous porous materials
is based on Darcy’s law, described as:
© The Textile Institute
86 Textile Progress doi:10.1533/tepr.2006.0001

∆P
q= – k (64)
η L
Thus, by substituting from equation (60) into Darcy’s law, the directional permeability
is given by [224, 225]:

2  
ST
k (θ ) = – 1 δ  π 

(65)
32 φ  Ω(α ){T cos 2 (θ – α ) + S sin 2 (θ – α )}dα 
 0 
The hydraulic radius theory treats the flow through a porous medium as a conduit
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

flow. The equivalent hydraulic capillary diameter, DH, in the direction θ in the nonwoven
fabrics will be:
δ (1 – φ )
DH = π
(66)
φ
∫ 0
Ω(α )/cos (θ – α )/dα

The magnitude of the capillary pressure, p, is commonly given by the Laplace


equation below:
2σ cos µ
P= – (67)
RC
where RC is the capillary radius, σ is the surface tension, and µ is the contact angle.
According to equation (66), the capillary pressure in the direction θ in the fabric
will be:
π

p(θ ) =


0
Ω (α )/cos (θ – α )/dα
σ cos µ (68)
δ (1 – φ )
Substituting k (θ) and p (θ) into one dimensional Darcy’s law, the rate of absorption
in the direction θ [224, 225]:

 ST 
1 δ2  π .
V (θ ) = –
32 φ 
 ∫0
Ω(α ){T cos 2 (θ – α ) + S sin 2 (θ – α )} dα 

π

×

∫0
Ω (α )/cos (θ – α )/dα
σ cos µ 1 (69)
δ (1 – φ ) ηL
From the rate of absorption, equation (70) is derived below for predicting the
spreading length:

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 87

 2 
ST 
– 2 δ  π .
 32 φ 
  ∫ 0
Ω(α ){T cos (θ – α ) + S sin (θ – α )} dα 
2 2

L (θ ) =
π

×

∫0
Ω (α )/cos (θ – α )/dα  .t
σ cos µ 1  (70)
δ (1 – φ ) η

The results from experimental and simulation are compared. A reasonable agreement
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

has been found between the two. The experimental data indicate that the spreading
rate is faster than the one predicted by simulation. This may be in part due to surface
finish that may be present on the surface of the fibers. Additionally, this difference
may be partly because of the slight differences between the simulation and the actual
experimental set up. The opening in the instrument plate connected to the tube is
about 1 cm diameter in comparison with that of the simulation, which is assumed to
be a point. Regardless of this however, the anisotropy of the flow is predicted correctly
[223].
Rajagopalan et al. [226] have discussed a mathematical model of capillary flow in
complex geometries representative of the void spaces formed between fibers in a
textile yarn. Moisture transport in textile yarns and fabrics is an important factor
affecting physiological comfort. Rajagopalan et al. [226] have extended an existing
analytical model for capillary flow in circular tubes to more complex geometries and
they have also validated this model using detailed computational fluid dynamic
simulations of this flow. These models are used to understand the effect of geometric
and material parameters on moisture transport. In vertical wicking in a bundle of
filaments, the model predicts that as the non-roundness of the filaments increases, or
the void area between the filaments decreases, the maximum liquid height increases
while the initial rate of penetration decreases.
A numerical model has been developed for simulating the heat and mass transfer
in fabrics during the wicking process. The model was applied to two different fabrics,
cotton and polypropylene [227]. The model showed that, as the water is wicked
through the fabric specimens, two temperature zones are formed. The structure of the
wet fabrics consists of the fiber matrix, liquid water, and a gaseous phase of water
vapor and air. Predicting the heat and mass transfer in such a system is a difficult task.
The problem is complicated by the fact that the transport properties involved are non-
linear functions of temperature and moisture content, and that fiber swelling may
occur due to changes in the degree of the saturation. This swelling may affect the
porosity and the available void space for the liquid and the gaseous mixture transport
inside the solid matrix. Because of the complexity of the problem, the following
approximations are imposed on the transport process.
(i) The fibrous system is represented by an ideal continuum medium and is divided
into volume fractions of liquid, solid, and liquid–vapor and air mixture at a
particular location. The inhomogeneity scale of the three-phase, solid, liquid,
and gas system is far smaller than the characteristic length over which an
appreciable change in moisture content occurs.
© The Textile Institute
88 Textile Progress doi:10.1533/tepr.2006.0001

(ii) The solid phase is considered to be non-deformable. Based on this assumption,


mechanical swelling and shrinkage of the solid phase are not included. Some
fibers, particularly natural materials, do swell. However, in actual fabrics the
fiber is usually a small portion of the total volume. Therefore, any swelling that
may occur has a relatively small effect on the void space for moisture movement.
Also, experimental data for permeability will reflect the effects of swelling.
(iii) Darcy’s equation is used to describe the transient flow through the saturated
and unsaturated regions in the fibrous medium, since the liquid is flowing at a
very low Reynolds number.
(iv) There is no convective flow of the vapor–air mixture.
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

(v) The liquid component is incompressible and the vapor–air mixture is considered
to be an ideal gas.
(vi) The fiber system, liquid, and gaseous phases of the fibrous media are in local
thermodynamic equilibrium.
(vii) Compressional work, viscous dissipation, and diffusive energy are considered
negligible.
(viii) The mass concentration of water vapor is small compared to that of liquid water
[227].
According to Ghali et al. [227], energy is transported in four different forms in the
wicking process: (i) conduction; (ii) diffusion of moisture in the plane of the fabric;
(iii) convection of moisture in the plane of the fabric; and (iv) evaporation of moisture
to the atmosphere surrounding the fabric. Each form is addressed in the following
with respect to the experiments conducted in their study. The first form is very small,
except at the boundary condition where the fabric is in contact with the water, because
of the low conductivity of the water and fabric. The second form is negligible because
of the small temperature changes that exist during the wicking process; the temperature
is nearly constant throughout the fabric except at the wetting front where there is a
change of about 2°C. The third form is also negligible because the water is transported
at a very low Reynolds number, except in the initial stage of wetting where the
capillary pressure is very high. Finally, the last form is the most dominant. The
numerical model is capable of simulating the temperature distribution of the fabric
specimen during the wicking process and it can predict the temperature front. In the
initial stage of wetting, the convective energy is high and the water at the beaker
temperature is wicked very rapidly into the fabric. As wicking proceeds, the degree
of saturation increases, which causes a decrease in the capillary pressure and thus a
decrease in the flow rate and the resulting convective energy. When the degree of
saturation becomes high enough, the mass transfer coefficient will attain a significant
value, causing evaporation of some of the moisture to the atmosphere. This evaporative
energy becomes the dominant energy effect, causing the temperature of the fabric to
decrease except in the region close to the wetted boundary, where conduction and
convection liquid are significant [227].
A model for the transverse (normal) impregnation of viscous liquids into a regular
hexagonally packed fiber lattice has been developed [228]. The calculation takes into
account the continuous and discontinuous changes in the capillary pressure as the
contact line penetrates into the fiber matrix. The rates of impregnation are calculated
as a function of the liquid contact angle, viscosity, and interfacial tension, as well as
© The Textile Institute
doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 89

the fiber radius and the fiber separation. The model predicts normal impregnation
rates which are up to two orders of magnitude slower as compared with axial
impregnation in the same fiber matrix.
The problems of mathematically modeling fluid spreading in composite web
structures using mass transfer equations are discussed by Yarlagadda and Yoganathan
[229]. An analogous three-dimensional transient heat transfer model with varying
thermal conductivities and heat capacities is used to represent the actual phenomenon
of fluid spreading in a composite web. A web structure with varying porosities in the
layers and pores with certain preferred orientations is considered, and the dynamics
of fluid spread are modeled. The results predicted by the model match well with
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

previous experimental data [229].


Letelier et al. [130] presented a mathematical analysis of the phenomenon of
capillary penetration. The analysis is based on the assumption of laminar parallel
flow of an incompressible Newtonian fluid in a rigid circular conduit, and constant
angle of contact that leads to a nonlinear differential equation requiring numerical
integration. Solutions depend on a single non-dimensional parameter (Ω). For Ω = 0,
the solution reduces to the classical Washburn [17] equation which is shown to be
limited in its utility to inertialess penetration processes.

6. APPLICATION AREAS
Wetting and wicking of fibrous materials is an important issue in a range of areas
including textile and composite manufacturing. Examples of textile applications are
sports clothes, hygiene disposable materials, medical products, and geotextiles, where
the wetting and wicking behavior of the fibrous structures is a critical aspect of the
performance of the products. In addition, wetting and wicking process occurring
especially during dyeing, finishing, and the wearing of clothes all have a practical
significance in controlling the quality of the textile processing and clothing comfort.
Another example where the wetting process is considered extremely important is in
the application of fibrous filters, where wetting of the fiber surface is the key mechanism
for the separation of two different liquids from their mixture; as an instance, separation
of oil from sea water during a cleaning process after an oil spillage.

6.1 Comfort
Engineered or structured fabrics have become popular in many commercial applications.
Moisture management fabrics are one such example. In applications such as active
sportswear, exercise garments, work clothing, intimate apparel, and footwear, the
concept of moisture management is utilized to prevent or minimize the collection of
liquids on the skin of the wearer due to perspiration. This is done by quickly wicking
or diffusing the liquid through a hydrophobic fiber inner layer to an outer hydrophilic
layer and then evaporated it to the atmosphere. Moisture management fabrics are
two-sided fabrics produced from a minimum of two yarns using weaving, or warp or
circular knitting machines. A few patents and papers in the area of comfort related to
liquid and moisture transports are reviewed in the following:
A composite textile fabric has been described by Yeh [230] for wicking moisture
away from the skin of the wearer and transporting it to the outer surface. The inner
layer fabric is made from relatively high denier yarn and is rendered hydrophobic so

© The Textile Institute


90 Textile Progress doi:10.1533/tepr.2006.0001

that it quickly wicks perspiration or body fluids from the skin. The outer layer of the
fabric is hydrophilic, made from low denier filaments and is processed by peach
finish sueding. The low denier filaments in the outer layer pull perspiration and other
body fluids from the inner fabric layer.
A continuous process of making tubular knitted arm warmer comprising arm
portion, wrist cuff, and a binder portion joining the arm portion is reported in which
the arm warmer is made from nylon, spandex, and wicking yarns [231]. The wicking
yarn is plaited to the inner surface whereas the spandex-covered nylon yarn is plaited
to the outer surface of the arm warmer.
A composite yarn made using air-jet texturing for use as a single-layer moisture
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

management fabric is reported [232]. Hydrophobic filaments are fed to the air jet at
higher overfeed than that of hydrophilic filaments. The preferred hydrophilic filaments
are modified nylon having high moisture regain such as ‘Hydrofil’, ‘Quup’ and
‘Hygra’. It is envisaged that a two-layered fabric with an inner layer made from
hydrophobic yarns and an outer layer constructed from this composite yarn or only
this composite yarn can be used as a single-layer fabric for moisture management
applications. The disadvantages of using two-layered moisture management fabrics
are listed. The advantage of this composite yarn claimed over that of a previously
patented composite yarn (a hydrophilic core yarn tightly wrapped by hydrophobic
sheath fibers) is the larger interfacial area between both types of filaments because of
mixing and migration of filaments. This increases the kinetics of absorption so that
moisture transfer becomes more rapid and effective in the moisture-management
fabrics.
A process of rendering polyester yarn hydrophilic during dyeing by adding
hydrophilic agents in the dye bath is claimed for wicking away moisture from the
body [233].
A sock for managing moisture and reducing friction includes a mohair wool yarn
contacting the skin and an acrylic yarn away from the skin. Hydrophobic mohair
wool wicks moisture away from the foot of a wearer, and hydrophilic acrylic yarn
absorbs moisture wicked from the foot by the mohair yarn and holds it away from the
skin. As a result, the undesirable friction-promoting effects of moisture on a foot are
decreased. A blend of mohair wool and acrylic is placed in areas of the sock likely to
have the greatest moisture build-up and the most friction of the foot against the sock
shoe [234]. Another sock construction for moisture-distribution, wicking, evaporation,
and other phases of control has been patented. This moisture-management sock has
the toe and heel portions knitted predominantly of hydrophilic yarn, and in between
them alternating rings of hydrophobic and hydrophilic yarns. Moisture absorbed
from the wearer’s foot by the hydrophilic yarn is transferred by wicking into the
hydrophobic rings and then to the leg portion to be evaporated [235].
Acrylic and olefin socks do not absorb moisture but have better wicking ability.
These socks are often worn next to the skin, with a wool or cotton sock over them to
absorb moisture. Some blends of socks are made so that the fibers with wicking
ability are next to the skin and the absorbent fiber forms the outer layer. This
accomplishes the same result as wearing double socks, but is less bulky [236].
Anand and Higgins [161] discussed the merits and demerits of various fibers
(natural and synthetic), yarns, and fabrics which can be used in activewear and

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 91

sportswear, where wicking plays a key role for clothing comfort. Fabrics made from
natural fibers such as cotton are considered comfortable for use under normal conditions;
their properties make them unsuitable for use during strenuous activity. Cotton fibers
absorb high levels of moisture, leading to a feeling of wetness and cling. It has a very
slow wicking rate from inner fabric to outer fabrics. This makes cotton fibers unsuitable
for use against the skin during strenuous activity, as in the case of sportswear. Wool
fibers have good wicking ability, but these fibers are slow to dry and have high wet
surface coefficient of friction, leading to potential skin abrasions [161, 196]. Silk has
good wicking ability and can absorb water up to one third of its weight without
feeling wet, but the easy-care characteristics of silk are poor, which is a disadvantage
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

for sportswear that is worn regularly. The irregular surface of viscose fibers contributes
to comfort when worn next to the skin, but their laundering characteristics are poor,
making them less suitable for use in strenuous activity [161].
Polyester fibers, which have faster wicking rates than cotton, are widely used in
activewear and sportswear. Where polyester is used against the skin, a hydrophilic
coating is applied to the polyester filaments to improve their wicking ability. An
alternative method is by changing the surface chemistry of polyester filaments by
molecular modifications. Nylon 6 and nylon 66 have higher moisture absorption
capabilities than polyester, have better wicking ability, but have a slower drying rate.
They are mainly used in swimwear or cycling wear, or as reinforcing fibers in blends
in sports socks. Polypropylene fibers have very low moisture absorbency but excellent
moisture vapor permeability and wicking capabilities. Liquid perspirations are
transported away from the skin without being absorbed, making them ideal for
sportswear. As polypropylene does not become wet, its thermal insulation is retained
during and after strenuous activity. Further, it is lightweight and does not produce
static. However, soil removal is difficult and fabrics made from polypropylene may
shrink if washed at high temperatures [161].
Some specialized synthetic fibers are available to improve thermophysiological
and sensorial properties, for examples microfibers with linear densities of much less
than 1 dtex per filament. These fibers enable very dense fabrics to be made in which
the fiber surface area is significantly increased and the spaces between fibers are
reduced and hence, an increased capillary action leads to better thermoregulation.
The introduction of voids within polyester fibers has been used to improve the
wicking and thermal resistance. Welkey by Teijin Ltd is a fiber, used in sportswear,
with a hollow core and a proliferation of smaller holes throughout the body of the
fiber, thus increasing capillary action and wicking away sweat from skin [161].
Fabrics made from synthetic fiber are better than those of wool or cotton for
sportswear because the moisture pick-up is nil but the vapor molecules pass quickly
to the atmosphere through the interstices, and through the layer [196]. Knitted fabrics
made from polyester/wool blends or polypropylene/wool blends can improve wicking
and thermal insulation properties within a single layer of fabric structures. One such
example is the DriRelease yarns made by Optimer, which are made from a blend of
85–90% of polyester, and 10–15% of cotton. The resultant fabrics have a soft handle,
are wettable and wick but have low absorbency. Sportwool is an example of a two-
layer sportswear used in many items of sports clothing, having good moisture
management properties. The inner layer, which is next to skin, consists of extremely

© The Textile Institute


92 Textile Progress doi:10.1533/tepr.2006.0001

fine Merino wool fibers (<20 µm) that are chemically modified. The outer layer is
made from 100% polyester filament yarn. The wool fiber has good water vapor
permeability, which transfers heat and moisture from the skin to the outer surface
where heat and moisture escape, aided by wind speed and body movement [161].
Special engineered polyester fibers have been developed for specific sports activities.
Coolmax fabrics developed by Invista/Dupont for warmer weather possess superior
wicking and evaporation of perspiration moisture [236, 237]. The double scallop
shaped fibers used to make these fabrics have 20% more surface area than regular
fibers so that they allow water to spread over a greater area in the fabric for faster
evaporation (Fig. 50a). Many other brand names in this market are Dryline, Dryfit,
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

Powerdry, Hydrofil, Capeline, BiPolar, and Stretch Supplex. For colder weather, fabrics
such as Thermolite have been developed, for use in thermal underwear, gloves, socks,
and turtleneck T-shirts, which contain hollow core fibers with convoluted or twisted
surfaces [238]. These enhance wicking properties (Fig. 50b).

Skin Insulating layer of warm air


©
The CoolMax Comfort System butween fabric and skin

Skin CoolMax©
Fabric Thermolite©
Base
Air Fabric
Moisture

Entrapped
insulating air

Moisture
Close-up of four-
channel fiber used
in CoolMax© fabric Close-up of
Channels speed hollow-care
moisture to outer fibers used in
surface.
Thermolite©
Base fabric
(a)
(b)

Fig. 50 (a) Close-up of four-channel fiber; (b) Close-up of hollow core fibers
[236 http://ohioline.osu.edu 237 http://coolmax.invista.com]

Blends of Teflon and Coolmax fibers are used in areas of sports socks most prone
to friction and chaffing. Pieffe Sports, an Italian manufacturer of sports socks, says
its latest product is engineered for both professional and amateur cyclists. The sock
contains a polytetrafluoroethylene (PTFE) fiber, which reduces friction as well as
keeping the wearer’s feet cool [238]. The DuPont Coolmax fiber is designed to
transport moisture away from the skin and stimulate evaporation [236–237].
4DG fibers are made from polyethylene terephthalate and other polymers, with an
eight legged cross-sectional shape (Fig. 51) [239]. Because of their unique grooved
shape, these fibers spontaneously move fluids, store and trap substances, and the

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 93
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

Fig. 51 4DG fibers cross-sections [239 www.clemson.edu]

large surface area per unit volume provides more bulk and cover than traditional
round fibers. The diverse characteristics of such capillary surface materials have
potential in a wide area of end-use applications including comfort, medical hygiene,
filtration, and industrial [239–241].
An ideal sportswear is a multi-layer structure [242, 243]. According to D’Silva
and Anand [242], in a two-layer structure, the layer close to the skin is of the wicking
type comprising synthetic fibers, e.g. micro-denier polyester, and the outer layer
usually is cotton or rayon that absorbs and evaporates. The three-layer structure has
a middle layer that wicks and spreads perspiration. Micro-denier polyester is ideal for
wicking perspiration away from the skin. Cellulose fibers assist in spreading and
evaporating the moisture vapor. Fabric structures, knit, and appropriate finishes can
influence the absorption and evaporation of moisture.
Ishtiaque [243] has shown that, with cotton as the upper layer and polyester or
acrylic as lower layer, the moisture in both layers is nearly the same after 30 minutes.
The use of superfine or microfiber yarn enables production of dense fabrics leading
to capillary action that gives the best wicking properties.
Several other substitutes for polyester can be considered for use in the bottom
wicking layer of sportswear. They include Acrysorb, a hydrophilic microporous fiber
developed by conjugate spinning and blending with cellulose acetate and acrylic
microfibers, namely Myoliss and Leacryl Micro. The outer knitted hydrophilic layer
of the twin layer sportswear can be of cotton or viscose rayon bended with a small
proportion of spandex fiber. The two layers of sportswear can be inter-knitted with
spandex filament or glued with polyurethane copolymer to produce highly comfortable
form-fitting sportswear [244]. This is illustrated in Fig. 52.

U
M
B
B = Bottom layer next to skin (Soft and
high wicking action)
M = Middle Layer (Highly absorbant)
U = Upper layer (High spreading action
for rapid evaporation of sweat)

Fig. 52 Multilayer sportswear for comfort [244 Bardhan]

© The Textile Institute


94 Textile Progress doi:10.1533/tepr.2006.0001

According to Brownless et al. [245], thermo physiological comfort entails both


thermal insulation and moisture management and can be argued to be the most
influential comfort feature. Achieving true comfort requires the optimization, not the
maximization, of these two properties. A series of fabrics has been identified as
responsive or ‘dynamic’. A ‘dynamic’ fabric is one that is cool when the wearer is
perspiring, aiding in maximum heat loss, with increasing thermal insulation once
perspiration has stopped. These fabrics can be used for sports and activewear.
The development of moisture vapor transmission (MVT) technology has been
reviewed, requiring control of the personal microclimate for athletics and active
leisure activities. This applies to cold/wet and warm/humid conditions. The function
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

of wicking in fabrics is explained with two caveats that apply [246].


A cover for a wheelchair seat discussed by Adams [247] is made from a cloth
having a first ply of wicking material for drawing moisture away from an occupant
of the wheelchair and a second ply of absorbent material for drawing moisture out of
the first ply and dispersing it. A third ply of impermeable material is to prevent
moisture from passing from the second ply into contact with the wheelchair
surface.
An article describes a transport moisture management system in a diverse number
of high-performance applications, including thermal underwear, socks, shirts, and
liners for bullet resistant vests [248]. The transport treatment enhances the inherent
hydrophobic properties of polyolefins, hence giving treated polyolefin fabrics up to
seven times less moisture absorption than standard polyolefin fabrics. Vapor and
moisture exit through the pores or spaces in the fabric into the next fabric layer or the
air. It is already being used in thermal underwear, body armor, and socks, with its
potential being almost limitless.

6.2 Medical and Hygiene Products


Many healthcare and personal hygiene products need to control liquid flow in terms
of both liquid capacity and flow rate factors, appropriate to the specific end use.
Incorporating effective capillary systems into the end product enables controlled
flow of liquid. Hygiene products comprise distinct layers of materials having appropriate
characteristics so that liquid or moisture drains quickly away from the inner surface
of material in contact with the body of the wearer. Liquid is retained in an outer layer.
A multi layered wound dressing is reported comprising an adsorbent layer having
high absorbency but low lateral wicking rate, and a transmission layer having a high
moisture vapor transmission rate overlaying the side of said absorbent layer furthest
from the wound [249].
An odor-absorbing wound dressing for use on infected wounds, especially malodorous
wounds such as infected ulcers, was disclosed in a US Patent [250]. The absorbent
fibrous layer can comprise any biodegradable fiber, preferably highly absorbent fibers
such as Oasis, modified cellulosic fibers or alginate fibers. This layer is placed
closest to the wound, in contact with the wound and surrounding skin. The barrier
layer retards movement of exudates to the odor-absorbing layer while permitting the
movement of odor. This so-called ‘one way wicking’ layer is perforated with small
cones (micro-funnels), which allow the passage of odors but deter the transfer of
exudates.

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 95

Absorbent laminates enhance wicking in hygiene and medical products. The laminate
described in another US patent [251] consists of a fluid-permeable first layer formed
from a hydrophobic fibrous material, an absorbent hydrophilic second layer, and
fibers that extend as wicks from within the first layer into the second layer. A pre-
needled nonwoven fleece made from a polyolefin, a vinyl polymer, a polyamide or
a polyester, and possibly hydrophilic fibers rendered hydrophobic by surface treatment,
performs best. Materials such as fluff pulp, peat moss board, swellable acrylate
polymers or cellulosic polymers are suitable for the hydrophilic second layer. The
wicks may be formed by stitch bonding the two layers together or by hydro entanglement,
using very fine, high-pressure water jets. However, better results are achieved when
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

the fibers of the first layer are used to form the wicks by needling the two layers
together. The density of the needling is chosen to give the strength and fluid transport
properties desired in the finished laminate.
By selecting a precursor laminated nonwoven web with certain properties and
post-drawing the web under certain conditions, the synthetic fibers of the web are
restructured to provide the laminated web with unique measures of pore size, directional
absorption, wicking, liquid absorption capacity, and other properties. These attributes
make its products particularly suitable for medical applications such as protective
apparel, face masks, diapers or sanitary napkin parts, wound dressings, respirators,
wipes, chemical reservoirs, wicks, and surgical drapes [252].
Biodisintegratable nonwoven materials having improved fluid management properties
have been disclosed [253, 254]. These materials demonstrate a higher contact angle
hysteresis, quicker intake times, high wetting rates, and improved skin dryness as
compared to other nonwoven materials. The nonwoven material may be produced
using thermoplastic compositions with polyolefin microfibers as a discontinuous
phase encased within an aliphatic polyester. These nonwoven materials may be used
in a disposable absorbent product for the absorption of body fluids [253].
A composite absorbent structure and method are discussed to provide preferred
liquid transport and liquid retention properties. The first wicking layers are in a
preferred contact with the second liquid retention layers made from hydro-gel forming
material [255].
A facing fabric for reusable incontinence products has been described in a US
Patent [256]. The fabric described in this patent has a hydrophobic body-facing
surface. This is said to be of relatively low cost to produce and provides the desired
handle and feel characteristics, together with the required wicking properties, and
without the need to nap or brush the fabric’s surface [256].
An invention is reported which relates to composite textile materials for applications
in moisture management in diapers and incontinence apparel. The composite textile
materials comprise distinct layers of materials having appropriate characteristics so
that moisture, or liquid, migrates or drains quickly away from an inner surface of the
material in contact with the body of the wearer [257].
Development of cotton-based composite structures for nonwoven hygiene applications
is reported [151]. Cotton surfaced nonwovens (CSN) are made by placing bleached
cotton/PP staple fiber webs on one or both sides of spun bonded PP filament webs
prior to the spun bonding calender. The CSNs have good surface hand and wettability
and may serve well as a back sheet or combined top sheet and acquisition layer for

© The Textile Institute


96 Textile Progress doi:10.1533/tepr.2006.0001

diapers and sanitary napkins. Cotton-core nonwoven (CCN) are thermally bonded
laminates having bleached cotton cores with outer layers of melt blown and/or
spunbonded webs. These have been characterized for their versatility, strength, wetting,
and wicking properties. Cellulosic fibers are particularly capable of providing the
capillary systems necessary for these products. Among the cellulosic fibers used in
absorbent products, only wood pulp and cotton are used to a great extent. Cotton
offers properties that are most suitable for constructing effective capillary control in
absorbent nonwoven healthcare and personal hygiene products. Internal microscopic
pore spaces of cotton contribute to its ability to swell during the transition from the
dry to the wet state; such properties are advantageous for these applications. Cotton
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

fiber retains its gross capillary advantage because of its irregular cross-sectional
shape and surface texture, even in its lowest micronaire range. By virtue of its gross
capillarity and hydrophilicity, cotton has the capacity to absorb, store, and subsequently
release water vapor.

6.3 Industrial and Household Applications


A unique filtration device that continuously removes odor contaminants from an air
stream through a hollow fiber material having very fine solid powder in its channels
and a contaminant-removing chemisorptive agent package is reported to be effective
in removing a wide range of odors. The liquid chemisorptive agent contains sodium
permanganate in combination with sodium carbonate or sodium phosphate into which
other reagents can be added for additional odor removal capabilities [258]. The
hollow wicking fibers have internal longitudinal cavities, each with a longitudinal
opening extending to the outer fiber surface. The powder particles such as carbon,
zeolites, baking soda, cyclodextrins, and/or PTFE are forced into the longitudinal
cavities and immobilized. After entrapping the powder, the wicking fiber is impregnated
with a liquid chemisorptive agent to increase the overall filtrating performance of the
media. Undesirable odors and toxic gas molecules are removed from the air steam by
the liquid chemisorptive agent which has the affinity for them. The longitudinal
extending openings in the wicking fibers retain the liquid in the wicking fiber cavities
to interact with the air stream so as to capture the unwanted odor vapors and noxious
gas molecules. Since this filter does not rely on adsorption, wherein particles to be
removed attach to the surface of the filter material elements, but rather absorption,
molecular motion mobility can be harnessed to move unwanted gases from one zone
to another in a non-mechanical manner. The open space between the wicking fibers
remain so that the pressure differential problem is minimized and airflow restrictions
are not increased by continuous use of the fine particles, chemical reagent liquid
package and the odor vapors which are absorbed.
Wicking and wetting principles are applied to multilayer protective clothing. The
outer layer of protective clothing consists of permeable fabric or a liner or a layer
made of carbon loaded foam. To reduce the wetting due to body fluid, the outer layer
is treated with a fluoro-chemical and this reduces the surface energy and thereby the
wetting is reduced. The spreading due to liquid wetting is controlled by the outer
fabric layer. The vapor break through is due to the non-uniformity in inter layer vapor
concentration and this leads to saturation of the absorbent fabric and makes vapor
break through. By means of diffusion or convection the vapor passes through the

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 97

fabric to the atmosphere. The wicking of organic liquids on the cover fabric is slow;
this provides water repelling and chemical protection [196].
Development of a multifunctional sorbent fibrous sheet material is disclosed in a
patent. It is claimed to have excellent electrical resistivity and good absorption properties
for a broad range of liquids including paraffin oil, sulfuric acid, and sodium hydroxide
[259].
A patent discloses a multilayer floor mat [260] having good wicking properties.
A wicking phenomenon is reported by Zhou et al. [261] with mats of nanoscale
fibers. A droplet containing a surfactant solution was placed on top of a well-aligned
mat of carbon nanotubes: wicking was then observed as a film of liquid propagating
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

within the nanocarpet, such as a stain or drop absorbed into a textile fabric. The
nanoscale wicking process in carbon nano-arrays offers a simple and enabling technology
for the processing (transport, mixing, and filtering) of picoliters of fluids without any
need for confinement (nanochannel) or bulky driving pressure apparatus.
In the case of awnings and canopies, sewing threads used for seams are finished
to impede liquids from migrating through the seam [262]. They are referred as ‘anti-
wicking’ or ‘non-wicking’ threads, made from continuous polyester filament yarns,
either bonded or coated.

7. ACKNOWLEDGEMENTS
The authors and publishers acknowledge the help of those sources who give permission
to reproduce several figures and illustrations in this issue of Textile Progress.
AATCC, AIChE Journal, American Institute of Physics, American Chemical Society,
ASTM, Autex Research Journal, Colloid and Polymer Science, Colloids and Surfaces,
Colloids and Surfaces A, Coolmax®, Composites Science and Technology, Croatica
Chemica Acta, Dept. of Textile Technology, Indian Institute of Technology Delhi, Dr.
P.K. Chatterjee, Discussions of the Faraday Society, DuPont, Elsevier, Eric Slater,
INDA, Indian Journal of Fibre & Textile Researchy, Industrial & Engineering Chemistry,
Industrial Fabrics Association International, International Journal of Adhesion and
Adhesives, International Journal of Heat and Mass Transfer, Invista, John Wiley &
Sons Inc, Journal of the American Chemical Society, Faraday Trans I, Journal of
Colloid and Interface Science, Journal of Applied Polymer Science, Journal of Institute
of Engineers (India), Textile Chemistry, Journal of Statistical Physics, Journal of
Physical Chemistry, Kluwer Academic Publishiers, Krüss GmbH, Langmuir, Man-
Made Textiles in India, Marcel Dekker, Melliand Textilberichte, Polymer Composites,
Pure and Applied Chemistry, United States Patent and Trademark Office, SASMIRA,
Springer, Taylor & Francis Journals, Technical Textiles International, Textile Chemist
and Colorist, Textile Horizons, Textile Month, Textile Research Journal, Textile Trend,
Textile World, The Indian Textile Journal, The Journal of the Textile Institute, The
Society of Fiber Science and Technology, Japan (Sen-I Gakkaishi), Thermolite®, The
Physics and Chemistry of Surfaces, The Royal Society of Chemistry, The Textile
Institute, Transactions of the Faraday Society, World Sports Activewear,
www.publish.csiro.au/journals/ajc,
www.ksvinc.com,www.trc.ucdavis.edu/textiles/ntc%20projects/M02-CD03-
04panbvrief.htm.
www.ohioline.osu.edu/hyg-fact/5000/5544.html,

© The Textile Institute


98 Textile Progress doi:10.1533/tepr.2006.0001

www.cottoninc.com/NonWovenTechnical/CottonBasedNonwovens
Extract with permission, from STP 937-Toughened Composites, copyright ASTM
International, 100 Barr Harbor Drive, West Conshohocken, PA 1942.
Material originallty presented at INDA-TEC 1996, (International Nonwovens
Conference), September 11-13, 1996, Hyatt Regency Crystal City, Crystal City, Virginia,
USA; INDA-TEC 1974, (International Nonwovens Conference), March 5-6, 1974,
Shoreham American, Washington D.C., USA; International Nonwovens Journal spring,
1999; International Nonwovens Journal 2003, 12, No. 2. Copyrights held by INDA,
Association of the Nonwoven Fabrics Industry, P.O. Box 1288, Cary, North Carolina
27512–1288, USA. Tel: (919) 233–1210 Fax: (919) 233–1282. Internet; www.inda.org.
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

8. REFERENCES
[1] Y.-L. Hsieh and B. Yu, Liquid Wetting, Transport, and Retention Properties of Fibrous Assemblies, Part
I: Water Wetting Properties of Woven Fabrics and Their Constituent Single Fibers, Text. Res. J., 1992,
62, No. 11, 677–685.
[2] M. T. Ghannam and M. N. Esmail, Experimental Study on Wetting of Fibers with Non-newtonian
Liquids, AIChE Journal, 1997, 43, No. 6, 1579–1588.
[3] E. Kissa, Wetting and Wicking, Text. Res. J., 1996, 66, No. 10, 660–668.
[4] B. Miller, The Wetting of Fibers, in Surface Characteristics of Fibers and Textiles Part II (edited by M.
J. Schick), Marcel Dekker, New York, USA, 1977, p. 417.
[5] P. R. Harnett and P. N. Mehta, Survey and Comparison of Laboratory Test Methods for Measuring
Wicking, Text. Res. J., 1984, 54, No. 7, 471–478.
[6] N. K. Adam, The Chemical Structure of Solid Surfaces as Deduced from Contact Angles, Advan. Chem.
Series, 1964, 43, 52–56.
[7] A.W. Adamson and I. Ling, The Status of Contact Angles as a Thermodynamic Property, Advan. Chem.
Series, 1964, 43, 57–73.
[8] B. Miller and R. A. Young, Methodology for Studying the Wettability of Filaments, Text. Res. J., 1975,
45, No. 5, 359–365.
[9] B. P. Saville, Comfort in Physical Testing of Textiles, The Textile Institute, Woodhead Publishing
Limited, Cambridge, UK, 1999, p. 227.
[10] E. Kissa, Repellent Finishes in Handbook of Fiber Science and Technology, Part II B (edited by B. M.
Lewin and S. B. Sello), Marcel Dekker, New York, USA, 1984, p. 144.
[11] T. Kawase, S. Sekoguchi, T. Fujii, and M. Minagawa, Spreading of Liquids in Textile Assemblies, Part
I: Capillary Spreading of Liquids, Text. Res. J., 1986, 56, No. 7, 409–414.
[12] T. Gillespie, The Spreading of Low Vapor Pressure Liquids in Paper, J. Coll. Interface Sci., 1958, 13,
32–50.
[13] E. Kissa, Capillary Sorption in Fibrous Assemblies, J. Coll. Interface Sci., 1981, 83, No. 1, 265–272.
[14] I. J. Gruntfest, The Mechanism of the Wetting of Textiles, Text. Res. J., 1951, 21, No. 12, 861–866.
[15] H. S. Whang and B. S. Gupta, Surface Wetting Characteristics of Cellulosic Fibers, Text. Res. J., 2000,
70, No. 4, 351–358.
[16] R. Lucas, Ueber das Zeitgesetz des Kapillaren Aufstiegs von Flussigkeiten, Kolloid Z., 1918, 23, 15.
[17] E.W. Washburn, The Dynamics of Capillary Flow, Phys. Rev., 1921, 27, 273–283.
[18] N. K. Adam, Solid Surfaces: General Properties, in The Physics and Chemistry of Surfaces, Dover, New
York, USA, 1968, p. 179.
[19] J. R. Huntsberger, Wetting and Adhesion, Chem. Eng. News, 1964, 42, No. 44, 82–87.
[20] A. Marmur, Penetration and Displacement in Capillary Systems in Modern Approaches to Wettability:
Theory and Applications (edited by M. E. Schrader and G. I. Loeb), Plenum Press, New York, USA,
1992, p. 327.
[21] A. W. Adamson, The Solid–Liquid Interface–Contact Angle, in Physical Chemistry of Surfaces, John
Wiley, New York, USA, 1990, p. 379.
[22] R. Shuttleworth and G. L. J. Bailey, The Spreading of a Liquid Over a Rough Solid, Discuss. Faraday
Soc., 1948, 3, 16–22.
[23] http://www.ksvinc.com/contact_angle.htm
[24] J. I. Yamaki and Y. Katayama, New Method of Determining Contact Angle Between Monofilament and
Liquid, J. App. Poly. Sci., 1975, 19, No. 10, 2897–2909.
[25] B. J. Carroll, The Accurate Measurement of Contact Angle, Phase Contact Areas, Drop Volume, and
Laplace Excess Pressure in Drop-on-Fiber Systems, J. Coll. Interface Sci., 1976, 57, No. 3, 488–495.
[26] R. H. Dettre and R. E. Johnson, Contact Angle Hysteresis Part II: Contact Angle Measurements on
Rough Surfaces in Contact Angle, Wettability and Adhesion (edited by R. F. Gould), Advances in
Chemistry Series, Vol. 43, Am. Chem. Soc., Washington, D. C., USA, 1964, p. 136.
© The Textile Institute
doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 99

[27] R. E. Johnson and R. H. Dettre, Contact Angle Hysteresis Part I: Study of an Idealized Rough Surface
in Contact Angle, Wettability and Adhesion (edited by R. F. Gould), Advances in Chemistry Series, Vol.
43, Am. Chem. Soc., Washington, D. C., USA, 1964, p. 112.
[28] J. F. Oliver, C. Huh, and S. G. Mason, An Experimental Study of Some Effects of Solid Surface
Roughness on Wetting, Coll. Surfaces, 1980, 1, No. 1, 79–104.
[29] M. Tagawa, K. Gotoh, A. Yasukawa, and M. Ikuta, Estimation of Surface Free Energies and Hamaker
Constants for Fibrous Solids by Wetting Force Measurements, Coll. Poly. Sci., 1990, 268, No. 6, 589–
594.
[30] L. S. Penn and B. Miller, Advancing, Receding, and ‘Equilibrium’ Contact Angles, J. Coll. Interface
Sci., 1980, 77, No. 2, 574–576.
[31] L. S. Penn and B. Miller, Study of the Primary Cause of Contact Angle Hysteresis on Some Polymeric
Solids, J. Coll. Interface Sci., 1980, 78, No. 1, 238–241.
[32] K. S. Birdi, Contact Angle Hysteresis on Some Polymeric Solids, J. Coll. Interface Sci., 1982, 88, No.
1, 290–293.
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

[33] B. Miller, Experimental Aspects of Fiber Wetting and Liquid Movement Between Fibers, in Absorbency
(edited by P. K. Chatterjee), Elsevier, New York, USA, 1985, p. 121.
[34] L. Grader, On the Modeling of the Dynamic Contact Angle, Coll. Poly. Sci., 1986, 264, No. 8, 719–726.
[35] A. A. Jeje, Rates of Spontaneous Movement of Water in Capillary Tubes, J. Coll. Interface Sci., 1979,
69, No. 3, 420–429.
[36] F. W. Minor, A. M. Schwartz, E. A. Wulkow, and L. C. Buckles, Migration of Liquids in Textile
Assemblies, Part III: The Behavior of Liquids on Single Textile Fibers, Text. Res. J., 1959, 29, 940–949.
[37] T. H. Grindstaff, Simple Apparatus and Technique for Contact-angle Measurements on Small-denier
Single Fibers, Text. Res. J., 1969, 39, No. 10, 958–962.
[38] T. H. Grindstaff and H. T. Patterson, Common Error in Contact-angle Measurements of Fiber Using the
Level-surface Method, Text. Res. J., 1975, 45, 760–761.
[39] J. Wilhelmy, Über die Abhängigkeit der Capilaritätsconstanten des Alkohols vonSubstanz und Gestalt
des Benetzten Festen Körpers, Ann. Physik, 1863, 119, 177–217.
[40] B. Miller, L. S. Penn, and S. Hedvat, Wetting Force Measurements on Single Fibers, Coll. Surfaces,
1983, 6, No. 1, 49–61.
[41] E. Bayramli, T. G. M. Van De Ven, and S. G. Mason, Tensiometric Studies on Wetting, Part III: Low and
High Energy Surfaces, Coll. Surfaces, 1981, 3, No. 2, 131–145.
[42] J. E. Lane and D. O. Jordan, Vertical-plate Balance: Comments, Aust. J. Chem., 1971, 24, No. 6, 1297–
1299.
[43] J. F. Padday, A. R. Pitt, and R. M. Pashley, Menisci at a Free Liquid Surface: Surface Tension from the
Maximum Pull on a Rod, J. Chem. Soc. Faraday Trans. I., 1975, 71, No. 10, 1919–1931.
[44] R. L. Bendure, Dynamic Adhesion Tension Measurement, J. Coll. Interface Sci., 1973, 42, No. 1, 137–144.
[45] Y.-L. Hsieh, M. Wu, and D. Andres, Wetting Characteristics of Poly (p-phenyleneterephthalamide)
Single Fibers and Their Adhesion to Epoxy, J. Coll. Interface Sci., 1991, 144, No. 1, 127–144.
[46] K. Van de Velde and P. Kiekens, Wettability and Surface Analysis of Glass Fibers, Indian J. Fiber Text.
Res., 2000, 25, No. 1, 8–13.
[47] H. J. Barraza, M. J. Hwa, K. Blakley, E. A. O’Rear, and B. P. Grady, Wetting Behavior of Elastomer-
modified Glass Fibers, Langmuir, 2001, 17, No. 17, 5288–5296.
[48] S. K. Li, R. P. Smith, and A. W. Neumann, Wilhelmy Technique and Solidification Front Technique to
Study the Wettability of Fibers, J. Adhesion, 1984, 17, No. 2, 105–122.
[49] D. H. Kaelble, Dispersion – Polar Surface Tension Properties of Organic Solids, J. Adhesion, 1970, 2,
66–81.
[50] E. Mäder, Study of Fiber Surface Treatments for Control of Interphase Properties in Composites, Comp.
Sci. Tech., 1997, 57, No. 8, 1077–1088.
[51] B. S. Gupta and H. S. Whang, Surface Wetting and Energy Properties of Cellulose Acetate, Polyester,
and Polypropylene Fibers, Int. Nonwovens J., 1999, 8, No.1, 36–45.
[52] O. N. Tretinnikov and Y. Ikada, Dynamic Wetting and Contact Angle Hysteresis of Polymer Surfaces
Studied with the Modified Wilhelmy Balance Method, Langmuir, 1994, 10, 1606–1614.
[53] Y. K. Kamath, C. J. Dansizer, S. Hornby, and H.-D. Weigmann, Surface Wettability Scanning of Long
Filaments by a Liquid Membrane Method, Text. Res. J., 1987, 57, No. 4, 205–213.
[54] W. C. Jones and M. C. Porter, A Method for Measuring Contact Angles on Fibers, J. Coll. Interface Sci.
1967, 24, No. 1, 1–3.
[55] B. J. Carroll, Deposition of Liquid Drops on a Long Cylindrical Fiber, Text. Res. J., 1988, 58, No. 9,
495–500.
[56] C. V. Le, N. G. Ly, and M. G. Stevens, Measuring the Contact Angles of Liquid Droplets on Wool Fibers
and Determining Surface Energy Components, Text. Res. J., 1996, 66, No. 6, 389–397.
[57] T. Fort Jr and H. T. Patterson, Simple Method for Measuring Solid–Liquid Contact Angles, J. Coll.
Interface Sci., 1963, 18, 217–222.
[58] N. K. Adam and R. S. Morrell, The ‘Bloom’ of Varnish Films, Part I: Measurements of the Water-
attracting Power of Varnished Surfaces, J. Soc. Chem. Ind., 1934, 53, 255T–260T.

© The Textile Institute


100 Textile Progress doi:10.1533/tepr.2006.0001

[59] W. A. Zisman, Relation of Equilibrium Contact Angle to Liquid and Solid Constitution, in Contact
Angle, Wettability, and Adhesion (edited by R. F. Gould), Advances in Chemistry Series, Vol. 43, Am.
Chem. Soc., Washington D. C., USA, 1964, p. 1.
[60] W. D. Bascom, The Wetting Behavior of Fibers, in Modern Approaches to Wettability: Theory and
Applications (edited by M. E. Schrader and G. I. Loeb), Plenum Press, New York, USA, 1992. p. 359.
[61] K. M. Byrne, M. W. Roberts, and J. R. H. Ross, Critical Surface Tension of Wool, Text. Res. J., 1979,
49, No. 1, 34–40.
[62] R. V. Dyba and B. Miller, Dynamic Measurements of the Wetting of Single Filaments, Text. Res. J.,
1970, 40, No. 10, 884–890.
[63] G. E. P. Elliott and A. C. Riddiford, Dynamic Contact Angles, Part I: The Effect of Impressed Motion,
J. Coll. Interface Sci., 1967, 23, No. 3, 389–398.
[64] A. H. Ellison and J. B. Tejada, Dynamic Liquid/Solid Contact Angles, in NASA Report No. CR72441,
Gillette Research Institute Inc., Rockville, USA, 1968, July 23.
[65] G. Inverarity, Dynamic Wetting of Glass Fiber and Polymer Fiber, Brit. Poly. J., 1969, 1, No. 6, 245–251.
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

[66] A. M. Schwartz and S. B. Tejada, Studies of Dynamic Contact Angles on Solids, J. Coll. Interface Sci.,
1972, 38, No. 2, 359–375.
[67] R. S. Hansen and M. Miotto, Relaxation Phenomena and Contact-angle Hysteresis, J. Am. Chem. Soc.,
1957, 79, 1765.
[68] T. D. Blake and J. M. Haynes, Kinetics of Liquid/Liquid Displacement, J. Coll. Interface Sci., 1969, 30,
No. 3, 421–423.
[69] R. V. Dyba and B. Miller, Dynamic Wetting of Filaments in Solution, Text. Res. J., 1971, 41, No. 12,
978–983.
[70] Y. T. Chen, H. T. Davis, and C. W. Macosko, Wetting of Fiber Mats for Composite Manufacturing, Part
I: Visualization Experiments, AIChE Journal, 1995, 41, No. 10, 2261–2273.
[71] Q. F. Wei, R. R. Mather, A. F. Fotheringham, and R. D. Yang, Dynamic Wetting of Fibers Observed in
an Environmental Scanning Electron Microscope, Text. Res. J., 2003, 73, No. 6, 557–561.
[72] Q. F. Wei, R. R. Mather, A. F. Fotheringham, R. Yang, and J. Buckman, ESEM Study of Oil Wetting
Behavior of Polypropylene Fibers, Oil & Gas Sci. Tech.-Rev. IFP., 2003, 58, No. 5, 593–597.
[73] Y. Nakamura, K. Kamada, and T. Kondo, Wetting Behavior of Fiber Assemblies Under Flowing Solution,
Part I: Hydrodynamic Measurement of Wettability of Fiber Assemblies, Sen-I Gakkaishi, 1970, 26, No.
6, 254–262.
[74] J. P. Mutchler, J. Menkart, and A. M. Schwartz, Estimation of the Critical Surface Tension of Fibers,
Advan. Chem. Series, 1969, 86, 7–14.
[75] J. R. Dann, Forces Involved in the Adhesive Processes, Part I: Critical Surface Tensions of Polymeric
Solids as Determined with Polar Liquids, J. Coll. Interface Sci., 1970, 32, No. 2, 302–320.
[76] T. L. Ward and R. R. Benerito, Testing Based on Wettability to Differentiate Washed and Unwashed
Cotton Fibers, Text. Res. J., 1985, 55, No. 1, 40–45.
[77] A. H. Ellison, H. W. Fox, and W. A. Zinsman, Wetting of Fluorinated Solids by Hydrogen-bonding
Liquids, J. Phys. Chem., 1953, 57, No. 7, 622–627.
[78] A. G. Pittman, Polymer Surface Coatings for Wool and Mohair Fibers – Influence of Wetting Behavior,
App. Poly. Symp., 1971, 18, 593–599.
[79] B. O. Bateup, J. R. Cook, H. D. Feldtman, and B. E. Fleischfresser, Wettability of Wool Fibers, Text. Res.
J., 1976, 46, No. 10, 720–723.
[80] M. Acar, R. K. Turton, and G. R. Wray, An Analysis of the Air-jet Yarn Texturing Process, Part V: The
Effect of Wetting the Yarns, J. Text. Inst., 1986, 77, No. 6, 359–370.
[81] M. Pociute, ˙ B. Lehmann, and A. Vitakauskas, Wetting Behavior of Surgical Polyester Woven Fabrics,
Mat. Sci., 2003, 9, No. 4, 410–413.
[82] Y.-L. Hsieh, Liquid Transport in Fabric Structures, Text. Res. J., 1995, 65, No. 5, 299–307.
[83] Testing of Nonwoven Fabrics, ASTM D1117-80, 1978.
[84] Water Testing of Bibulous Papers, TAPPI Standard T 432–72, 1972.
[85] R. N. Wenzel, Resistance of Solid Surfaces to Wetting by Water, Ind. Eng. Chem., 1936, 28, No. 8, 988–
994.
[86] B. Miller and D. B. Clark, Liquid Transport Through Fabrics: Wetting and Steady State Flow, Part I: A
New Experimental Approach, Text. Res. J., 1978, 48, No. 3, 150–155.
[87] Drop Shape Analysis System G10/DSA 10, KRÜSS GmbH, Company Information, www.Kruss.de
[88] H. Herlinger, B. Kuster, and W. Volz, Methods for Characterizing the Wetting Behavior of Textile
Fabrics, Melliand Textilber.(Eng. Edn.), 1986, 67, No. 11, 807–811, E330–E332.
[89] N. R. S. Hollies, M. M. Kaessinger, and H. Bogaty, Water Transport Mechanisms in Textile Materials,
Part I: The Role of Yarn Roughness in Capillary-type Penetration, Text. Res. J., 1956, 26, 829–835.
[90] N. R. Bertoniere and S. P. Rowland, Spreading and Imbibition of Durable Press Reagent Solutions in
Cotton-containing Fabrics, Text. Res. J., 1985, 55, No. 1, 57–64.
[91] J. J. De Boer, The Wettability of Scoured and Dried Cotton Fabrics, Text. Res. J., 1980, 50, No. 10, 624–
631.
[92] S. P. Rowland, O. Cirino, and A. L. Bullock, How the Structure of Crosslinked Cotton Affects Fabric
Performance, Text. Chem. Color., 1969, 1, No. 21, 450–457.
© The Textile Institute
doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 101

[93] S. P. Rowland and N. R. Bertoniere, Some Interactions of Water-soluble Solutes with Cellulose and
Sephadex, Text. Res. J., 1976, 46, No. 10, 770–775.
[94] S. P. Rowland, D. J. Stanonis, and W. D. King, Penetration–Sorption of Cotton Fibers Measured by
Immersed Weight, J. App. Poly. Sci., 1980, 25, No. 10, 2229–2234.
[95] D. J. Stanonis and S. P. Rowland, Interaction of Carbamates and Their N-Methylol Derivatives with
Cotton and Sephadex, Measured by Gel Filtration, Text. Res. J., 1979, 49, No. 2, 72–74.
[96] E. Kissa, Wetting and Detergency, Pure Appl. Chem., 1981, 53, No. 11, 2255–2268.
[97] E. Kissa, in Detergency: Theory and Technology (edited by G. Cutler and E. Kissa), Vol. 20, Marcel
Dekker, New York, USA, 1987, p. 194.
[98] C. Heinrichs, S. Dugal, G. Heidemann, and E. Schollmeyer, Effect of Residual Layers on the Printability
of Polyester Fabric, Text. Prax. Int., 1982, 37, No. 5, 515–518.
[99] C. Heinrichs, S. Dugal, G. Heidemann, and E. Schollmeyer, Effect Value of the Drop Method as a
Measurement for the Wetting Properties of Textile Sheets, Melliand Textilber., 1982, 63, No. 12, 892–
893.
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

[100] A. M. Schwartz, Surface Tension and Surface Energy, in Absorbency (edited by P. K. Chatterjee),
Elsevier, New York, USA, 1985, p. 86.
[101] J. C. Berg, The Role of Surfactants, in Absorbency (edited by P. K. Chatterjee), Elsevier, New York,
1985, USA, p. 179.
[102] A. M. Schwartz, Capillarity: Theory and Practice – The Engineering Point of View of Macroscopic
Motion or Flow of a Liquid Under the Influence of its own Surface and Interfacial Forces, Ind. Eng.
Chem., 1969, 61, No. 1 10–21.
[103] S. Chwastiak, Wicking Method for Measuring Wetting Properties of Carbon Yarns, J. Coll. Interface
Sci., 1973, 42, No. 2, 298–309.
[104] T. Gillespie and T. Johnson, The Penetration of Aqueous Surfactant Solutions and Non-newtonian
Polymer Solutions into Paper by Capillary Action, J. Coll. Interface Sci., 1971, 36, No. 2, 282–285.
[105] B. Miller and I. Tyomkin, Spontaneous Transplanar Uptake of Liquids by Fabrics, Text. Res. J., 1984, 54,
No. 11, 706–712.
[106] W. B. Palmer, The Advances of a Liquid Front Along a Glass Yarn, J. Text. Inst., 1953, 44, T391–T400.
[107] P. K. Chatterjee and H. V. Nguyen, Mechanism of Liquid Flow and Structure Property Relationships, in
Absorbency (edited by P. K. Chatterjee), Elsevier, New York, USA, 1985, p. 29.
[108] K. T. Hodgson and J. C. Berg, The Effect of Surfactants on Wicking Flow in Fiber Networks, J. Coll.
Interface Sci., 1988, 121, No. 1, 22–31.
[109] P. R. Lord, A Comparison of the Performance of Open-end and Ring Spun Yarns in Terry Toweling, Text.
Res. J., 1974, 44, 516–522.
[110] F. E. Bartell and J. W. Shepherd, The Effect of Surface Roughness on Apparent Contact Angles and on
Contact Angle Hysteresis, Part I: The System Paraffin–Water–Air, J. Phys. Chem., 1953, 57, No. 2, 211–
215.
[111] D. D. Eley and D. C. Pepper, A Dynamical Determination of Adhesion Tension, Trans. Faraday Soc.,
1946, 42, 697–702.
[112] F. W. Minor, A. M. Schwartz, E. A. Wulkow, and L. C. Buckles, The Migration of Liquids in Textile
Assemblies, Part II: The Wicking of Liquids in Yarns, Text. Res. J., 1959, 29, No. 12, 931–939.
[113] http:trc.ucdavis.edu/textiles/ntc%20projects/M02-CD03–04panbrief.htm.
[114] A. K. Sengupta, V. K. Kothari, and R. S. Rengasamy, Wicking Behavior of Air-jet Textured Yarns, Indian
J. Fiber Text. Res., 1991, 16, No. 2, 123–127.
[115] Y. K. Kamath, S. B. Hornby, H. D. Weigmann, and M. F. Wilde, Wicking of Spin Finishes and Related
Liquids into Continuous Filament Yarns, Text. Res. J., 1994, 64, No. 1, 33–40.
[116] R. Chattopadhyay and A. Chauhan, Wicking Behavior of Compact and Ring Spun Yarns and Fabrics, in
One Day Seminar on Comfort in Textiles, Dept. of Textile Technology, I I T Delhi, New Delhi, India,
2004, October 16, p. 20.
[117] A. K. Sengupta and H. V. S. Murthy, Wicking in Ring-spun vis-a-vis Rotor-spun Yarns, Indian J. Text.
Res., 1985, 10, No. 4, 155–157.
[118] V. Subramaniam, K. S. Natarajan, A. P. Mohammed, and S. N. Many, Wicking in Siro Spun Yarns, Text.
Res. J., 1988, 58, No. 5, 302–303.
[119] http://www.ksvinc.com/powder_wetting.htm
[120] Z. Persšin, K. S. Kleinschek, and T. Kreže, Hydrophilic/Hydrophobic Characteristics of Different Cellulose
Fibers Monitored by Tensiometry, Croat. Chem. Acta, 2002, 75, No. 1, 271–280.
[121] N. Ansari and M. H. Kish, The Wicking of Water in Yarns as Measured by an Electrical Resistance
Technique, J. Text. Inst., 2000, Part 1, 91, No. 3, 410–419.
[122] X. Chen, K. G. Kornev, Y. K. Kamath, and A. V. Neimark, The Wicking Kinetics of Liquid Droplets into
Yarns, Text. Res. J., 2001, 71, No. 10, 862–869.
[123] A. Perwuelz, P. Mondon, and C. Caze, Experimental Study of Capillarity Flow in Yarns, Text. Res. J.,
2000, 70, No. 4, 333–339.
[124] H. Ito and Y. Muraoka, Water Transport Along Textile Fibers as Measured by an Electrical Capacitance
Technique, Text. Res. J., 1993, 63, No. 7, 414–420.

© The Textile Institute


102 Textile Progress doi:10.1533/tepr.2006.0001

[125] H. Tagaya, J. Haikata, K. Nakata, and K. Nishijawa, Measurement of Capillary Rise in Fabrics by
Electric Capacitance Method, Sen-I Gakkaishi, 1987, 47, 422–430.
[126] G. L. Batten Jr., Liquid Imbibition in Capillaries and Packed Beds, J. Coll. Interface Sci., 1984, 102, No.
2, 513–518.
[127] R. V. Dyba and B. Miller, Evaluation of Wettability from Capillary Rise Between Filaments, Text. Res.
J., 1969, 39, No. 10, 962–970.
[128] Y.-L. Hsieh, B. Yu, and M. M. Hartzell, Liquid Wetting, Transport, and Retention Properties of Fibrous
Assemblies, Part II: Water Wetting and Retention of 100% and Blended Woven Fabrics, Text. Res. J.,
1992, 62, No. 12, 697–704.
[129] K. Kurematsu and M. Koishi, Kinetic Studies on Liquid Penetration in Polyester Nonwoven Fabric, J.
Coll. Interface Sci., 1984, 101, No. 1, 37–45.
[130] M. F. S. Letelier, H. J. Leutheusser, and C. Z. Rosas, Refined Mathematical Analysis of the Capillary
Penetration Problem, J. Coll. Interface Sci., 1979, 72, No. 3, 465–470.
[131] A. R. Winch, Theoretical Analysis of Rate of Liquid Absorption by Capillary Absorbing Media, Text.
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

Res. J., 1959, 29, No. 3, 193–199.


[132] M. E. Trounson, J. R. McLaughlin, and P. W. Robinson, Surfactant Solution Transport in Wool Yarn, Part
II: Influence of Scouring, Prewetting, and Rewetting Surfactants, Text. Res. J., 1988, 58, No. 8, 455–462.
[133] D. A. Hangey, Wetting and Wettability of Fourth-generation Nylon Carpet Fibers, Text. Chem. Color.,
1986, 18, No. 8, 17–25.
[134] F. M. Fowkes, Role of Surface Active Agents in Wetting, J. Phys. Chem., 1953, 57, No. 1, 98–103.
[135] M. J. Rosen, in Surfactant and Interfacial Phenomena, John Wiley & Sons, New York, USA, 1989.
[136] M. J. Rosen and X. Y. Hua, Dynamic Surface Tension of Aqueous Surfactant Solutions, Part II: Parameters
at 1 s and at Mesoequilibrium, J. Coll. Interface Sci., 1990, 139, No. 2, 397–407.
[137] J. R. McLaughlin, M. E. Trounson, R. G. Stewart, and A. J. McKinnon, Surfactant Solution Transport
in Wool Yarn, Part III: Immersion Absorption, Text. Res. J., 1988, 58, No. 9, 501–506.
[138] V. B. Holland, A Comparison of Methods for the Determination of Water Absorbency by Terry Towels,
Text. Rec., 1943, 61, No. 727, 38, 40, 42.
[139] K. K. Wong, X. M. Tao, C. W. M. Yuen, and K. W. Yeung, Wicking Properties of Linen Treated with Low
Temperature Plasma, Text. Res. J., 2001, 71, No. 1, 49–56.
[140] B. Miller, H. L. Friedman, R. A. Johnson, and C. E. Holmes, Pro- and Anti-gravity Wicking Compared,
in INDA-TEC 96 Book of Papers, 1996, p. 13.1.
[141] Y.-L. Hsieh and D. A. Timm, Relationships of Substratum Wettability Measurements and Initial
Staphylococcus aureus to Films and Fabrics, J. Coll. Interface Sci.,1998, 123, No. 1, 275–286.
[142] P. Lennox-Kerr, Leaglor: Super Absorbent Acrylic from Italy, Text. Inst. Ind., 1981, 19, 83–84.
[143] R. Phukon, Effect of Gummy Substance on Wicking Behavior of Silk Fabrics, Text. Trend, 1998, 40, No.
11, 29–32.
[144] N. R. S. Hollies, M. M. Kaessinger, and H. Bogaty, Water Transport Mechanisms in Textile Materials,
Part II: Capillary Type Penetration in Yarns and Fabrics, Text. Res. J., 1957, 27, No. 1, 8–13.
[145] H. Rhee, R. A. Young, and A. M. Sarmadi, The Effect of Functional Finishes and Laundering on Textile
Materials, Part II: Characterization of Liquid Flow, J. Text. Inst., 1993, 84, No. 3, 406–418.
[146] G. L. Madan, A. M. Dave, T. K. Das, and T. S. Sarma, Hydrophilicity of Textile Fibers, Text. Res. J.,
1978, 48, No. 11, 662–663.
[147] V. Ravichandran, M. Wilde, D. R. Gadoury, A. T. Lemley, and S. K. Obendorf, Some Observations on
the Effects of Selected Dye Carriers on Poly (ethylene terephthalate), Text. Chem. Color., 1987, 19, No.
11, 35–37.
[148] B. M. Lichstein, in Second Technical Symposium, Sponsored by the International Nonwovens and
Disposable Association (INDA), Washington D. C., USA, 1974, March 5–6, p. 129.
[149] D. E. Hirt, R. K. Prud’homme, B. Miller, and L. Rebenfeld, Dynamic Surface Tension of Hydrocarbon
and Fluorocarbon Surfactant Solutions Using the Maximum Bubble Pressure Method, Coll. Surfaces,
1990, 44, 101–117.
[150] E. M. Buras Jr., C. F. Goldthwait, and R. M. Kraemer, Measurement and Theory of Absorbency of
Cotton Fabrics, Text. Res. J., 1950, 20, No. 4, 239–248.
[151] www.cottoninc.com/NonWovenTechnical/ CottonBasedNonwovens
[152] A. A. Burgeni and C. Kapur, Capillary Sorption Equilibria in Fiber Masses, Text. Res. J., 1967, 37, No.
5, 356–366.
[153] B. S. Gupta, Study of Absorbency in Nonwovens: The Role of Structural Factors and Fluid Characteristics,
in International Conference on Nonwovens, NISTI, New Delhi, India, 1992, December 5, p. 166.
[154] J. F. Oliver, L. Agbezuge, and K. Woodcock, A Diffusion Approach for Modeling Penetration of Aqueous
Liquids into Paper, Coll. Surfaces A, 1994, 89, No. 2–3, 213–226.
[155] Method of Test for Wettability of Textile Fabrics, BS 4554, 1970.
[156] S. A. Heap, Consideration of the Critical Add-on and the Uniformity of Crosslinking, Text. Res. J., 1979,
49, No. 3, 150–155.
[157] W. B. Achwal and N. Sanghavi, Horizontal Ring Test to Assess Water Transport Properties of Fabrics,
Indian Text. J., 1983, 93, No. 9, 115–117.

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 103

[158] T. Kawase, S. Sekoguchi, T. Fujii, and M. Minagawa, Spreading of Liquids in Textile Assemblies, Part
II: Effects of Softening on Capillary Spreading, Text. Res. J., 1986, 56, No. 10, 617–621.
[159] T. Kawase, Y. Morimoto, T. Fujii, and M. Minagawa, Spreading of Liquids in Textile Assemblies, Part
III: Application of an Image Analyzer System to Capillary Spreading of Liquids, Text. Res. J., 1988, 58,
No. 5, 306–308.
[160] A. P. D’Silva, C. Greenwood, S. C. Anand, D. H. Holmes, and N. Whatmough, Concurrent Determination
of Absorption and Wickability of Fabrics: A New Test Method, J. Text. Inst., 2000, Part 1, 91, No. 3,
383–396.
[161] S. C. Anand and L. Higgins, Textile Materials and Products for Activewear and Sportwear, in 2nd
International Conference of NISTI, New Delhi, India, 2004, December 2–3, p. 204.
[162] K. Ghali, B. Jones, and J. Tracy, Experimental Techniques for Measuring Parameters Describing Wetting
and Wicking in Fabrics, Text. Res. J., 1994, 64, No. 5, 106–111.
[163] P. V. D. Meeren, J. Cocquyt, S. Flores, H. Demeyere, and M. Declercq, Quantifying Wetting and
Wicking Phenomena in Cotton Terry as Affected by Fabric Conditioner Treatment, Text. Res. J., 2002,
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

72, No. 5, 423–428.


[164] R. Phukon and A. Phukan, Wicking Behavior, Indian Text. J., 1997, 107, No. 9, 60–62.
[165] R. Phukon and A. Phukan, Wicking Behavior, Indian Text. J., 1997, 108, No. 1, 40–43
[166] Q. Zhuang, S. C. Harlock, and B. D. Brook, Transfer Wicking Mechanisms of Knitted Fabrics Used as
Undergarments for Outdoor Activities, Text. Res. J., 2002, 72, No. 8, 727–734.
[167] J. L. Spencer-Smith, The Physical Basis of Clothing Comfort, Part IV: The Passage of Heat and Water
Through Damp Clothing Assemblies, Cloth. Res. J., 1977, 5, 116–128.
[168] M. M. Adler and W. K. Walsh, Mechanism of Transient Moisture Transport Between Fabrics, Text. Res.
J., 1984, 54, No.5, 334–343.
[169] R. M. Crow and R. J. Osczevski, The Interaction of Water with Fabrics, Text. Res J., 1998, 68, No. 4,
280–288.
[170] M. Howaldt and A. P. Yoganathan, Laser-Doppler Anemometry to Study Fluid Transport in Fibrous
Assemblies, Text. Res. J., 1983, 53, No. 9, 544–551.
[171] B. S. Gupta and C. J. Hong, Absorbent Characteristics of Nonwovens Containing Cellulosic Fibers, Int.
Nonwovens J., 1995, 7, No.1, 34–43.
[172] P. Luner and M. Sandell, The Wetting of Cellulose and Wood Hemicelluloses, J. Poly. Sci. Part C, 1969,
28, 115–142.
[173] G. Giannotta, M. Morra, E. Occhiello, F. Garbassi, L. Nicolais, and A. D’Amore, Dynamic Wetting of
Carbon Fibers by Viscous Fluids, J. Coll. Interface Sci., 1992, 148, No. 2, 571–578.
[174] M. Weinberg, Surface Energy Measurements of Graphite and Glass Filaments, ASTM Specl. Tech. Publ.,
1987, 937, 166–178.
[175] A. M. Cazabat and M. A. C. Stuart, Dynamics of Wetting: Effects of Surface Roughness, J. Phys. Chem.,
1986, 90, No. 22, 5845–5849.
[176] E. M. Sanders and S. H. Zeronian, An Analysis of the Moisture-related Properties of Hydrolyzed
Polyester, J. App. Poly. Sci., 1982, 27, No. 11, 4477–4491.
[177] A. K. Chesters, A. Elyousfi, A. M. Cazabat, and S. Vilette, The Influence of Surfactants on the
Hydrodynamics of Surface Wetting, J. Petrol. Sci. Eng., 1998, 20, No. 3–4, 217–222.
[178] A. Perwuelz, T. N. D. Olivera, and C. Caze, Study of Wetting at the Silicone Oil/Water/Fiber Interface,
Coll. Surfaces A, 1999, 147, No. 3, 317–329.
[179] Y.-L. Hsieh, M. M. Hartzell, and G. Barrall, Moisture Sorption on the Wetting Behavior of Poly
(p-Phenyleneterephthalamide) Fibers in Water and Epoxy Resin, J. App. Poly. Sci., 1992, 44, 1457–
1464.
[180] M. M. H. Lawson and Y.-L. Hsieh, Characterizing the Noncellulosics in Developing Cotton Fibers, Text.
Res. J., 2000, 70, No. 9, 810–819.
[181] I. A.-Askargorta, T. Lampke, and A. Bismarck, Wetting Behavior of Flax Fibers as Reinforcement for
Polypropylene, J. Coll. Interface Sci., 2003, 263, No. 2, 580–589.
[182] J. Hautojarvi and S. Laaksonen, On-line Surface Modification of Polypropylene Fibers by Corona
Treatment During Melt Spinning, Text. Res. J., 2000, 70, No. 5, 391–396.
[183] T. Wakida and S. Tokino, Surface Modification of Fiber and Polymeric Materials by Discharge Treatment
and its Application to Textile Processing, Indian J. Fiber Text. Res., 1996, 21, No. 1, 69–78.
[184] B. J. Mullins, I. E. Agranovski, R. D. Braddock, and C. M. Ho, Effect of Fiber Orientation on Fiber
Wetting Processes, J. Coll. Interface Sci., 2004, 269, No. 2, 449–458.
[185] T. L. Staples and D. G. Shaffer, Wicking Flow in Irregular Capillaries, Coll. Surfaces A., 2002, 204,
No.1–3, 239–250.
[186] H. Bogaty, N. R. S. Hollies, J. C. Hintermaier, and M. Harris, Some Properties of Serges Made from
Blends of Wool with Acrylic-type Synthetics, Text. Res. J., 1953, 23, 536–544.
[187] R. E. Wolfrom and A. C. Nuessle, The Wetting of Griege Cotton, Text. Res. J.,1952, 22, 246–253.
[188] S. J. McNeil, Comments on Surfactant Solution Transport in Wool Yarn, Text. Res. J., 1990, 60, No. 4,
244–245.

© The Textile Institute


104 Textile Progress doi:10.1533/tepr.2006.0001

[189] Y. -L. Hsieh, J. Thompson, and A. Miller, Water Wetting and Retention of Cotton Assemblies as Affected
by Alkaline and Bleaching Treatments, Text. Res. J., 1996, 66, No. 7, 456–464.
[190] Y.-L. Hsieh, A. Miller, and J. Thompson, Wetting Pore Structure and Liquid Retention of Hydrolyzed
Polyester Fabrics, Text. Res. J., 1996, 66, No. 1, 1–10.
[191] W. Barthlott and C. Neinhuis, Purity of the Scared Lotus, or Escape from Contamination in Biological
Surfaces, Plant, 1997, 202, 1–8.
[192] A. Das and S. M. Ishtiaque, Study on Comfort Characteristics of Fabrics Made of Specialty Yarn
Structure, in One Day Seminar on Comfort in Textiles, Dept. of Textile Technology, I I T Delhi, New
Delhi, India, 2004, October 16, p. 30.
[193] G. S. Bhargava, Effect of Kind of Terry Weave on Strength, Water Absorbency, and Wettability of Terry
Fabrics, Indian Text. J., 1985, 95, No. 9, 95–101.
[194] H. N. Yoon and A. Buckley, Improved Comfort Polyester, Part I: Transport Properties and Thermal
Comfort of Polyester/Cotton Blend Fabrics, Text. Res. J., 1984, 54, No.5, 289–298.
[195] N. Tarafder, S. Chakraborty, and M. M. Hossain, Laboratory Study for Measurement of Moisture
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

Transport of Terry Fabrics, Man-made Text. India, 2002, 45, No. 3, 98–102.
[196] T. Ramachandran and N. Kesavaraja, A Study on Influencing Factors for Wetting and Wicking Behavior,
Inst. Eng. India J(Text)., 2004, 84, No. 2, 37–41.
[197] Z. Weiyuan, L. Jun, C. Wenfei, and S. Long, Wetness Comfort of Fine Polypropylene-fiber Fabrics, J.
Text. Inst., 1999, 90, Part 1, No. 2, 252–263.
[198] C. Kim and Y. -L. Hsieh, Wetting and Absorbency of Nonionic Surfactant Solutions on Cotton Fabrics,
Coll. Surfaces A, 2001, 187–188, 385–397.
[199] F. H. Steiger and C. Kapur, Absorption of Liquids by Compressed Fiber Systems, Text. Res. J., 1972, 42,
No. 8, 443–449.
[200] A. S. Basra and C. P. Malik, Development of the Cotton Fiber, Int. Rev. Cytology, 1984, 89, 65–111.
[201] R. Freytag and J. J. Donzé, Akali Treatment of the Cellulose Fibers, in Handbook of Fiber Science and
Technology, Part I (edited by B. M. Lewin and S. B. Sello), Marcel Dekker, New York, USA, 1983, p.
111.
[202] M. M. Hartzell and Y. -L. Hsieh, Enzymatic Scouring to Improve Cotton Fabric Wettability, Text. Res.
J., 1998, 68, No. 4, 233–241.
[203] Y.-L. Hsieh and L. A. Cram, Enzymatic Hydrolysis to Improve Wetting and Absorbency of Polyester
Fabrics, Text. Res. J., 1998, 68, No. 5, 311–319.
[204] P. P. Tsai, L. C. Wadsworth, and J. R. Roth, Surface Modification of Fabrics Using a One-atmosphere
Glow Discharge Plasma to Improve Fabric Wettability, Text. Res. J., 1997, 67, No. 5, 359–369.
[205] S. H. Zeronian and M. J. Collins, Surface Modification of Polyester by Alkaline Treatments, Text. Prog.,
1989, 20, No. 2, 1–34.
[206] D. Nandy, J. H. Thakker, A. K. Sengupta, and S. D. Bhattacharya, Effect of Hydrolytic Action of NaOH
in Presence of Methanol on Aesthetic and Comfort Properties of Polyester Fabric, Indian J. Fiber Text.
Res., 1999, 24, No. 4, 279–283.
[207] C. J. Hawkyard, M. R. B. Lavasani, and P. Singh, A Comparison of Manual and Automated Test Methods
for Wettability, in Proceedings of the 80th World Conference of the Textile Institute, Manchester, UK,
2000, April 16–19, p. 1.
[208] M. A. M. -Morán, A. M. -Alonso, J. M. D. Tascón, and R. J. Young, Effect of Plasma Oxidation on the
Surface and Interfacial Properties of Ultra-high Modulus Carbon Fibers, Composites A, 2001, 32, No.
3–4, 361–371.
[209] Y. A. Kwon, Influence of the CF4 Plasma Treatments on the Wettability of Polypropylene Fabrics, Fibers
Poly., 2002, 3, No. 4, 174–178.
[210] M. N. Micheal and N. A. E. Zaher, Efficiency of Ultraviolet/Ozone Treatments in the Improvement of
the Dyeability and Light Fastness of Wool, J. App. Poly. Sci., 2003, 90, No. 13, 3668–3675.
[211] M. Yatagai, Correlation Between Wettability and Water Absorbency of Soiled Fabrics, Text. Res. J.,
1992, 64, No. 8, 461–465.
[212] J. N. Etters, The Influence of Polymeric Electrolyte on Flow of Water and Particulate Dye in Textile
Capillary Assemblies, Text. Chem. Color., 1989, 21, No. 6, 23–25.
[213] L. Yi, Fabric Wetting Factors, Text. Asia, 1999, 30, No. 6, 39–41.
[214] S. Rebouillat, B. Letellier, and B. Steffenino, Wettability of Single Fibers – Beyond the Contact Angle
Approach, Int. J. Adh. Adhes., 1999, 19, No. 4, 303–314.
[215] A. A. Kirsch, Increase of Pressure Drop in a Model Filter During Mist Filtration, J. Coll. Interface Sci.,
1978, 64, No.1, 120–125.
[216] D. Lukas, E. Glazyrina, and N. Pan, Computer Simulation of Liquid Wetting Dynamics in Fiber Structures
Using the Ising Model, J. Text. Inst., 1997, 88, Part 1, No. 2, 149–161.
[217] S. S. Manna, H. J. Herrmann, and D. P. Landau, A Stochastic Method to Determine the Shape of a Drop
on a Wall, J. Stat. Phys., 1992, 66, No. 3/4, 1155–1163.
[218] F. Brochard, Spreading of Liquid Drops on Thin Cylinders: the ‘Manchon/ Droplet’ Transition, J. Chem.
Phys., 1986, 84, No. 8, 4664–4672.

© The Textile Institute


doi:10.1533/tepr.2006.0001 Wetting and Wicking in Fibrous Materials 105

[219] W. Zhong, X. Ding, and Z. L. Tang, Modeling and Analyzing Liquid Wetting in Fibrous Assemblies,
Text. Res. J., 2001, 71, No. 9, 762–766.
[220] Y. T. Chen, C. W. Macosko, and H. T. Davis, Wetting of Fiber Mats for Composites Manufacturing, Part
II: Air Entrapment Model, AIChE Journal, 1995, 41, No.10, 2274–2281.
[221] W. Zhong, X. Ding, and Z. L. Tang, Analysis of Fluid Flow Through Fibrous Structures, Text. Res. J.,
2002, 72, No. 9, 751–755.
[222] J. Wiener and P. Dejlová, Wetting and Wicking in Textiles, Autex Res. J., 2003, 3, No. 2, 64–71.
[223] H. S. Kim and B. Pourdeyhimi, In-plane Liquid Distribution in Nonwoven Fabrics, Part II: Simulation,
Int. Nonwovens J., 2003, 12, No. 2, 29–33.
[224] M. Mao and S. J. Russell, Directional Permeability in Homogenous Nonwoven Structures, Part I: The
Relationship Between Directional Permeability and Fiber Orientation, J. Text. Inst., 2000, 91, Part 1, No.
2, 235–243.
[225] M. Mao and S. J. Russell, Directional Permeability in Homogenous Nonwoven Structures, Part II:
Permeability in Idealized Structures, J. Text. Inst., 2000, 91, Part 1, No. 2, 244–258.
Downloaded By: [Majumdar, Abhijit][Indian Institute of Technology] At: 12:19 21 April 2009

[226] D. Rajagopalan, A. P. Aneja, and J. M. Marchal, Modeling Capillary Flow in Complex Geometries, Text.
Res. J., 2001, 71, No. 9, 813–821.
[227] K. Ghali, B. Jones, and J. Tracy, Modeling Heat and Mass Transfer in Fabrics, Int. J. Heat. Mass
Transfer, 1995, 38, No.1, 13–21.
[228] E. Bayramli and R. L. Powell, The Normal (Transverse) Impregnation of Liquids into Axially Oriented
Fiber Bundles, J. Coll. Interface Sci., 1990, 138, No. 2, 346–353.
[229] A. P. Yarlagadda and A. P. Yoganathan, A Simplified Model for Fluid Spreading in Composite Web
Structures, Text. Res. J., 1990, 60, No. 1, 23–32.
[230] P. Yeh. USP 6 432 504 (13 August, 2002)
[231] S. Cooper. USP 6 223 565 (1 May, 2001)
[232] T. Peters and W. L. Fay Sr. USP Publication No. 2003/0182922 (2 October, 2003)
[233] J. Wallace and S. E. Tolley & Co. USP 6 238 441 (29 May, 2001)
[234] R. T. Penley. USP 6 308 337 (30 October, 2001)
[235] R. E. Dahlgren. USP 6 341 505 (29 January, 2002)
[236] http://ohioline.osu.edu/hyg-fact/5000/5544.html
[237] http://coolmax.invista.com/why_coolmax.html
[238] E. Peter, Non-Stick Socks for Cyclists. Tech. Text. Int., 1999, 8, No. 9, 7.
[239] www.clemson.edu/cucsm/CONCEPT/4DGFIBER/Introd.htm
[240] V. K. Kothari, Fiber Innovations in Technical Textiles Applications, in 2nd International Conference of
NISTI, New Delhi, India, 2004, December 2–3, p. 13.
[241] B. C. Goswami, Current and Emerging Developments in Fibers, in Emerging Trends in Polymers &
Textiles, New Delhi, India, 2005, January 7–8, p. 89.
[242] A. P. D’Silva and S. C. Anand, Responsive Garments for Sportswear, in Proceedings on Smart Textile,
NIFT, New Delhi, India, 2000, November 29, p. 32.
[243] S. M. Ishtiaque, in One Day Seminar on Automotive Textiles, SASMIRA, Mumbai, India, 2000, February
19, p. 67.
[244] M. K. Bardhan and A. D. Sule, Anatomy of Sportswear and Leisurewear: Scope for Spandex Fibers,
Man-made Text. India, 2001, 44, No. 3, 81–86.
[245] N. J. Brownless, S. C. Anand, D. A. Holmes, and T. Rowe, The Dynamics of Moisture Transportation,
Part I: The Effect of Wicking on the Thermal Resistance of Single and Multi-layer Fabric Systems, J.
Text. Inst., 1996, 87, No. 1, 172–182.
[246] D. Draper, The Quickwick Papers, World Sports Actw., 2002, 8, No. 2, 13–19.
[247] C. L. Adams. USP 6 312 051 (6 November, 2001)
[248] A. Dockery, Dry Fiber’s Transport Goes Beyond Wicking, Amer. Text. Int., 1998, 27, No 1, 78.
[249] E. Jacques and S. M. Bishop & Co. USP 6 552 244 (22 April, 2003)
[250] B. Griffiths, E. Jacques, and S. Bishop. USP 6 348 423 (19, February, 2002)
[251] N. J. Brassington. USP 5 296 290 (22 March, 1994)
[252] C. B. Hassenboehler Jr and L. C. Wadsworth. USP 5 599 366 (4 February, 1997)
[253] F.-J. D. Tsai and B. C. Wertheim. USP 6 309 988 (30 October, 2001)
[254] F.-J. D. Tsai and B. C. Wertheim. USP 6 194 483 (27 February, 2001)
[255] J. K. Dutkiewicz and K. A. G. Wood & Co. USP 6 329 565 (11 December, 2001)
[256] M. J. Heiman. USP 5 759 662 (2 June, 1998)
[257] Y. Li and K. Wing & Co. USP Publication No 2002/0094740 (18 July, 2002)
[258] R. Rohrbach and L. Xue & Co. US P 5 902 384 (11 May, 1999)
[259] A. Yahiaoui and G. H. Adam. USP 6 355 583 (12 March, 2002)
[260] G. M. Krotine. USP 5 962 350 (5 October, 1999)
[261] J. Zhou, E. Sansom, M. Gharib, and F. Noca, Nanoscale Wicking, in 56th Annual American Physical
Society/Division of Fluid Dynamics (APS/DFD) Meeting, East Rutherford, USA, November 23–25,
2003.
[262] K. Vesley, Uncommon Thread, Ind. Fab. Prod. Rev., 2001, 77, No. 11, 32–36.

© The Textile Institute

View publication stats

You might also like