Moisture Vapor Permeability and Thermal Wear Comfort

Download as pdf or txt
Download as pdf or txt
You are on page 1of 24

materials

Article
Moisture Vapor Permeability and Thermal Wear Comfort of
Ecofriendly Fiber-Embedded Woven Fabrics for
High-Performance Clothing
Hyun-Ah Kim

Korea Research Institute for Fashion Industry, 45-26, Palgong-ro, Dong-gu, Daegu 41028, Korea; [email protected]

Abstract: This study examined the moisture vapor permeability and thermal wear comfort of
ecofriendly fiber-embedded woven fabrics in terms of the yarn structure and the constituent fiber
characteristics according to two measuring methods. The moisture vapor permeability measured
using the upright cup (CaCl2 ) method (JIS L 1099A-1) was primarily dependent on the hygroscopicity
of the ecofriendly constituent fibers in the yarns and partly influenced by the pore size in the fabric
because of the yarn structure. On the other hand, the moisture vapor resistance measured using the
sweating guarded hot plate method (ISO 11092) was governed mainly by the fabric pore size and
partly by the hygroscopicity of the constituent ecofriendly fibers. The difference between the two mea-
suring methods was attributed to the different mechanisms in the measuring method. The thermal
conductivity as a measure of the thermal wear comfort of the composite yarn fabrics was governed
primarily by the pore size in the fabric and partly by the thermal characteristics of the constituent
fibers in the yarns. Lastly, considering market applications, the Coolmax® /Tencel sheath/core

fabric appears useful for winter warm feeling clothing because of its the good breathability with
 low thermal conductivity. The bamboo and Coolmax® /bamboo fabrics are suitable for summer
Citation: Kim, H.-A. Moisture Vapor clothing with a cool feel because of their high thermal conductivity with good breathability. Overall,
Permeability and Thermal Wear ecofriendly fibers (bamboo and Tencel) are of practical use for marketing environmentallyfriendly
Comfort of Ecofriendly high-performance clothing.
Fiber-Embedded Woven Fabrics for
High-Performance Clothing. Keywords: moisture vapor permeability; ecofriendly fiber; Tencel; bamboo; KES-F7; pore diameter
Materials 2021, 14, 6205. https://
doi.org/10.3390/ma14206205

Academic Editor: Dubravko Rogale 1. Introduction


The environmental impact of human beings has taken various forms, some familiar
Received: 19 August 2021
Accepted: 14 October 2021
and others not generally recognized. The former includes energy consumption and pol-
Published: 19 October 2021
lution, together with global warming, melting icecaps, rising sea levels, and increasing
frequency of adverse weather conditions [1] (p. 171). Albeit a minor contributor, the textile
Publisher’s Note: MDPI stays neutral
industry is exerting some impact, and the contribution of synthetic fibers such as PET and
with regard to jurisdictional claims in
nylon must be taken into account. The term ecofriendly has been coined to define a process
published maps and institutional affil- that is effective without harming the environment. Environmentally friendly fiber materials
iations. in textiles are divided into three areas: (organic) natural fibers, biodegradable synthetic
fibers, and recycled fibers. An organic natural fiber implies organic cotton and bamboo,
as well as Tencel as a regenerated cellulose fiber. Bamboo fibers made from bamboo pulp
have a noncircular cross-section and impart good wear comfort with superior absorption
Copyright: © 2021 by the author.
and breathability while wearing clothing. It has 100% biodegradable characteristics (de-
Licensee MDPI, Basel, Switzerland.
caying after 3 or 4 months in the soil) and does not cause environmental pollution [2].
This article is an open access article
Tencel fibers developed by Lenzing AG in Austria originated from wood. The biobased
distributed under the terms and Tencel (brand name of Lyocell) fibers are certified as compostable and biodegradable, and
conditions of the Creative Commons they can degrade after 3–4 months in the soil [2]. Biodegradable synthetic fibers include
Attribution (CC BY) license (https:// polylactic acid (PLA). PLA made from corn starch degrades after 3–5 years in the atmo-
creativecommons.org/licenses/by/ sphere and degrades after 2–4 months in landfill [2]. Recycled fibers are commercialized
4.0/). from recycled PET bottle and ECO CIRCLE® developed by Teijin Fibers Limited in Japan.

Materials 2021, 14, 6205. https://doi.org/10.3390/ma14206205 https://www.mdpi.com/journal/materials


Materials 2021, 14, 6205 2 of 24

PET is manufactured from petrochemicals and will not decompose naturally. One manner
in which this problem is tackled is through recycling [2]. The consumption of synthetic
fibers such as PET, PP, and nylon is growing steadily every year with the appearance of
highly functional new synthetic fibers. Two concerns in the textile industry are how the
consumption of synthetic fibers can be reduced by substituting them with ecofriendly
fibers with improved moisture transmission while wearing clothing, and how they can be
achieved using various yarn manufacturing technology.
Moisture transmission through textile materials is divided into two areas: moisture
liquid transmission and moisture vapor transmission. Moisture liquid transmission in-
volves a two-stage process: initially wetting and then wicking. In contrast, moisture vapor
transmission is governed by the diffusion of moisture vapor through the inter-yarn and
inter-fiber air spaces of the fabrics, called breathability. The moisture vapor transmission
behavior of fabric materials plays a vital role in maintaining the clothing wear comfort.
In particular, breathable fabrics that maintain high breathability with good perspiration
absorption and fast-drying properties are needed in sportswear, work wear, and various
types of protective clothing. In addition, high breathability in clothing allows the human
body to provide cooling due to perspiration and evaporation. Moreover, minimizing sweat
build-up in clothing is also important in cold environments. On the other hand, when
examining the mechanism of moisture vapor transmission behavior from a human body
wearing clothing, the first behavior is the diffusion of moisture vapor by sweating through
the air spaces in the fabric. The drying of moisture and the moisture vapor absorbed
by perspiration from the human body coincides, which is considered the second type of
moisture vapor transmission behavior but is slightly complicated.
Therefore, many studies [3–7] related to the moisture vapor permeability (MVP) of
various fabrics have been carried out. They reported the breathable characteristics of fabrics
according to the fiber materials and fabric structural parameters with various measuring
methods of breathability. Rego et al. [3] examined the thermo-physiological wear comfort of
cotton/polyester fabrics using a sweating guarded hot plate method. Gorjanc et al. [4] used
a water cup measuring method to examine the effects of the fabric structural parameters on
the thermal and moisture vapor resistance of cotton fabrics. Lee et al. [5] reported the effects
of fiber materials and fabric structural parameters on the MVP using statistical modeling.
Kim et al. [6] examined the relationship between the clothing performance of fabrics made
from artificial and natural fibers and the dynamic moisture vapor transfer in a microclimate.
Cubric et al. [7] explored the technical parameters affecting the moisture vapor resistance
of knitted fabrics using a sweating guarded hot plate and a thermal manikin. Thus far,
most studies have used traditional staple yarns and fabrics made from natural fibers and
blended yarns using different methods to measure the MVP.
On the other hand, some studies [8–14] on the moisture vapor transmission of water
proof breathable fabrics have been performed under various conditions, such as steady-
state, rainy, windy, and rainy and windy. In particular, Ruckman and coworkers [11–14]
examined the condensation phenomena of waterproof breathable fabrics. Yoo and cowork-
ers [15,16] analyzed the condensation of the inner surface of the fabrics at ambient tem-
peratures below 0 ◦ C in cold weather. The textile materials used in previous studies were
divided into two areas: one on traditional fabrics using natural fibers, such as wool, cot-
ton, and their blended fibers, and the other on waterproof breathable fabrics made from
synthetic filaments, such as nylon and polyester. Moreover, the method for measuring
breathability in each study was different, making it difficult to compare the breathability of
waterproof breathable fabrics.
In particular, many studies [17–29] have examined breathability characteristics accord-
ing to the measuring method. Lomax [17] reported a difference in the MVP between ISO
11092 and BS 7209 methods using coated nylon and PET fabrics. Salz [18] pointed out that
the moisture vapor transmission rates are often difficult to compare because of various
test methods. He developed a laboratory method for measuring the MVP using a heated
cup method combined with an artificial rain condition. Yoo et al. [19] explored simulated
Materials 2021, 14, 6205 3 of 24

heat and moisture transport characteristics in fabrics and garments determined using a
vertical plated sweating skin model. Several studies [20–29] compared the performance of
moisture vapor transmission of waterproof breathable fabrics using different measuring
methods. Gibson et al. [20–22] examined the parameters affecting the steady-state heat
and moisture vapor transmission measurements for clothing using a hydrophobic and
hydrophilic membrane laminated with two- and three-layer fabrics. They used the most
common techniques, such as a sweating guarded hot plate and the cup-type method to
measure moisture vapor resistance and moisture vapor transmission rate.
Dolhan [23] reported a correlation between the upright cup and Canadian control
dish methods when comparing the measuring apparatus for the moisture vapor resistance.
Congalton [24] examined the heat and moisture transport of clothing ensembles and re-
ported a strong correlation between the Hohenstein measuring method (ISO 11092) and
the evaporative dish method (BS 7209), which was in contrast to Lomax [17], who reported
an inverse correlation between the two measuring methods. Gretton et al. [25] reported a
linear correlation between moisture vapor resistance measured using the Gore-modified
desiccant method and the MVP index measured by BS 7209. McCullogh et al. [26] examined
the correlation among five measuring methods of breathability using 26 waterproof breath-
able fabrics. They reported that the upright cup method showed the lowest water vapor
transmission rate, followed by the dynamic cell method, inverted cup method, and desic-
cant inverted cup method. In addition, the correlation coefficient between the sweating
guarded hot plate and desiccant inverted cup methods showed a high inverse correlation.
Huang [27] and Huang and coworkers [28,29] examined the factors affecting the
moisture vapor resistance obtained using the ISO 11092 method and compared them with
the existing water vapor transmittance method. Many studies carried out thus far have
reported that the breathability of the fabrics differed according to the fiber materials, fabric
structural factors, and surface modification method, such as coating and laminating, as well
as the measuring method. Measuring the breathability of fabrics can be achieved using two
methods. The first is to measure the water vapor transmittance rate (WVTR), which is a
simple method used for quality control and marketing purposes. The second is to measure
the moisture vapor resistance using the wet thermal resistance method, which is more
precise (accurate) and used mainly for fabric development and research. On the other hand,
many studies [30–36] focusing on improving wear comfort by sweating focused on the yarn
and fabric manufacturing technologies by combining hydrophobic and hydrophilic yarns.
Several wear comforts of woven fabrics made from various yarns were examined using a
different yarn structure and various constituent fiber characteristics, such as Coolmax® ,
Tencel, Bamboo, and other ecofriendly fibers [37–43].
Despite these studies, there are few reports on the MVP (breathability) of the woven
fabrics made from composite yarns, such as siro, siro-fil, and sheath/core yarns, using
Coolmax® , Tencel, bamboo, PET, and polypropylene (PP) filaments. This study used
bamboo and Tencel fibers to produce ecofriendly yarns with PET and PP filaments using
siro and sheath/core spinning systems. The PET and PP filaments in the sheath/core or
siro-fil yarn structures play a very important role in passing water and moisture vapor
as drainage in the yarns and fabrics. Accordingly, PET and PP filaments were used to
enhance the wear comfort characteristics with superior water absorption and moisture
vapor permeability, even though they are not ecofriendly fibers.
There are no reports on the difference in breathability characteristics according to the
yarn structure and the measuring method of breathability. Therefore, the main concern
of this study is how the breathability of different composite yarn fabrics is influenced
by the yarn structures and constituent fiber characteristics, and how it is associated with
the thermal and absorption properties of the fabrics regarding the pore size of the fabrics
according to the measuring method. Accordingly, in this study, two types of breathability,
i.e., WVTR and moisture vapor resistance, by the wet thermal transmission of 15 fabric
specimens made from different composite yarns were measured and compared in terms
of the yarn structure and constituent fiber characteristics according to the two measuring
Materials 2021, 14, 6205 4 of 24

methods. In addition, the moisture vapor transmission characteristics were compared with
the thermal conductivity of the fabric specimens measured using the KES-F7 in terms of
the pore size of the fabric and the thermal characteristics of constituent fibers.

2. Materials and Methods


2.1. Yarn Preparation
Coolmax® (Dupont, Torrance, CA, USA) and Tencel (Gemeindeverwaltung, Lenzing,
Austria) as ecofriendly fibers are used widely in the textile market. Various composite
yarns have been made and commercialized using hi-multi PET and PP filaments in the
functional and work wear market. In particular, a novel polypropylene (PP) filament with
good absorption properties was applied to achieve clothing with thermo-physiological
comfort [1–3]. On the other hand, few studies have investigated the moisture and heat
transport of Tencel/Coolmax® -incorporated yarns and their fabrics according to the yarn
structure with their measuring method. In this study, seven types of composite yarn
were spun using ring, siro, and sheath/core spinning systems. Two existing hi-multi PET
(75d/144f) and PP (100d/48f) filaments were used. Table 1 provides details of the yarn
specimens. Sheath/core composite yarn specimens (no. (1), 147.5 dtex) were made by
feeding PP DTY (draw-textured yarn) filament (30d/24f) between the double roving of
Tencel. The siro-fil composite yarn (no. (2), 147.5 dtex) was prepared by replacing one of
the siro components with a PET DTY filament (55d/216f) inserted at the back of the front
rollers on the siro-spinning machine. The siro-spun yarn specimen (no. (3), 147.5 dtex)
was made using Tencel roving on the siro-spinning frame. These were used as warp yarns.
Sheath/core composite yarn specimen (no. (4), 196.7 dtex) was made by feeding Coolmax®
filament (50d/36f) between the double roving of Tencel. A ring-spun yarn specimen
(no. (5), 196.7 dtex) was prepared using bamboo and Coolmax® rovings on a ring spinning
frame (Zinzer MAT 670, Krefeld, Germany).The siro-fil composite yarn specimen (no. (6),
196.7 dtex) was spun by replacing one of the siro components with a PET DTY filament
(55d/216f) inserted atthe back of the front rollers on the siro-spinning frame. The bamboo
spun yarn specimen (no. (7), 196.7 dtex) was made using bamboo roving. In addition,
the hi-multi PET filament (no. (8), 83.3 dtex) and PP filament (no. (9), 111.1 dtex) were
prepared as existing commercial yarns. These were used as weft yarns. Each warp yarn
specimen was prepared as a 147.5 dtex and a 196.7 dtex for each weft yarn specimen by
regulating the draft ratio on the ring (siro) and sheath/core spinning frames. Table 1 lists
their twist multiplier (TM), spindle rpm, and blend ratio.
Table 2 presents a schematic diagram of the yarn specimens in this study, which
was drawn with reference to the images of yarn cross-sections obtained by SEM and
optical microscopy. As shown in Table 2, the yarn structure of the sheath/core composite
yarns (no. (1) and (4)) was composed of filaments in the core and staple fibers in the
sheath, respectively. The siro-fil yarns (no. (2) and (6)) were drawn as a side-by-side
cross-section, which was assumed to be formed as two parts caused by the centrifugal
force by the traveler rotation on the ring frame. The siro-spun yarn (no. (3)) was twisted
using two Tencel rovings on the siro-spinning frame and had a compact yarn structure with
uniformly distributed Tencel fibers in the yarn cross-section. The Coolmax® /bamboo spun
yarn (no. (5)) was twisted by a traveler on the ring-spinning system and had a relatively
bulky yarn structure with noncircular cross-sections of Coolmax® and bamboo fibers.
The bamboo spun yarn (no. (7)) had a slightly compact yarn structure with bamboo fibers
distributed uniformly in the yarn cross-section. PET and PP filaments (no. (8) and (9)) were
composed of parallel filament bundles and had yarn structures with many pores in the
yarn cross-section.
Materials 2021, 14, 6205 5 of 24

Table 1. Details of the yarn specimens used in this study.

Twist Period of Biological


Yarn BlendRatio Spinning Yarn No. Fiber (Filament)
Yarn Type TM Spindle Decay in Soil
No. (%) Method ( dtex) √ Used
tpi/ Ne) (rpm) (Month)
PP: 39.3/ PP DTY 30d/24f PP: no decay
(1) PP/Tencel S/C Sheath/core 147.5 4.53 7000
T: 60.7 and Tencel S/F T: 3-4
PET DTY
PET/Tencel P: 44.4/ P: no decay
(2) Siro-fil 147.5 4.12 9000 55d/216f and
Siro-fil T: 55.6 T: 3-4
Tencel S/F
Tencel
(3) T: 100 Siro-spun 147.5 4.42 11,000 Tencel S/F T: 3-4
Siro-spun
Coolmax
C: 39.3 C: no decay
(4) Coolmax/TencelS/C Sheath/core 196.7 4.34 9000 50d/36f and
T: 60.7 T: 3-4
Tencel S/F
Coolmax/ C: 48.6/ Coolmax/bamboo C: no decay
(5) Ring-spun 196.7 3.82 12,000
BambooSpun B: 51.4 S/F B: 3-4
PET DTY
PET/Tencel P: 44.4/ P: no decay
(6) Siro-fil 196.7 4.12 9000 55d-216f and
Siro-fil T:55.6 T: 3-4
Tencel S/F
(7) Bamboo spun B: 100 Ring-spun 196.7 3.82 12,000 Bamboo S/F B: 3-4
PET DTY
(8) * Hi-multi PET P: 100 - 83.3 - - no decay
75d/144f
PP DTY
(9) * PP filament PP: 100 - 111.1 - - no decay
100d/48f
Note: S/F: staple fiber, T: Tencel, *: existing filament. PP: polypropylene, C: Coolmax. P: PET, B: bamboo, DTY: draw textured yarn.

Table 2. Composite warp and weft yarn models [44].

Model

PP DTY 30d/24f + Tencel PET DTY 55d/216f + Tencel siro-fil Tencel + Tencel staple fibres
Spec.
sheath/core yarn for yarn siro-spun yarns
Yarn PP/Tencel S/C 147.5 dtex PET/Tencel siro-fil 147.5 dtex Tencel siro-spun 147.5 dtex
specimens No (1) No (2) No (3)

Model

Coolmax 50d/36f+ Tencel Coolmax/bamboo staple fibers PET DTY 55d/216f + Tencel
Spec.
sheath/core yarn ring-spun yarns siro-fil yarn
Coolmax/bamboo spun yarn 196.7
Yarn Coolmax/Tencel S/C 196.7 dtex PET/Tencel siro-fil 196.7 dtex
dtex
specimens No (4) No (6)
No (5)
Materials 2021, 14, 6205 6 of 24

Table 2. Cont.

Model

bamboo staple fibees


Spec. PET DTY 75d/144f filament PP DTY 100d/48f
ring-spun yarns
Yarn Hi-multi PET 83.3 dtex PP 111.1 dtex
bamboo spun yarn 196.7 dtex No (7)
specimens No (8) No (9)

2.2. Fabric Preparation


Fifteen types of fabric specimens were woven on a rapier loom (GTX 4-R, Picanol,
Belgium), which was divided into three groups according to the warp beams. Three types
of warp beams were prepared on a single warping machine (ROM2, Karl Mayer, Germany)
using yarn specimens (1), (2), and (3) (Table 1), with which three groups of fabric specimens
were woven using five types of weft yarn specimens ((4) to (8) in Table 1). Yarn specimen
(9) was alternatively inserted as second weft yarn. Table 3 lists these 15 fabric specimens.
Group A was composed of five different fabric specimens ((1) to (5)) using five weft
yarn specimens ((4) to (8) in Table 1), with the first warp beam made from PP/Tencel
core/sheath yarn (no. (1) in Table 1). Group B was composed of five types of fabric
specimens ((6) to (10)) using the same five weft yarn specimens ((4) to (8) in Table 1) with
the second warp beam made from PET/Tencel siro-fil yarn (no. (2) in Table 1). Group C was
composed of five fabric specimens ((11) to (15)) using the same five weft yarn specimens
((4) to (8) in Table 1) with the third warp beam made from Tencel siro-spun yarn (no. (3) in
Table 1). PP DTY (111.1 dtex/48f) was inserted alternatively for all fabric specimens as a
second weft yarn for a plain weave pattern.
The warp density of all fabric specimens was 36.0 ends/cm and 24.6 picks/cm for
the weft. The fabric weight was calculated using yarn linear density and fabric density of
the fabric specimen. The fabric thickness was measured at a pressure of 2 gf/cm2 using
a FAST-1 compression meter [42]. Fifteen types of gray fabric specimens, 20 m long each,
were prepared and followed by dyeing and finishing processes. Gray fabric specimens
were scoured on a CPB scouring machine (BPB, Kuester, Germany) and washed at a speed
of 50 m/min on a continuous drying machine (Extra-CTA 2400, Benninger, Switzerland).
A preset was done at 40 m/min at 150 ◦ C. Dyeing was performed on a rapid dyeing
machine (Cut-MF-1, Hisaka work Ltd., Osaka, Japan) at 120 ◦ C for 60 min, followed by
a drying treatment on a continuous dryer machine (Shrink dryer, Ilsung Ltd. Co., Seoul,
Korea).The final setting was performed on a stenter machine (Sun-super, Ilsung Ltd. Co.,
Seoul, Korea) at a speed of 50 m/min at 130 ◦ C.
Materials 2021, 14, 6205 7 of 24

Table 3. Specification of the fabric specimens.

Fabric Density
Fabric Weft Yarn Weight
(Ends, Picks/cm) Thickness
Group Warp Yarn
Specimen No. (g/y) (10−3 m)
Yarn 1 Yarn 2 Wp Wf
Coolmax/Tencel
1 162 0.368
S/C (196.7 dtex)
Coolmax/bamboo
2 162 0.345
spun (196.7 dtex)
PP/Tencel
PET/Tencel PP
A 3 Sheath/core 36.0 24.6 162 0.341
Siro-fil (196.7 dtex) (111.1 dtex)
(147.5 dtex)
4 Bamboo spun (196.7 dtex) 162 0.364
5 Hi-multi PET (83.3 dtex) 137 0.352
Coolmax/Tencel
6 160 0.396
S/C (196.7 dtex)
Coolmax/bamboo
7 161 0.337
spun (196.7 dtex)
PET/Tencel
PET Tencel PP
B 8 Siro-fil 36.0 24.6 161 0.345
Siro-fil (196.7 dtex) (111.1 dtex)
(147.5 dtex)
9 Bamboo spun (196.7 dtex) 161 0.345
10 Hi-multi PET (83.3 dtex) 137 0.294
Coolmax/Tencel
11 158 0.380
S/C (196.7 dtex)
Coolmax/bamboo
12 160 0.356
spun (196.7 dtex)
Tencel
PET Tencel PP
C 13 Siro-spun 36.0 24.6 160 0.345
Siro-fil (196.7 dtex) (111.1 dtex)
(147.5 dtex)
14 Bamboo spun (196.7 dtex) 161 0.349
15 Hi-multi PET (83.3 dtex) 133 0.301

2.3. Measurement of Pore Size of the Fabric Specimens


The moisture and heat transport of woven fabrics were strongly dependent on
the constituent fiber characteristics, pore size, and fabric structural parameters [45–49].
Kim and Kim [50] reported that the thermal conductivity, drying rate, and air permeability
of hollow filament-embedded woven fabrics were strongly dependent on the porosity and
pore size of the fabrics. In this study, the primary concern of pore size measurements was
how the pore size is influenced by the constituent yarn structure and how it affects the
moisture vapor transport of fabrics according to the measuring method. The pore diameter
(D, µm) was measured using a capillary flow porometer (CFP-1200 AE PMI Co., Ithaca, NY,
USA) according to the ASTM measuring method. Figure 1 presents the porometer used in
this study.
Materials 2021, 14, 6205 USV Symbol Macro(s) 8 of 24

01A0 Ơ \Ohorn
USV Symbol Macro(s) \textrighthorn{O}Description
01A0 Ơ \Ohorn 01A1 ơ \ohorn LATIN CAPITAL LETTER O WITH HORN
\textrighthorn{O} \textrighthorn{o}
01A1 ơ \ohorn 01A4 Ƥ \m{P} LATIN SMALL LETTER O WITH HORN
\textrighthorn{o} \textPhook
01A4 Ƥ \m{P} 01A5 ƥ \m{p} LATIN CAPITAL LETTER P WITH HOOK
\textPhook \texthtp
\textphook
01A5 ƥ \m{p} LATIN SMALL LETTER P WITH HOOK
\texthtp 01A9 Ʃ \ESH
\textphook \textEsh
01A9 Ʃ \ESH 01AA ƪ \textlooptoprevesh
LATIN CAPITAL LETTER ESH
\textEsh \textlhtlongi
01AA ƪ \textlooptoprevesh 01AB ƫ \textpalhookbelow{t}
LATIN LETTER REVERSED ESH LOOP
\textlhtlongi \textlhookt
01AB ƫ \textpalhookbelow{t} 01AC Ƭ \m{T} LATIN SMALL LETTER T WITH PALATAL HOOK
\textlhookt \textThook
01AC Ƭ \m{T} 01AD ƭ \m{t} LATIN CAPITAL LETTER T WITH HOOK
\textThook \texthtt
\textthook
Figure ƭ
01AD 1. Capillary \m{t}
flow porometer machine. LATIN SMALL LETTER T WITH HOOK
\texthtt 01AE Ʈ \M{T}
\textthook \textTretroflexhook
A fabric specimen 47 mm in diameter 01AF was
Ư placed
\Uhornin the specimen holder shown
01AE Ʈ \M{T} LATIN CAPITAL LETTER T WITH RETROFLEX HOOK
in Figure 1. The\textTretroflexhook
specimen holder was closed, and slight gas pressure was applied to
\textrighthorn{U}
eliminate
01AF Ưthe possible
\Uhorn liquid backflow.01B0 The gas ư pressure
\uhorn was increased slowly.
LATIN CAPITAL LETTER U WITHFinally,
HORN
\textrighthorn{U} \textrighthorn{u}
the lowest pressure at which a steady stream of bubbles rises from the central area of the
ư
01B0 reservoir \uhorn 01B1 Ʊ \textupsilon LATIN SMALL LETTER U WITH HORN
liquid was recorded. The maximum pore\m{U} diameter (D) was calculated using
\textrighthorn{u}
Equation
01B1
(1)
Ʊ from the median
\textupsilon
value of
01B2 the graph
Ʋ between
\m{V} the airflow and pressure.
LATIN CAPITAL LETTER UPSILON
\m{U} \textVhook
01B2 Ʋ \m{V} 01B3D = C Ƴ /p, \m{Y} LATIN CAPITAL LETTER V WITH HOOK (1)
\textVhook \textYhook
01B3 D, Ƴ , and\m{Y}
where p are the maximum01B4 ƴ
pore diameter \m{y}
(µm),the surface tension
LATIN CAPITAL ofY WITH
LETTER theHOOK
liquor
\textYhook \textyhook
(dynes/cm), and pressure (psi); C = 0.415 when p is in psi units. The yarn and fabric cross-
01B4 ƴ measured
\m{y} 01B5 Ƶ \B{Z} LATIN SMALL LETTER Y WITH HOOK
sections were \textyhook
by field-emission scanning electron
\Zbar microscopy (FE-SEM, S-4100,
Hitachi
01B5 Co.,
Ƶ Omori,
\B{Z} Japan) and optical
01B6 microscopy
ƶ (i-Camscope-305A,
\B{z} Seoul, Korea).
LATIN CAPITAL LETTER Z WITH STROKE
\Zbar 01B7 Ʒ \m{Z}
2.4.
01B6Measurement
ƶ of the WVTR of the Fabric Specimens \EZH
\B{z} LATIN SMALL LETTER Z WITH STROKE
\textEzh
Ʒ
01B7 The resistance \m{Z}to moisture vapor diffusion (i.e., moisture vapor resistance)
LATIN CAPITAL LETTER EZH depends
\EZH 01B9 ƹ \textrevyogh
mainly on the air\textEzh
permeability of the 01BA
fabric andƺ indicates its ability to transfer perspira-
\textbenttailyogh
tion
01B9 vaporƹ leaving of human skin. 01BB
\textrevyogh In this study,
ƻ the main
\B{2} concerns of moisture vapor
LATIN SMALL LETTER EZH REVERSED
transmission
01BA ƺ measurements are how the moisture vapor
\textbenttailyogh resistance
\textcrtwo isSMALL
LATIN associated with
LETTER EZH WITH TAIL the

absorption
01BB ƻ rate of
\B{2} 01BEto theƾconstituent
the fabrics according yarn structure
LATIN LETTER (porosity)
\textcrinvglotstop TWO WITH STROKEin the

fabric, and how the \textcrtwo 01BF of theƿ fabric\wynn


thermal conductivity is related to the moisture vapor resis-
01BE and ƾis dependent
tance \textcrinvglotstop 01C0
on the yarn structure ǀ
and \textpipe
thermal conductivity of the constituent
LATIN LETTER INVERTED GLOTTAL STOP WITH STROKE

01BF Theƿ WVTR


fibers. (g/m2 ·h) was measured using the\textpipevar
\wynn JIS L LATIN LETTER WYNN
1099A-1, which was based on
\textvertline
01C0 ǀ \textpipe
BS7209, similar to the upright cup method, asǁ shown LATIN LETTER DENTAL CLICK
in Figure 2a. Five aluminum cups,
\textpipevar 01C1 \textdoublepipe
6 cm in diameter\textvertline
and 2.5 cm in height, were prepared, and the cup inside was heated to
\textdoublepipevar
40 ◦ 01C2 ǂ \textdoublebarpipe
01C1C by heating
ǁ in an air-conditioned
\textdoublepipe room (container), then filled
LATINwith 33 g of
LETTER LATERAL CaCl2 as
CLICK

a desiccating agent. Five fabric specimens,7 cm in \textdoublebarpipevar


\textdoublepipevar diameter, were prepared and condi-
01C2 at 20 ǂ ± 2 ◦\textdoublebarpipe 01C3 ǃ \textrclick LATIN LETTER ALVEOLAR CLICK
tioned C and an RH of 65 ± 2% for 24 h. The fabric specimen was laid 3 mm
\textdoublebarpipevar01C4 DŽ \v{\DZ}
away from ǃ CaCl in an aluminum
2\textrclick cup; its surface was faced toward the CaCl 2 in the cup.
01C3 01C5 Dž \v{\Dz} LATIN LETTER RETROFLEX CLICK
Packing rubber
DŽ with a circular covering was clamped and sealed over the fabric specimen
01C4 \v{\DZ} 01C6 dž \v{\dz} LATIN CAPITAL LETTER DZ WITH CARON
on the mouth
Dž of\v{\Dz}the aluminum cup to prevent the leakage of moisture vapor between
01C5 01C7 LJ \LJ LATIN CAPITAL LETTER D WITH SMALL LETTER Z WITH CAR
the
01C6fabricdž specimens
\v{\dz}and the aluminum
01C8
cup. LjFive aluminum
\Lj
cup assemblies with fabric
LATIN SMALL LETTER DZ WITH CARON
specimens were placed in the conditioning room at 40 ± 2 ◦ C and 90 ± 5% RH for 1 h.
01C7 LJ \LJ 01C9 lj \lj LATIN CAPITAL LETTER LJ
The
01C8waterLj vapor \Lj
transmission rate was 01CAcalculated
NJ using \NJ EquationLATIN (2).CAPITAL LETTER L WITH SMALL LETTER J
01C9 lj \lj LATIN SMALL LETTER LJ

01CA NJ \NJ WVTR (g/m2 ·h) = 10(W2 − W1 )/(A × t),LATIN CAPITAL LETTER NJ (2)

9
Materials 2021, 14, 6205 9 of 24

where, WVTR is water vapor transmission rate (g/m2 ·h), W2 is the mass of the fabric
specimen (mg) after the test, W1 is the mass of the fabric specimen (mg) before the test, A
is the specimen area (cm2 ), and t is the testing time (h).

Figure 2. Schematic diagram of the measuring equipment of (a) upright cup method (JIS L 1099A-1), (b) sweating guarded
hot plate method (ISO 11092), and (c) KES-F7 system [44].

2.5. Measurement of Moisture Vapor Resistance of the Fabric Specimens


Moisture vapor resistance (Ref , m2 Pa/W) of the fabric was measured using a sweating
guarded hot plate (Therm DAC, London, UK) according to the ISO 11092 method. Figure 2b
presents a schematic diagram of this apparatus. A fabric specimen, 50.8 cm × 50.8 cm in
size, was prepared and conditioned in a standard atmosphere with an RH of 65% and a
temperature of 20 ◦ C. The specimen was placed over the PTFE membrane on perforated
metal on a hot plate, which was used to prevent water on the perforated metal of the hot
plate from wetting the fabric specimen. The temperatures of the guarded hot plate and air
in the chamber were kept at 35 ± 0.5 ◦ C (i.e., the temperature of human skin) with an RH
of 40% and an air speed of 1 m/s. The moisture vapor resistance (Ref ) of the fabric was
determined by measuring the evaporative heat loss (H) under the steady-state condition,
using Equations (3) and (4). 
ps − pa A
Re,t = , (3)
H
where Re,t is the total resistance to evaporative heat transfer provided by the fabric system
and air layer (m2 ·Pa·W−1 ), A is the area of the plate test section (m2 ), ps is the water vapor
pressure at the plate surface (Pa), pa is the moisture vapor pressure in the air (Pa), and H is
the power input (heat loss)(W).
Re,f = Re,t − Re,a , (4)
where Re,f is the resistance to evaporative heat transfer provided by the fabric (i.e., moisture
vapor resistance of fabric), and Re,a is the resistance measured for the air layer and liquid
barrier. The arithmetic mean of five readings from each fabric specimen was calculated.
Materials 2021, 14, 6205 10 of 24

2.6. Measurement of the Thermal Conductivityof the Fabric Specimens


Thermal transmission through textile materials is divided into two methods: wet and
dry heat transmission. Moisture vapor transmission by wet heat transport is governed by
diffusion and convection, whereas dry heat transport occurs through conduction, convec-
tion, and radiation from the human body to the atmosphere. Moisture vapor resistance
measurements using the ISO 11092 method were assessed using the principle of wet heat
transmission, which means the movement of wet heat particles evaporated by perspiration
sweated from human skin. In this study, one concern of the thermal transport measurement
was how the wet heat transmission related to the moisture vapor resistance is associated
with the dry heat transmission, and how they are influenced by the yarn structure and
measuring method. Accordingly, the thermal conductivity of the fabric specimens as a
measure of dry heat transport was measured to determine how it is influenced by the
constituent yarn structure and then how it affects the moisture vapor resistance of the fabric
specimens. The thermal conductivity of the fabric specimens was measured using the
KES-F7 system (Kato Tech. Co., Ltd., Kyoto, Japan), of which a schematic diagram is shown
in Figure 2c. First, the B.T. Box temperature was set to 30 ◦ C, and water was circulated
at a constant temperature of 20 ◦ C in a water bath. A fabric specimen was placed on the
water bath. Heat flowed from a high temperature (B.T. Box 30 ◦ C) to a low temperature
(water bath, 20 ◦ C) in the apparatus through a plate and specimen. The B.T. Box (composed
of an electrical system equipped with temperature sensors) then measured the heat loss
emanating from the plate as watts (W) from the change in electrical voltage. The heat
loss (W/10−4 m2 ) with the fabric specimen placed on the water bath is H in Equation (5).
The thermal conductivity (K) was calculated using Equation (5) as follows:

H D
K= × , (5)
t A·T

where, K, H, and D are the thermal conductivity (W/10−2 m·◦ C), dry heat loss (W/10−4 m2 ),
and fabric thickness (10−2 m), respectively. A and t are the fabric area (10−4 m2 ) and
time (h), respectively. ∆T is the temperature difference (◦ C).

2.7. Measurement of the Absorption Rateof the Fabric Specimens


The moisture vapor particles sweated from the human body move throughout the
fabric and are partly adsorbed and wetted, after which drying will occur. The absorp-
tion rate of the fabric specimens was measured using a drying rate measuring apparatus
(IT-ACD, INTEC Co. Ltd., Tokyo, Japan), as shown in Figure 3. A square fabric spec-
imen (40 cm × 40 cm) was conditioned at 20 ◦ C and 65% RH in the conditioning room
(JIS L 1096, 2010), and the initial mass (m1 ) was then measured. A square fabric specimen
was submerged in distilled water for 20 min at 27 ± 2 ◦ C in a water bath. The speci-
men in the distilled water bath was passed through a mangle at 25 cm/s, and the mass
(m2 ) was weighed. The absorption rate (%) of each fabric specimen was calculated using
Equation (6).
R (%) = (m2 − m1 )/m1 , (6)
where R is the absorption rate of fabric (%), m1 is the initial mass (g) of the fabric specimens,
and m2 is fabric mass (g) after passing through the mangle.
Materials 2021, 14, 6205 11 of 24

Figure 3. Schematic diagram of measuring equipment of the absorption rate.

3. Results and Discussion


3.1. Pore Size of the Fabric Specimens with SEM Images of the Cross-Sections of Yarns
The fabric porosity affecting clothing wear comfort is divided into two types: micro
and macro porosity [51,52]. In particular, the moisture vapor and heat permeabilities are
incorporated with both micro and macro porosities. In this study, the calculated porosity
does not apply because the fabric specimens prepared in this study were made from the
same yarn count and fabric sett, which means that the calculated porosity may not be
available to examine the difference in the MVP among the fabric specimens prepared
using the same yarn count and fabric sett. Therefore, in this study, the pore diameter was
considered a measure to examine the WVP according to the yarn structure and measuring
method. Table 3 lists the measured physical properties of the fabric specimens with
the measured pore diameters. ANOVA (F-test) was carried out to verify the statistical
significance of the experimental data shown in Table 4. ANOVA was performed between
the mean value of the physical properties of each specimen (five specimens) in each group
with the 95% confidence limit (5% significance level). Table 5 lists the ANOVA analysis of
the physical properties of 15 fabric specimens. As shown in Table 5, the significance test
between each mean pore diameter among the five specimens in each group A, B, and C
was statistically significant, as F0 (V/Ve) > F (4, 20, 0.95) and p < 0.05. Similarly, WVTR,
Ref , and thermal conductivity were statistically significant, as F0 (V/Ve) > F (4, 20, 0.95)
and p < 0.05, as shown in Table 5. Figure 4 presents a diagram of the pore diameters
(mean) with the deviation of the fabric specimens listed in Table 4. The deviation in Table 4
denotes the difference between maximum and minimum values of five experimental data
of each specimen.
Materials 2021, 14, 6205 12 of 24

Table 4. Physical properties of the fabric specimens.

Moisture Vapor Permeability Thermal Conductivity


Pore Absorption
Diameter Water Vapor Moisture Vapor
Rate
Fabric D Transmission Rate Resistance K
Group R
Specimen No. (µm) WVTR Ref (m2 ·Pa/W)
(%)
(g/m2 ·h) (m2 ·Pa/W)
Mean Dev. Mean Dev. Mean Dev. Mean Dev.(10−3 ) Mean
1 3.82 0.100 425.4 11.1 1.91 0.103 0.0445 1.66 28.3
2 3.19 0.088 413.2 10.0 2.80 0.109 0.0458 1.67 27.2
A 3 2.85 0.112 410.6 9.2 3.42 0.118 0.0489 1.64 26.4
4 2.98 0.110 430.2 11.1 2.91 0.101 0.0473 1.49 30.2
5 3.56 0.124 385.5 10.0 1.74 0.107 0.0423 1.74 26.3
6 3.20 0.104 348.2 10.1 1.56 0.101 0.0439 1.57 30.2
7 2.85 0.115 340.5 10.0 2.85 0.109 0.0437 1.67 28.5
B 8 2.30 0.096 337.6 7.1 3.24 0.107 0.0481 1.58 28.1
9 2.70 0.121 375.4 9.0 2.88 0.101 0.0453 1.67 32.4
10 3.11 0.116 326.2 9.1 1.34 0.109 0.0395 1.64 26.2
11 3.45 0.100 369.8 11.0 2.12 0.118 0.0453 1.80 34.4
12 2.95 0.116 358.6 11.1 2.43 0.103 0.0461 1.81 34.2
C 13 2.55 0.106 355.4 10.3 3.32 0.101 0.0478 1.80 34.0
14 2.76 0.094 378.5 9.1 2.86 0.118 0.0472 1.54 36.7
15 3.25 0.111 337.7 11.2 1.78 0.101 0.0421 1.45 32.1
Note: dev = max − min.

Table 5. ANOVA analysis of the fabric physical properties.

Physical Properties F-Value(F0 ) F(4, 20, 0.95) p-Value


Group A 318.0 2.87 8.57 × 10−18 (p < 0.05)
Pore diameter Group B 221.9 2.87 2.91 × 10−16 (p < 0.05)
Group C 279.3 2.87 3.06 × 10−17 (p < 0.05)
Group A 83.7 2.87 3.36 × 10−12 (p < 0.05)
WVTR Group B 129.7 2.87 5.29 × 10−14 (p < 0.05)
Group C 67.4 2.87 2.54 × 10−11 (p < 0.05)
Group A 1119.5 2.87 3.31 × 10−23 (p < 0.05)
Ref Group B 1971.7 2.87 1.18 × 10−25 (p < 0.05)
Group C 872.7 2.87 3.94 × 10−22 (p < 0.05)
Group A 64.4 2.87 3.90 × 10−11 (p < 0.05)
K Group B 124.8 2.87 7.66 × 10−14 (p < 0.05)
Group C 54.7 2.87 1.72 × 10−10 (p < 0.05)

Figure 4. Diagram of pore diameters of the fabric specimens.


Materials 2021, 14, 6205 13 of 24

As shown in Figure 4, of the three fabric specimen groups (A, B, and C), the pore
diameters of the fabric specimens in group A ((1) to (5)) showed higher values than those in
groups B ((6) to (10)) and C ((11) to (15)). This suggests that the pore size of the PP/Tencel
sheath/core yarns used as a warp yarn of group A was larger than that of the PET/Tencel
siro-fil and Tencel siro-spun yarns used as the warp yarns of groups B and C. Hence, the
sheath/core yarn has larger pores and voids. By contrast, the siro-fil and siro-spun yarns
have compact yarn structures, resulting in relatively small pore diameters in the yarns,
even though the fabric specimens were made from the same yarn count and fabric density.
These were verified by SEM and optical microscopy of the constituent warp yarns used
in the fabric specimens. Table 6 presents SEM and optical microscopy images of the yarn
specimens shown in Table 1. As shown in yarn specimen (1), which was used as a warp
yarn of fabric specimens (1) to (5) (group A), many air voids were observed in the sheath
and core. Round-shaped capillary channels at the border between the Tencel fibers and
the filaments in the core of the PP/Tencel sheath/core yarn were found, resulting in a
higher pore diameter of fabric group A ((1) to (5)) than fabric groups B ((6) to (10)) and
C ((11) to (15)). On the other hand, as shown in yarn specimen (2) in Table 6, a compact
yarn cross-section was observed in the PET/Tencel siro-fil yarn, resulting in a smaller
pore diameter of the fabric specimen group B ((6) to (10)). In yarn specimen (3) in Table 6,
small air voids were observed in the Tencel siro-spun yarn, resulting in relatively small
pore diameters of the fabric specimen group C ((11) to (15)). Regarding the pore size
of the fabric specimens according to the yarn structure in the weft direction, of the five
types of fabric specimens ((1) to (5), (6) to (10), and (11) to (15)), as shown in Figure 4,
the Coolmax® /Tencel sheath/core fabric (specimen (1)) and Coolmax® /bamboo spun
fabric (specimen (2)) exhibited larger pore diameters than the PET/Tencel siro-fil fabric
(specimen (3)). By contrast, the pore diameter of the siro-fil fabric (specimen (3)) was
smaller than that of the bamboo spun fabric (specimen (4)) and hi-multi PET filament
fabric (specimen (5)). These results suggest that sheath/core and spun yarn fabrics have
larger pores in the fabrics, whereas siro-fil has a compact yarn structure, which results
in small pores in the fabric. Relatively large air voids in the Coolmax® /bamboo spun
yarns observed in yarn specimen ((5)) in Table 6 were noted in the weft direction in fabric
specimen (2), resulting in relatively large pore diameters for fabric specimens (2), (7), and
(12), as shown in Table 4 and Figure 4. On the other hand, a compact yarn cross-section
was observed in yarn specimen (6) in Table 6, resulting in small pore diameters for fabric
specimens (3), (8), and (13), as shown in Table 4 and Figure 4. Similar to yarn specimen (5),
yarn specimen (7) showed a relatively compact yarn cross-section, as shown in Table 6.
The pore sizes in yarn specimen (7) were smaller than those of yarn specimen (5), resulting
in smaller pore diameters of fabric specimens (4), (9), and (14) than those of fabric specimens
(2), (7), and (12), as shown in Table 4 and Figure 4. Regarding the hi-multi PET filament of
the yarn specimen (8), many air voids were observed in the SEM and optical microscopy
images of the yarn cross-section. In addition, non-twisted parallel filament bundles were
observed in the SEM image of the yarn surface shown in Table 6, which produced fine
capillary channels along the filament bundles with many pores in the yarn cross-section, as
shown in the SEM and optical microscopy images of yarn specimen (8) in Table 6, resulting
in a high pore diameter of fabric specimens (5), (10), and (15) compared to siro-fil and
spun yarn fabrics made from yarn specimens (5), (6), and (7) in Table 6. According to
previous studies [45,46], the MVP is incorporated with micro and macro porosities, and
fine voids cause microporosity among the fibers in the yarns. By contrast, macro porosity
is produced from the void spaces among the threads in the fabric. The pore diameters
measured in this study were assessed as a measure of the fabric porosity considering both
micro voids among the fibers in the yarns and macro voids among the threads in the
fabrics. Therefore, the MVP was examined in relation to the pore diameter measured from
SEM images of the air void and capillary channels formed according to the different yarn
structures. In addition, the MVP of the fabrics made from the different yarn structures
Materials 2021, 14, 6205 14 of 24

was compared and discussed with its two methods (WVTR and Ref ) for measuring the
breathability, as shown in the next section.

Table 6. SEM images of cross-sections (×500) and surfaces (×150) of the yarns and optical microscopy (×300) [53].

Yarn
Optical Microscopy
Specimen Yarns SEM (Cross-Section) SEM (Surface)
(Cross-Section)
No

PP/Tencel
(1)
Sheath/core

PET/Tencel
(2)
Siro-fil

Tencel
(3)
Siro-spun

Coolmax/
(4) Tencel
Sheath/core

Coolmax/
(5) Bamboo
Spunyarn

PET/
(6) Tencel
Siro-fil
Materials 2021, 14, 6205 15 of 24

Table 6. Cont.

Yarn
Optical Microscopy
Specimen Yarns SEM (Cross-Section) SEM (Surface)
(Cross-Section)
No

Bamboo
(7)
Spunyarn

Hi-multi PET
(8)
75d/144f

PP DTY
(9)
100d/48f

3.2. WVTR of the Fabric Specimens Using Upright Cup Method


Figure 5 presents the WVTR of the 15 fabric specimens. The mean value of the five
specimens for the WVTR in groups A, B, and C was statistically significant, as shown
in Table 5. The mean values of the 15 specimens were plotted with the maximum and
minimum values of the five experimental data of each specimen, respectively. A comparison
of the WVTR according to the warp yarn structure (i.e., groups A, B, and C) revealed the
WVTR of group A to be higher than that of groups B and C because of the larger pore
diameter by the warp yarn (PP/Tencel sheath/core) of group A than groups B (PET/Tencel
siro-fil) and C (Tencel siro-spun), as shown in Figure 4. This was verified by SEM images
(Table 6), i.e., larger pores with a capillary channel between PP filament in core and Tencel
fibers in the sheath were observed in the yarn specimen (1) (a warp yarn of group A fabric
specimens), which resulted in a higher WVTR of group A fabrics. The WVTR of group C
was slightly higher than group B because of the larger pore diameter by the warp yarn
(Tencel siro-spun) of group C than group B (PET/Tencel siro-fil), as shown in Figure 4.
These results were consistent with Fohr et al. [54], who reported that the WVTR was
strongly dependent on the porosity and diffusion characteristics of the moisture vapor
particles, which is in agreement with the current findings.
Materials 2021, 14, 6205 16 of 24

Figure 5. WVTR of the specimens using the upright cup method.

The WVTR of the fabric specimens was examined according to the yarn structure in
the weft direction. As shown in Figure 5, of five fabric specimens ((1) to (5)) in group A,
the WVTR of fabric specimen (1) was higher than that of specimens (2) and (3). That of
fabric specimen (2) was slightly higher than that of specimen (3), because of the larger
pore diameter by the weft yarn (Coolmax/Tencel sheath/core yarn) of fabric specimen (1)
than fabric specimen (2) (Coolmax/bamboo spun yarn) and 3 (PET/Tencel siro-fil yarn),
and of fabric specimen (2) than (3). Similar results were shown in groups B ((6) to (10))
and C ((11) to (15)). On the other hand, the WVTR of fabric specimen (4) (as well as (9)
and (14)) inserted with bamboo spun yarn was the highest compared to the other fabric
specimens, which was attributed to the high absorption rate and the diffusivity of bamboo
spun yarn fabric with the appropriate pores in the yarns. As shown in Table 3, of the
15 fabric specimens, specimens (4), (9), and (14) composed of bamboo spun yarns in the
weft direction exhibited the highest absorption rate, which was partly responsible for the
highest WVTR of the fabrics. These results are in accordance with previous studies [55,56]
reporting that the water vapor transmission of hygroscopic fibrous materials was higher
than that of the materials that do not absorb moisture. This reduces the moisture built
up in the microclimate, enhancing moisture vapor transmission from human skin to the
environment. According to Li et al. [57], an increase in the WVTR by the absorption of
moisture vapor is mainly because the heat of sorption increases the temperature of the
fibrous assemblies, which in turn affects the moisture vapor transmission rate. The effect
of water vapor absorption on the WVTR can be explained in fabric specimens (5), (10), and
(15). As shown in Figure 5, the WVTR of fabric specimens (5), (10), and (15) showed the
lowest value compared to the other specimens. This is because the PET 75d/144f filaments
inserted in fabric specimens (5), (10), and (15) are hydrophobic and do not absorb moisture,
resulting in a low WVTR. Summarizing the WVTR measured by the upright cup method
according to the yarn structures in the warp and weft directions, the WVTR of the fabric
specimens divided into groups A, B, and C was dependent on the pore diameter of the
fabric, i.e., the WVTR of group A fabrics (specimens (1) to (5)) was higher than that of
group B (specimens (6) to (10)) and C (specimens (11) to (15)) fabrics, which was attributed
to the larger pore diameter of the group A fabrics. On the other hand, a study of the WVTR
of fabric specimens according to yarn structure in the weft direction revealed the WVTR to
be primarily dependent on the absorption rate of the constituent fibers in the fabric and
partly on the pore size of the fabric. The fabric specimens composed of bamboo fibers with
hygroscopic characteristics in the weft direction (i.e., high absorption rate) exhibited the
highest WVTR. In contrast, the fabric specimens composed of hydrophobic PET filaments
Materials 2021, 14, 6205 17 of 24

showed the lowest WVTR. Hence, in this study, the WVTR of the fabric measured using
the upright cup method according to the weft yarn structure and fiber characteristics was
strongly dependent on the hygroscopicity of the constituent fibers. On the other hand,
the WVTR of the fabric according to the warp yarn structure was governed by the pore
diameter of the fabric, i.e., dependent on the warp yarn structure.

3.3. Moisture Vapor Resistance of the Fabric Specimens by ISO 11092 Method
The ISO 11092 method uses a sweating guarded hot plate apparatus to simulate
moisture transport through the textile when worn next to the human skin. This model
measures the moisture vapor resistance of the fabric by measuring the evaporative heat
loss in the steady state. Its measuring mechanism is different from the upright cup method.
Figure 6 shows the moisture vapor resistance (Ref ) of the fabric specimens. The mean value
of the moisture vapor resistance of the 15 fabric specimens was statistically significant, as
shown in Table 5.

Figure 6. Ref of the specimens using the ISO 11092 method.

The moisture vapor resistance of the 15 fabric specimens according to the weft yarn
structure showed a distinctive result, and a similar trend among fabric groups A, B, and
C was observed, i.e., proportional to the pore diameters. As shown in Figure 6, fabric
specimens (1) and (5) in group A, (6) and (10) in group B, and (11) and (15) in group
C exhibited low moisture vapor resistance, i.e., superior breathability to other fabric
specimens. These results were attributed to the larger pore sizes (specimens (1), (5), (6), (10),
(11), and (15) in Figure 4) in the fabrics depending on the weft yarn structure. Hence, the
superior moisture vapor transmittance of fabric specimens (1), (6), and (11) was due to the
following: high pore diameters with the capillary channels (yarn specimen (4) in Table 6)
between the Coolmax® noncircular filaments in the core and Tencel fibers in the sheath
of the yarns, and the fine capillary channels (yarn specimen (8) in Table 6) between the
hi-multi PET filaments of fabric specimens (5), (10), and (15), which enable more moisture
vapor to be transmitted from the fabric toward the outside. On the other hand, fabric
specimens (3), (8), and (13) showed the highest moisture vapor resistance, i.e., inferior
breathability to that of the other fabric specimens. This was attributed to the low pore
diameters (specimens (3), (8), and (13) in Figure 4) in the fabrics because of the compact yarn
structure of the PET/Tencel siro-fil yarns (yarn specimen (6) in Table 6), which prevents
moisture vapor from being pushed by the moisture vapor pressure. In addition, the
moisture vapor resistance of the fabric specimens composed of spun yarns in the weft
direction ((2) and (4) in group A, (7) and (9) in group B, and (12) and (14) in group C)
was higher, i.e., showed inferior breathability to that of the fabric specimens with the
Materials 2021, 14, 6205 18 of 24

sheath/core and hi-multi PET filament. This was attributed to the smaller pore diameters
(Figure 4) of the spun yarn fabrics than the sheath/core and hi-multi PET fabrics and partly
to the higher hygroscopicity of the bamboo fibers. According to previous studies [1,58],
the correlation between the diffusion process and moisture vapor transmission can be
explained by the swelling of the fibers due to the affinity of the hydrophilic fiber molecules.
Hence, as moisture vapor diffuses through the fibers in the fabric, it is absorbed by the
fibers, causing fiber swelling and reducing the size of the air (void) spaces between the
fibers. This delays the diffusion process, which reduces the rate of moisture vapor particle
movements [58]. This explains why hygroscopic spun yarn fabrics (specimens (2) and (4)
in group A, (7) and (9) in group B, and (12) and (14) in group C) exhibited higher moisture
vapor resistance than the filament fabrics (specimens (5), (10), and (15)). This explains
why the Ref measured by the sweating guarded hot plate method differs from the WVTR
measured by the upright cup method, which is due to the difference of mechanism between
the two measuring methods, i.e., the transmission of moisture vapor by forced convection
due to CaCl2 in the upright cup and the diffusion process by the free convection of wet heat
particles in a sweating guarded hot plate apparatus. Furthermore, these results indicate
that the sweating guarded hot plate method is appropriate to measure the breathability of
coated (or laminated) nylon (or PET) fabrics, whereas the upright cup method is suitable
for non coated ordinary fabrics.

3.4. Thermal Conductivity of the Fabric Specimens


The sweating guarded hot plate (ISO 11092 method) applies the transmission (dif-
fusion, or convection) of wet thermal particles to measure the moisture vapor resis-
tance, which is similar to the transmission by the conduction of dry thermal particles.
Understanding how the moisture vapor resistance of the fabrics measured using the diffu-
sion of wet thermal particles (sweating) is associated with the thermal resistance (conduc-
tivity) measured by the conduction of dry thermal particles is very important for examining
dry and wet thermal wear comforts and their correlations according to the yarn structure
of the fabric and the thermal characteristics of constituent fibers. Figure 7 presents the
thermal conductivity of 15 fabric specimens. The mean value of the thermal conductivity
of the 15 fabric specimens was statistically significant, as shown in Table 5.

Figure 7. Thermal conductivity of the fabric specimens.

As shown in Figure 7, the thermal conductivity of fabric specimens ((3), (8), and
(13)) with PET/Tencel siro-fil in the weft direction was the highest, followed by the fabric
Materials 2021, 14, 6205 19 of 24

specimens ((4), (9), and (14)) composed of bamboo spun yarns. These fabrics have smaller
pore diameters (Figure 4) than the other fabrics. Small air voids in the low porosity fabric
due to the compact yarn structure of the PET/Tencel siro-fil and bamboo spun yarns cannot
entrap the neighboring air in the compact yarns and their fabrics. This enables easy heat
conduction from the inner layer to the outer one of the fabrics, resulting in higher thermal
conductivity than the other fabrics. On the other hand, fabric specimens (5), (10), and (15)
in groups A, B, and C composed of the hi-multi PET in the weft direction showed the lowest
thermal conductivity because of the larger pore diameter in the filament bundles (Figure 4),
which entraps the neighboring air and prevents heat flow from the inner layer to the outer
layer of the fabrics, resulting in lower thermal conductivity than the other fabrics. The lower
thermal conductivity of these fabrics was attributed partly to the lower thermal conductivity
of the PET filament than that of the Tencel and bamboo fibers (p. 150) [59,60]. In addition,
the thermal conductivity of the Coolmax® /Tencel sheath/core fabrics (specimens (1),
(6), and (11)) was lower than that of the Coolmax® /bamboo spun, bamboo spun, and
PET/Tencel siro-fil fabrics (specimens (2) to (4), (7) to (9), and (12) to (14)) in groups A, B,
and C, which was also attributed to the larger pore diameters of the sheath/core fabrics than
the other fabrics. In particular, fabric specimens (1), (6), and (11) with Coolmax® /Tencel
sheath/core yarns in the weft direction exhibited higher thermal conductivity than the hi-
multi PET fabric specimens (5), (10), and (15), despite the sheath/core fabrics having much
larger pore diameters than the hi-multi PET fabrics. The higher thermal conductivity
of the Tencel fibers than the PET filament was considered the cause [59,60], because
it may compensate for the thermal conductivity of the two fabrics, resulting in higher
thermal conductivity of the Coolmax® /Tencel sheath/core fabrics. Overall, the thermal
conductivity of the composite yarn fabrics is influenced mainly by the pore size according to
the yarn structure and partly by the thermal characteristics of the constituent fibers. These
results are in agreement with previous findings [61–63]. Das et al. [61,62] reported that
cotton/acrylic blend fabric exhibited low thermal conductivity because of its high porosity.
Das and Istiaque [63] published a similar result for the hollow yarn fabric. The thermal
properties of the composite yarn fabrics were influenced mainly by the fabric porosity; the
sheath/core yarn fabric with a noncircular cross-sectional filament exhibited low thermal
conductivity. Furthermore, the bulky yarn fabric with noncircular cross-sectional fibers
showed low thermal conductivity because of the high porosity and hairy and crimpy
constituent fibers. Thus, the thermal conductivity of the composite yarn fabrics was
governed partly by the thermal characteristics of the constituent fibers. Finally, considering
the market application of ecofriendly fiber-embedded fabrics for high-performance clothing
with good wear comfort, the Coolmax® /Tencel sheath/core fabrics are useful for winter
clothing with a warm feel due to good breathability with low thermal conductivity. Bamboo
and Coolmax® /bamboo spun yarn fabrics are suitable for summer clothing with a cool
feeling because of their high thermal conductivity and good breathability.

3.5. Correlation Analysis between Wear Comfort Characteristics of Fabrics


Heat and moisture vapor transmission is essential for characterizing the thermo-
physiological behavior of fabrics to assess their wear comfort for the design of high-
performance clothing. They are characterized by two parameters: resistance to dry heat
(inverse of thermal conductivity) and resistance to moisture vapor. Therefore, correlation
analysis was carried out to determine the interrelationships between the thermal conduc-
tivity and moisture vapor resistance according to the yarn structure and constituent fiber
characteristics with the measuring method, as well as the fabric absorption rate. Table 7
lists the correlation coefficient between each parameter. The correlation between the mois-
ture vapor resistance (Ref ) and pore diameter was significant at the 99% confidence level.
In addition, the correlation between Ref and thermal conductivity (K) was significant at the
99% confidence level. Therefore, the results obtained with significant correlation at the 99%
confidence level were graphed. A trend line was added to the graph shown in Figure 8.
Materials 2021, 14, 6205 20 of 24

Table 7. Correlation coefficient between each parameter of the wear comfort characteristics of the fabrics.

Water Vapor Moisture Vapor Thermal Absorption


Pore Diameter
Transmission Rate Resistance Conductivity Rate
(µm)
(g/m2 ·h) (m2 ·Pa/W) (m·Pa/W) (%)
Pore diameter (D) 1
Water vapor transmission rate(WVTR) 0.354 1
Moisture vapor resistance (Ref ) −0.734 a 0.264 1
Thermalconductivity (K) −0.545 b 0.412 b 0.872 a 1
Absorptionrate (A) −0.226 −0.159 0.161 0.301 1
Note: a significant at the 0.01 level; b significant at the 0.05 level.

Figure 8. Correlation diagram between each wear comfort characteristic of the fabrics: (a) moisture
vapor resistance and pore diameter; (b) moisture vapor resistance and thermal conductivity.

Figure 8 presents a diagram showing the interrelationships among the moisture vapor
resistance, thermal conductivity, and pore diameter of the fabric specimens.
As shown in Table 7 and Figure 8a correlation coefficient between the moisture
vapor resistance (Ref ) and pore diameter was observed as an inverse correlation (−0.73).
Hence, the moisture vapor resistance is influenced by the pore diameters of the fabrics and
partly by the hygroscopicity of the constituent fibers, which is consistent with previous
results [54–58]. Fohr et al. [54] and Lomax [64] reported that moisture vapor through a
textile medium diffuses in two ways: through the air space (pore) and along the fiber itself.
In addition, the diffusion rate along the textile material depends on the porosity, and the
moisture vapor diffusivity of the fiber is affected by the absorption and its hygroscopicity.
Materials 2021, 14, 6205 21 of 24

Barnes et al. [55] and Hong et al. [56] reported that the moisture vapor transmission
of the hygroscopic fibrous materials was higher than that of non-absorbing materials.
On the other hand, the correlation between Ref and WVTR was very low (0.26) because
the measuring mechanisms of the two methods were different. The upright cup method
by CaCl2 measures the moisture vapor transport by Fick’s law, whereas Ref measures
the wet thermal transport by the difference in vapor pressure. This is in accordance with
McCullogh et al. [26], who reported that the correlations between Ref and the upright and
inverted cup methods of WWB (waterproof, windproof, and breathable) shell fabrics made
of nylon and PET were very low (−0.59 and 0.1, respectively). The results are in contrast
to those by Huang and Qian [29] and Gorjanc et al. [4]. Huang and Qian [29] reported
that the correlation coefficients between Ref and the upright and inverted cup methods
using ordinary and breathable fabrics were −0.87 and −0.66, respectively. Gorjanc et al. [4]
reported that the correlation coefficient between the water cup method and moisture
vapor resistance by the Permetest method using cotton and cotton-stretch fabrics was >0.9.
Previous findings [4,26,29] suggested that the breathability characteristics according to the
constituent yarn structure were dependent on the measuring method. In addition, the
correlation coefficient between Ref and thermal conductivity (K) was 0.87, meaning that
the K values measured by dry heat transmission and Ref measured by wet heat movement
are governed by a similar mechanism: the movement of heat and moisture vapor particles,
as shown in Figure 8b. Moreover, they are dependent on the pore size of the fabrics and
the hygroscopicity and thermal characteristics of the constituent fibers.

4. Conclusions
The MVP and thermal wear comfort of the ecofriendly fiber-embedded woven fabrics
were examined according to the measuring method in relation to their absorption and
thermal properties with the constituent yarn structure and fiber characteristics. Fifteen
fabric specimens composed of sheath/core, siro-fil, siro-spun, and ring-spun yarns were
prepared using bamboo and Tencel as ecofriendly fibers, as well as PP, PET, and Coolmax®
as core filaments. The WVTR and moisture vapor resistance (Ref ) were measured using
the upright cup method by CaCl2 and a sweating guarded hot plate method, respectively.
The WVTR measured by the upright cup method was dependent primarily on the hy-
groscopicity of the eco-friendly constituent fibers (bamboo and Tencel) in the yarns and
partly influenced by the pore size in the fabric depending on the yarn structure. Of the 15
fabric specimens, the bamboo spun yarn fabrics exhibited superior WVTR, followed by the
Coolmax® /Tencel sheath/core fabrics.
On the other hand, the hi-multi PET fabrics showed inferior WVTR, followed by the
PET/Tencel siro-fil fabrics. The moisture vapor resistance (Ref ) by the sweating guarded
hot plate method was governed mainly by the pore size in the fabric and partly by the
hygroscopicity of the constituent ecofriendly fibers. These results suggest that ecofriendly
fibers, bamboo, and Tencel can contribute to environmental improvement and wear comfort
related to water and moisture vapor transmission. The moisture vapor resistance of hi-
multi PET filament fabrics was the lowest, i.e., best, followed by the Coolmax® /Tencel
sheath/core fabrics. In contrast, the PET/Tencel siro-fil fabric was the highest, i.e., showed
inferior breathability. Ref measured by the sweating guarded hot plate method differed
from the WVTR measured by the upright cup method due to the difference in measuring
mechanism between the two methods.
The thermal conductivity of the composite yarn fabrics was influenced by pore size in
the fabric and the thermal characteristics of the constituent ecofriendly fibers in the yarns.
The hi-multi PET filament fabrics exhibited the lowest thermal conductivity, followed by
the Coolmax® /Tencel sheath/core fabrics, whereas the PET/Tencel siro-fil fabrics showed
the highest thermal conductivity, followed by the bamboo spun yarn fabrics. These results
were verified by correlation analysis. The correlation coefficient between the moisture vapor
resistance and pore diameter was −0.73. The correlation coefficients between the moisture
vapor resistance (Ref ) and thermal conductivity (K) and between K and pore diameter
Materials 2021, 14, 6205 22 of 24

were 0.87 and −0.55, respectively. On the other hand, the correlation coefficient between
the WVTR and moisture vapor resistance was very low (0.27), which was attributed
to the different mechanisms between the two measuring methods, i.e., transmission of
moisture vapor by forced convection in an upright cup and the diffusion process viathe
free convection of wet heat particles in a sweating guarded hot plate apparatus.
Lastly, considering the market application for high-performance ecofriendly cloth-
ing with good wear comfort, the Coolmax® /Tencel sheath/core yarn fabrics are useful
for winter clothing with a warm feeling due to the good breathability with low thermal
conductivity. Bamboo and Coolmax® /bamboo spun yarn fabrics are suitable for sum-
mer clothing with a cool feeling because of their high thermal conductivity and good
breathability. Although based on the MVP and thermal wear comfort obtained in this
study, the market application of Tencel fibers for winter outdoor clothing and bamboo
fibers for summer outdoor clothing is of practical use for engineering high-performance
fabrics. These results suggest that an increase in the consumption of ecofriendly fibers with
a decrease in synthetic fibers can reduce environmental pollution in the textile industry.

Funding: This research is supported by Ministry of SMEs and Startups (Project Number: S240111).
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: The data presented in this study are available on request from the
corresponding author.
Conflicts of Interest: The author declares no conflict of interest.

References
1. Deopura, B.L.; Alagirusamy, R.; Joshi, M.; Gupta, B. Polyesters and Polyamides, 1st ed.; Woodhead Publishing Limited: Cambridge,
UK, 2008; pp. 171–199.
2. Ozgen, B. New biodegradable fibers, yarn properties and their applications in textiles: A review. Ind. Text. 2012, 63, 3–7.
3. Rego, J.M.; Verdu, P.; Nieto, J.; Blanes, M. Comfort Analysis of Woven Cotton/Polyester Fabrics Modified with a New Elastic
Fiber, Part 2: Detailed Study of Mechanical, Thermo-Physiological and Skin Sensorial Properties. Text. Res. J. 2009, 80, 206–215.
[CrossRef]
4. Gorjanc, D.; Dimitrovski, K.; Bizjak, M. Thermal and water vapor resistance of the elastic and conventional cotton fabrics.
Text. Res. J. 2012, 82, 1498–1506. [CrossRef]
5. Lee, S.; Obendorf, S.K. Statistical modeling of water vapor transport through woven fabrics. Text. Res. J. 2012, 82, 211–219.
[CrossRef]
6. Kim, J.; Spivak, S. Dynamic Moisture Vapor Transfer Through Textiles. Text. Res. J. 1994, 64, 112–121. [CrossRef]
7. Cubric, I.S.; Skenderi, Z.; Havenith, G. Impact of raw material, yarn and fabric parameters, and finishing on water vapor
resistance. Text. Res. J. 2013, 83, 1215–1228. [CrossRef]
8. Ruckman, J. Water vapour transfer in waterproof breathable fabrics: Part I: Under steady-state conditions. Int. J. Cloth. Sci. Technol.
1997, 9, 10–22. [CrossRef]
9. Ruckmann, J.E. Water vapour transfer in waterproof breathable fabrics: Part II: Under windy conditions. Int.J. Cloth. Sci. Technol.
1997, 9, 23–33. [CrossRef]
10. Ruckmann, J.E. Water vapour transfer in waterproof breathable fabrics: Part III:under rainy and windy conditions. Int.J. Cloth.
Sci. Technol. 1997, 9, 141–153. [CrossRef]
11. Ruckman, J.; Murray, R.T.; Choi, H.S. Engineering of clothing systems for improved thermophysiological comfort. Int. J. Cloth.
Sci. Technol. 1999, 11, 37–52. [CrossRef]
12. Ren, Y.J.; Ruckman, J.E. Effect of condensation on water vapour transfer through waterproof breathable fabrics. J. Coat.Fabr. 1999,
29, 20–36.
13. Ren, Y.J.; Ruckman, J.E. Water Vapour Transfer in Wet Waterproof Breathable Fabrics. J. Ind. Text. 2003, 32, 165–175. [CrossRef]
14. Ren, Y.J.; Ruckman, J.Y. Condensation in three-layer waterproof breathable fabrics for clothing. Int. J. Cloth. Sci. Technol. 2004, 16,
335–347. [CrossRef]
15. Yoo, S.J.; Kim, E.A. Wear trial assessment of layer structure effects on vapor permeability and condensation in a cold weather
clothing ensemble. Text. Res. J. 2012, 82, 1079–1091. [CrossRef]
16. Yoo, H.S.; Kim, E.A. Effects of multilayer clothing system array on water vapor transfer and condensation in cold weather
clothing ensemble. Text. Res. J. 2008, 78, 189–197. [CrossRef]
17. Lomax, G.R. Hydrophilic polyurethane coatings. J. Coat. Fabr. 1990, 20, 88–107. [CrossRef]
Materials 2021, 14, 6205 23 of 24

18. Salz, P. Testing the Quality of Breathable Textiles. Performance of Protective Clothing: Second Symposium; ASTM Special Technical
Publication, 989; Mandorf, F.Z., Sagar, R., Bielson, A.P., Eds.; American Society for Testing and Materials: Philadelphia, PA, USA,
1988; p. 295.
19. Yoo, H.; Hu, Y.; Kim, E. Effects of Heat and Moisture Transport in Fabrics and Garments Determined with a Vertical Plate
Sweating Skin Model. Text. Res. J. 2000, 70, 542–549. [CrossRef]
20. Gibson, P. Factors Influencing Steady-State Heat and Water Vapor Transfer Measurements for Clothing Materials. Text. Res. J.
1993, 63, 749–764. [CrossRef]
21. Gibson, P.; Kendrick, C.; Rivin, D.; Sicuranza, L.; Charmchi, M. An Automated Water Vapor Diffusion Test Method for Fabrics,
Laminates, and Films. J. Coat. Fabr. 1995, 24, 322–345. [CrossRef]
22. Gibson, P.; Kendrick, C.; Rivin, D.; Charmchii, M. An Automated Dynamic Water Vapor Permeation Test Method. In Performance
of Protective Clothing: Sixth Volume; ASTM International: West Conshohocken, PA, USA, 1997; pp. 93–107. [CrossRef]
23. Dolhan, P.A. A Comparison of Apparatus Used to Measure Water Vapour Resistance. J. Coat. Fabr. 1987, 17, 96–109. [CrossRef]
24. Congalton, D. Heat and moisture transport through textiles and clothing ensembles utilizing the Hohenstein skin model. J. Coat.
Fabr. 1999, 28, 183–196.
25. Gretton, J.; Brook, D.; Dyson, H.; Harlock, S. A Correlation between Test Methods Used to Measure Moisture Vapour Transmission
through Fabrics. J. Coat. Fabr. 1996, 25, 301–310. [CrossRef]
26. McCullough, E.A.; Kwon, M.; Shim, H. A comparison of standard methods for measuring water vapour permeability of fabrics.
Meas. Sci. Technol. 2003, 14, 1402–1408. [CrossRef]
27. Huang, J. Review of test methods for measuring water vapor transfer properties of fabrics. Cell.Polym. 2007, 26, 167–191.
[CrossRef]
28. Huang, J. Sweating guarded hot plate test method. Polym. Test. 2006, 25, 709–716. [CrossRef]
29. Huang, J.; Qian, X. Comparison of Test Methods for Measuring Water Vapor Permeability of Fabrics. Text. Res. J. 2008, 78, 342–352.
[CrossRef]
30. Chen, Q.; Fan, J.T.; Sarkar, M.K.; Bal, K. Plant-based biomimetic branching structures in knitted fabrics for improved comfort-
related properties. Text. Res. J. 2011, 81, 1039–1048. [CrossRef]
31. Chen, Q.; Fan, J.; Sarkar, M.; Jiang, G. Biomimetics of Plant Structure in Knitted Fabrics to Improve the Liquid Water Transport
Properties. Text. Res. J. 2009, 80, 568–576. [CrossRef]
32. Okada, H. Sweat-Absorbent Textile Fabric. U.S. Patent 4,530,873A, 1985.
33. Strauss, I.; Rankin, S.A., Jr. Fabric for recreational clothing. U.S. Patent 5,050,406A, 1991.
34. Lee, Y.K. Method for making fabric with excellent water transition ability. U.S. Patent 6,381,994B1, 2002.
35. Yeh, P. Fabric for moisture management. U.S. Patent 6,509,285B1, 2003.
36. Burrow, T.R.; Firgo, H. Wicking fabric and garment made therefrom. U.S. Patent 8127575B2, 2012.
37. Kim, H.A.; Kim, S.J. Physical Properties and Wear Comfort of Bio-Fiber-Embedded Yarns and their Knitted Fabrics According to
Yarn Structures. Autex Res. J. 2019, 19, 279–287. [CrossRef]
38. Kim, H.A. Physical Property of PTT/Wool/Modal Air Vortex Yarns for High Emotional Garment. J. Korean Soc. Cloth. Text. 2015,
39, 877–884. [CrossRef]
39. Kim, H.A.; Kim, S.J. Flame retardant, anti-static and wear comfort properties of modacrylic/Excel® /anti-static PET blend yarns
and their knitted fabrics. J. Text. Inst. 2019, 110, 1318–1328. [CrossRef]
40. Kim, H.A. Tactile hand and wear comfort of flame-retardant rayon/anti-static polyethylene terephthalate imbedded woven
fabrics. Text. Res. J. 2019, 89, 4658–4669. [CrossRef]
41. Kim, H.A.; Kim, S.J. Hand and Wear Comfort of Knitted Fabrics Made of Hemp/Tencel Yarns Applicable to Garment. Fibers Polym.
2018, 19, 1539–1547. [CrossRef]
42. Kim, H.A.; Kim, S.J. Mechanical Properties of Micro Modal Air Vortex Yarns and the Tactile Wear Comfort of Knitted Fabrics.
Fibers Polym. 2018, 19, 211–218. [CrossRef]
43. Kim, H.A.; Kim, S.J. Flame-Retardant and Wear Comfort Properties of Modacrylic/FR-Rayon/Anti-static PET Blend Yarns and
Their Woven Fabrics for Clothing. Fibers Polym. 2018, 19, 1869–1878. [CrossRef]
44. Kim, H.A. Water/moisture vapor permeabilities and thermal wear comfort of the Coolmax® /bamboo/tencel included PET and
PP composite yarns and their woven fabrics. J. Text. Inst. 2020. [CrossRef]
45. Saricam, C.; Kalaoglu, F. Investigation of the wicking and drying behaviour of polyester woven fabrics. Fibers Text. East Eur. 2014,
22, 73–78. [CrossRef]
46. Tashkandi, S.; Wang, L.; Kanesalingam, S. An investigation of thermal comfort properties of Abaya woven fabrics. J. Text. Inst.
2013, 104, 830–837. [CrossRef]
47. Varshney, R.K.; Kothari, V.K.; Dhamija, S. A study on thermophysiological comfort properties of fabrics in relation to constituent
fiber fineness and cross-sectional shapes. J. Text. Inst. 2010, 101, 495–505. [CrossRef]
48. Vimal, J.T.; Murugan, R.; Subramaniam, V. Effect of Weave Parameters on the Air Resistance of Woven Fabrics. Fibres Text. East. Eur.
2016, 24, 67–72. [CrossRef]
49. Wei, J.; Xu, S.; Liu, H.; Zheng, L.; Qian, Y. Simplified modal for predicting fabric thermal resistance according to its micro-structural
parameters. Fibers Text. East Eur. 2015, 23, 57–60. [CrossRef]
Materials 2021, 14, 6205 24 of 24

50. Kim, H.A.; Kim, S.J. Moisture and thermal permeability of the hollow textured PET imbedded woven fabrics for high emotional
garments. Fibers Polym. 2016, 17, 427–438. [CrossRef]
51. Beskisiz, E.; Ucar, N.; Demir, A. The Effects of Super Absorbent Fibers on the Washing, Dry Cleaning and Drying Behavior of
Knitted Fabrics. Text. Res. J. 2009, 79, 1459–1466. [CrossRef]
52. Ucar, N.; Beskisiz, E.; Demir, A. Design of a Novel Filament with Vapor Absorption Capacity Without Creating Any Feeling of
Wetness. Text. Res. J. 2009, 79, 1539–1546. [CrossRef]
53. Kim, H.A. Wear Comfort of Woven Fabrics for Clothing Made from Composite Yarns. Fibers Polym. 2021, 22, 2344–2353. [CrossRef]
54. Fohr, J.; Couton, D.; Treguier, G. Dynamic Heat and Water Transfer Through Layered Fabrics. Text. Res. J. 2002, 72, 1–12.
[CrossRef]
55. Barnes, J.C.; Holcombe, B.V. Moisture sorption and transport in clothing during wear. Textile Research Journal 1996, 66, 777–786.
[CrossRef]
56. Hong, K.; Hollies, N.R.S.; Spivak, S.M. Dynamic moisture vapour transfer through textile. Textile Research Journal 1988, 68, 697–706.
[CrossRef]
57. Li, Y.; Holcombe, B.V.; Scheider, A.M.; Apcar, F. Mathematical modeling of the coolness to the touch of hygroscopic fabrics.
J. Text. Inst. 1993, 84, 267–273. [CrossRef]
58. Pause, B. Measuring the Water Vapor Permeability of Coated Fabrics and Laminates. J. Coat. Fabr. 1996, 25, 311–320. [CrossRef]
59. Bona, M. Textile Quality: Physical Methods of Product and Process Control; Texilia: Biella, Italy, 1994; pp. 83–150.
60. Lavate, S.S.; Burji, M.C.; Patil, S. Study of yarn and fabric properties produced from modified viscose Tencel, Excel, Modal and
Their Comparison against cotton. Text. Today 2016, 9, 36–42.
61. Das, A.; Kothari, V.K.; Balaji, M. Studies on cotton-acrylic bulked yarns and fabrics. Part I: Yarn characteristics. J. Text. Inst. 2007,
98, 261–267. [CrossRef]
62. Das, A.; Kothari, V.K.; Balaji, M. Studies on cotton–acrylic bulked yarns and fabrics. Part II: Fabric characteristics. J. Text. Inst.
2007, 98, 363–376. [CrossRef]
63. Das, A.; Ishtiaque, S.M. Comfort characteristics of fabrics containing twist-less and hollow fibrous assemblies in weft. J. Text. App.
Technol. Manag. 2004, 3, 1–7.
64. Lomax, G.R. The design of waterproof, water vapour-permeable fabrics. J. Coat. Fabr. 1985, 15, 40–49. [CrossRef]

You might also like