3D Composite Flexural Properties

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Materials & Design 212 (2021) 110267

Contents lists available at ScienceDirect

Materials & Design


journal homepage: www.elsevier.com/locate/matdes

Research on development of 3D woven textile-reinforced composites


and their flexural behavior
Yajun Liu b, Canyi Huang b, Hong Xia c, Qing-Qing Ni a,c,⇑
a
Key Laboratory of Advanced Textile Materials and Manufacturing Technology, Ministry of Education, Zhejiang Sci-Tech University, Hangzhou 310018, Zhejiang, China
b
Interdisciplinary Graduate School of Science and Technology, Shinshu University, Ueda 386-8567, Japan
c
Department of Mechanical Engineering and Robotics, Shinshu University, Ueda 386-8567, Japan

h i g h l i g h t s g r a p h i c a l a b s t r a c t

 A new weaving technology with a


modified heddle position system
based on a self-built three-
dimensional weaving loom is
designed.
 The new weaving technology has
great potential for manufacturing
various three-dimensional woven
structures effectively and efficiently.
 Four typical three-dimensional
woven textiles are fabricated
successfully based on this weaving
system for the first time.
 The relationship among woven
structure, flexural properties and
failure modes of the three-
dimensional woven textile-reinforced
composites are studied
comprehensively.

a r t i c l e i n f o a b s t r a c t

Article history: In this study, a new weaving technology with a modified heddle position system based on a self-built
Received 23 August 2021 three-dimensional (3D) weaving loom is designed, and four typical 3D woven-structure textile groups
Revised 31 October 2021 are manufactured: layer-to-layer orthogonal weaving, through-thickness orthogonal weaving, layer-to-
Accepted 20 November 2021
layer angle-interlock weaving, and through-thickness angle-interlock weaving. The new weaving tech-
Available online 22 November 2021
nology has great potential for manufacturing various 3D woven structures effectively and efficiently.
The fabricated 3D woven textile-reinforced epoxy-resin composites undergo quasi-static three-point
Keywords:
bending tests to study the influence of the woven structure on the flexural performance and failure
Three-dimensional textile weaving
technology
modes along the textile warp and weft directions. The composites along the weft direction (weft-
Textile structure direction beams) have a larger flexural modulus but smaller failure strain compared with the warp direc-
Fiber-reinforced composite tion (warp-direction beams) for all woven-structure types. Among the designed 3D textile composites,
Flexural test the angle-interlock woven structures have a larger flexural strength (50%), modulus (40%), and failure
Failure mode resistance than have the orthogonal-woven structures. Overall, the through-thickness angle-interlock
woven structure has the best flexural-failure resistance among all textile structures, and is the optimal
structural design based on this modified weaving technology.
Ó 2021 Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://
creativecommons.org/licenses/by-nc-nd/4.0/).

⇑ Corresponding author at: Department of Mechanical Engineering and Robotics, Shinshu University, Ueda 386-8567, Japan
E-mail address: [email protected] (Q.-Q. Ni).

https://doi.org/10.1016/j.matdes.2021.110267
0264-1275/Ó 2021 Published by Elsevier Ltd.
This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
Y. Liu, C. Huang, H. Xia et al. Materials & Design 212 (2021) 110267

1. Introduction ral response of 3D woven textile composite panels with two


woven-structure types using the end notch flexural test and found
For many years, fiber-reinforced polymer composites have been that textile structure has a clear influence on crack propagation
widely used in automobile, aerospace, civil, and many other fields and failure mode under flexural loading. Turner [18] studied the
owing to their light weight, excellent mechanical performance, and effect of through-thickness reinforcement density on the collapse
easy manufacturing process. Traditional laminated composites behavior of a 3D TTOW composite under out-of-plane bending.
have a weak inter-laminar connection when subjected to transver- Gerlach [19] studied the effect of binder-yarn volume fraction on
sal loading, and, as such, fiber or yarn is introduced in the through- the through-thickness properties of a 3D through-thickness
thickness direction to improve mechanical strength. Many tech- angle-interlock woven (TTAIW) composite and found that a high
niques have been developed to achieve this, such as stitching binder-yarn volume fraction increases the delamination resistance
[1,2], tufting [3], and Z-pinning [4]. Three-dimensional (3D) textile under transversal loading. Behera [20,21] studied the mechanical
structures with an additional yarn group in the through-thickness behavior of 3D woven composites and two-dimensional (2D) lam-
direction can be used to improve both the through-thickness inated composites and found that the former has considerably
strength and delamination resistance of composites [5–7]. superior out-of-plane mechanical properties; however, only 3D
In essence, 3D textiles are defined as textiles with a 3D shape or through-thickness angle-interlock and 3D through-thickness
with 3D inner-fiber interlacement [7]. On the basis of the latter, 3D orthogonal structures were compared with their 2D laminate
woven textiles are defined as textiles developed with weaving counterparts, and the layer-to-layer woven structure was not
principle, having a substantial dimension in the thickness direction reported. Zhang [22] examined various hybrid structures with dif-
formed by layers of textiles or yarns, and containing a set of yarns ferent percentages and layups of the constituent fibers in a 3D
(binder yarn or Z yarn) running in the thickness direction, binding orthogonal-woven composite subjected to flexural loading. Other
or interlacing the layers into a structural integrity. These 3D woven researchers [23,24] have examined small sections of 3D woven
structures can be classified into several structure groups, i.e., structures to clarify the relationship between the woven structure
orthogonal, angle interlock, multilayer, through thickness, and and its mechanical properties under transversal loading. Many
layer-to-layer, and they can be fabricated with a conventional or researchers have studied the flexural performance of 3D textile
modified weaving loom as well as with a 3D weaving loom[8]. composites with various structure and weaving parameters, the
Unfortunately, manufacturing 3D woven textiles has a number of results of which suggest that weaving pattern seriously influences
limitations due to the problems associated with weaving technolo- mechanical properties. Unfortunately, research is lacking concern-
gies, fiber properties, and textile structures [6,9]. For a traditional ing systematic comparisons of the four typical 3D woven struc-
weaving loom, two yarn groups are mainly used to create textiles, tures, i.e., layer-to-layer orthogonal weaving (LLOW), layer-to-
i.e., warp and weft yarn, and part of the warp or weft yarns served layer angle-interlock weaving (LLAIW), TTOW, and TTAIW, and
as binder yarns to form the 3D woven textile structure, whereas, the flexural properties and failure modes under transversal loading
for a 3D weaving loom, three yarn groups are used, i.e., warp, weft, among the four main 3D woven structures both need to be
and binder yarn. Various 3D woven structures have been manufac- clarified.
tured on the basis of a variety of weaving looms. In some research In this study, a new weaving technology with a modified heddle
[10–12], 3D woven textiles were manufactured on a traditional position system based on a self-built 3D weaving loom is proposed,
weaving loom, and, as such, yarn selection, density, and whole tex- and four typical 3D woven textile structures with different binder-
tile thickness and structure were limited. With respect to structure, yarn paths and undulations are designed and manufactured based
through-thickness orthogonal-woven (TTOW) textiles are mostly on the said system for the first time. The flexural properties of the
fabricated with a 3D weaving loom, whereas angle-interlock hybrid glass/aramid 3D woven textile-reinforced epoxy-resin com-
woven textiles are mostly fabricated with a traditional or modified posites are systematically investigated as well as the relationship
weaving loom owing to their relative structural complexity. Partic- between woven structure, flexural mechanical performance, and
ularly, 3D woven composites reinforced with various and complex failure mode.
textile structures have drawn much attention from researchers due
to their specific engineering applications, for example, one of the
pioneering applications is aircraft engine fan blade [13,14]. The 2. Experimental
3D woven preforms applied for the composite fan blade have
near-net shape structure, large thickness, and a complex geometry 2.1. Three-dimensional weaving loom
with different woven structures in different sections. Jacquard
loom is now used to produce complex 3D woven textiles due to In a conventional 2D weaving process, as shown in Fig. 1(a), the
its specific shedding mechanism [15,16]. Research is currently warp yarn runs in the direction of the weaving loom, passes
lacking concerning complex woven structures such as angle- through the heddle eyes, and moves upwards or downwards with
interlock woven structures fabricated with a 3D loom because of the heddle frames to form the shed, whereas the weft yarn runs
weaving technology limitations. transversely from one side to the other side of the loom and passes
For 3D textile-reinforced composites, the failure mode is com- through the shed to interlace with the warp yarn to form a 2D tex-
plex because of the anisotropic characteristics. Different failure tile. Only one shed is opened in each shedding process, and one
modes can yield simultaneously or successively under various group of weft yarn passes through in each weft insertion process.
loadings until the composite structure fails. Out-of-plane loading In contrast, for a traditional 3D weaving process, as shown in
is more likely than any other loading type to introduce delamina- Fig. 1(b), the warp yarn and binder yarn run in the loom direction,
tion failure in fiber-reinforced composites, and thus it is important and the weft yarn runs transversely from one side to the other side
to study the transversal loading behavior and its failure modes on of the loom; the binder yarn passes through the heddle eyes and
3D woven-structure textile-reinforced composites. Many research- moves upwards or downwards with the heddle frames to form
ers have examined the influence of design parameters on flexural sheds. During the shedding process, the warp yarn is stable; more-
performance and failure modes, such as reinforcement density, over, there are several warp-yarn layers, and, as such, several sheds
binder-yarn path and undulation, number of warp- and weft- can be formed simultaneously. In the weft insertion process, weft
yarn layers, and yarn arrangement. Pankow [17] studied the flexu- yarn passes through the sheds to form a 3D textile.

2
Y. Liu, C. Huang, H. Xia et al. Materials & Design 212 (2021) 110267

Fig. 1. Illustration of (a) two-dimensional and (b) three-dimensional weaving. (c)


Self-built 3D weaving loom and (d) heddles in different positions in the modified
heddle position system. (e) Schematic view of the self-built 3D weaving loom.

On a traditional 3D weaving loom without heddle position


modification, the heddle frames only have two statuses, i.e., lifted
and nonlifted, which correspond to the same two positions in the
shedding process of a 2D weaving loom. In this weaving system,
only parts of 3D woven structures can be manufactured owing to
the limitations associated with heddle-frame positioning. For
example, through-thickness orthogonal woven structures are usu- Fig. 2. Heddle position in the first three shedding processes for through-thickness
ally manufactured based on the said 3D weaving loom, but layer- angle-interlock woven.
to-layer interlock structures and angle-interlock structures could
not be manufactured[25,26]. For the present study, a modified
heddle-frame position system based on a self-built 3D weaving modified 3D weaving may be more suitable for manufacturing 3D
loom was developed, which is shown in Fig. 1(c) and (d). The sche- woven textiles with greater thickness and larger structural integ-
matic view of the 3D weaving loom is depicted in Fig. 1(e), eight rity due to its primary 3D weaving benefits. In a traditional weav-
heddle frames were used, and the frames could stop at five posi- ing loom, warp yarn passes through the heddle eyes and moves
tions to form four sheds in each shedding process for the designed upwarps and downwards to form the shed for angle-interlock
3D woven structures, four layers of weft yarn could pass through and layer-to-layer structure manufacturing, but, in the modified
the sheds at the same time and allow the binder yarns to bind sev- weaving system, it is not necessary for the warp yarn to pass
eral layers of weft yarns to form different woven structures. The through the heddle eyes and the warp yarns keep stable in weaving
weaving technology for the TTAIW structure is shown in Fig. 2. process for all designed woven structures, and thus, yarn damage/
The eight heddle frames were in different positions in each shed- degradation decreases, especially for brittle yarn such as glass or
ding process to ensure that the eight binder-yarn groups were lay- carbon fiber. Moreover, compared with traditional weaving sys-
ered appropriately. Different from the traditional weaving tem, less heddle frames are needed based on this modified weaving
technology, which is based on a weave diagram that defines the system to weave complicated structures such as angle-interlock
warp- and weft-yarn interlacing, heddle position information is structures and layer-to-layer structures, due to that the warp yarns
required for 3D weaving to define the binder-yarn position in each do not need to pass through the heddle eyes during weaving as
shedding process. For a TTAIW structure, the heddle position infor- mentioned previously.
mation of eight shedding in one weave repeat is listed in Table 1.
There are several advantages for the proposed modified heddle 2.2. Three-dimensional textile structures
position system. More types of 3D woven structures can be effec-
tively and efficiently manufactured. It provides possibilities for Six woven structures, i.e., LLOW-1, LLOW-2, TTOW, LLAIW-1,
manufacturing complex 3D woven textiles such as textile preforms LLAIW-2, and TTAIW, were fabricated under similar weaving con-
for composite fan blade applications based on the modified heddle ditions (such as warp-, weft-, and binder-yarn tension) and with
position system. Compared with Jacquard weaving technology, the a comparable yarn density. For all designed 3D woven structures,
3
Y. Liu, C. Huang, H. Xia et al. Materials & Design 212 (2021) 110267

Table 1
Heddle position information in one weave repeat of a through-thickness angle-interlock woven.

Shedding Heddle 1 position(Binder Heddle 2 position(Binder Heddle 3 position(Binder Heddle 4 Heddle 5 Heddle 6 Heddle 7 Heddle 8
1) 2) 3) position(Binder position position position position
4) (Binder (Binder (Binder (Binder
5) 6) 7) 8)
First 1 2 3 4 5 4 3 2
Second 2 1 2 3 4 5 4 3
Third 3 2 1 2 3 4 5 4
Fourth 4 3 2 1 2 3 4 5
Fifth 5 4 3 2 1 2 3 4
Sixth 4 5 4 3 2 1 2 3
Seventh 3 4 5 4 3 2 1 2
Eighth 2 3 4 5 4 3 2 1

there are three layers of warp yarn and four layers of weft yarn, are 24 warp yarns, 32 weft yarns, and 8 binder yarns in one unit for
and the yarn arrangement in the warp direction is three warp TTAIW and 6 warp yarns, 8 weft yarns, and 2 binder yarns in one
yarns (one warp yarn in each layer) followed by one binder yarn. unit for TTOW. Surface top-view photographs of the six 3D woven
In orthogonal-woven structures, the binder yarn is perpendicular textiles are shown in Fig. 4.
to the warp and weft yarn theoretically, and, in angle-interlock
woven structures, binder yarns have an angle smaller than 90° 2.3. Three-dimensional textile composites
from the warp direction as designed. Moreover, in through-
thickness woven structures, all four weft-yarn layers are bounded E-glass fiber (1150tex, RS 110 QL-520, Nitto Boseki Co., Japan)
from the top to the bottom layer by the binder yarn, whereas, in was used for the warp and weft yarn, and aramid fiber (330tex,
layer-to-layer woven structures, two or three weft-yarn layers KevlarÒ 29, Du Pont-Toray Co., Japan) was used for the binder yarn.
are bounded. Accordingly, in this study, for the layer-to-layer The 3D hybrid woven textiles were produced with comparable
woven structures, both layer numbers are used: two adjacent yarn density and textile-area density; the warp-yarn density is
weft-yarn layers are bonded together for LLOW-1 and LLAIW-1, roughly 3.5 ends/cm/layer, the binder-yarn density is roughly 3.5
and three adjacent weft-yarn layers are bonded together for ends/cm, and the weft-yarn density is roughly 2.0–3.6 picks/cm/
LLOW-2 and LLAIW-2. Fig. 3 shows 3D models, weft-z and warp- layer. Thermoset epoxy resin and a hardener (DENATITE
z diagrams in one unit cell of the six 3D woven textiles. The unit XNR6815 and DENATITE XNH6815, Nagase ChemteX Co., Japan)
cell sizes vary because of structural differences. For instance, there were used for the composite matrix. The textiles were infused

Fig. 3. Three-dimensional models, weft-z and warp-z schematic diagrams: (a) LLOW-1, (b) LLOW-2, (c) TTOW, (d) LLAIW-1, (e) LLAIW-2, and (f) TTAIW.

4
Y. Liu, C. Huang, H. Xia et al. Materials & Design 212 (2021) 110267

Fig. 4. Surface top-view photographs of (a) LLOW-1, (b) LLOW-2, (c) TTOW, (d) Fig. 5. (a) Schematic and (b) experimental setup for three-point bending test.
LLAIW-1, (e) LLAIW-2, and (f) TTAIW.

3. Results and discussion

within a vacuum bag using vacuum-assisted resin-transfer mold- 3.1. Internal geometry of three-dimensional woven composites
ing (VARTM) technology. The epoxy resin and hardener were
heated separately at 40 °C in an oven and then mixed with a Optical microscopy was applied to observe composite cross sec-
weight ratio of 100:27. The mixed resin was pumped into the vac- tions and internal geometry. The internal geometry of the 3D
uum bag with a vacuum pump, and the pressure was about one woven composites is shown in Figs. 6 and 7. Different colors were
bar. After infusion, the composite was cured within the sealed vac- used to separate the warp, weft, and binder yarns as well as resin
uum bag for 24 h at room temperature. The normalized parameters matrix. Obvious geometrical differences can be observed for the
of the six 3D woven textile-reinforced composites are summarized following parameters: binder-yarn path, undulation angle, and
in Table 2. The composite fiber volume fraction was calculated by a cross-section shape; warp- and weft-yarn waviness, distribution,
resin burn-off method. and cross-section shape; and resin-pocket shape and distribution.
All woven structures have relative uniform weft-yarn distribution.
This is mainly because the binder yarn is perpendicular to the weft
2.4. Mechanical testing yarn, and, therefore, the weft yarn does not exhibit large shifting,
at the same time, binder yarn is parallel to the warp yarn, and adja-
The 3D woven textile-reinforced composite beam specimens cent warp yarns shift and attach to each other easily if there is no
underwent three-point bending tests at room temperature, the binder yarn between them in layer-to-layer woven group. The
schematic and experimental setup of which are shown in Fig. 5. warp-yarn shift has an influence on the weft-yarn straightness.
The beam specimens were cut from composite plate along textile For instance, for LLOW and LLAIW, the weft yarn, especially the
warp and weft directions and named as warp-direction beam central two layers, has an obvious waviness owing to warp-yarn
and weft-direction beam, respectively. For example, in warp- shifting. It should be noted that the binder yarn also influences
direction beams, warp and binder yarns run in beam length direc- the weft-yarn waviness, particularly at the interlaced part. Among
tion and weft yarns run in beam width direction. The beam speci- all woven structures, the through-thickness woven structures have
mens had a length of 110 mm, a width of 25 mm, and a thickness of relative uniform warp- and weft-yarn distribution and cross-
roughly 2.4–2.8 mm. Three warp-direction beam specimens were section shape, which is mainly because the binder yarn has a con-
tested as well as three weft-direction beams for each textile- sistent path in these 3D woven textile systems: it moves in the
structure composite. A universal testing machine (AG-20kND, Shi- through-thickness direction and binds weft-yarn layers from top
madzu Co., Japan) with a 20-kN load cell was used. The support to bottom, thereby securing structural uniformity.
span was 80 mm long with a loading-nose diameter of 10 mm, As shown in Fig. 6, the binder-yarn peak span corresponds to
and a loading-nose displacement speed of 5 mm/min was applied one unit cell along the warp direction (Lx) in each woven structure.
until beam failure occurred. Optical microscopy was applied to In Fig. 7, the arrows correspond to one unit cell along the weft
observe and analyze internal structural failure after flexural direction (Ly) in each woven structure. The binder-yarn weave
testing. length (L) and height (h), the undulation degree is measured, in

Table 2
Normalized parameters of the three-dimensional woven textiles and reinforced composites.

Textile group LLOW TTOW LLAIW TTAIW


Textile ID LLOW-1 LLOW-2 TTOW LLAIW-1 LLAIW-2 TTAIW
Dry textile thickness 3.26 3.04 2.72 2.98 3.25 3.20
(mm)
Dry textile-area density (kg/m2) 3.00 2.53 2.49 3.32 3.51 3.24
Composite thickness 2.73 2.73 2.83 2.40 2.71 2.49
(mm)
Composite fiber volume fraction 0.40 0.36 0.34 0.48 0.49 0.47

5
Y. Liu, C. Huang, H. Xia et al. Materials & Design 212 (2021) 110267

Fig. 6. Internal geometry of three-dimensional woven composites (parallel to the warp direction).

which undulation degree of binder-yarn is defined as the largest processes, which reduces the binder-yarn tension and increases
angle form warp direction in this study; a larger angle means a lar- the weft-yarn density. The angle-interlock woven structures have
ger binder yarn undulation. Table 3 shows the internal geometry of a weft-yarn density that is roughly 1.5 times larger than that of
the six 3D woven composites, from which it is evident that the orthogonal-woven structures, as well as a composite fiber volume
TTAIW structure has the largest real unit cell size, followed by fraction that is roughly 1.3 times larger.
LLAIW, LLOW, and TTOW. Moreover, the TTOW structure has the
largest binder-yarn undulation, followed by LLOW-2, TTAIW, 3.2. Mechanical properties
LLOW-1, LLAIW-2, and LLAIW-1. For both LLOW and LLAIW, the
binder yarn that binds three weft-yarn layers has a larger undula- In this study, the support span-to-thickness ratio of the beam
tion than has the binder yarn binding two weft-yarn layers. specimens was greater than 16, and the deflection was larger than
The angle-interlock woven structures have the largest weft- 10% of the support span. Stress was calculated by Eq. (1) to correct
yarn density values and thus exhibit a larger composite fiber vol- the end forces at the support noses, and strain was calculated by
ume fraction than do the orthogonal-woven structures, which Eq. (2) in accordance with ASTM D790 [27]:
can be attributed to structural differences and weaving condition  
limits. For TTOW, the binder yarn binds weft-yarn layers from r ¼ 3PL=2bd2 ½1 þ 6ðd=LÞ2  4ðd=LÞðd=LÞ ð1Þ
top to bottom, and two adjacent binder yarns move in opposite
directions in each shedding process, which means it undergoes
the largest tension during the weaving process, and, as such, it is
e ¼ 6dd=L2 ð2Þ
difficult to obtain a large weft-yarn density. Alternatively, for where r denotes the stress on the outer surface at the midspan, P
layer-to-layer and angle-interlock structures, the binder yarn binds denotes the applied force, L denotes the support span, b denotes
two or three adjacent weft-yarn layers in two continuous shedding the beam width, d denotes the beam thickness, e denotes the max-
6
Y. Liu, C. Huang, H. Xia et al. Materials & Design 212 (2021) 110267

Fig. 7. Internal geometry of three-dimensional woven composites (parallel to the weft direction).

Table 3
Internal geometry information of three-dimensional woven composites.

Textile group LLOW TTOW LLAIW TTAIW


Textile ID LLOW-1 LLOW-2 TTOW LLAIW-1 LLAIW-2 TTAIW
Warp-yarn density 3.6 3.5 3.5 3.5 3.6 3.6
(ends/cm/layer)
Binder-yarn density 3.6 3.5 3.5 3.5 3.6 3.6
(ends/cm)
Weft-yarn density 2.8 2.4 2.0 3.6 3.6 3.4
(picks/cm/layer)
Real unit size 7.1  16.6 8.2  11.3 9.8  5.6 11.2  22.8 16.5  22.4 23.5  22.3
(mm2)
Binder-yarn weave length (mm) 7.1 8.2 9.8 11.2 16.5 23.5
Binder-yarn weave height (mm) 1.2 2.0 2.8 1.0 2.2 2.5
Binder-yarn 19.6° 29.6° 46.9° 8.6° 15.6° 20.1°
undulation angle

7
Y. Liu, C. Huang, H. Xia et al. Materials & Design 212 (2021) 110267

yarns in the beam length direction mainly carry tensile load at


outer layers and carry compression load at inner layers, yarns in
the beam width direction mainly carry shear load. Compared with
the warp-direction beam, the weft-direction beam has an addi-
tional yarn layer in the beam length direction, at the same time,
two of the beam-length yarn layers are on the beam surface where
the maximum stress and strain occurred, and, as such, it has an
enhanced load-carrying ability, especially for bending loads.
The flexural modulus is the modulus at the initial loading stage,
during which the beam undergoes minimal deflection, and the
yarns in composite beam length direction may contribute more
to the modulus. The weft-direction composite beams have a larger
flexural modulus than have the warp-direction composite beams
for all woven structures, which can be attributed to geometrical
differences between the warp- and weft-direction beams as men-
tioned previously. The angle-interlock woven structures have a lar-
ger flexural modulus than have the orthogonal-woven structures
for both warp- and weft-direction beams, as shown in Fig. 10(a),
and this is mainly due to that there is a larger weft yarn density
and fiber volume fraction accordingly for angle-interlock woven
structures. As shown in Fig. 10 (e), the flexural modulus increases
with an increase in weft-yarn density for both weft- and warp-
direction beams; this tendency is more evident for the
orthogonal-woven structures. For weft-direction beams, a higher
weft-yarn density corresponding to the increased number of
load-bearing yarns in the beam length direction, increases the flex-
ural modulus. For warp-direction beams with a similar warp-yarn
density, the increased weft-yarn density corresponds to the
increased composite fiber volume fraction and enhances the flexu-
ral modulus. In angle-interlock woven group, LLAIW-1 and LLAIW-
2 have similar weft-yarn density and composite fiber volume frac-
tion, but LLAIW-1 has a larger flexural modulus, which can be
attributed to its superior structure, and the binder yarn may also
Fig. 8. Representative flexural stress–strain curves of six woven composites under
three-point bending: (a) warp-direction beam results and (b) weft-direction beam play a key role. The binder yarn for LLAIW-1 has a smaller undula-
results. tion angle (8.6°) compared with LLAIW-2 (15.6°), which means
that the binder yarn for LLAIW-1 is in more laying in in-plane
imum strain in the outer surface, and d denotes the midspan direction and undergoes increased tension during bending.
deflection. The flexural strength is the largest load-carrying ability until
Fig. 8 shows representative flexural stress–strain curves of the beam failure occurs, at which the beam undergoes relatively large
six 3D woven composites along warp and weft directions. It is evi- deflection, and all yarn layers may contribute to the flexural
dent that all types of 3D woven composites exhibit linear stress– strength. A similar failure mechanism was observed during the
strain behavior at the beginning of loading; nonlinear behavior three-point bending test for fiber-reinforced composites:
can be observed after a strain of roughly 0.005 and 0.02 for the compression-side yarn kink-band failure developed first, followed
warp- and weft-direction beams, respectively. Warp-direction by fiber-tow rupture failure in the tension side, where yarn kink-
beams exhibit ductile behavior, whereas weft-direction beams band failure refers to small load drops and tensile-side fiber-tow
exhibit brittle behavior. This may be due to geometrical differences rupture failure refers to a final large load drop [28]. There are sev-
between the warp- and weft-direction beams. In the warp- eral obvious small load drops before the final large load drop for
direction composite beam shown in Fig. 9(a), three layers of warp weft-direction composites, as shown in Fig. 8(b); this may because
and binder yarns are in the beam length direction, and four layers of that in a weft-direction beam, two of the weft yarn layers are on
of weft yarn are in the beam width direction. In the weft-direction the beam surface, inner surface weft-yarn kinking failures hap-
composite beam shown in Fig. 9(b), four layers of weft yarn are in pened first and then followed by the outer surface weft-yarn crack-
the beam length direction, and three layers of warp and binder ing failure. For warp-direction beams, three layers of warp yarn and
yarns are in the beam width direction. During three-point bending, binder yarn are in the beam length direction, and the warp yarn is

Fig. 9. TTAIW composite deformation model under three-point bending test: (a) warp-direction beam and (b) weft-direction beam.

8
Y. Liu, C. Huang, H. Xia et al. Materials & Design 212 (2021) 110267

Fig. 10. Flexural property information under three-point bending: (a) flexural modulus, (b) flexural strength, (c) failure strain, (d) deformation energy, (e) relationship
between flexural modulus and weft-yarn density, and (f) relationship between flexural strength and weft-yarn density.

not on the beam surface, which means that yarn kink-band failure is Overall, LLAIW-1, LLAIW-2, and TTAIW have the smallest
unlikely compared with weft-direction beams, and, thus, there are binder-yarn undulation angle and the largest composite fiber vol-
no obvious small load drops before the final load drop in the flexural ume fraction, and, as such, compared with those of LLOW-1,
stress–strain figures shown in Fig. 8(a). The angle-interlock woven LLOW-2, and TTOW, their flexural strength and modulus are larger
structures have a larger flexural strength than have the orthogonal- by 50% and 40%, respectively. The flexural modulus is more sensi-
woven structures, perhaps owing to the higher weft-yarn density tive to weft-yarn density and composite fiber volume fraction,
and structural superiority. As shown in Fig. 10(f), there is no clear whereas the flexural strength is more sensitive to woven structure.
tendency between flexural strength and weft-yarn density, espe- Indeed, a smaller binder-yarn undulation angle may have
cially for weft-direction beams; thus, the woven structure may con- increased the tensile strength for warp-direction composite beams
tribute more to the flexural strength. during three-point bending [29].
The energy required for beam failure is referred to as the
composite-failure resistance and was calculated from the area under 3.3. Failure modes
the load–deflection curves. The warp-direction beams exhibit a
smaller bending load but larger deflection than do weft-direction Fig. 11 shows strong-light background photographs of warp-
beams for all structures, and accordingly larger failure strain for and weft-direction beam specimens of six woven types after flexu-
warp-direction beams, as shown in Fig. 10(c). Basically, angle- ral test. Sunlight was used as the strong-light source positioned
interlock woven structures have larger deformation energy than behind the specimens and the concave surface of the composite
have orthogonal woven structures, as shown in Fig. 10(d). TTAIW beams was facing the camera. The transparent part is the area with-
has the largest composite-failure resistance along warp direction. out failures, and the opaque part (dark part) is the area with failures
9
Y. Liu, C. Huang, H. Xia et al. Materials & Design 212 (2021) 110267

(dark dots enclosed by red squares in Fig. 11(b)). This is mainly


owing to the binder-yarn load-transfer ability, thus woven struc-
ture with specific binder-yarn path has an influence in failure map.
A detailed failure map for the warp- and weft-direction beams
of the TTAIW structure is shown in Fig. 12(a), from which it is evi-
dent that the propagating-failure area and center-failure area in
the warp direction are relatively large, which is similar for all other
woven structures as shown in Fig. 11(a). This may be because, for
warp-direction composite beams, the binder yarn runs in the beam
length direction, which means it could transfer loads to adjacent
weft yarn and thereby increase the weft-yarn debonding-failure
area. There are similar failure modes for weft-direction beams
among all six woven structures. For instance, in the center part
of the beam for structure TTOW shown in Fig. 12(b), weft-yarn
kink-band failure occurs at the inner surface (compression side)
owing to compression loads, and warp-yarn tow-splitting failure
mainly occurs at the outer layers (tension side) owing to shear
loads; at the same time, delamination failure occurs mainly at
the beam center for the outer yarn layers. This failure modes are
associated with the structure of weft-direction beams, as four lay-
ers of weft yarn run in the beam length direction and two of these
layers run on the beam surface, which mean that kink-band failure
and fiber-breakage failure both likely occur. In contrast, for warp-
direction beams, three layers of warp yarn run in the beam length
direction, but none run at the beam surface, which means that
warp-yarn kink-band failure and fiber-breakage failure rarely
occur; rather, surface weft-yarn tow-splitting failure in the beam
center-failure area more likely occurs, as shown in Fig. 12(a), due
to the fact that the weft-yarn layers run in the beam width direc-
tion and thus bear large shear loads during bending.
Overall, there are obvious differences in the failure modes of
Fig. 11. (a) Strong-light background photographs of warp- and weft-direction beam warp- and weft-direction beams for the six types of woven com-
specimens of six woven types after flexural test. (b) Comparation of warp-direction
beams of TTOW and TTAIW structures. The transparent part is the area without
posites, and woven structure or binder-yarn path influences the
failure, and the dark part is the area with failure.

due to the transparent characteristic of glass fiber and epoxy resin


matrix as well as refraction and reflection of light. Failures such as
yarn debonding and delamination will introduce serious refraction
and reflection of light, corresponding to the dark pixels in strong-
light background photographs, the darker pixels, the more serious
failures. Obvious failures can be observed at the center of the
warp- and weft-direction beams for all woven types as shown in
Fig. 11(a). This is expected as the beam center where was the con-
tact area for the loading nose during the three-point bending test,
and, as such, tensile and compressive stress are both at a maximum
and failures easily occur. For the warp-direction composite beams,
obvious propagating failures can be observed beyond the center
areas, spreading along the weft-yarn direction (beam-width direc-
tion) as shown in Fig. 11(b); for weft-direction composite beams,
the failures are mostly localized at the center area, beyond which
there are no obvious propagating failures. On the one hand, the
orthogonal-woven structures, which have a larger binder-yarn
undulation angle and more interlaced parts between surface-layer
weft and binder yarn, exhibit serious weft-yarn debonding failure
(thick dark lines in Fig. 11(b)). On the other hand, the angle-
interlock woven structures, which have a smaller binder-yarn
undulation angle and less interlaced parts between surface-layer
weft and binder yarn, exhibit slight weft-yarn debonding failure
(thin dark lines in Fig. 11(b)). The different failures could be attrib-
uted to the internal geometrical differences: there is a relative
homogeneous weft-yarn arragement in angle-interlock woven
group, whereas the weft yarns have obvious gaps between two
adjacent weft-insetions in orthogonal woven group, which could
be observed in Fig. 6. Moreover, for warp-direction beams of
orthogonal woven structures, propagating failures along weft yarns Fig. 12. Failure modes under flexural loading: (a) TTAIW in-plane failure map and
are more serious at the interlaced part of the binder and weft yarn (b) TTOW through-thickness failure map along the weft direction.

10
Y. Liu, C. Huang, H. Xia et al. Materials & Design 212 (2021) 110267

failure map for warp-direction beams. In essence, for weft- Data availability: The raw/processed data required to reproduce
direction beams, failure mainly occurred at the center part of the these findings cannot be shared at this time as the data also forms
tested beams; for warp-direction beams, obvious propagating fail- part of an ongoing study.
ures occurred because of the binder-yarn load-transfer ability.
Yarn kink-band failure and fiber-breakage failure are likely for References
weft-direction beams, and surface-yarn tow-splitting failure is
likely for warp-direction beams. Moreover, the angle-interlock [1] K. Dransfield, C. Baillie, Y.-W. Mai, Improving the delamination resistance of
CFRP by stitching—a review, Compos. Sci. Technol. 50 (3) (1994) 305–317.
woven structures have slighter propagating weft-yarn debonding [2] A.P. Mouritz, B.N. Cox, A mechanistic interpretation of the comparative in-
failure compared with orthogonal-woven structures. plane mechanical properties of 3D woven, stitched and pinned composites,
Compos. A Appl. Sci. Manuf. 41 (6) (2010) 709–728.
[3] G. Dell’Anno, J.W.G. Treiber, I.K. Partridge, Manufacturing of composite parts
4. Conclusion reinforced through-thickness by tufting, Rob. Comput. Integr. Manuf. 37
(2016) 262–272.
[4] A.P. Mouritz, Review of z-pinned composite laminates, Compos. A Appl. Sci.
In this paper, a new weaving technology with modified heddle
Manuf. 38 (12) (2007) 2383–2397.
position system based on a self-built 3D weaving loom was [5] R. Kamiya, B.A. Cheeseman, P. Popper, T.-W. Chou, Some recent advances in the
designed and used to develop four typical 3D woven structures: fabrication and design of three-dimensional textile preforms: a review,
LLOW, TTOW, LLAIW, and TTAIW. The new weaving technology Compos. Sci. Technol. 60 (1) (2000) 33–47.
[6] X. Chen, Advances in 3D textiles, Elsevier (2015).
has great potential for manufacturing various 3D woven structures [7] X. Chen, L.W. Taylor, L.-J. Tsai, An overview on fabrication of three-dimensional
effectively and efficiently. VARTM was applied to fabricate the 3D woven textile preforms for composites, Text. Res. J. 81 (9) (2011) 932–944.
woven composites. The flexural properties were experimentally [8] M. Ansar, W. Xinwei, Z. Chouwei, Modeling strategies of 3D woven composites:
A review, Compos. Struct. 93 (8) (2011) 1947–1963.
characterized with a three-point bending test, and optical micro- [9] N. Khokar, 3D-weaving: theory and practice, J. Text. Inst. 92 (2) (2001) 193–
scopy was used to observe failures after bending. 207.
Based on the proposed weaving system, the angle-interlock [10] B.K. Behera, B.P. Dash, An experimental investigation into structure and
properties of 3D-woven aramid and PBO fabrics, The Journal of The Textile
woven structures have larger weft-yarn density (roughly 1.5 times Institute. 104 (12) (2013) 1337–1344.
larger than that of the orthogonal structures) and composite fiber [11] B. Zahid, X. Chen, Manufacturing of single-piece textile reinforced riot helmet
volume fraction (roughly 1.3 times larger than that of the orthog- shell from vacuum bagging, J. Compos. Mater. 47 (19) (2013) 2343–2351.
[12] El-Dessouky HM, Snape AE, Turner JL, Saleh MN, Tew H, Scaife RJ. 3D weaving
onal structures). The through-thickness woven structures have a for advanced composite manufacturing: from research to reality, in:
relative uniform yarn distribution than have the layer-to-layer Proceedings of the SAMPE Conference2017.
woven structures. TTAIW is the optimal design with a larger [13] G. Marsh, Aero engines lose weight thanks to composites, Reinf. Plast. 56 (6)
(2012) 32–35.
weft-yarn density as well as a uniform warp, weft, and binder yarn
[14] T. Huang, Y. Wang, G. Wang, Review of the mechanical properties of a 3D
distribution. A further modification is needed to achieve higher woven composite and its applications, Polymer-Plastics Technol. Eng. 57 (8)
weft-yarn density for orthogonal woven structures based on this (2018) 740–756.
weaving system. [15] C. Harvey, E. Holtzman, J. Ko, B. Hagan, R. Wu, S. Marschner, D. Kessler,
Weaving objects: spatial design and functionality of 3D-woven textiles,
Different mechanical behavior is shown along textile warp and Leonardo. 52 (4) (2019) 381–388.
weft directions for the designed 3D woven structures: weft- [16] M. Amirul, 3D woven fabrics, structures, and methods of manufacture. Woven
direction composite beams have a higher flexural modulus but Textiles: Elsevier; 2020. p. 329-91.
[17] M. Pankow, A. Salvi, A.M. Waas, C.F. Yen, S. Ghiorse, Resistance to delamination
lower failure strain than have warp-direction beams. Moreover, of 3D woven textile composites evaluated using End Notch Flexure (ENF) tests:
the angle-interlock woven structures have a larger flexural Experimental results, Compos. A Appl. Sci. Manuf. 42 (10) (2011) 1463–1476.
strength and modulus than that of orthogonal structures: 50% [18] P. Turner, T. Liu, X. Zeng, Collapse of 3D orthogonal woven carbon fibre
composites under in-plane tension/compression and out-of-plane bending,
and 40% larger, respectively. Woven structure or binder-yarn path Compos. Struct. 142 (2016) 286–297.
has an obvious influence on the composite failures. Warp-direction [19] R. Gerlach, C.R. Siviour, J. Wiegand, N. Petrinic, In-plane and through-thickness
beams have obvious propagating failure beyond beam center com- properties, failure modes, damage and delamination in 3D woven carbon fibre
composites subjected to impact loading, Compos. Sci. Technol. 72 (3) (2012)
pared with weft-direction beams due to the load-transfer ability of 397–411.
binder yarn which runs in the beam length direction. The angle- [20] B.K. Behera, B.P. Dash, Mechanical behavior of 3D woven composites, Mater.
interlock woven structures have slight weft-yarn debonding fail- Des. 67 (2015) 261–271.
[21] B.K. Behera, B.P. Dash, An experimental investigation into the mechanical
ures whereas the orthogonal woven structures have serious weft-
behaviour of 3D woven fabrics for structural composites, Fibers Polym. 15 (9)
yarn debonding failures in the propagating-failure area, mainly (2014) 1950–1955.
due to the internal geometrical differences of the two woven [22] D. Zhang, A.M. Waas, C.-F. Yen, Progressive damage and failure response of
groups. Proper structural design of 3D woven composite is key hybrid 3D textile composites subjected to flexural loading, part I:
Experimental studies, Int. J. Solids Struct. 75-76 (2015) 309–320.
issue to achieve better flexural performance. [23] R. Umer, H. Alhussein, J. Zhou, W.J. Cantwell, The mechanical properties of 3D
woven composites, J. Compos. Mater. 51 (12) (2017) 1703–1716.
Declaration of Competing Interest [24] D. Zhang, A. Waas, M. Pankow, C. Yen, S. Ghiorse, Flexural behavior of a layer-
to-layer orthogonal interlocked three-dimensional textile composite, J. Eng.
Mater. Technol. 134 (2012).
The authors declare that they have no known competing finan- [25] M. Mohamed, A. Bogdanovich, Comparative analysis of different 3D weaving
cial interests or personal relationships that could have appeared processes, machines and products, in: Proceedings of the 17TH international
conference on composite materials (ICCM-17)2009. p. 27-31.
to influence the work reported in this paper. [26] H.M. El-Dessouky, M.N. Saleh, in: Recent Developments in the Field of Carbon
Fibers, InTech, 2018, https://doi.org/10.5772/intechopen.74311.
Acknowledgement [27] ASTM S. Standard test methods for flexural properties of unreinforced and
reinforced plastics and electrical insulating materials. ASTM D790. Annual
book of ASTM Standards, 1997.
This work was supported by the Japan Society for the Promotion of [28] B.N. Cox, M.S. Dadkhah, W.L. Morris, J.G. Flintoff, Failure mechanisms of 3D
Science (JSPS KAKENHI 20H00288 and 26420721) and National woven composites in tension, compression, and bending, Acta Metall. Mater.
42 (12) (1994) 3967–3984.
Natural Science Foundation of China (No.52073259).
[29] G. Stegschuster, K. Pingkarawat, B. Wendland, A.P. Mouritz, Experimental
determination of the mode I delamination fracture and fatigue properties of
Compliance with ethical standards.
thin 3D woven composites, Compos. A Appl. Sci. Manuf. 84 (2016) 308–315.
Declaration of Competing Interest: The authors declare that they
have no conflict of interest.

11

You might also like