You are on page 1of 34

Journal Pre-proof

Self-sensing Impact Damage in and Non-destructive Evaluation of


Carbon Fiber-Reinforced Polymers using Electrical Resistance and the
Corresponding Electrical Route Models

Hyung Doh Roh, So Young Oh, Young-Bin Park

PII: S0924-4247(21)00225-9
DOI: https://doi.org/10.1016/j.sna.2021.112762
Reference: SNA 112762

To appear in: Sensors and Actuators: A. Physical

Received Date: 18 January 2021


Revised Date: 7 April 2021
Accepted Date: 16 April 2021

Please cite this article as: Roh HD, Oh SY, Park Y-Bin, Self-sensing Impact Damage in and
Non-destructive Evaluation of Carbon Fiber-Reinforced Polymers using Electrical Resistance
and the Corresponding Electrical Route Models, Sensors and Actuators: A. Physical (2021),
doi: https://doi.org/10.1016/j.sna.2021.112762

This is a PDF file of an article that has undergone enhancements after acceptance, such as
the addition of a cover page and metadata, and formatting for readability, but it is not yet the
definitive version of record. This version will undergo additional copyediting, typesetting and
review before it is published in its final form, but we are providing this version to give early
visibility of the article. Please note that, during the production process, errors may be
discovered which could affect the content, and all legal disclaimers that apply to the journal
pertain.

© 2020 Published by Elsevier.


Self-sensing Impact Damage in and Non-destructive Evaluation of

Carbon Fiber-Reinforced Polymers using Electrical Resistance

and the Corresponding Electrical Route Models

Hyung Doh Roh1, So Young Oh1, and Young-Bin Park*

Mechanical Engineering, Ulsan National Institute of Science and Technology,

of
UNIST-gil 50, Eonyang-eup, Ulju-gun, Ulsan 44919 Republic of Korea

ro
*To whom correspondence should be addressed:
Phone: +82-52-217-2314, Email: ypark@unist.ac.kr
1
These authors contribute equally to this article.
-p
re
Graphical Abstract
lP
na
ur
Jo

Impact damage of a uni-directional carbon fiber reinforced plastic: Perspective (a) along
the fiber length-wise and (b) perpendicular to the fiber length-wise. (c) Largest
electrical resistance change ratio at the impacted area. (d) Electrical route modeling
detouring the puncture.

1
Highlights

 The location and the size of the impact damage changed electrical resistance.

 Non-destructive evaluation was available using electrical resistance.

 Self-sensing mechanism was analyzed by proposing electrical route modeling.

of
 Ultimate electrical network chose the easiest path such as along the carbon fiber.

ro
Abstract
-p
re
Carbon fiber-reinforced plastics (CFRPs) made of uni-directional carbon fibers

(UDCFs) are used in various applications such as construction, aerospace, and


lP

automobiles. Therefore, their structural health monitoring (SHM) and non-destructive

evaluation (NDE) are important to ensure safety during operation. While there is
na

literature on self-sensing of CFRPs to realize various properties, there is no information

on their impact self-sensing properties. Therefore, in this study, CFRPs in several


ur

orientations were investigated in terms of their mechanical fracture and

electromechanical behavior. Changes in their electrical resistance due to impact damage


Jo

can be utilized for SHM using the corresponding electrically equivalent circuit models.

The circuit models constructed consisted of electrical resistors that described the

UDCFs. In addition to converting CFRPs into 2D circuits, 3D electrical routes between

electrodes were proposed for NDE. Calculating the detour length of the electrical routes

2
using the proposed models helps in assessing the severity of the impact damage.

Therefore, the models for CFRPs developed in this study not only provide support for

SHM but also for NDE using electrical resistance.

Keywords: carbon fiber; functional composite; smart materials; self-sensing; non-


destructive testing

of
1. Introduction

The need for lightweight materials with superior mechanical properties is driving

ro
market demand for carbon fiber-reinforced plastics (CFRPs) in various applications

such as automobiles [1-3], aerospace [4-6], and wind-power generation [7, 8]. Most
-p
CFRPs are made of cross-ply uni-directional carbon fibers (UDCFs) to ensure structural
re
stiffness [9, 10] and durability against fatigue loading [11-13]. Therefore, the structural

health monitoring (SHM) and non-destructive evaluation (NDE) of CFRPs is necessary


lP

to ensure high safety and control maintenance costs.

Several studies have reported on self-sensing in CFRPs using electrical resistance


na

to replace additional sensors such as fiber Bragg grating sensors [14, 15], strain gauges

[16], and piezoelectric sensors [17, 18] with CFRP structures themselves. Todoroki et al.
ur

[19-23] investigated the electrical-resistance behavior of CFRPs in various directions

(for e.g., in the direction of tensile elongation) to realize real-time strain self-sensing. In
Jo

addition, Xie’s group [23, 24] and Okabe’s group [25, 26] proposed electrically

equivalent circuit models to analyze the electromechanical behavior of CFRPs. In these

investigations, carbon fibers were converted into electrical resistors and the contact

between the adjacent fibers resulted in an electrical network; the authors suggested a

3
global load-sharing model based on the effective length of carbon fibers according to

the Weibull model.

Recently, research has focused on the composite design for superior impact

resistance [5, 27], finite element analysis (FEA) for impact-behavior analysis [28-30],

and NDE of internal impact damage [31-33]. However, there is limited information on

impact self-sensing in cross-ply CFRPs. In addition, the exact parameters of electrically

equivalent circuit models are not available (only conceptual models are known).

Targeting these research gaps, we describe the electrically equivalent circuit models of

of
cross-ply CFRPs considering intra-tow and inter-tow networks in this study; these

ro
models were built by investigating the electromechanical properties of the used CFRPs.

The developed models can be applied to not only SHM, which localizes the impact
-p
damage, but also to NDE, which identifies the size of the impact puncture.
re
lP

2. Experimental

2.1 Materials
na

UDCFs (12K, T700SC, 305 g/m2, and 0.35 mm thick; TEI Composites, Taiwan)

were purchased from Keun Young Industry (Seoul, Korea). A 24K carbon fiber spread-
ur

tow woven fabric (STWF) (25-mm-wide, 0.22-mm-thick, and 90 g/m2 weight; R-Tex,

Korea) was procured from RIST (Pohang, Korea). A TCG-662 epoxy polymer (Total C
Jo

& G, Korea) supplied by Jet Korea (Changwon, Korea) was used as the matrix. The

curing agent used was co-mixed with 20 wt.% of the epoxy resin. The densities of the

epoxy resin and curing agent were 1.16 and 0.98 g/cm3, respectively, while the

corresponding viscosities were 3000–6000 and 20–60 cps.

4
Electrodes were installed on the carbon fibers with a 30 AWG copper wire and silver

paste (P-100, Elcoat, USA) to minimize contact resistance between the fiber and wire.

2.2 Sample preparation

Electrodes were installed on dry carbon fibers with a copper wire and silver paste.

The silver paste at the conjunction was annealed for 5 min using a heating gun. Eight

plies of UDCFs were symmetrically stacked according to the orientations [08], [03/901]s,

of
[02/902]s, [01/903]s, and [908]. In the STWF, 16 plies were stacked to achieve a thickness

similar to the aforementioned samples, which consist of eight UDCF plies. The CFRPs

ro
were manufactured by vacuum-assisted resin-transfer molding, in which resin was

infused into the mold using the difference between the atmospheric pressure and

vacuum.
-p
re
To analyze the distance between the electrodes, the CFRPs were manufactured into

a rectangular shape with dimensions of 300 × 100 mm2 (Fig. 1(a)). The electrodes were
lP

aligned at the top and bottom plies to investigate changes in the electrical resistance in

terms of the electrode distance against an identical impact. The surface with the
na

embedded electrical wires was placed on the inner side of the sample to dampen the

electrical noise at the conjunction. The conjunctions were embedded inside the fabric
ur

and enclosed by the matrix to further minimize the electrical noise due to external

stimulus. The distance between the electrodes was set to 70, 120, 170, and 220 mm. The
Jo

electrodes were arranged as per the alignments of both the tows and stitches in the

fibers. The stitches, which were made of polyester, were bonded to the carbon fiber

tows perpendicular to the lengthwise fiber. The electrodes were installed along either

the same carbon fiber tows or the same stitches to secure the alignments.

5
The CFRP used for damage localization was a 200-mm-square sample (Fig. 1(b)).

Electrodes were embedded at a distance of 60 mm inside the samples from the edges

and the distance between the adjacent electrodes was set to 80 mm. The electrodes were

located on the top and bottom UDCF plies.

For electrical-resistivity analysis, 50-mm-square CFRP samples were manufactured

with four electrodes in series (Fig. 1(d)). The sample shown in Fig. 1(d) was made of

eight UDCF plies, similar to the samples shown in Figs. 1(a) and 1(b). The fiber

of
stacking configurations were either [08] or [908] to compare electrical resistivity of the

inter-tow and intra-tow paths. The purpose of the sample shown in Fig. 1(d) is to

ro
measure inter-tow electrical resistivity. The electrodes were embedded either along the

fiber or perpendicular to the fiber, as CFRPs made of UDCF exhibit orthotropic


-p
electrical and mechanical properties. Further, similar to the samples introduced in Figs.
re
1(a) and 1(b), the electrodes were embedded in the top ply. The direct current (DC)

electrical resistance was measured using the four-probe method, which was then divided
lP

by the length. The calculated resistance value per unit length can be applied to the other

CFRP samples because the fiber stacking configurations of all the samples are similar.
na

Electrical resistivity was measured to build an electrically equivalent circuit model.


ur

2.3 Piezoresistive characterization

Impact testing was carried out with a load of 40 J on a drop-weight impact testing
Jo

tower (CEAST 9350, USA) (Fig. 2(a)). The pneumatic clamping force was set to 60 N

with a circular jig of 40 mm in inner diameter (Fig. 2(b)). The corresponding impact area

is shown in Fig. 1(c) and represents the mid-point of the left-handed side electrode pairs.

The electrical resistance of the CFRPs was simultaneously measured with impact

6
testing using a multimeter (Keithley 2002, USA) and a switching system (Keithley 7001,

USA) (Fig. 2(a)). The measuring rate of the electrical resistance was 5 Hz with six

significant digits. The square-shaped CFRPs contained a total of 16 electrode pairs, four

in the bottom ply in length and width, four in the top ply (same as the bottom), four in the

thickness direction, and four along the diagonal connecting the top and bottom plies. The

electrode numbers are shown in Fig. 1(c). For electromechanical analysis, the electrical

resistance change ratio was investigated. The electrical resistance change ratio is the

of
resistance change after the impact divided by the initial resistance. For electromechanical

analysis, the electrical resistance change ratio was calculated to normalize the data. The

ro
electrical resistance change ratio is the resistance change after the impact divided by the

initial resistance. This minimized the differences between the CFRP specimens and
-p
enabled us to compare the data more effectively/accurately.
re
2.4 Computer-aided analysis for electromechanical behavior
lP

The built electrical circuits were analyzed using a commercial software for analog

electronic circuit simulators, LTspice. A 100 V DC source, with similar settings to that
na

of a multimeter, was connected to the resistors and the electrical resistance was

calculated using Ohm’s law by monitoring the current. The commercial software
ur

MATLAB was employed to solve the partial differential equations (PDEs)

corresponding to the electrical current density by using FEA. One electrode was
Jo

connected to a 1 A current source while the other indicated –1 A with four insulating

edges. The effective electrical paths in the CFRP were investigated using the MATLAB

PDE Toolbox and the PDEs of conservation law (Eq. (1)),

𝜕𝑓
𝑑
𝑑𝑡
∫Ω(t) 𝑓𝑑𝑉 = ∫Ω(t) 𝜕𝑡 𝑑𝑉 + ∫∂Ω(t) (𝑣 𝑏 ∙ 𝑛)𝑓𝑑𝐴 (1)

7
where t indicates time, Ω(t) represents the time-dependent region with the boundary

condition 𝜕Ω(𝑡), and f may represent tensor, vector, or scalar values with the

aforementioned boundary. While f is designated as the vector for the electrical current

density in this study, it can be neglected in this step. Eq. (1) can be reduced to a simpler

form with constrained boundary conditions. In addition, dV and dA indicate volume and

surface elements, respectively, n indicates the outward-pointing unit vector, and vb

denotes the velocity of the area element. Using a constant Ω at vb = 0 and Ohm’s law,

of
the following relationship was obtained,

𝜕𝜑
𝐽 = 𝜎𝐸 ∙ 𝐸 = −𝜎𝐸 𝜕𝑋 (2)

ro
where J, 𝜎E, E, 𝜑, and X denote electrical current density, electrical conductivity,

-p
electrical field, electrical potential, and displacement, respectively. Eq. (2) can be

inserted into Eq. (1) by replacing f with J. Then, Maxwell’s equation can be reduced and
re
transformed with respect to electrical analysis as shown below,

∫𝑆 𝐽 ∙ 𝑛 𝑑𝐴 = ∫𝑉 𝑟𝐶 𝑑𝑉 (3)
lP

where 𝑟𝐶 and V denote current source per unit volume and the predefined volume,

respectively. In addition, the differential form of Gauss’ law, which is a subset of


na

Maxwell’s equations, can be written as shown in Eq. (4),

𝛻∙𝐷 = 𝜌 (4)
ur

where D and 𝜌 denote electric displacement field and free electric-charge density,
Jo

respectively. The electrical status of the CFRP samples was automatically analyzed by

the MATLAB Toolbox for the electrical potential and electrical current density using

Eqs. (3) and (4), respectively.

3. Results and discussion

8
3.1 Mechanical analysis

The mechanical damage due to the drop-weight impact propagated along the

UDCFs (Fig. 3) occurs because the reinforcing material absorbs the energy. Even

though circular jigs clamped the samples, the impact energy propagated along the fiber

resulting in breaking the sample with 0º UDCFs into two parts (Fig. 3(a)). Magnified

photographs of the sample from the top and bottom views are shown in Figs. 3(b) and

3(c), respectively. The matrix between the adjacent reinforcing materials failed and so

of
did the impact-struck UDCFs. However, some UDCFs around the targeted area were

able to endure the absorbed energy (Figs. 3(b) and (c)); hence, only half the sample split

ro
near the targeted area.

Similarly, the sample with 90º UDCFs underwent mechanical failure along the fiber
-p
(Figs. 3(d)–(f)). The UDCFs in the impact-struck area completely failed while those
re
around the struck area were split, similar to the 0º UDCFs. However, the UDCFs were

sufficiently long along the length of the sample; hence, the sample did not completely
lP

fracture. Square CFRP samples represented similar failures (Figs. 3(g)–(i)). The width

of the damaged section was similar to the diameter of the impact striker in the 90º
na

direction (Fig. 3(h)), while the width in the 0º direction was larger (Fig. 3(i)).

The maximum absorbed impact energy by the 200-mm-square CFRP samples was
ur

similar to the average value of 34.81 J (Fig. 4). All the CFRP samples were punctured

when impacted with 40 J loads, similar to the case in Fig. 3. The only difference
Jo

observed was that the orientation of the bottom UDCFs determined the broken shape.

Therefore, the fiber stacking configuration was not directly related to the maximum

absorbed impact energy.

9
3.2 Electromechanical analysis of electrode distance

Changes occurring in the electrical resistance of the CFRPs with the 0º and 90º

UDCFs in Fig. 3 are shown in Figs. 5(a)–(d) and Figs. 5(e)–(h), respectively. All the

electrical pairings crossing the impact-struck-area showed an increased electrical

resistance in both cases (Figs. 5(a), (b), (d)–(f), and (h)). However, the ratio of change

decreased as the electrode distance across the struck area increased from 70 to 220 mm.

This is because volumetric change in terms of mechanical damage is directly related to

of
change in the electrical resistance. Therefore, the electrode pair with the shortest

distance experienced the highest increase in electrical resistance. Thus, the distance

ro
between the electrodes in square CFRPs was optimized to 80 mm.

Electrode pairs in the through-thickness direction also showed resistance changes.


-p
The magnitude of the change decreased as the distance from the impact-struck area
re
increased (Figs. 5(c) and (g)). The resistance of the CFRP with 0º UDCFs decreased

(Fig. 5(c)), while that of the CFRP with 90º UDCFs increased (Fig. 5(g)). The decrease
lP

observed in the former is attributed to the collapse of the adjacent UDCFs. However, the

relative magnitude of the resistance change was approximately 4.8% of the CFRP,
na

which consists of 90˚ UDCF (Figs. 5(c) and (g)).

The CFRP with 90º UDCFs showed a positive resistance change in the through-
ur

thickness direction (Fig. 5(g)). Energy absorption via the UDCF crossing the impact-

struck-area led to the internal matrix cracking, which resulted in an increased resistance.
Jo

In addition, the internal damage and changes in the resistance decreased as mechanical

energy dissipated, and the distance between the struck area and through-thickness

pairings increased.

10
3.3 Damage localization with self-sensing

Sixteen electrode pairs were used to analyze the changes in the electrical resistance

of the square CFRPs (Fig. 1(c)), and the observed results are shown in Fig. 6. The

largest increase in the electrical resistance was observed in the pairs crossing the impact

spot in the [08] CFRP (Figs. 6(a) and (b)). Energy was disseminated through the 0º

UDCFs and internal damage was generated, which resulted in an increase in the

electrical resistance of the electrode pairs in the through-thickness direction (pairs 1-5

of
and 4-8), (Fig. 6(c)). However, the diagonal pairs exhibited minimal change (Fig. 6(d))

because the impact-struck-area was not along the electrical path.

ro
The square CFRP with the 90º UDCFs also showed large resistance changes when

the impact spot was along the electrical path (for e.g., pairs 4-1 and 8-5 in Figs. 6(e) and
-p
(f)). Meanwhile, the pairs on the opposite sides (2-3 and 6-7) showed similar change
re
ratios. At the impact spot, the 90º UDCFs transferred the impact energy and internal

damage was generated along the UDCFs, which led to an increase in resistance. In
lP

addition, diagonal pairs also showed higher resistance (Fig. 6(h)) when compared to

their counterparts in 0º UDCFs, which can be attributed to the mechanical damage


na

sustained by the 90º UDCFs as mentioned above. In contrast, the pairs in the through-

thickness direction showed minimal change, as there were no UDCFs to disseminate the
ur

impact energy in this direction (Fig. 6(g)).

Damage at other impact spots can be subsequently localized because the square
Jo

CFRPs were symmetric. Localized damage on the left-hand side guarantees damage

localization on the right-hand side. The localized damage in the upper side, in regard to

the top view, can be verified by another sample. For example, the localized damage in

the upper side of the [08] CFRP can be examined by localizing the damage in the left-

11
hand side of the [908] CFRP. Therefore, it can be concluded that the method of

monitoring the electrical resistance by analyzing multiple channels successfully

achieved damage localization in the samples.

When the stacking configuration of the square CFRP was changed to [03/901]s, the

changes observed in the electrical resistance varied from those observed in the [08] and

[908] CFRPs. To illustrate these differences, the electromechanical trends of the [08] and

[908] CFRPs were compared to that of the [03/901]s CFRPs in Fig. 7.

of
The largest change in the electrical resistance was observed in the pairs 4-1 and 8-5

at the impact-struck-area (Figs. 7(a) and (b)). In addition, the pairs 1-2 and 3-4 also

ro
showed remarkable differences in their resistance because of energy dissemination by

the 90º UDCFs (similar to the case in Fig. 6(e)). However, the changes in pairs 5-6 and
-p
7-8 were not as significant as those in the pairs in the bottom ply (Fig. 7(b)) because the
re
direction of the impact-energy dissipation was radially downward.

Changes in the resistance along the through-thickness were also observed as the 0º
lP

UDCFs transferred the impact energy (Fig. 7(c)). The resistance of the electrodes (pairs

1-7 and 4-6) located in the bottom ply of the impacted side increased, while there were
na

only small changes in the resistance of the electrodes in the top ply (pairs 2-8 and 3-5)

(Fig. 7(d)). This phenomenon can be explained based on the direction of the impact-
ur

energy dissemination, i.e., radial, downward, and along the fiber.

Even though the number of the 90º UDCFs in different samples varied, their
Jo

electromechanical behavior was similar to that of the [03/901]s CFRP. For example, the

electromechanical behavior of the [02/902]s and [01/903]s CFRPs is illustrated in Figs.

8(a)–(d) and (e)–(h), respectively. In the impact-struck area, the pairs 4-1 and 8-5

showed the largest change in resistance; meanwhile, the through-thickness pairs 1-5 and

12
4-8 showed significant changes in their resistance due to impact-energy transfer via the

0º UDCFs.

3.4 Equivalent circuit modeling

3.4.1 Electrical-resistor-grid modeling

The size of impact damage in CFRPs can be analyzed in terms of electrical

resistance, as CFRPs are electrical resistors. Thus, electrically equivalent circuit models

of
of CFRPs can be used to evaluate their electromechanical behavior using the measured

electrical resistance (Table 1).

ro
Electrical resistivity can be calculated from the average values in Table 1 using the

sample geometry shown in Fig. 1(d). Orthogonality in the CFRPs was reflected in the
-p
electrical circuit models such that resistivity along the perpendicular direction of the
re
UDCF was 42 times larger than that along the fiber.

Electrical potential and electric field were analyzed by FEA (Figs. 9(a) and 9(b)).
lP

The impact damage, which is represented by the hole in Fig. 9(b), resulted in changes in

the electrical potential and electric field, owing to which the electrical resistance
na

changed. However, there was no local concentration of the electrical potential and

electric field. The inherent principles used were introduced in Eqs. (3) and (4).
ur

Therefore, an electrically equivalent circuit model of the [08] CFRP was constructed

(Fig. 9(c)). The values of the electrical resistance of each resistor were selected
Jo

considering the orthogonality of the CFRP, using the values listed in Table 1. Resistance

in the lengthwise direction (along the fiber) was found to be 0.125575 Ω, while that

along the width was 5.24115 Ω (perpendicular to the fiber) (Table 1). Thus, the two

electrodes on the UDCFs were converted into a ground and voltage source, as described

13
in Figs. 9(c) and (d).

The effective width of the electrical network and grid size of the circuit between

the electrodes were evaluated (Table 2). The unit block size of the grid and number of

the resistor rows were correlated with the electrical resistance. When the inter-tow

distance increased, which is the distance between the adjacent tows, the resistivity

increased because there is insufficient electrical network. The higher electrical

resistivity decreases the effective width of the network because the electrons prefer the

of
easier intra-tow network to the inter-tow network. Alternatively, when the intra-tow

distance decreases, the electrical resistivity also decreases. Thus, the effective width of

ro
the network decreases because the electrons are more prone to go through the single

tow.
-p
Therefore, the measured and theoretically calculated resistance values were used to
re
select the appropriate models. It was assumed that the [08] CFRP consists of a 2D

electrically equivalent circuit with repetitive electrical resistors in a grid (Fig. 9(c)). The
lP

size of the equivalent circuit was determined by iteratively estimating the electrical

resistance and comparing it to the measured resistance (Table 2). Subsequently, an


na

equivalent circuit with resistor-grids was modeled as repetitive nine rows of 4-mm-

square blocks.
ur

The impact damage in Figs. 3(d)–(f) and the square [08] CFRP was 40 and 16 mm

along the fiber and perpendicular to the fiber, respectively. The measured damage is
Jo

depicted in the equivalent model (Fig. 9(d)) in terms of the size of the unit box. Finally,

the discrepancy between the measured and calculated changes in resistance due to

impact damage decreased (Table 3).

14
3.4.2 Electrical route modeling and non-destructive evaluation

Cross-ply CFRPs have a 3D electrical path such as in-plane and through-thickness

routes (Fig. 10), whereas CFRPs made of either the 0º or 90º UDCFs have in-plane

paths (Figs. 9(a) and 9(b)). The initial path before impact is demonstrated by the yellow

highlighted lines. The effective paths of the cross-ply CFRPs prefer intra-tow routes

with the shortest inter-tow route as electrons prefer the path of least resistance.

The equation corresponding to the electrical route model of the cross-ply CFRPs is

of
proposed as follows:

∆R = 𝜌𝑡ℎ 𝑖𝑐𝑘 × 𝑛 𝑝𝑙𝑦 × 2 + 𝜌𝑡ℎ 𝑖𝑐𝑘 × 1 𝑝𝑙𝑦 × 2 + 𝜌𝑖𝑛𝑡𝑟𝑎 × 𝐷𝑒𝑡𝑜𝑢𝑟 × 2 (5)

ro
where ∆R indicates the measured change in resistance, 𝜌𝑡ℎ 𝑖𝑐𝑘 is the electrical resistivity

in the through-thickness direction, 𝑛 is an integer representing the number of 90º


-p
UDCFs required to minimize the inter-tow network, 𝜌𝑖𝑛𝑡𝑟𝑎 indicates electrical resistivity
re
along the fiber, and 𝐷𝑒𝑡𝑜𝑢𝑟 indicates the half-width of the impact damage

perpendicular to the fiber. The steps to estimate the damage size, which is twice the size
lP

of 𝐷𝑒𝑡𝑜𝑢𝑟, using Eq. (5) are as follows: (1) Measure ∆R from one CFRP sample, (2)

use 𝑛, 𝜌𝑖𝑛𝑡𝑟𝑎, and 𝐷𝑒𝑡𝑜𝑢𝑟 from the actual sample and Table 1 to evaluate 𝜌𝑡ℎ 𝑖𝑐𝑘, (3) in
na

the new CFRP sample, measure ∆R with the known 𝜌𝑡ℎ 𝑖𝑐𝑘, 𝑛 and 𝜌𝑖𝑛𝑡𝑟𝑎, and (4)

calculate 𝐷𝑒𝑡𝑜𝑢𝑟 for NDE. As the electrical network is conformed along the path with
ur

the lowest electrical resistance, the autonomously developed model, Eq. (5), can be

rationalized.
Jo

For example, a database can be constructed to obtain 𝜌𝑡ℎ 𝑖𝑐𝑘 from the [03/901]s

sample (Fig. 10(a)), where 𝑛 = 2 and the 𝐷𝑒𝑡𝑜𝑢𝑟 is measured. Here, Eq. (5) is used with

the acquired 𝜌𝑡ℎ 𝑖𝑐𝑘 of a new CFRP, such as [02/902]s (Fig. 10(b)). The final result yields

the 𝐷𝑒𝑡𝑜𝑢𝑟; thus, NDE can be performed by monitoring the changes in the electrical

15
resistance of the cross-ply CFRPs. The average error of the suggested electrical route

model was ~2.53%.

The electromechanical sensitivity of the 4-1 electrode pair in the CFRPs was

analyzed (Fig. 11). The largest change in the electrical resistance was observed in the

[08] CFRPs, while other samples with 90º UDCFs and spread-woven fabrics exhibited

relatively lower changes in resistance. In particular, the resistance of the spread-woven

CFRP varied by 0.86%, which is ~40 times lesser than the sensitivity of [08], i.e.,

of
electrical conductivity was secured even though there was a puncture in between the

electrodes.

ro
Moreover, the [03/901]s, [02/902]s, [03/901]s, and [908] samples exhibited relatively

lower sensitivity than the [08] CFRP. However, there was no correlation between the
-p
number of 90º UDCFs and sensitivity. Irrespective of their number, the presence of the
re
90º UDCFs lowered the ratio of the resistance change, as they replaced the inter-tow

network with intra-tow and through-thickness networks. As shown in Table 1, the intra-
lP

tow network was 42 times more conductive than the inter-tow network. In addition, the

through-thickness network was more conductive than the in-plane inter-tow network, as
na

the total path length in the through-thickness direction was much shorter. Therefore, the

suggested equivalent model proves the electromechanical phenomenon observed in Fig.


ur

11.
Jo

4. Conclusions

In this study, the electromechanical behavior of CFRPs with various stacking

orientations was investigated using drop-weight impact testing. Changes in electrical

resistance were analyzed in terms of electrode distance when samples were punctured.

16
Using an optimal electrode distance, square CFRPs were investigated with eight-

electrode and 16-electrode pairings. The impact-struck-area represented the largest

change in electrical resistance and the UDCFs disseminated the impact energy, owing to

which the fiber-connected areas also showed comparatively large changes.

Electrically equivalent circuit models were proposed by converting the UDCFs and

their contacts into electrical resistors. One model consisted of a grid with electrical

resistors for the [08] and [908] CFRPs, while another model used route modeling in

of
which the easiest path for electron movement in the cross-ply CFRPs was employed.

Using the proposed models, damage severity could be estimated. In addition, the models

ro
subsequently determined the electromechanical sensitivity of the system. The

representative novelty of this study is the electrically equivalent circuit models of the
-p
CFRPs with various fiber configurations that can estimate the size of a damage.
re
Author statement
lP

Hyung Doh Roh: Sample manufacturing, mechanical testing, electromechanical


analysis, writing, reviewing, and editing
na

So Young Oh: Hyung Doh Roh and So Young Oh equally contributed to this paper:
Sample manufacturing, mechanical testing, chemical analysis, writing, reviewing, and
ur

editing

Young-Bin Park: Supervision as a corresponding author.


Jo

All authors read and approved the final manuscript.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal

17
relationships that could have appeared to influence the work reported in this paper

Declaration of interests

The authors declare that they have no known competing financial interests or personal

relationships that could have appeared to influence the work reported in this paper.

Acknowledgments

This work was supported by the National Research Foundation of Korea (NRF) grant

of
funded by the Ministry of Science and ICT, Korea (NRF-2017R1A5A1015311).

ro
-p
re
lP
na
ur
Jo

18
References
[1] C. Wu, Y. Gao, J. Fang, E. Lund, Q. Li, Discrete topology optimization of ply
orientation for a carbon fiber reinforced plastic (CFRP) laminate vehicle door, Materials
& Design, 128 (2017) 9-19. https://doi.org/10.1016/j.matdes.2017.04.089.

[2] P. Feraboli, A. Masini, L. Taraborrelli, A. Pivetti, Integrated development of CFRP


structures for a topless high performance vehicle, Composite Structures, 78 (2007) 495-
506. https://doi.org/10.1016/j.compstruct.2005.11.011.

[3] W. Wu, Q. Liu, Z. Zong, G. Sun, Q. Li, Experimental investigation into transverse
crashworthiness of CFRP adhesively bonded joints in vehicle structure, Composite
Structures, 106 (2013) 581-9. https://doi.org/10.1016/j.compstruct.2013.07.009.

of
[4] Y. Lin, M. Gigliotti, M.C. Lafarie-Frenot, J. Bai, D. Marchand, D. Mellier,
Experimental study to assess the effect of carbon nanotube addition on the through-
thickness electrical conductivity of CFRP laminates for aircraft applications, Composites

ro
Part B: Engineering, 76 (2015) 31-7. https://doi.org/10.1016/j.compositesb.2015.02.015.

[5] F. Ahmad, J.-W. Hong, H.S. Choi, M.K. Park, Hygro effects on the low-velocity

Composite Structures, 135


https://doi.org/10.1016/j.compstruct.2015.09.040.
-p
impact behavior of unidirectional CFRP composite plates for aircraft applications,
(2016) 276-85.
re
[6] K. Shetty, R. Bojja, S. Srihari, Effect of hygrothermal aging on the mechanical
properties of IMA/M21E aircraft-grade CFRP composite, Advanced Composites Letters,
lP

29 (2020) 2633366X20926520. 10.1177/2633366X20926520.

[7] W.-X. Wang, T. Matsubara, J. Hu, S. Odahara, T. Nagai, T. Karasutani, et al.,


Experimental investigation into the influence of the flanged diffuser on the dynamic
na

behavior of CFRP blade of a shrouded wind turbine, Renewable Energy, 78 (2015) 386-
97. https://doi.org/10.1016/j.renene.2015.01.028.

[8] R. Yang, Y. He, H. Zhang, Progress and trends in nondestructive testing and evaluation
ur

for wind turbine composite blade, Renewable and Sustainable Energy Reviews, 60 (2016)
1225-50. https://doi.org/10.1016/j.rser.2016.02.026.

[9] H.A. Maples, S. Wakefield, P. Robinson, A. Bismarck, High performance carbon fibre
Jo

reinforced epoxy composites with controllable stiffness, Composites Science and


Technology, 105 (2014) 134-43. https://doi.org/10.1016/j.compscitech.2014.09.008.

[10] Y. Yu, B. Zhang, Z. Tang, G. Qi, Stress transfer analysis of unidirectional composites
with randomly distributed fibers using finite element method, Composites Part B:
Engineering, 69 (2015) 278-85. https://doi.org/10.1016/j.compositesb.2014.09.035.

19
[11] Y. Miyano, M. Nakada, K. Nishigaki, Prediction of long-term fatigue life of quasi-
isotropic CFRP laminates for aircraft use, International Journal of Fatigue, 28 (2006)
1217-25. https://doi.org/10.1016/j.ijfatigue.2006.02.007.

[12] M. Meng, H. Le, S. Grove, M. Jahir Rizvi, Moisture effects on the bending fatigue
of laminated composites, Composite Structures, 154 (2016) 49-60.
https://doi.org/10.1016/j.compstruct.2016.06.078.

[13] L. Liu, X. Wang, Z. Wu, T. Keller, Optimization of multi-directional fiber


architecture for resistance and ductility of bolted FRP profile joints, Composite Structures,
248 (2020) 112535. https://doi.org/10.1016/j.compstruct.2020.112535.

[14] S.-i. Takeda, Y. Aoki, Y. Nagao, Damage monitoring of CFRP stiffened panels under
compressive load using FBG sensors, Composite Structures, 94 (2012) 813-9.

of
https://doi.org/10.1016/j.compstruct.2011.02.020.

[15] Z. Zhou, Z. Wang, L. Shao, Fiber-Reinforced Polymer-Packaged Optical Fiber Bragg

ro
Grating Strain Sensors for Infrastructures under Harsh Environment, Journal of Sensors,
2016 (2016) 3953750. 10.1155/2016/3953750.
-p
[16] R. Ruzek, M. Kadlec, K. Tserpes, E. Karachalios, Monitoring of compressive
behaviour of stiffened composite panels using embedded fibre optic and strain gauge
sensors, International Journal of Structural Integrity, 8 (2017) 134-50. 10.1108/IJSI-11-
re
2015-0052.

[17] S. Park, J.-W. Kim, C. Lee, S.-K. Park, Impedance-based wireless debonding
lP

condition monitoring of CFRP laminated concrete structures, NDT & E International, 44


(2011) 232-8. https://doi.org/10.1016/j.ndteint.2010.10.006.

[18] C. Andreades, G.P. Malfense Fierro, M. Meo, A nonlinear ultrasonic SHM method
na

for impact damage localisation in composite panels using a sparse array of piezoelectric
PZT transducers, Ultrasonics, 108 (2020) 106181.
https://doi.org/10.1016/j.ultras.2020.106181.
ur

[19] A. Todoroki, K. Omagari, Y. Shimamura, H. Kobayashi, Matrix crack detection of


CFRP using electrical resistance change with integrated surface probes, Composites
Science and Technology, 66 (2006) 1539-45.
Jo

https://doi.org/10.1016/j.compscitech.2005.11.029.

[20] A. Todoroki, M. Tanaka, Y. Shimamura, Measurement of orthotropic electric


conductance of CFRP laminates and analysis of the effect on delamination monitoring
with an electric resistance change method, Composites Science and Technology, 62 (2002)
619-28. https://doi.org/10.1016/S0266-3538(02)00019-2.

[21] A. Todoroki, The effect of number of electrodes and diagnostic tool for monitoring
the delamination of CFRP laminates by changes in electrical resistance, Composites

20
Science and Technology, 61 (2001) 1871-80. https://doi.org/10.1016/S0266-
3538(01)00088-4.

[22] T. Yamane, A. Todoroki, Doublet analysis of changes in electric potential induced


by delamination cracks in carbon-fiber-reinforced polymer laminates, Composite
Structures, 176 (2017) 217-24. https://doi.org/10.1016/j.compstruct.2017.05.019.

[23] J. Yin, R.-g. Liu, J.-j. Huang, G. Liang, D. Liu, G.-h. Xie, Comparative study on
piezoresistive properties of CFRP tendons prepared by two different methods,
Composites Part B: Engineering, 129 (2017) 124-32.
https://doi.org/10.1016/j.compositesb.2017.07.064.

[24] R.-g. Liu, Z.-h. Xu, J. Yin, J.-j. Huang, D. Liu, G.-h. Xie, A coupled mechanical and
electrical model concerning piezoresistive effect of CFRP materials, Composites Part B:

of
Engineering, 96 (2016) 125-35. https://doi.org/10.1016/j.compositesb.2016.04.010.

[25] J.B. Park, T. Okabe, N. Takeda, New concept for modeling the electromechanical

ro
behavior of unidirectional carbon-fiber-reinforced plastic under tensile loading, Smart
Materials and Structures, 12 (2003) 105-14. 10.1088/0964-1726/12/1/312.
-p
[26] Z. Xia, T. Okabe, J.B. Park, W.A. Curtin, N. Takeda, Quantitative damage detection
in CFRP composites: coupled mechanical and electrical models, Composites Science and
Technology, 63 (2003) 1411-22. https://doi.org/10.1016/S0266-3538(03)00083-6.
re
[27] M.C. Serna Moreno, S. Horta Muñoz, Mechanical response of ±45° angle-ply CFRP
plates under low-velocity impact and quasi-static indentation: Influence of the
lP

multidirectional strain state, Composites Science and Technology, 194 (2020) 108145.
https://doi.org/10.1016/j.compscitech.2020.108145.

[28] O.T. Topac, B. Gozluklu, E. Gurses, D. Coker, Experimental and computational


na

study of the damage process in CFRP composite beams under low-velocity impact,
Composites Part A: Applied Science and Manufacturing, 92 (2017) 167-82.
https://doi.org/10.1016/j.compositesa.2016.06.023.
ur

[29] X.C. Sun, S.R. Hallett, Barely visible impact damage in scaled composite laminates:
Experiments and numerical simulations, International Journal of Impact Engineering, 109
(2017) 178-95. https://doi.org/10.1016/j.ijimpeng.2017.06.008.
Jo

[30] A. Elias, F. Laurin, M. Kaminski, L. Gornet, Experimental and numerical


investigations of low energy/velocity impact damage generated in 3D woven composite
with polymer matrix, Composite Structures, 159 (2017) 228-39.
https://doi.org/10.1016/j.compstruct.2016.09.077.

[31] Y. Duan, H. Zhang, X.P.V. Maldague, C. Ibarra-Castanedo, P. Servais, M. Genest, et


al., Reliability assessment of pulsed thermography and ultrasonic testing for impact

21
damage of CFRP panels, NDT & E International, 102 (2019) 77-83.
https://doi.org/10.1016/j.ndteint.2018.11.010.

[32] A. Trellu, G. Pichon, C. Bouvet, S. Rivallant, B. Castanié, J. Serra, et al., Combined


loadings after medium velocity impact on large CFRP laminate plates: Tests and enhanced
computation/testing dialogue, Composites Science and Technology, 196 (2020) 108194.
https://doi.org/10.1016/j.compscitech.2020.108194.

[33] S.T. Rakotonarivo, C. Payan, J. Moysan, C. Hochard, Local damage evaluation of a


laminate composite plate using ultrasonic birefringence of shear wave, Composites Part
B: Engineering, 142 (2018) 287-92. https://doi.org/10.1016/j.compositesb.2018.01.006.

of
ro
-p
re
lP
na
ur
Jo

22
Biographies

Hyung Doh Roh is a post-doctor at the Ulsan National Institute of Science and

Technology, Republic of Korea. He received his B.S. degree in February 2014 and

Ph.D. degree in August 2020 at the Ulsan National Institute of Science and Technology.

The research area of his Ph.D. thesis was structural health self-sensing of carbon fiber-

reinforced plastics using electrical resistance.

of
So Young Oh is a graduate student at the Ulsan National Institute of Science and

ro
Technology, Republic of Korea. She received her B.S. degree in February 2020 at the

Ulsan National Institute of Science and Technology. Her research interests include
-p
structural health monitoring of carbon fiber-reinforced plastics and analyzing the
re
electrical network of the carbon fiber.
lP

Young-Bin Park is a full-time professor at the Department of Mechanical Engineering


na

in Ulsan National Institute of Science and Technology, Republic of Korea. He received

his B.S. and M.S. degrees at the Seoul National University, Republic of Korea, in 1995
ur

and 1997, respectively. His Ph.D. degree was received at the Georgia Institute of

Technology, USA in 2003. His research field is multi-scale functional composites and
Jo

composite manufacturing.

23
List of Figures

Figure 1. Photographs of CFRP samples for (a) the analysis of electrode distance and
(b) damage localization, (c) Schematic of the square CFRP sample with electrode
numbers, (d) CFRP for electrical resistivity measurement

Figure 2. Photographs of (a) impact testing and (b) clamping of a CFRP sample

Figure 3. Photographs of impact-struck-CFRPs: (a) Top view of the CFRP with 0˚


UDCFs, (b) Zoomed in top view of the CFRP with 0˚ UDCFs, (c) Zoomed in bottom
view of the CFRP with 0˚ UDCFs, (d) Top view of the CFRP with 90˚ UDCFs, (e)
Zoomed in top view of the CFRP with 90˚ UDCFs, (f) Zoomed in bottom view of the
CFRP with 90˚ UDCFs, (g) Top view of the CFRP with 0˚ square CFRP sample, (h)

of
Zoomed in front view of the CFRP with 0˚ UDCFs, (i) Zoomed in side view of the
CFRP with 0˚ UDCFs

ro
Figure 4. Energy-time graphs of the square CFRP samples whose fiber stacking
configurations were (a) [08], (b) [03/901]s, (c) [02/902]s and (d) [01/903]s
-p
Figure 5. Electromechanical results of (a)-(d) [08] and (e)-(h) [908] CFRPs
re
Figure 6. Electromechanical results of square CFRPs: (a)-(d) [08], (e)-(h) [908]

Figure 7. Electromechanical results of cross-ply CFRP: [03/901]s


lP

Figure 8. Electromechanical results of UD CFRPs: (a)-(d) [02/902]s, (e)-(h) [01/903]s

Figure 9. FEA results of [08] CFRP (a) without damage and (b) with damage. Electrically
equivalent circuit model (c) without damage and (d) with damage
na

Figure 10. Schematics of electrical route modeling of (a) [03/901]s and (b) [02/902]s
CFRPs
ur

Figure 11. Comparative analysis for electromechanical sensitivities of CFRP samples

Table 1. Electrical resistance monitoring for electrically equivalent circuit modeling


Jo

Table 2. Electrical resistance monitoring for electrically equivalent circuit modeling

Table 3. Comparison of the theoretical electrical resistance from the model with the
measured resistance change from the sample.

24
of
Figure 1. Photographs of the CFRP samples for (a) the analysis of the electrode

ro
distance and (b) damage localization. (c) Schematic of the square CFRP sample with the
electrode numbers. (d) CFRP for electrical resistivity measurement.
-p
re
lP
na
ur
Jo

Figure 2. Photographs of (a) impact testing and (b) clamping of a CFRP sample.

25
of
ro
-p
re
lP

Figure 3. Photographs of the impact-struck-CFRPs: (a) Top view of the CFRP with 0˚
UDCFs, (b) zoomed in top view of the CFRP with 0˚ UDCFs, (c) zoomed in bottom
view of the CFRP with 0˚ UDCFs, (d) top view of the CFRP with 90˚ UDCFs, (e)
na

zoomed in top view of the CFRP with 90˚ UDCFs, (f) zoomed in bottom view of the
CFRP with 90˚ UDCFs, (g) top view of the CFRP with 0˚ square CFRP sample, (h)
zoomed in front view of the CFRP with 0˚ UDCFs, and (i) zoomed in side view of the
ur

CFRP with 0˚ UDCFs.


Jo

26
of
Figure 4. Energy-time graphs of the square CFRP samples whose fiber stacking

ro
configurations were (a) [08], (b) [03/901]s, (c) [02/902]s, and (d) [01/903]s.

-p
re
lP
na
ur
Jo

Figure 5. Electromechanical results of (a)–(d) [08] and (e)–(h) [908] CFRPs.

27
of
ro
-p
Figure 6. Electromechanical results of square CFRPs: (a)–(d) [08], (e)–(h) [908].
re
lP
na
ur
Jo

Figure 7. Electromechanical results of cross-ply CFRP: [03/901]s.

28
of
ro
-p
re
Figure 8. Electromechanical results of UD CFRPs: (a)–(d) [02/902]s, (e)–(h) [01/903]s.
lP
na
ur
Jo

29
of
ro
-p
re
lP

Figure 9. Finite element analysis results of the [08] CFRP (a) without damage and (b)
na

with damage. Electrically equivalent circuit model (c) without damage and (d) with
damage.
ur
Jo

30
Figure 10. Schematics of electrical route modeling of (a) [03/901]s and (b) [02/902]s

of
CFRPs.

ro
-p
re
lP
na
ur

Figure 11. Comparative analysis of the electromechanical sensitivities of the CFRP


Jo

samples.

31
Table 1. Electrical resistance monitoring for electrically equivalent circuit modeling.

Along Perpendicular
the fiber (Ω) to the fiber (Ω)
0.158 4.9934
0.1065 5.6851
0.1253 5.1831

of
0.128 6.6084
0.1425 5.103

ro
0.1177 4.8488
Average (Ω)
0.125575 -p5.24115
re
lP

Table 2. Electrical resistance monitoring for electrically equivalent circuit modeling.


na

Block length 3 rows (Ω) 5 rows (Ω) 7 rows (Ω) 9 rows (Ω)

20 mm 1.0066 None None None


ur

10 mm 0.8656 0.8611 0.8610 None

8 mm 0.8036 0.7942 0.7939 0.7939


Jo

4 mm 0.6144 0.5682 0.5615 0.5605

32
of
ro
Table 3. Comparison of the theoretical electrical resistance from the model with the
measured resistance change from the sample.

Rinitial (Ω) Rfinal (Ω)


-p
∆Rcalculated (Ω) ∆Rmeasured (Ω) Error (%)
re
0.5605 1.1509 0.5904 0.5806 1.6879
lP
na
ur
Jo

33

You might also like