You are on page 1of 35

Accepted Manuscript

In-Situ Reduction of Graphene Oxide-Wrapped Porous Polyurethane Scaffolds:


Synergistic Enhancement of Mechanical Properties and Piezoresistivity

Yumeng Tang, Quanquan Guo, Zhenming Chen, Xinxing Zhang, Canhui Lu

PII: S1359-835X(18)30422-6
DOI: https://doi.org/10.1016/j.compositesa.2018.10.025
Reference: JCOMA 5225

To appear in: Composites: Part A

Received Date: 23 August 2018


Revised Date: 8 October 2018
Accepted Date: 17 October 2018

Please cite this article as: Tang, Y., Guo, Q., Chen, Z., Zhang, X., Lu, C., In-Situ Reduction of Graphene Oxide-
Wrapped Porous Polyurethane Scaffolds: Synergistic Enhancement of Mechanical Properties and Piezoresistivity,
Composites: Part A (2018), doi: https://doi.org/10.1016/j.compositesa.2018.10.025

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. 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.
In-Situ Reduction of Graphene Oxide-Wrapped Porous

Polyurethane Scaffolds: Synergistic Enhancement of

Mechanical Properties and Piezoresistivity

Yumeng Tang‡,1 Quanquan Guo‡,1 Zhenming Chen,2 Xinxing Zhang,1* and Canhui

Lu,1

1
State Key Laboratory of Polymer Materials Engineering, Polymer Research Institute

of Sichuan University, Chengdu 610065, China

2
Guangxi Key Laboratory of Calcium Carbonate Resources Comprehensive

Utilization,College of Materials & Environmental Engineering, Hezhou University,

Hezhou 542899, China

Abstract:

Lightweight conductive graphene@polyurethane (PU) sponge with unique 3D

structure and prominent piezoresistivity is a promising candidate for highly sensitive

pressure sensor. However, the poor mechanical properties of PU skeleton and the use

of toxic reductant for the reduction of graphene oxide (GO) greatly hampered its

large-scale application. Here, we demonstrated for the first time that aramid

Corresponding authors.
E-mail addresses: xxzwwh@ scu.edu.cn (X. Zhang), canhuilu@scu.edu.cn (C. Lu).
1
These authors contributed equally to this work.
nanofibers (ANFs) can be used as desirable building blocks to endow the

layer-by-layer (LBL) assembled GO@PU nanocomposites with excellent mechanical

properties. At the same time, the potassium hydroxide existed in ANFs solution could

in-situ reduce the GO without extra reductants. The as-prepared rGO@PU/ANFs

conductive nanocomposites exhibited excellent piezoresistive sensitivity and

repeatability, enabling the nanocomposites to serve as flexible sensors for the

application of human motions monitoring. This work opens up new opportunity for

the facile fabrication of high performance graphene-based piezoresistive sensors and

other electronic devices.

Keywords: aramid nanofibers; graphene; in-situ reduction; piezoresistive sensors.

1. Introduction

Piezoresistive pressure sensors have attracted considerable interests in scientific,

technological and commercial fields for a variety of promising applications, e.g.

human motion detection [1-15],personal healthcare [16, 18], energy harvesting [19]

and human-machine interactions [20-26] etc. Recently, conductive porous

polyurethane (PU) nanocomposites have been employed to fabricate flexible and

low-cost piezoresistive sensors with prominent sensitivity [27-30]. The sensing

mechanism of these sensors is based on the increased connection of the mechanical

microcrack junctions of PU sponges under pressure stress, which lead to the decrease

of resistivity. For example, Yao et al. utilized graphene-coated PU sponge to prepare


highly sensitive pressure sensor [31]. The piezoresistivity of conductive sponge is

determined by the variation rate of contact among backbones with compressive stress.

Similarly, Zhang et al. constructed a temperature- and pressure-sensor by depositing

organic thermoelectric material on the deformable microstructured PU frame [32].

Recently, inspired by the spider sensory system, our group has also fabricated a

flexible, highly sensitive, and versatile pressure sensing platform based on

microcrack-designed conductive PU sponge via layer-by-layer (LBL) assembly [33].

However, few studies paid attention to the mechanical weaknesses of the

sponge-based pressure sensors, arising from the intrinsic porous structure of PU

sponges. When these sensors were used in practical application, they had poor

durability and thus couldn’t maintain the accuracy of electric signals. On the other

hand, toxic reducing agents (i.e. hydrazine hydrate, hydroiodic acid, etc.) were usually

employed in these methods to endow the graphene oxide (GO)-decorated PU sponge

nanocomposites with electric conductivity, which is harmful to human health and the

environment [34]. Therefore, improving the mechanical properties of these sensors

and developing more sustainable reducing approach for GO are highly desired for

their practical application.

Aramid fibers (AFs), such as Kevlar, have been widely used as the reinforcing

fillers of advanced composites for aerospace, automotive, marine and construction

industries due to their outstanding mechanical properties [35-39]. Recently, it was

reported that aramid macroscale fibers could be split into nanofibers by deprotonation

of the amide groups in a solvent system consisting of dimethylsulfoxide (DMSO) and


potassium hydroxide (KOH) [40]. Owing to their abundant functional groups and

mechanical properties, aramid nanofibers (ANFs) are widely employed as

high-performance polymeric building blocks for advanced PU composites [41]. For

example, Kuang et al. assembled PU with ANFs through LBL assembly [42]. The

resultant nanocomposites exhibited significant mechanical improvement with

record-high stiffness and ultimate strength, owing to the multiple intermolecular

interactions between PU and ANFs.

Here, we demonstrated for the first time that ANFs can be used as desirable

building blocks to endow the LBL assembled GO@PU porous nanocomposites with

excellent mechanical properties. At the same time, the KOH existed in ANFs solution

could in-situ reduce the GO without extra reductants. In this study, a facile

water-based LBL assembly method was used to prepare GO@PU sponge combined

with in-situ reduction of GO with KOH. No extra reducing agents or stabilizers were

involved in this elaborate strategy, which avoid the complicated process of removing

reducing agents from the resulted reduced graphene oxide (rGO)-decorated

nanocomposites. The as-prepared rGO@PU/ANFs nanocomposite exhibited

remarkably improved mechanical properties compared with the original GO@PU

nanocomposites. Owing to the great electric conductivity and high piezoresistive

sensitivity, it can be used as a pressure sensor for human motion detection. The

proposed fabrication process provides a convenient and green approach for scalable

fabrication of high performance flexible electronic devices with improved mechanical

and multifunctional properties.


2. Materials and Methods

2.1 Materials

Graphite powder was purchased from J&K Chemical, Inc. (China). Sulfuric acid

(H2SO4), phosphoric acid (H3PO4), hydrochloric acid (HCl), dimethyl sulfoxide

(DMSO), potassium hypermanganate (KMnO4) and chitosan (CS) were all purchased

from Chengdu Kelong Chemical Reagent Company (China) and used without further

purification. PU sponges were cut from commercial cleaning sponges (3M Company,

USA). The Kevlar fiber (solid content: 58 wt%) was provided by Qingdao Benzo

Advanced Materials Co., Ltd (China).

2.2 Preparation of GO@PU composites

GO was prepared according to the modified Hummer’s method [43]. The graphite

powder (3.0 g, 1 wt eqiv) and KMnO4 (18.0 g, 6 wt eqiv) were mixed and slowly

added into a mixture of concentrated H2SO4/H3PO4 (360:40 vol./vol.) solution. Then

the obtained blends were heated to 50 oC and stirred for 12 h. After that, the blends

were cooled to room temperature and kept it in the ice bath with 30 % H 2O2 (3 mL).

The resulted products were purified by repeated centrifuge and dialysis in deionized

water.

The GO@PU composites were fabricated via LBL assembly by alternately dipping

PU sponges into positively charged CS solution and negatively charged GO

suspensions according to our previous studies [8, 33]. For the preparation of positively

charged CS solution, 1.0 g glacial acetic acid was added into 100 mL deionized water

drop by drop, and then 1.0 g CS was added with agitation intensively. The CS
solution was diluted to 1000 mL (1.0 mg/mL) when used as a charge mediator. In a

typical fabrication procedure for negatively charged GO suspensions, 200 mg of GO

was dispersed in 200 mL of deionized water with stirring vigorously. First, the

cleaned PU sponge was dipped into CS solution to form positively charged surface.

Then the positively charged PU sponge was dipped into negatively charged GO

suspension for deposition of GO on the surface of PU sponge. This process was

repeated to allow the LBL assembly. During these LBL cycles, the GO@PU sponge

was dried after every five LBL cycles. Finally, the obtained GO@PU sponge was

dried at 80 oC for 6 h to remove residual water. GO@PU sponges with different GO

content were prepared with the same procedure mentioned above.

2.3 Preparation of rGO@PU/ANFs nanocomposites

For the preparation of ANFs solution, 1 g Kevlar pulp (from Thread Exchange,

right twist) and 1.5 g KOH were added into 500 mL of DMSO, which were

subsequently magnetically stirred for 1 week at room temperature according to Kotov

et al.’s study. Then, GO@PU sponges with different GO content were added into a

desired amount of ANFs dispersion, followed by heating to 90 oC. After reacted for

10 h, the resulted composites were taken out and purified by deionized water. Then

the composites were dried at 60 oC for 12 h until the weight remained unchanged.

2.4 Characterization

Transmission electron microscopy (TEM) was performed to observe the

morphology of ANFs using a transmission electron microscope (JEOL JEM-100CX,

Japan). Diluted ANFs/DMSO suspension (0.05 mg/mL) was directly dropped on a


copper grid for observation. Before the test, ANFs/DMSO mixture was stirred

vigorously to obtain a homogenous suspension with a concentration of 2 mg/mL.

Attenuated total reflectance Fourier transform infrared (ATR/FTIR) spectroscopy

was performed to characterize chemical construction of AFs and ANFs using a

Nicolet 560 spectrophotometer (USA) ranging from 400 to 4000 cm-1 with a

resolution of 2 cm-1. All the samples were dried in vacuum oven at 60 oC for 24 h

before testing.

Scanning electron microscopy (SEM, JEOL JSM-5600LV, Japan) was utilized to

characterize the morphology of PU sponges, GO@PU and rGO/ANFs@PU

nanocomposites. The samples were cryo-fractured in liquid nitrogen and the fractured

surfaces of the cross-section were sputtered with a thin layer of Au for better

observation.

Laser scanning confocal microscopy (LSCM, LSM700 Carl Zeiss, Germany) was

carried out with reflection mode to detect the morphology evolution of the

nanocomposites.

X-ray diffraction (XRD) measurements were performed to characterize the layered

structure of the nanocomposites by a Philips Analytical X’Pert X-diffractometer

(Philips Co., Netherlands) operating in the continuous scanning mode. Diffraction

profiles were recorded by a copper K-alpha X-ray source at 40 kV/200 mA. XRD

spectra were collected in the 2θ range from 6o to 35o with a resolution of 0.02o.

Elemental composition analyses of GO and rGO were carried out using X-ray

photoelectron spectroscopy (XPS, XSAM800, Kratos Inc., Britain). The obtained


XPS spectra were analyzed via XPS peak4.1 software and Shirley was used as

background when fitting the peaks.

Raman spectroscopy was carried out to further evaluate the reduction efficiency of

GO, using a Horiba LabRAM HR apparatus (Horiba Co., France) with an excitation

laser wavenumber of 532 nm.

Thermal gravimetric analysis (TGA) was performed on TA-2000

thermogravimetric analyzer (TA Instruments, USA) at a heating rate of 5 ºC/min

under a steady nitrogen atmosphere from 100 to 800 oC.

Mechanical properties measurements were carried out on a versatile testing

machine (Instron-5560 electronic universal material testing machine, USA) at room

temperature according to ASTM standards D 3574. Cubic samples underwent uniaxial

compression tests at a crosshead rate of 2 mm/min by imposing a constant rate of 1

k/N.

For the electrical conductivity measurement, a resistance meter (UT61, Uni-trend,

China) was employed to measure the electric resistance of all samples by a two-point

measurement. In order to make sure the good contact, silver paint was used to connect

the tiny gap between samples and the electrodes. And all sensing performances were

characterized in real-time by a two-point measurement with a Keithley 2601B source

meter (USA).

3. Results and discussions

3.1 Materials design


ANFs have been proved as a pronounced reinforcement material for

nanocomposites due to its intrinsic high strength and modulus [44]. Dark yellow

solution of ANFs (Fig. S1, ESI†) was obtained by adding the aramid macroscale

fibers into a solvent system containing DMSO and KOH. Detailed structure

characterization of ANFs was conducted and discussed in ESI (Fig. S2-4†). On the

other hand, GO nanosheets with huge specific surface area and graphitized basal

plane structure allow them to have strong π-π interactions with ANFs in solution [45,

46 ]. Thus, the incorporation of ANFs and GO into PU sponge is expected to obtain a

mechanically robust GO@PU/ANFs nanocomposite with strong interfacial

interactions.

Recently, Fan et al. reported a novel way to prepare rGO by simply heating an

exfoliated GO suspension in the presence of strong alkaline at moderate temperatures

(50–90 oC) [47]. Its underlying mechanism might be the exfoliated GO deoxygenated

under alkaline conditions, appeared as the reverse of the oxidation reaction of graphite

in strong acids. Chemical reduction process of GO under strong alkali has been

considered as an effective route to prepare rGO nanocomposites as no additional

reductant is involved during the synthesis process [48]. Inspired by this pioneering

work, for the first time, KOH existed in ANFs dispersion was used as an efficient

in-situ reductant for GO, which could be removed easily by deionized water after the

reduction.

The fabrication procedure of the rGO@PU/ANFs nanocomposites is schematically

illustrated in Fig. 1. Chitosan is a natural, biocompatible and biodegradable


polycationic polysaccharide. It is generally obtained from chitin which is the second

most plentiful natural biopolymers on the earth [49, 50]. According to our previous

work [8, 33], the CS aqueous solution is positively charged, which can improve the

interfacial adhesion between GO and PU sponge through electrostatic interactions.

Thereby, LBL assembly process was employed to fabricate GO@PU nanocomposites

via alternative dipping PU sponges to positively charged CS solution and negatively

charged GO suspensions. In this study, we varied the contents of GO by the different

deposition cycles of GO.

3.2 Characterization of rGO@PU/ANFs nanocomposites

SEM and LSCM measurements were carried out to intuitively observe the

assembled structure and evaluate the change of the surface morphology of the

sponges. The surface morphology of original PU sponge is shown in Figs. 2a, d, and

g. The pristine PU sponge presents three-dimensional and cellular-like networks, and

the surface of PU bone is very smooth. After LBL assembly process, layers of

tulle-like GO nanosheets were coated on the surface of the cellular-like networks of

PU sponge, indicating the successful deposition of CS chains and GO nanosheets

layers (Figs. 2b, e, and h). We characterized the cross-sectional surface morphology of

GO@PU nanocomposites. As shown in Fig. S6†, GO was uniformly wrapped on each

side of the PU skeleton due to the excellent water dispersibility of GO, which is

similar with GO-coated sponges reported in other works [51, 52]. In addition, we

evaluated the porosity of PU sponge before and after LBL assembly. The mean cell

size of the pristine PU sponge is 295.3 μm, and it decreased to 230.2 μm after
decorated with 10 layers of GO. Morphology images of PU sponges after reacted with

ANFs suspension are illustrated in Figs. 2c, f, and i. It can be found that plenty of

tulle-like structured ANFs nanorods/particles are absorbed on the PU sponge skeleton

owing to the π-π interactions between rGO and ANFs, and the sizes of which are

consistent with the TEM images of ANFs (Fig. S2, ESI†). These results explicitly

indicate the successful construction of a connected conductive network on the surface

of PU sponge.

As shown in Fig. 3a, the amount of epoxide and hydroxyl groups decreased

significantly after in-situ reduction of GO. XRD patterns was used to illustrate the

layered structure of GO and rGO/ANFs (Fig. S5, ESI†). Raman spectra were further

performed to prove the successful reduction of GO nanosheets and the detailed

analysis are involved in Fig. 3b. The Raman spectrum of the original GO has two

main prominent peaks at 1349 and 1582 cm-1 corresponding to the D and G bands,

respectively. It is worth noting that the ratio of D to G band intensity (ID/IG) of rGO

(0.99) shows an obvious increase comparing to that of the original GO (0.92),

indicating that smaller in-plane sp2 domains were formed during the in-situ reduction

of GO.

The XPS survey spectra of GO and rGO are shown in Figs. 3c and d. The results

intuitively indicate the reduction effect since the content of oxygen atoms of rGO

remarkably reduces compared with GO. The high resolution C1s XPS spectrum of

GO (Fig. 3e) presents two prominent peaks centered at 284.6 and 286.8 eV, which can

be further divided into three Gaussian peaks with binding energy of 284.6 (Csp 2),
286.8 (C-O) and 288.0 eV (C=O), respectively. The C/O atomic ratio of the original

GO is 2.36, suggesting oxygen-containing groups were introduced into the graphite

during oxidation. Compared to the C1s spectrum of GO, the C1s spectrum of

rGO/ANFs displays bands at 284.7, 285.7, and 288.6 eV, due to Csp 2, C-O, and C=O,

respectively (Fig. 3f). The C/O atomic ratio of rGO increases to 5.25, demonstrating

that most oxygen containing functional groups in GO were removed, in line with

Raman analysis.

3.3 Enhanced mechanical and thermal properties

Owing to the existence of the intrinsic porous structure of PU and nanoscale

components, the nanocomposite is ultra-light (Fig. 4a) with the density as low as

0.101 g/cm3. The sample can be curled and twisted more than 90o, and then recover to

its original shape without fracture (Fig. 4b), indicating that the as-prepared

rGO@PU/ANFs nanocomposite has a superb flexibility.

The mechanical properties of the rGO@PU/ANFs nanocomposite are shown in Fig.

4c. In the compressing measurement, the nanocomposite was compressed at the strain

deformation of 60%. It can be found that the compressive strength of pristine sponge

is as low as 7 kPa. After LBL assembly with 5 and 10 layers of GO nanosheets, the

compressive strength of GO@PU nanocomposites increased to 11 and 16 kPa,

respectively. After reinforced by ANFs, the compressive strength of rGO@PU/ANFs

nanocomposites decorated with 5 and 10 layers of GO nanosheets remarkably

increased to 55 and 58 kPa, respectively, showing an approximately 5-fold

improvement compared with GO@PU composites. The results demonstrate that the
introduction of ANFs into GO@PU composites can significantly improve their

mechanical performance, which are ascribed to the inherent stiffness of ANFs and

their strong interfacial interaction with GO@PU matrix. When the nanocomposites

were under pressure, the nanoparticles wrapped on the sponge skeleton could function

as a load transfer to improve the mechanical properties [53, 54]. As the

nanocomposite coated with 10 layers of GO presents better compressive strength, it

was chosen to do the next characterization analysis.

Thermogravimetric analysis (TGA) was used to characterize the thermal stability of

the rGO@PU/ANFs nanocomposites (Fig. 4d). Impressively, neat ANFs have a high

thermal stability with mass loss at around 450 oC. In contrast, pristine PU sponge

began to decompose at 250 oC and was decomposed completely before the

temperature reached 400 oC. For GO@PU composites, there is a mass loss around 200

o
C, which is caused by the pyrolysis of labile oxygen-containing groups from GO.

After reduction, the rGO@PU/ANFs nanocomposite exhibits no mass loss around 200

o
C, as most of the oxygen containing functional groups have been removed by the

treatment with KOH. Meanwhile, it has a major mass loss around 350 oC, showing

improved thermal stability compared to the original composites, which might be

attributed to the high thermal stability of the ANFs decorated on the graphene

nanosheets.

3.4 Piezoresistive sensing applications of rGO@PU/ANFs nanocomposite

As shown in the inset of Fig. 4a, the rGO@PU/ANFs nanocomposite was

integrated into the circuit of light-emitting diode (LED) lamp. The lamp was lighted
up as soon as the power supply was turned on, which showed the good electrical

conductivity of the as-prepared nanocomposites after the reduction of GO. The

conductivity of different nanocomposites with fixed size (10 mm x 10 mm x10 mm)

were given in Figs. S7 and the detailed discussion could be found in supporting

information. In addition, mechanical deformation can lead to breakage-reconstruction

of the conductive network of the nanocomposites, causing the variation of the

conductivity. Owing to the elaborately designed conductive nanostructure,

rGO@PU/ANFs nanocomposites show great potential application in fabricating

electronic devices.

The detection of the pressure in real-time was realized using an electrical analyzer.

Upon a constant voltage, the electrical current changes of rGO@PU/ANFs

nanocomposite were recorded with the variation of the applied pressure (Figs. 5a and

b). During the compression process, it is evident that the electrical resistance increases

progressively with the increasing strain (Fig. 5c). The piezoresistive-sensing

mechanism of the conductive sponge is based on the variation of contact among

conductive nanofibers during compressive deformation. As shown in Fig. 5c, external

pressure causes variation of the percolated conductive network in PU sponges, giving

electrical signal output and making them possible to detect external stimuli. It can be

seen that the current signal experienced a mild increase in the strain range of 0–56 %

(domain A). This is due to the counteraction between disconnection of the mechanical

microcrack junctions in rGO layers and the increased contact junctions between

rGO@PU conductive backbones. When the strain reached to 56% (domain B), the
current began to rise rapidly, due to the fact that the increasing contact area between

rGO@PU backbones plays a decisive role for the resistance change. Here, the gauge

factor (GF), defined as the ratio of relative changes in electrical resistance to the

compression strain, was calculated to evaluate the sensitivity of rGO@PU/ANFs

nanocomposite. As shown in picture of Fig. 5d, the GF in domain A (0-56% strain)

was calculated to be -0.81. However, the average GF value increased to -5.27 in

domain B (56-60% strain). In addition, we prepared other two specimens to study the

repeatability of responsive current signals along with corresponding GF variation in

Fig. S8. The sensitivity of this pressure sensor was shown in Fig. 5e, which defined as

the slope of the resistance response curve under a monotonously increased pressure.

The pressure sensitivity of our pressure sensor is as high as 1.06 kPa−1 in the low

pressure range (below 40 kPa), and then increases to 2.82 kPa−1 in the relatively

higher pressure range (40–58 kPa). In addition, we have listed the corresponding

value of current reported sensing materials to quantitatively compare the properties.

As shown in the table S1, our sensor shows an excellent mechanical property and high

sensitivity compared to that of some recently reported sensors. Moreover, the

current-voltage (I–V) characteristics of rGO@PU/ANFs nanocomposite under

different pressure were also studied (Fig. 5f). The results show that rGO@PU/ANFs

nanocomposite exihibited good linear I–V characteristics under different pressure, and

the slope of I–V curves increases as the pressure applied owing to the corresponding

decrease of resistance.

The electrical cyclic curves of rGO@PU/ANFs nanocomposite under different


strain/stress loads were recorded and plotted in Fig. 5g. It shows that this pressure

sensor can quantitatively detect the pressure and give corresponsding output signals in

real time. The intensity as well as the shape of these signal peaks varies with each

other at different strain values. The higher the strain reaches, the sharper the peaks

becomes. These distinguishing responsive signals in large strain region make the

rGO@PU sponges capable of detecting and differentiating pressure. Moreover, to

verified the durability and stabilities of as-prepared pressure sensor, the sensor was

compressed for 10,000 times under 2 kPa, and the current signal of the sensor was

recorded in real time (Fig. 5h). Cyclic tests of the sensor were conducted with large

amplitude (60% strain). Through more than 10 000 cyclic loading-unloading tests, the

electrical response of the piezoresistive sensor was nearly unchanged. The excellent

reproducibility of this sensor can be ascribed to the positively charged chitosan used

in LBL assembly, which improves the interfacial adhesion between GO and PU

sponge through electrostatic interactions. The result reveals that the piezoresistive

sensor possesses high stability and durability without obvious creep. The performance

of the sensor provides an essential condition to ensure its practical applications. In

addition, we also studied the piezoresistive properties of rGO@PU nanocomposites in

Fig. S7and detailed discussion can be found in supporting information.

The piezoresistive response of the rGO@PU/ANFs nanocomposite to repeated

pressure loading and unloading cycles was recorded in Fig. 6a. When the pressure

was loaded, the sensor was compressed and thus the current rised owing to the

decrease of resistance, which is consistent with the mechanism of previously reported


piezoresistive sensors. When the load was removed, the current gradually recovered

to its original state. At the large repeated strain region, a series of characteristic and

noise-free signal outputs were obtained. The intensity as well as the shape of these

signal peaks varied with each other at different strain values. Along with the repeated

pressure loading and unloading, the current patterns recorded are nearly invariable,

indicating that the response behavior of our rGO@PU/ANFs nanocomposites is stable

and repeatable (Movie S1†).

Benefit from the excellent mechanical properties and high piezoresistive sensitivity,

our nanocomposite sensor revealed good potential application in indetecting human

motions. To monitor the muscle motions, the insulating side of a sensor was attached

to the forefinger (insets of Fig. 6b), wrist (insets of Fig. 6c) and arm (insets of Fig. 6d)

of the volunteer with medical adhesive. As shown in Fig. 6b, when the volunteer

bended the forefinger downward, the sensor was compressed and thus the current

increased abruptly due to the increase of resistance. While the forefinger became

straight, the current gradually recovered to its original state. The current value

exhibited clear changes during the bending/releasing sequences (Movie S2†). We

further evaluated the capability of our sensors in monitoring the motions of the wrist

and the arm (as depicted in Fig. 6c and d) the sensor can clearly distinguish the

bending and relaxation of the wrist and the arm according to the output current, and

the response behaviors are repetitive. The piezoresistive sensor was compressed 10

000 time under versatile testing machine. And we tested the piezoresistive cycles

manually before and after experiencing 10 000 compression cycles, and then
representatively chose two different actions (finger and wrist bending) to carry out the

durability experiments. It can be seen that there is no obvious change in piezoresistive

sensing performance of our sensor even after bending 10 000 cycles, which confirms

the good durability of our strain sensor (Fig. S9, ESI†). The excellent durability of

this sensor can be ascribed to the positively charged chitosan used in LBL assembly,

which can improve the interfacial adhesion between GO and PU sponge through

electrostatic interactions according to our previous work. According to the current

signals, we can precisely distinguish the human movements via the rGO@PU/ANFs

sensor, which makes it promising for wearable electronics, soft robotics and artificial

intelligence applications.

4. Conclusions

In this work, we developed a facile route to obtain rGO@PU/ANFs nanocomposite

after LBL assembly. It was proved that ANFs remarkably enhanced the mechanical

properties of the nanocomposites. Interestingly, KOH in ANFs solution acted as

reductant to realize the in-situ reduction of GO via deoxygenation. This effective

strategy avoided the use of toxic reducing agents and incorporated ANFs as a

reinforcement into rGO@PU composites at the same time. The resulted porous

nanocomposite presented great electrical conductivity and excellent piezoresistive

sensing properties. Owing to the excellent properties and the advantage of easy

fabrication, our rGO@PU/ANFs nanocomposite showed great potential in the field of

flexible pressure sensor and other electronic devices applications.


ACKNOWLEDGMENT

This work was supported by National Natural Science Foundation of China

(51673121 and 51473100) and China Postdoctoral Science Foundation

(2017M610601). And we would like to thank Dr. Guiping Yuan from Analytical and

Testing Center of Sichuan University for providing TEM measurement facility.

REFERENCE

[1] Liao M, Wan P, Wen J, Gong M, Wu X, Wang Y, et al. Wearable, healable, and

adhesive epidermal sensors assembled from mussel‐inspired conductive hybrid

hydrogel framework. Adv Funct Mater 2017;27(48):1703852.

[2] Hou C, Wang H, Zhang Q, Li Y, Zhu M. Highly conductive, flexible, and

compressible all-graphene passive electronic skin for sensing human touch. Adv

Mater 2014;26(29):5018-24.

[3] Yu X, Li Y, Zhu W, Huang P, Wang T, Hu N, et al. A wearable strain sensor based

on a carbonized nano-sponge/silicone composite for human motion detection.

Nanoscale 2017;9(20):6680-5.

[4] Xie D, Qian D, Song F, Wang X, Wang Y. A fully biobased encapsulant

constructed of soy protein and cellulose nanocrystals for flexible

electromechanical sensing. ACS Sustainable Chem Eng 2017;5(8):7063-70.


[5] Chen S, Wei Y, Yuan X, Lin Y, Liu L. A highly stretchable strain sensor based on a

graphene/silver nanoparticle synergic conductive network and a sandwich

structure. J Mater Chem C 2016;4(19):4304-11.

[6] Guo Q, Cao J, Han Y, Tang Y, Zhang X, Lu C. Biological phytic acid as a

multifunctional curing agent for elastomers: towards skin-touchable and flame

retardant electronic sensors. Green Chem 2017;19:3418-27.

[7] Guo Q, Luo Y, Liu J, Zhang X, Lu C. A well-organized graphene nanostructure

for versatile strain-sensing application constructed by a covalently bonded

graphene/rubber interface. J Mater Chem C 2018;6:2139-47.

[8] Wu X, Han Y, Zhang X, Lu C. Highly sensitive, stretchable, and wash-durable

strain sensor based on ultrathin conductive layer@polyurethane yarn for tiny

motion monitoring. ACS Appl Mater Interface 2016;8(15):9936-45.

[9] Yoon SG, Koo HJ, Chang ST. Highly stretchable and transparent microfluidic

strain sensors for monitoring human body motions. ACS Appl Mater Interface

2015;7(49):27562-70.

[10] Fan X, Xu B, Wang N, Wang J, Liu S, Wang H, et al. Highly conductive

stretchable all-plastic electrodes using a novel dipping-embedded transfer method

for high-performance wearable sensors and semitransparent organic solar cells.

Adv Electron Mater 2017;3(5):1600471.


[11] Wang H, Liu Z, Ding J, Lepro X, Fang S, Jiang N, et al. Downsized sheath-core

conducting fibers for weavable superelastic wires, biosensors, supercapacitors,

and strain sensors. Adv Mater 2016;28(25):4998-5007.

[12] Gong S, Lai DT, Wang Y, Yap LW, Si KJ, Shi Q, et al. Tattoo like polyaniline

microparticle-doped gold nanowire patches as highly durable wearable sensors.

ACS Appl Mater Interface 2015;7 (35):19700-8.

[13] Wang J, Zhang H, Xie Y, Yan Z, Yuan Y, Huang L, et al. Smart network node

based on hybrid nanogenerator for self-powered multifunctional sensing. Nano

Energy 2017;33:418-26.

[14] Zhang S, Wang S, Wang Y, Fan X, Li D, Xuan S, et al. Conductive shear

thickening gel/polyurethane sponge: a flexible human motion detection sensor

with excellent safeguarding performance. Compos Part A Appl Sci Manuf

2018;(112):197-206.

[15] Zhou J, Yu H, Xu XZ, Han F, Lubineau G. Ultrasensitive, stretchable strain

sensors based on fragmented carbon nanotube papers. ACS Appl Mater Interfaces

2017; 9(5):4835-42.

[16] Trung TQ, Lee NE. Flexible and stretchable physical sensor integrated platforms

for wearable human-activity monitoring and personal healthcare. Adv Mater

2016;28(22):4338-72.
[17] Rito RL, Crocombe AD, Ogin SL. Health monitoring of composite patch repairs

using CFBG sensors: experimental study and numerical modelling. Compos Part

A Appl Sci Manuf 2017;100:255-68.

[18] Yao Q, Fan B, Xiong Y, Wang C, Wang H, Jin C, et al. Stress sensitive electricity

based on Ag/cellulose nanofiber aerogel for self-reporting. Carbohyd Polym

2017:168;265-73.

[19] Alluri NR, Selvarajan S, Chandrasekhar A, Saravanakumar B, Jeong JH, Kim SJ.

Piezoelectric BaTiO3/alginate spherical composite beads for energy harvesting

and self-powered wearable flexion sensor. Compos Sci Technol 2017;142:65-78.

[20] Xin Y, Zhou J, Xu XZ, Lubineau G. Laser-engraved carbon nanotube paper for

instilling high sensitivity, high stretchability, and high linearity in strain sensors.

Nanoscale 2017; 9:10897-905.

[21] Cao J, Lu C, Zhuang J, Liu M, Zhang X, Yu Y, et al. Multiple hydrogen bonding

enables the self-healing of sensors for human-machine interactions. Angew Chem

Int Edit 2017;129(30):8795-800.

[22] Wu X, Han Y, Zhang X, Lu C. Spirally structured conductive composites for

highly stretchable, robust conductors and sensors. ACS Appl Mater Inter

2017;9(27): 23007-16.
[23] Wu C, Fang L, Huang X, Jiang P. Three-Dimensional highly conductive

graphene-silver nanowire hybrid foams for flexible and stretchable conductors.

ACS Appl Mater Inter 2014;6(23):21026-34.

[24] Lee D, Lee H, Jeong Y,Ahn Y, Nam G, Lee Y. Highly sensitive, transparent, and

durable pressure sensors based on sea-urchin shaped metal nanoparticles. Adv

Mater 2016;28(42):9364-.

[25] Choong CL, Shim MB, Lee BS, Jeon S, Ko DS, Kang TH, et al. Highly

stretchable resistive pressure sensors using a conductive elastomeric composite on

a micropyramid array. Adv Mater 2014;26(21):3451-8.

[26] Liu X, Su G, Guo Q, Lu C, Zhou T, Zhou C. et al. Hierarchically structured

self-healing sensors with tunable positive/negative piezoresistivity. Adv Funct

Mater 2018;28.

[27] Hodlur RM, Rabinal MK. Self assembled graphene layers on polyurethane foam

as a highly pressure sensitive conducting composite. Compos Sci Technol

2014;90(90):160-5.

[28] Liu H, Huang W, Gao J, Dai K, Zheng G, Liu C, et al. Piezoresistive behavior of

porous carbon nanotube-thermoplastic polyurethane conductive nanocomposites

with ultrahigh compressibility. Appl Phys Lett 2016;108(1):918-24.


[29] Wu Y, Liu H, Chen S, Dong X, Wang P, Liu S, et al. Channel crack-designed

gold@PU sponge for highly elastic piezoresistive sensor with excellent

detectability. ACS Appl Mater Interface 2017;9(23):20098-105.

[30] Jeong YR,Park H, Jin SW, Hong SY, Lee SS, Ha JS. Highly stretchable and

sensitive strain sensors using fragmentized graphene foam. Adv Funct Mater

2015;25(27):4228-36.

[31] Yao H, Ge J, Wang C, Wang X, Hu W, Zheng Z, et al. A flexible and highly

pressure-sensitive graphene-polyurethane sponge based on fractured

microstructure design. Adv Mater 2013;25(46):6692-8.

[32] Zhang F, Zang Y, Huang D, Di C, Zhu D. Flexible and self-powered

temperature-pressure dual-parameter sensors using

microstructure-frame-supported organic thermoelectric materials. Nat Commun

2015;6:8356.

[33] Wu X, Han Y, Zhang X, Zhou Z, Lu C. Large-area compliant, low-cost, and

versatile pressure-sensing platform based on microcrack-designed carbon

black@polyurethane sponge for human-machine interfacing. Adv Funct Mater

2016;26(34):6246-56.

[34] Barman BK, Nanda KK. An ultrafast-versatile-domestic-microwave-oven based

graphene oxide reactor for the synthesis of highly efficient graphene based hybrid

electrocatalysts. ACS Sustainable Chem Eng 2018;6(3):4037–45.


[35] Cao K, Siepermann C, Yang M, Waas AM, Kotov NA, Thouless MD, et al.

Reactive aramid nanostructures as high-performance polymeric building blocks

for advanced composites. Adv Funct Mater 2013;23(16):2072-80.

[36] Li Z, Cheng X, He S, Shi X, Gong L, Zhang H. Aramid fibers reinforced silica

aerogel composites with low thermal conductivity and improved mechanical

performance. Compos Part A Appl Sci Manuf 2016;84(3):316-25.

[37] Zhu J, Yuan L, Guan Q, Liang G, Gu A. A novel strategy of fabricating high

performance UV-resistant aramid fibers with simultaneously improved surface

activity, thermal and mechanical properties through building polydopamine and

graphene oxide bi-layer coatings. Chem Eng J 2017;310:134-47.

[38] Qi G, Zhang B, Du S, Yu Y. Estimation of aramid fiber/epoxy interfacial

properties by fiber bundle tests and multiscale modeling considering the fiber

skin/core structure. Compos Struct 2017;167:1-10.

[39] Dai Y, Yuan Y, Luo L, Liu X. A facile strategy for fabricating aramid fiber with

simultaneously high compressive strength and high interfacial shear strength

through crosslinking promoted by oxygen. Compos Part A Appl Sci Manuf

2018;113:233–41.

[40] Yang M, Cao K, Sui L, Qi Y, Zhu J, Waas A, et al. Dispersions of aramid

nanofibers: a new nanoscale building block. ACS Nano 2011;5(9):6945-54.


[41] Little BK, Li Y, Cammarata V, Broughton R, Mills G. Metallization of kevlar

fibers with gold. ACS Appl Mater Interface 2011;3(6):1965-73.

[42] Kuang Q, Zhang D, Yu J, Chang Y, Yue M, Hou Y, et al. Toward record-high

stiffness in polyurethane nanocomposites using aramid nanofibers. J Phys Chem

C 2015;119(49):27467-77.

[43] Marcano DC, Kosynkin DV, Berlin JM, Sinitskii A, Sun Z, Slesarev A, et al.

Improved synthesis of graphene oxide. ACS Nano 2010;4(8):4806-14.

[44] Arrieta C, David E, Dolez P, Vu-Khanh T. X-ray Diffraction, Raman, and

Differential Thermal Analyses of the thermal aging of a Kevlar (R)-PBI blend

fabric. Polym Compos 2011;32(3):362-7.

[45] Fan J, Shi Z, Zhang L, Wang J, Yin J. Aramid nanofiber-functionalized graphene

nanosheets for polymer reinforcement. Nanoscale 2012;4(22):7046-55.

[46] Fan J. Shi Z, Tian M, Yin J. Graphene-aramid nanofiber nanocomposite paper

with high mechanical and electrical performance. RSC Adv 2013;3(39):17664-7.

[47] Fan, X, Peng W, Li Y, Li X, Wang S, Zhang G, et al. Deoxygenation of exfoliated

graphite oxide under alkaline conditions: a green route to graphene preparation.

Adv Mater 2010;20(23):4490-3.

[48] Yazid SNAM, Isa IM, Hashim N. Novel Alkaline-reduced cuprous

oxide/graphene nanocomposites for non-enzymatic amperometric glucose sensor

application. Mat Sci Eng C Mater 2016;68:465-73.


[49] Desai K, Kit K, Li J, Zivanovic S. Morphological and surface properties of

electrospun chitosan nanofibers. Biomacromolecules 2008; 9(3):1000-6.

[50] Yang X, Tu Y, Li L, Shang S, Tao X. Well-dispersed chitosan/graphene oxide

nanocomposites. ACS Appl Mater Interface 2010;2(6):1707-13.

[51] Zhu H, Chen D, Yang S, Li N, Xu Q, Li H, et al. A versatile and cost-effective

reduced graphene oxide-crosslinked polyurethane sponge for highly effective

wastewater treatment. RSC Adv 2016;6(44):38350-5.

[52] Zhang X, Liu D, Sui G. Superamphiphilic polyurethane foams synergized from

cellulose nanowhiskers and graphene nanoplatelets. Adv Mater Interfaces

2018;5:1701094.

[53] Gibson RF. A review of recent research on mechanics of multifunctional

composite materials and structures. Compos Struct 2010;92,(12):2793-2810.

[54] Cao X, Lee L, Widya T, Macosko C. Polyurethane/Clay nanocomposites foams:

processing, structure and properties. Polymer 2005;46(3):775-83.


Figure 1. Schematic diagram for the preparation of rGO@PU/ANFs nanocomposite.

Figure 2. SEM images of neat PU sponge (a and d), GO@PU (b and e), and

rGO@PU/ANFs (c and f) nanocomposites. LCSM images of neat PU sponge (g),

GO@ PU (h), and rGO@PU/ANFs (i) nanocomposites.

Figure 3. (a)The in-situ reduction progress of GO. Raman images (b) and XPS

spectra (c-f) of GO and rGO.

Figure 4. (a) The lightweight foam standing on a flower. And the inset shows the

electrical conductivity of the nanocomposites. (b) Photographs showing the excellent

flexibility of rGO@PU/ANFs nanocomposite during bending and twisting; (c)

Mechanical properties of neat PU, 5 layers GO@PU composite, 10 layers GO@PU

composite, and rGO@PU/ANFs nanocomposite accordingly. (d) TGA of neat PU,

ANFs, GO@PU and rGO@PU/ANFs nanocomposites.

Figure 5. (a) The compressive test of the rGO@PU/ANFs nanocomposite under

loading-unloading of 60% strain. (b) Schematic diagram of the compressive test.

(c) Responsive current signals, (d) corresponding GF variation, (e) the sensitivity

of the pressure sensor, (f) the current-voltage (I–V) characteristics, (g) the electrical

cyclic curves under different strain/stress loads and (h) the reproducibility test for 10

000 cycles under 2 kPa of rGO@PU/ANFs nanocomposite.

Figure 6. Characterization of the piezoresistive sensing performance of rGO@PU/ANFs nanocomposites. (a)

The piezoresistive responses of the rGO@PU/ANFs nanocomposites to repeated pressure loading and unloading.

Application of the piezoresistive sensors in human motion monitoring: the corresponding current response to the
forefinger (b), wrist (c) and arm (d) motion of the virgin and bent (inset: digital pictures of the released and bent

states of the finger, wrist and arm).

You might also like