imperial college london - electrically-responsive graphene ... · web viewdiglycidylether of...

20
Electrically-responsive graphene-based shape- memory composites D’Elia Eleonora a,, Hanaë Said Ahmed a , Ezra Feilden a , Eduardo Saiz a,a Centre of Advanced Structural Ceramics, Department of Materials, Imperial College London, Prince Consort Road, SW7 2BP, UK Corresponding Authors; Email: [email protected] Tel: +44 (0)20 7594 0976 Abstract Shape memory materials can open new design opportunities in fields as diverse as healthcare, transportation or energy generation. In this respect, shape memory polymers (SMPs) have attracted much attention due to their advantages over metals in terms of weight and reliability. However, they are still marred by slow reaction times and poor mechanical performance. In this work we show how by integrating a graphene network in a SMP matrix it is possible to create composites with very low carbon contents (below 1 wt%) able to change shapes in short times (10s of seconds) in response to low electric voltages (<10 V). This is possible because the conductive network is highly interconnected at the microscopic scale, acting as a very efficient Joule heater. The composites exhibit excellent shape fixity (>0.95 ± 0.03) and shape recovery ratios (>0.98 ± 0.03). Due to the 2D nature of graphene, this network directs crack propagation during fracture resulting in materials that retain bending strengths close to 100 MPa and exhibit significant extrinsic toughening (with toughness that reach values up to 3 times the initiation value). Furthermore, changes in conductivity

Upload: others

Post on 23-Feb-2021

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Imperial College London - Electrically-responsive graphene ... · Web viewdiglycidylether of bisphenol A (DGEBA, Sigma, weight per epoxy group = 174.25 g/mol), ndodecylamine (DA,

Electrically-responsive graphene-based shape-memory composites D’Elia Eleonora a,∗, Hanaë Said Ahmed a, Ezra Feilden a, Eduardo Saiz a,∗

a Centre of Advanced Structural Ceramics, Department of Materials, Imperial College London, Prince Consort Road, SW7 2BP, UK

∗ Corresponding Authors; Email: [email protected] Tel: +44 (0)20 7594 0976

Abstract

Shape memory materials can open new design opportunities in fields as diverse as healthcare,

transportation or energy generation. In this respect, shape memory polymers (SMPs) have attracted

much attention due to their advantages over metals in terms of weight and reliability. However, they

are still marred by slow reaction times and poor mechanical performance. In this work we show how

by integrating a graphene network in a SMP matrix it is possible to create composites with very low

carbon contents (below 1 wt%) able to change shapes in short times (10s of seconds) in response to

low electric voltages (<10 V). This is possible because the conductive network is highly

interconnected at the microscopic scale, acting as a very efficient Joule heater. The composites

exhibit excellent shape fixity (>0.95 ± 0.03) and shape recovery ratios (>0.98 ± 0.03). Due to the 2D

nature of graphene, this network directs crack propagation during fracture resulting in materials that

retain bending strengths close to 100 MPa and exhibit significant extrinsic toughening (with

toughness that reach values up to 3 times the initiation value). Furthermore, changes in conductivity

can be used to follow the formation and growth of damage in the material before catastrophic failure,

allowing the use of this material as a damage sensor. These results provide practical guidelines for

the design of reliable shape memory composites for structural and sensing applications.

Keywords: shape-memory composites; graphene; strength; toughness; freeze casting

1. Introduction

In recent decades the field of smart or responsive materials has progressed rapidly. Now we can

fabricate structures able of responding to external stimuli, such as heat, light [1], pressure or

Page 2: Imperial College London - Electrically-responsive graphene ... · Web viewdiglycidylether of bisphenol A (DGEBA, Sigma, weight per epoxy group = 174.25 g/mol), ndodecylamine (DA,

humidity [2], in fascinating ways[3]. These structures can act as sensors, self-repair or change shape

on demand [4-6]. There is an increasing push to move these smart materials from the scientific to the

engineering scene in the hope to find new and comprehensive solutions from daily consumer

products, to aerospace or healthcare. However, responsive materials usually exhibit poor mechanical

behaviour (low strength and toughness) impairing their use in load-bearing applications. In addition,

“smart” properties often cannot be combined in the same material. So far much of the progress has

been based on the development of new compounds (e.g. supramolecular polymers or shape memory

alloys). Much less work has been done on use of composites to address these shortcomings. [7, 8]

Composites are a very interesting alternative as they can combine materials with different

functionalities. However, effective composite design requires structural control at different length

scales from the nano-level and up.

Shape memory materials (SMMs) are able to recover their set original shape after deformation.

Applying the right combination of mechanical strain and stimulus, they can exist in either a

permanent or a temporary shape that can be virtually maintained forever [3]. This phenomenon is

called shape memory effect (SME). Due to their high versatility in term of compositions and the

wide palette of external stimuli that can be used to trigger shape changes (heat [9], electricity [10-

14], magnetism [15], light [1] or moisture [16, 17]), research on shape memory polymers (SMPs) has

greatly surpassed that on shape-memory alloys[18]. SMPs can be used in a wide range of

applications from healthcare, in stitches capable of tightening autonomously around a wound [19], to

robotics, to build actuators capable of acting as synthetic muscles [20]. Triggering shape changes in

SMPs usually involves removing the material from the site of use in order to apply the required

stimuli (e.g. homogeneous heating to the structure). To overcome this inconvenience, recent studies

have focussed their attention on the development of shape memory composites (SMCs), particularly

on polymers containing a secondary phase, be it magnetic or conductive particles, to elicit a response

in the material, in situ, at command, without the need for site alteration or removal [2]. Internal joule

heating is an attractive alternative to induce shape changes in situ by applying a current. This

requires an electrically conductive material, and, for this reason, carbon-based additives (e.g. carbon

nanotubes, graphite…) have been used to obtain electrically-responsive shape-memory composites

Page 3: Imperial College London - Electrically-responsive graphene ... · Web viewdiglycidylether of bisphenol A (DGEBA, Sigma, weight per epoxy group = 174.25 g/mol), ndodecylamine (DA,

[10, 14, 20]. The limitations of this method are often linked to processing. The standard procedure to

reach the percolation threshold and obtain high electrical conductivities involves the thorough

homogenisation of the filler within the polymers. This leads to long mixing times and large filler

amounts [14]. Moreover, these materials often exhibit poor mechanical properties, with strengths of

the order of a few MPa to kPa and their electrical conductivities remain limited (lower than a few

S/m). Therefore, high voltages (that can be above 90V) are required for actuation [21-23].

With its low dimensionality and high electrical conductivity, graphene is a very attractive alternative

for the fabrication of shape memory composites. Li et al. produced polymer/graphene foams

exhibiting a shape memory effect [24] and good homogeneity of the conductive phase. However their

material didn’t possess structural capabilities, the compressive strengths were only ~50 kPa, and

actuation was limited. The standing challenge is to fabricate composites with the ability to react to

relatively low voltages while retaining the mechanical properties required for load bearing

applications. In this work we show a path to build composite materials able to change shape and with

the potential to act as load-bearing structures. These materials combine graphene and shape memory

polymers. By controlling the structure from nanoscale and up, the conductive phase provides the

required actuation and plays a role in enhancing the mechanical response while introducing

additional capabilities. We have used freeze-casting to integrate the graphene in the form of a

microscopic, highly interconnected and conductive network that can act as a very effective joule

heater. Using this approach this can be achieved with very low graphene contents (~0.5 wt%),

avoiding the problems associated with the use of large filler contents. This network also directs crack

propagation during fracture; promoting toughening and stable crack growth. Furthermore, changes in

electrical resistivity triggered by its disruption can be used to monitor damage formation and

propagation in situ. These properties can be used to avoid the catastrophic failure of shape changing

structures.

2. Results and discussion

Graphene Oxide (GO) foams were obtained via the freeze-casting of GO suspensions in water. The

foams were subsequently reduced at 900 °C in Ar/H2. The reduced graphene oxide (rGO) foams (GF)

Page 4: Imperial College London - Electrically-responsive graphene ... · Web viewdiglycidylether of bisphenol A (DGEBA, Sigma, weight per epoxy group = 174.25 g/mol), ndodecylamine (DA,

are formed by parallel channels ~ 18 µm ± 3 µm wide. The walls have a thicknesses of 20-30 nm and

are formed by the accumulation of GO flakes between the ice crystals during freeze casting. They

exhibit porosities above 99% and electrical conductivities of ~70 S/m (Figure 1a). Cyclic

compression tests confirm their high flexibility and stress recovery even after 6 cycles of

compression up to 20% strain (Figure 1b) [25]. Composites are fabricated by infiltrating these foams

with shape-memory epoxy resin precursors followed by in situ curing. The resulting composites have

a total rGO content of ~ 0.5 wt.% and retain high electrical conductivities in the order of 40 S/m. A

similar methodology was previously reported in our group to obtain self-healing sensors [26] and

ceramic-graphene composites [27].

Two different compositions of SMPs were used as matrices. SMP1, composed of an aromatic and an

aliphatic epoxy (EPON and NGDE) and of a diamine curing agent (Jeffamine) [28]; and SMP2,

composed of an aromatic epoxy (DGEBA), a monoamine (DA) and an aromatic diamine cross-linker

(MXDA) [29]. Typical stress vs. strain curves from flexural and tensile tests for both epoxies are

shown in Figures 1c and d. Dynamic modulus analysis (DMA) was performed on the epoxy

compositions with and without graphene in order to evaluate the differences in glass transition

temperature (Tg). The Tg in shape-memory polymers represents their shape-shifting temperature,

above which, they will return to their initially “set” shape. The glass transition temperature of SMP1

is ~60°C, around 10°C higher than that of SMP2. The introduction of the graphene network does not

affect the Tg of the polymers (Figure 3a). In addition, while the Tg of GF-SMP1 remained as sharp

as that of the epoxy, the one of GF-SMP2 was broader so the latter composite was expected to be

more sensitive to a wider range of temperature changes (Figure S2). SMP1 is slightly stronger in

tension (43.1 ± 4 MPa vs 38.7 ± 3 MPa for SMP2) and stiffer (3.1 ± 0.5 vs 2.3 ± 0.4 GPa for SMP2).

The two different epoxy compositions have been used to determine an optimal approach to shape

memory composites and to prove the flexibility of this approach. Polymer bars bent during flexural

tests recovered their original flat shape after immersion in hot water suggesting that no irreversible

damage occurred within their structure. The slow displacement rate (0.6 mm/min) may have enabled

the polymer chains to slide over each other without breaking chemical bonds.

Page 5: Imperial College London - Electrically-responsive graphene ... · Web viewdiglycidylether of bisphenol A (DGEBA, Sigma, weight per epoxy group = 174.25 g/mol), ndodecylamine (DA,

The composites (GF-SMP1 and GF-SMP2) exhibit highly interconnected conductive pathways

homogeneously embedded in the polymeric matrix (Figure 2a). Flexural tests show how this network

affects the mechanical properties (Figure 3). The addition of graphene increases the rubber and glass

moduli of SMP1 by ~14 MPa and more than 400 MPa respectively. This is probably due to the high

degree of confinement of the polymer within the graphene walls. The flexural modulus remains

similar (2.3 ± 0.3 GPa versus 2.8 ± 0.1 GPa for the polymer). However, it decreases the mechanical

properties of SMP2. For example, the flexural modulus decreases by more than half (1.0 ± 0.1 GPa

versus 2.2 ± 0.2 GPa) as confirmed by DMA (Figure S2). This is probably due to the higher viscosity

of the polymeric precursor solution that hampers full infiltration.

The graphene network also affects the fracture behaviour. The flexural strengths of the composites

remain of the order of those of the starting polymers and can go up to 100 MPa for GF-SMP1.

However, the epoxies did not break in a brittle manner and exhibited elongations of 30% before

fracture while the composites show a more brittle behaviour due to the introduction of the network,

which facilitates crack propagation along predetermined pathways. Still, their KIC remains of the

order of those typically reported for brittle epoxies and is higher for GF-SMP2 (KIC = 0.99 ± 0.3

MPam-1/2) than for GF-SMP2 (KIC = 0.65 ± 0.1 MPam-1/2). Within the composite, the graphene

foam formed an intricate network of weak interfaces that directs crack propagation following the 3D-

graphene path (Figure 4a). It has been shown in other systems that a controlled architecture of weak

interfaces can be used to increase fracture resistance [30]. The composites benefit from extrinsic

toughening mechanisms such as crack deflection and bridging (Figure 5c and d) from the epoxy

layers and exhibit an R-curve behaviour with the toughness increasing with crack length up to values

that can be three times those of KIC. These toughening mechanisms seem to saturate at distances <

500 µm.

The conductive network can be used to follow damage progression. To assess mechanical properties

in conjunction with electrical properties we have developed a novel approach that monitors the

variations in electrical resistance in situ during double cantilever bending tests carried out in the

SEM (Figure S1). This method allowed us to track and record the changes in sample electrical

Page 6: Imperial College London - Electrically-responsive graphene ... · Web viewdiglycidylether of bisphenol A (DGEBA, Sigma, weight per epoxy group = 174.25 g/mol), ndodecylamine (DA,

resistance with crack growth (Figure 4b). Because the network is highly interconnected is it possible

to sense small cracks independently of the matrix. Even a 100-200 µm crack can be detected as a

0.25 Ω resistance change. These results prove the potential of using the network to sense damage and

showcase the possibilities of in-situ testing in the SEM, coupling electrical measurements and

mechanical testing at the microscale.

Thanks to the graphene network the shape memory behaviour can be easily and effectively triggered

via joule heating (Figure 3b). Samples with a tight spiral shape were machined and subsequently

fixed in an “open ring” shape. The application of a small, 9V, voltage for a short time (1 minute) was

sufficient to produce a temperature increase of almost 40 °C. This led to the recovery of the initial

tight spiral shape. Furthermore, within only a few seconds of cooling the material regains its original

strength and it is able to lift an object of approximately 200 times its own weight. A more thorough

study on the joule heating behaviour of the samples GF-SMP1 and GF-SMP2 is shown in Figure 5.

Both composites were subjected to increasing potentials while the temperature reached was recorded

via infrared camera. Both samples are highly responsive to small voltages responding with a change

in temperature of almost 20°C at 4V. The highest temperatures recorded at 10 V were above 100°C,

proving the wide range of this technique. I vs V curves show how the resistivity of the samples is not

affected by the temperature in the range needed to trigger shape changes, making this a very stable

process. Furthermore, the network enables uniform heating in situ.

Both composites exhibit excellent shape fixity and shape recovery (Figure 6a) even after 4 cycles of

Joule-heating and cooling (Figure 6b). Indeed, the recovery ratios are always superior to 98 %. When

subjected to Joule-heating, the samples recover most of their shape in 10s of seconds. The shape

recovery of GF-SMP2 is 0.98 ± 0.03, which is slightly higher than that of GF-SMP1, 0.96 ± 0.03

(Table 3). GF-SMP2 samples achieved complete recovery of their original shape at a lower voltage

(7 ± 0.2 V versus 9 ± 0.2 V for GF-SMP1) and in shorter times (1 min 47 s versus 3 min 53 s with ±

4 s). It is worth noting that, when the structures are tested for recovery in hot water instead of joule

heating, while GF-SMP2 samples recover their shape instantaneously, GF-SMP1 samples take

several seconds more. Moreover, just after immersion, the latter were more difficult to bend than the

Page 7: Imperial College London - Electrically-responsive graphene ... · Web viewdiglycidylether of bisphenol A (DGEBA, Sigma, weight per epoxy group = 174.25 g/mol), ndodecylamine (DA,

former ones. This behaviour confirmed the DMA results that GF-SMP2 is more sensitive to

temperature changes than GF-SMP1. The process is highly repeatable and flexural tests carried out

before and after 5 joule heating and cooling cycles show a maximum strength loss of only 20-25%.

3. Conclusion

We have here demonstrated that we can use graphene to insert a microscopic conductive network in a

shape memory polymer matrix. The result is a composite with very low carbon content (<1 wt%)

where the highly interconnected network can trigger a shape memory effect by joule heating at low

voltages while the thermal behaviour and mechanical strength of the polymer remains almost

unaltered. The network directs crack propagation and promotes fracture resistance while its

disruption during fracture can be used to sense damage even for cracks in the order of few hundreds

of microns. Therefore, this approach can be used to build strong self-shaping structures that will

avoid catastrophic failure. These results highlight the importance of controlling the architecture of

composites in order to take full advantage of graphene’s 2D nature and build materials able to

combine a structural and functional role. The approach is extremely versatile as it can be applied to a

range of different polymeric matrices capable of performing different functions. These novel

materials can open new opportunities in biomedical applications as well as in sensor- and

actuator/robotic systems.

4. Materials and Methods :

GFs were prepared via freeze-casting 9.4 mg/mL GO suspensions, prepared using modified

Hummer’s method, at a cold finger cooling rate of - 3.0 °C/min. 10% wt PVA (Sigma Aldrich) is used

as a binder of the structure and sucrose (AnalaR NORMAPUR) as a binder and structural modifier.

The weight ratio of graphene:additives and PVA:Sucrose was maintained at 1:1. Samples were

subsequently reduced in Ar/H atmosphere at 900°C at a rate of 5°C/min.

Two main compositions of SMEPs were synthesised. The first one (SMP1) is obtained by the

reaction of an aromatic epoxy (EPON 826) with an aliphatic epoxy (Jeffamine D230, Hunstman) and

a curing agent (NGDE, Aldrich). The second one (SMP2) is obtained by the reaction of

Page 8: Imperial College London - Electrically-responsive graphene ... · Web viewdiglycidylether of bisphenol A (DGEBA, Sigma, weight per epoxy group = 174.25 g/mol), ndodecylamine (DA,

diglycidylether of bisphenol A (DGEBA, Sigma, weight per epoxy group = 174.25 g/mol),

ndodecylamine (DA, 98 wt.%, Aldrich) and m-xylylenediamine (MXDA, Aldrich). To synthesize

SMP1, 21.84 g of EPON were first heated at 70 °C and the two other chemicals (13.80 g of

Jeffamine and 12.96 g of NGDE) were added simultaneously under vigorous stirring. Curing was

carried out during 1 h 30 min at 100 °C and samples were subsequently annealed for 1 h at 130 °C.

To synthesise 32.41 mL of SMP2, 8.42 g of DA was first heated at 130 °C to remove its CO2 content

and cooled down to 100 °C. 23.73 g of DGEBA was then added and the mixture was mixed

energetically at 100 °C. 1.55 g of MXDA was quickly added to avoid the composition from

becoming viscous and a two ramp-and-hold curing process was carried out. Samples were first

heated at 60 °C during 1 h and complete curing was achieved after 2 h at 100 °C. Finally, according

to Leonardi and al. [29] and to increase the amount of physical crosslinks (dodecyl groups) annealing

was carried out at 100 °C during 72 h. For SMP2, the molar ratio was the same as for Leonardi and

al.[29] and DGEBA:DA:MXDA = 6:4:1. Both the weight and dimensions of the starting foams and

the final composites are measured and used to calculate densities and carbon content. Densities of the

composites are also measured using the Archimedes method.

Following ASTM D638-14, tensile tests were performed on SMP1, and SMP2 samples at room

temperature. Samples were cut to dogbone shape of type V with fixed dimensions and were

manufactured from SMP plates by using a water jet. Zwick/Roell® test-machine with a maximum-

load of 10 kN and a crosshead-speed of 1 mm/min was used.

Following ASTM D790-03, five specimens of GF-SMP1 and of GF-SMP2 were tested in 3-point

bending tests in the in-plane direction (direction parallel to the growth of the ice lamellae).

Zwick/Roell® test-machine with a maximum-load of 2 kN was used and the displacement rate was

fixed at either 0.6 mm/min. For each sample, the span-to-thickness ratio equalled 16 and the strain

rate 0.01 mm/mm/min.

Following ASTM D5045-14, the toughness of GF-SMPs and that of pristine polymers was evaluated

in the in-plane direction in bending. Samples were notched with a SiC blade. Finally, notches were

sharpened by razor blade hammering into the notch and the initial crack length was measured on

Page 9: Imperial College London - Electrically-responsive graphene ... · Web viewdiglycidylether of bisphenol A (DGEBA, Sigma, weight per epoxy group = 174.25 g/mol), ndodecylamine (DA,

Zeiss AxioCam MR microscope. A Zwick/Roell® test-machine with a maximum-load of 2 kN was

used at a displacement rate of 10 mm/min and the span was fixed at 27 mm.

DMA tests were performed on a Q800 dynamic mechanical analyser (DMA) from TA Instruments. A

single cantilever fixture was used and the samples SMP1, SMP2, GF-SMP1 and GF-SMP2 were

tested in a multi-frequency stress mode at 0.1 MPa and 1Hz. Each run followed the same procedure:

samples were heated and held at 30 °C for 0.5 min. Then, temperature was increased to 100 °C at a

rate of 2 °C/min.

The shape fixity and shape recovery ratios of GF-SMPs samples were calculated as shown in Figure

S4. Samples of GF-SMP1 and of GF-SMP2 were heated at 100°C and bent onto a 90 ° mould to a

maximum angle θ_max. Specimens were cooled down to room temperature, load was released and

the fixed shapes were scanned for further measuring of θ_fixed. For Joule heating effect, the samples

were connected to a PSU (Iso-thech® IPS-330 DD) and the voltage was varied until the shape-

memory effect is observed. The recovered shape angle (θ_rec) was recorded. The cyclic recovering

ability was evaluated over five cycles (Figure S4).

In-situ double cantilever beam testing inside an SEM was used to generate r curves by measuring

stress intensity as the crack grows. To achieve this a 120° wedge was used to open the arms of a

double cantilever beam sample of the material. This test allows long crack paths to be monitored

using the SEM during crack propagation. This was coupled with a multimeter to detect changes in

resistance of the sample with crack length (Figure S1).

Data availability

The raw data required to reproduce these findings are available on request. The processed data required to reproduce these findings are available on request.

Acknowledgements

The authors would like to thank DARPA and ONRG for funding of the project. EF would like to thank the CASC industrial consortium for financial support.

Page 10: Imperial College London - Electrically-responsive graphene ... · Web viewdiglycidylether of bisphenol A (DGEBA, Sigma, weight per epoxy group = 174.25 g/mol), ndodecylamine (DA,

References 1. Lendlein, A., H.Y. Jiang, O. Junger, and R. Langer, Light-induced shape-memory polymers.

Nature, 2005. 434(7035): p. 879-882.2. Meng, H. and G.Q. Li, A review of stimuli-responsive shape memory polymer composites.

Polymer, 2013. 54(9): p. 2199-2221.3. Hu, J.L., Y. Zhu, H.H. Huang, and J. Lu, Recent advances in shape-memory polymers:

Structure, mechanism, functionality, modeling and applications. Progress in Polymer Science, 2012. 37(12): p. 1720-1763.

4. Huang, W.M., Z. Ding, C.C. Wang, J. Wei, Y. Zhao, and H. Purnawali, Shape memory materials. Materials Today, 2010. 13(7-8): p. 54-61.

5. Li, G., Recent advances in smart self-healing polymers and composites. 2015, Boston, MA: Elsevier.

6. Choi, S.-B. and J. Kim, eds. Smart Materials Actuators Recent Advances in Characterization and Applications. Materials Science and Technology. 2015, Nova Science Publishers Incorporated: New York, USA.

7. Jani, J.M., M. Leary, A. Subic, and M.A. Gibson, A review of shape memory alloy research, applications and opportunities. Materials & Design, 2014. 56: p. 1078-1113.

8. Liu, C., H. Qin, and P.T. Mather, Review of progress in shape-memory polymers. Journal of Materials Chemistry, 2007. 17(16): p. 1543-1558.

9. Serrano, M.C., L. Carbajal, and G.A. Ameer, Novel Biodegradable Shape-Memory Elastomers with Drug-Releasing Capabilities. Advanced Materials, 2011. 23(19): p. 2211-2215.

10. Zhou, G.X., H. Zhang, S.P. Xu, X.C. Gui, H.Q. Wei, J.S. Leng, N. Koratkar, and J. Zhong, Fast Triggering of Shape Memory Polymers using an Embedded Carbon Nanotube Sponge Network. Scientific Reports, 2016. 6: 24148.

11. Liu, T.Z., T.Y. Zhou, Y.T. Yao, F.H. Zhang, L.W. Liu, Y.J. Liu, and J.S. Leng, Stimulus methods of multi-functional shape memory polymer nanocomposites: A review. Composites Part a-Applied Science and Manufacturing, 2017. 100: p. 20-30.

12. Mu, T., L.W. Liu, X. Lan, Y.J. Liu, and J.S. Leng, Shape memory polymers for composites. Composites Science and Technology, 2018. 160: p. 169-198.

13. Wang, X., J. Sparkman, and J.H. Gou, Electrical actuation and shape memory behavior of polyurethane composites incorporated with printed carbon nanotube layers. Composites Science and Technology, 2017. 141: p. 8-15.

14. Zhou, J., H. Li, R. Tian, R. Dugnani, H.Y. Lu, Y.J. Chen, Y.P. Guo, H.N. Duan, and H.Z. Liu, Fabricating fast triggered electroactive shape memory graphite/silvernanowires/epoxy resin composite from polymer template. Scientific Reports, 2017. 7: 5535.

15. Schmidt, A.M., Electromagnetic activation of shape memory polymer networks containing magnetic nanoparticles. Macromolecular Rapid Communications, 2006. 27(14): p. 1168-1172.

16. Erb, R.M., J.S. Sander, R. Grisch, and A.R. Studart, Self-shaping composites with programmable bioinspired microstructures. Nature Communications, 2013. 4: 1712.

17. Huang, W.M., B. Yang, Y. Zhao, and Z. Ding, Thermo-moisture responsive polyurethane shape-memory polymer and composites: a review. Journal of Materials Chemistry, 2010. 20(17): p. 3367-3381.

18. Leng, J. and S. Du, Shape-memory polymers and multifunctional composites. 2010, Boca Raton: CRC Press/Taylor & Francis.

19. Lendlein, A. and R. Langer, Biodegradable, elastic shape-memory polymers for potential biomedical applications. Science, 2002. 296(5573): p. 1673-1676.

20. Song, J.J., H.H. Chang, and H.E. Naguib, Biocompatible shape memory polymer actuators with high force capabilities. European Polymer Journal, 2015. 67: p. 186-198.

Page 11: Imperial College London - Electrically-responsive graphene ... · Web viewdiglycidylether of bisphenol A (DGEBA, Sigma, weight per epoxy group = 174.25 g/mol), ndodecylamine (DA,

21. Gong, X.B., L.W. Liu, Y.J. Liu, and J.S. Leng, An electrical-heating and self-sensing shape memory polymer composite incorporated with carbon fiber felt. Smart Materials and Structures, 2016. 25(3): 035036.

22. Park, J.H., T.D. Dao, H.I. Lee, H.M. Jeong, and B.K. Kim, Properties of Graphene/Shape Memory Thermoplastic Polyurethane Composites Actuating by Various Methods. Materials, 2014. 7(3): p. 1520-1538.

23. Wang, Y.K., W.C. Tian, J.Q. Xie, and Y. Liu, Thermoelectric Responsive Shape Memory Graphene/Hydro-Epoxy Composites for Actuators. Micromachines, 2016. 7(8): 145.

24. Li, C.W., L. Qiu, B.Q. Zhang, D. Li, and C.Y. Liu, Robust Vacuum-/Air-Dried Graphene Aerogels and Fast Recoverable Shape-Memory Hybrid Foams. Advanced Materials, 2016. 28(7): p. 1510-1516.

25. Ni, N., S. Barg, E. Garcia-Tunon, F.M. Perez, M. Miranda, C. Lu, C. Mattevi, and E. Saiz, Understanding Mechanical Response of Elastomeric Graphene Networks. Scientific Reports, 2015. 5: 13712.

26. D'Elia, E., S. Barg, N. Ni, V.G. Rocha, and E. Saiz, Self-Healing Graphene-Based Composites with Sensing Capabilities. Advanced Materials, 2015. 27(32): p. 4788-4794.

27. Picot, O.T., V.G. Rocha, C. Ferraro, N. Ni, E. D'Elia, S. Meille, J. Chevalier, T. Saunders, T. Peijs, M.J. Reece, and E. Saiz, Using graphene networks to build bioinspired self-monitoring ceramics. Nature Communications, 2017. 8: 14425.

28. Xie, T. and I.A. Rousseau, Facile tailoring of thermal transition temperatures of epoxy shape memory polymers. Polymer, 2009. 50(8): p. 1852-1856.

29. Leonardi, A.B., L.A. Fasce, I.A. Zucchi, C.E. Hoppe, E.R. Soule, C.J. Perez, and R.J.J. Williams, Shape memory epoxies based on networks with chemical and physical crosslinks. European Polymer Journal, 2011. 47(3): p. 362-369.

30. Mirkhalaf, M., A.K. Dastjerdi, and F. Barthelat, Overcoming the brittleness of glass through bio-inspiration and micro-architecture. Nature Communications, 2014. 5: 3166.

Page 12: Imperial College London - Electrically-responsive graphene ... · Web viewdiglycidylether of bisphenol A (DGEBA, Sigma, weight per epoxy group = 174.25 g/mol), ndodecylamine (DA,

Figures

Figure 1. Mechanical properties of the starting materials. a) Scanning electron micrograph of the freeze cased rGO foams (with average electrical conductivity values reported in the inset). The picture has been taken in the plane perpendicular to the direction of ice growth. b) Mechanical properties of the foams in compression up to a strain of 80%. The graph also shows the recovery strength of the foam when subjected to 6 compression cycles at 20% strain (blue dots with error bars), demonstrating a high retention of mechanical strength c) and d) typical tensile and flexural test behaviour of shape-memory polymers SMP1 and SMP2.

Page 13: Imperial College London - Electrically-responsive graphene ... · Web viewdiglycidylether of bisphenol A (DGEBA, Sigma, weight per epoxy group = 174.25 g/mol), ndodecylamine (DA,

Figure 2. Composite structure and mechanical properties. a) Optical microscopy of the GF-SMP composite structure. The graphene network shows as the black strands and the polymer as the violet filler. Inset shows a sample as prepared and cut to size. b) Typical stress-displacement curves of composite GF-SMP samples made with polymers SMP1 and SMP2 recorded in 3 point bending.

Figure 3. Mechanical properties and shape memory effect. a) Table summarizing the glass transition temperature (Tg) and the mechanical properties, ultimate tensile strength and modulus (T

U, ET), flexural strength and modulus (F

U, EF), and fracture initiation toughness (KIC), of the materials. b) Infographic showing the behaviour of an GF-SMP1 composite subject to a potential of 9V. The composite coils back to the pre-set shape after 60 seconds of joule heating. As shown in the thermograms, the effect increases the temperature of the sample from RT to around 58 ◦C, which is enough to activate the shape-memory effect.

Page 14: Imperial College London - Electrically-responsive graphene ... · Web viewdiglycidylether of bisphenol A (DGEBA, Sigma, weight per epoxy group = 174.25 g/mol), ndodecylamine (DA,

Figure 4. Fracture toughness and damage detection. a) Scanning electron micrograph of the fracture surface of a GF-SMP1 composite showing the texture created by the presence of the graphene network. b) Results from the DCB in situ SEM testing coupled with 2 point electrical measurements, The electrical resistance and fracture toughness of the GF-SMP1 composites rise with crack length. c) Optical image of the crack progression through the composite. d) Magnification of the crack path showing evidence of bridging.

Page 15: Imperial College London - Electrically-responsive graphene ... · Web viewdiglycidylether of bisphenol A (DGEBA, Sigma, weight per epoxy group = 174.25 g/mol), ndodecylamine (DA,

Figure 5. Thermal response as a function of applied voltage. Samples of GF-SMP1 and GF-SMP2 were Joule-heated and the temperature increase was recorded with increasing voltage input. The insets A and C represent the initial room temperature spectrum of the samples (respectively containing SMP1 and SMP2) and insets B and D show the same sample spectra when subjected to 10V. A typical I vs V graph (top right) shows the linear dependence of the intensity with voltage.

Page 16: Imperial College London - Electrically-responsive graphene ... · Web viewdiglycidylether of bisphenol A (DGEBA, Sigma, weight per epoxy group = 174.25 g/mol), ndodecylamine (DA,

Figure 6. Shape-memory behaviour. a) Photograph showing the initial state (dark) and the recovered state (faint) after joule heating with a voltage of 7-9 V. b) Table summarizing the recovery data for composite GFSMP samples. c) Graph showing the stable behaviour and reproducibility of the shape-memory effect after 4 cycles. d) Variation of the case flexural strength after 5 cycles of heating and cooling using hot Joule heating. Samples named –T indicate the samples tested in flexion after 5 cycles.