Abstract
Electroactive and magnetoactive shape memory polymer nanocomposites (SMCs) are multistimuli-responsive smart materials that are of great interest in many research and industrial fields. In addition to thermoresponsive shape memory polymers, SMCs include nanofillers with suitable electric and/or magnetic properties that allow for alternative and remote methods of shape memory activation. This review discusses the state of the art on these electro- and magnetoactive SMCs and summarizes recently published investigations, together with relevant applications in several fields. Special attention is paid to the shape memory characteristics (shape fixity and shape recovery or recovery force) of these materials, as well as to the magnitude of the electric and magnetic fields required to trigger the shape memory characteristics.
1. Introduction
Shape memory polymers (SMPs) and their composites (SMCs) are smart materials that are able to respond to certain external stimuli by changing their shape. In most cases, these shape memory materials (SMMs) are thermoresponsive, meaning that their shape memory properties are triggered depending on their temperature. Other response methods exist such as light, pH, and solvents,1 but these will not be covered in this review. The simplest shape memory cycle of a thermoresponsive SMM is composed of a shape programming step and a shape recovery step. The shape programming step starts with the SMM in its original shape, commonly termed the permanent shape, that is heated above its transition temperature (Ttrans). The Ttrans values of a SMM are those related to the phase transitions that control the shape memory process: i.e., the glass transition or the melting–crystallization transition. Once at a temperature passes Ttrans, a stress is applied, leading to a deformation from the permanent shape to the desired temporary shape. The shape programming step finishes by cooling the material under stress below Ttrans while maintaining said deformation in order to fix the temporary shape, after which the applied stress can be released. Reheating the material above Ttrans leads to the shape recovery from the temporary shape to the permanent shape due to stress relaxation.2
Conventionally, the shape recovery is triggered with direct heating by placing the SMM close to a heat source (e.g., inside an oven). This technique, which relies on an external heater, is occasionally impossible, problematic, or even dangerous. In recent years, much effort has been put into remotely inducing the shape recovery by indirectly heating the SMM using an electric or a magnetic field.3 However, pristine SMPs are not suitable for these remote actuation triggers because of their high electric resistivity and insensitivity to magnetic fields, akin to the properties of conventional polymers. In order to benefit from these alternative heating mechanisms, either high electrical conductivity or magnetic fillers are embedded within the SMPs, hence conforming a SMC. For conciseness, SMM is used in the remainder of the article to refer to SMPs and SMCs.
In this review, we address the topic of shape memory polymer nanocomposites activated by an electric or magnetic field. We start with a brief introduction and overview on the topic of shape memory and the different types of shape memory behavior: one-way, two-way, and multiple shape memory. We continue by addressing the phenomena responsible for electro- and magnetoactivation of SMMs. After this theoretical introduction, we discuss the typical nanofillers used to disperse within the polymer matrix in order to confer electroactive and magnetoactive properties to SMCs. Some of the numerous works published in the past few years are summarized, paying special attention to the shape memory characteristics and the magnitude of the electric or magnetic fields required for activation. As an additional tool for the reader, an extensive table of additional references on these SMMs is available in the Appendix. Some recent progress and applications of these SMCs are discussed next. Finally, future horizons and challenges in the investigation of SMMs are pointed out.
2. Basic Concepts on SMPs and Their Composites
The shape memory effect is a result of an adequate polymer molecular architecture including two parts: (i) the netpoints that build up the permanent shape of the SMP and (ii) the molecular-switchable segments that are responsible for fixing the temporary shape of the SMP.4 The netpoints are cross-links that can be either chemical (covalent bonds) or physical. The latter is related to interactions among the molecules of the polymer network that consists of crystallites, hydrogen bonds, or ionic interactions. Physically cross-linked netpoints are weaker than their covalent chemical counterpart; thus, thermoplastics are more prone to incomplete shape recovery.5 Regarding the switchable segments, they are reversible cross-links that can also be of physical or chemical nature.
Depending on the nature of both the netpoints and the switchable segments, a comprehensive SMP classification consisting of four classes was introduced by the research group of Mather6 and since reported by many others.1,4,7−10 The classes are (I) glassy covalently cross-linked thermosets, (II) semicrystalline covalently cross-linked polymer networks, (III) glassy physically cross-linked copolymers and blends, and (IV) semicrystalline physically cross-linked block copolymers.
The Ttrans of a SMM is the characteristic temperature related to the formation and destruction of the switchable segments. In general, the Ttrans of semicrystalline SMPs is related to the melting–crystallization phase transition, i.e. the melting temperature (Tm) during heating and the crystallization temperature (Tc) during cooling. On the other hand, the Ttrans of glassy SMPs is related to the glass transition temperature (Tg). These glassy SMPs usually have a broader Ttrans range than semicrystalline-network SMPs and therefore a slower recovery. Although this may be regarded as an undesirable feature, in combination with the biocompatibility of some of these polymer networks, it makes them excellent candidates for biomedical applications.11−13 Nevertheless, they are not ideal for applications where a fast shape recovery is needed. On the other hand, traditionally, the majority of covalent bonds are strong chemical cross-links that are stable and hard to break. Thus, most polymers that are covalently cross-linked cannot be recycled or reprocessed. In order to overcome the disadvantages associated with covalent cross-links while still having a strong bond among molecules, dynamic covalent bonds can be incorporated in SMPs. Dynamic covalent bonds are reversible and thus are able to be broken and formed when they are subjected to certain stimuli such as heat, light, or pH. In recent years more and more polymers have been produced incorporating dynamic covalent bonds in order to achieve materials that can self-heal, reshape for 3D printing purposes, and have shape memory.14 There is a wide choice of methods to obtain dynamic covalent bonds, and the corresponding literature14−18 is available.
2.1. The Basic Shape Memory Effect
The
shape memory effect of a SMM arises after the shape programming process
that is achieved in three steps. To aid in the explanation, the stress
evolution with applied strain of a SMP is illustrated in Figure 1. The shape programming
process starts by ① heating the material, initially in its
permanent shape, above a characteristic Ttrans (either Tg or Tm depending on the polymer). Second, ② a mechanical
deformation is applied on the material while it is above Ttrans. This mechanical deformation leads to a strain denoted
by in Figure 1. Finally, the programming process finishes by ③
cooling the material below Ttrans while
maintaining the desired load or the desired deformation. ④
Unloading the sample toward zero stress may result in a fixed strain εf slightly lower than
. After this programming process is finished,
the shape memory material is fixed in its new deformed temporary shape.
During the applied deformation, the SMP, with either an amorphous
or semicrystalline structure, is in the rubbery state and has a rubber-like
behavior. In other words, the SMP above its Ttrans exhibits entropic elasticity. Entropic elasticity exists
while deforming an elastomer (or a SMP in the amorphous state) because
the high mobility of the polymer chains permits the originally disordered
structure to align in the direction of deformation.19 Ordering the polymeric chains decreases the entropy of
the material, and this energetic state gets fixed by vitrification
or crystallization upon cooling the material below Ttrans.
Figure 1.
Illustration of the stress vs strain curve of a SMM during
two
conventional consecutive shape memory cycles performed under uniaxial
tension. The marked strain levels ε0, εp, εf, and are used to calculate the shape fixity
and shape recovery ratios.
During the shape memory programming step, the shape memory material has stored energy due to the freezing of the lower entropic polymer network. The new temporary deformed shape can be viewed as a metastable shape: when the material is exposed to a certain external stimulus, the stored energy will be released. ⑤ Reheating the SMM above Ttrans will cause a shape change back toward the permanent shape. This process, in which the material recovers the preferred high entropic state, is called shape recovery. Different temporary shapes can be attained by repeating the programming process with a different deformation step. Nevertheless, the primary permanent shape is always approximately recovered during the shape memory cycle. This is because Ttrans only affects the switchable links that are related to the temporary shape, whereas the permanent shape is dictated by the netpoints of the SMM. Any damage on the netpoints during the shape memory cycle, for instance breaking of the netpoints due to excessive deformation in the shape programming process, may cause the recovered permanent shape to slightly differ from the original permanent shape.
Two quantities are commonly used to evaluate the quality of a SMM: the shape fixity (RF) and shape recovery ratios (RR). They are calculated using eqs 1 and 2, respectively.
![]() |
1 |
![]() |
2 |
The shape fixity ratio quantifies the
ability of the SMM to fix
the temporary shape and it is calculated as the ratio of strain at
the end of ③ cooling (εf indicated in Figure 1) over the strain
at the end of ② elongation ( in Figure 1). On the other hand, the shape recovery ratio gives
an idea of how well the SMM “remembers” its permanent
shape by relating the strain during shape recovery to the strain due
to elongation.
2.2. Types of Shape Memory Behavior
The shape memories presented in this section include the simplest one-way shape memory (1W-SM) effect. This section will also define the more complex shape memory behaviors commonly found in the literature, namely the two-way shape memory (2W-SM) effect and the multiple shape memory effect. An illustration of these behaviors is shown in Figure 2, and an explanation is provided next.
Figure 2.
Illustration of the most common shape memory behaviors of SMMs: (a) one-way shape memory effect; (b) two-way shape memory effect; (c) multiple shape memory effect.
2.2.1. One-Way Shape Memory (1W-SM) Effect
The simple 1W-SM effect has already been explained in the previous sections. As illustrated in Figure 2a, it consists of a shape programming process composed by ① heating, ② deformation above Ttrans, and ③ fixing of the temporary shape by cooling below Ttrans. The programming process is followed by ④ reheating above Ttrans with no load to trigger shape recovery toward the permanent shape. Once the shape recovery has been completed, the SMP is no longer thermoresponsive. To benefit from the shape memory effect again, another shape programming process needs to be performed on the material. Most shape memory polymers and their composites show this type of memory effect. In fact, it was believed until the late 2000s that only shape memory alloys but neither polymers nor their composites could show a two-way shape memory effect. Many examples of SMPs exhibiting the 1W-SM effect can be found in the literature.20−25 Furthermore, a more exhaustive review on 1W-SM polymers and their composites has been given by Leng et al.26
2.2.2. Two-Way Shape Memory (2W-SM) Effect
The two-way shape memory (2W-SM) effect is present in some semicrystalline SMPs and their composites.27,28 A typical 2W-SM effect under tension is illustrated in Figure 2b. The programming process is the same as for the 1W-SM effect: ① heating above Ttrans, ② deformation above Ttrans, and ③ cooling below Ttrans. The 2W-SM effect in semicrystalline polymers arises, in general, by the cyclic ④ heating and ③ cooling above and below the respective Ttrans at a constant load. In this case, the SMP is not left free to deform to the permanent shape but, instead, the shape keeps varying between a primary and a secondary temporary shape. The primary temporary shape is that obtained at the end of ② loading and recovered every time that the SMP is ④ reheated above its Ttrans at constant load. The secondary temporary shape is that obtained by ③ cooling below Ttrans at a constant load. The change in shape between the primary and the secondary temporary shapes can be repeated several times without the need of any other additional programming process.13 The 2W-SM effect in semicrystalline polymers under constant tensile stress is characterized by a contraction of the material in the loading direction during heating above Ttrans. Conversely, elongation along the loading direction is observed upon cooling below Ttrans. Thus, the second temporary shape is longer than the first. Generally, if the reheating above Ttrans happens in a stress-free configuration, the permanent shape is recovered, following the conventional 1W-SM effect.
In 2008, Chung, Romo-Uribe, and Mather29 first reported the 2W-SM effect of a semicrystalline network of poly(cyclooctene) under constant stress. They showed that a lower cross-linking density would lead to a higher cooling-induced crystallization and, hence, a wider 2W-SM effect. Furthermore, Tm and Tc are found to be shifted to higher values on lowering the crystallization degree.
There is still disagreement on the mechanism behind the 2W-SM effect in semicrystalline polymer networks.30 Most authors seem to agree on a melting-induced contraction and a crystallization-induced elongation as the (main) mechanism: oriented crystallites are formed in the direction of loading during cooling.30,31 Chung, Romo-Uribe, and Mather postulated that the mechanism was a combination between crystallization-induced elongation and rubber elasticity.29
Generally, the 2W-SM effect is shown when the polymer is placed under nonzero stress.28,30,32−42 Nevertheless, a few freestanding 2W-SM polymers have also been reported. These reports ascribe the freestanding 2W-SM effect to different sources, including two melting transitions, one broad melting transition, and chemically heterogeneous structures by creating thermostable crystallites after deformation-induced crystallization.43−48 Furthermore, other types of materials such as liquid-crystalline elastomers49,50 and shape memory composite laminates51,52 have been described to exhibit 2W-SM effect.
2.2.3. Multiple Shape Effect
Triple shape memory materials are those that can utilize twice the 1W-SM effect in order to switch from a one temporary shape to another and from the latter to the permanent shape. A conventional 1W-SM effect is based on one Ttrans for shape recovery. In the case of triple SMMs, the two shape changes are related to either one broad temperature interval for the transition or to two separated Ttrans values. This is because a broad thermal transition can be regarded as an infinite amount of abrupt thermal transitions.53,54
For a triple SMM, two independent shape programming processes need to be followed. Figure 2c illustrates the typical procedure followed for a triple SMM based on the two transition temperatures Ttrans,1 and Ttrans,2 such that T0 > Ttrans,1 > Ttrans,2, where T0 is the starting temperature. The deformation to obtain the two different temporary shapes in the illustration is applied by uniaxial tension. The first programming process consists of ① heating the SMM in its permanent shape at T0 to a temperature above Ttrans,1, where the polymer chains gain mobility, and ② the deformation to the first temporary shape can be applied. The deformation produces an orientation of the polymer chains along the loading direction, which in turn reduces the entropy of the material. This first temporary shape can be fixed by ③ cooling to a temperature T such that Ttrans,1 > T > Ttrans,2, where the elastic energy is stored inside of the material thanks to the phase characterized by Ttrans,1. The second programming process can start by ④ deforming the material once again while the temperature is between both Ttrans values, which further decreases the entropy of the material. The second temporary shape is then fixed by ⑤ cooling below Ttrans,2. Once again, the newly generated phase fixes the shape of the material and stores the extra elastic energy generated during step ④. If the freestanding material is ⑥ reheated above Ttrans,2, its shape will change from the second to the first temporary shape due to the release of the energy stored during step ⑤. ⑦ A further increase in the temperature above Ttrans,1 will reactivate the mobility of all polymer chains in the network, and the remaining stored energy will be released. This causes the shape recovery from the first temporary shape to the permanent shape.
In the last few years, several polymers have been reported to exhibit triple shape memory. Some examples of these materials based on a broad temperature interval can be found in the literature.55−57 On the other hand, references on triple shape memory materials based on two individual Ttrans values are also available.58−60 Furthermore, the triple shape effect has also been observed in polymer laminates where each layer had a different Ttrans.61 A triple shape memory composite was created by Tobushi et al.62 that functions as a bending actuator. It is a laminate of a mix of SMP and shape memory alloy (SMA) that bends in two different directions depending on which Ttrans, that of the SMP or that of the SMA, was surpassed. As was already mentioned, conventional triple (or multiple) SMPs and SMCs are based on the 1W-SM effect,63 thus making the shape memory effect an irreversible feature requiring the application of the programming processes for further shape change. Nevertheless, a novel triple SMP exhibiting 2W-SM is reported.64
The same concept can be extended to multiple (quadruple, quintuple, ...) SMMs by chaining more than two programming processes, hence leading to more shape recoveries. In general, this is done in polymer networks that present very broad transition temperature intervals.56,65,66
A few references are available reporting the cyclic quality of SMMs during a high number (≥1000) of 1W-SM67 and 2W-SM68 cycles, still showing remarkable shape memory characteristics. However, we have not found reports regarding a high number of cycle repetitions involving multiple shapes. Some reports address the RF and RR of multiple shape memory materials for the first ≤6 cycles.55,58,59,64,69 Generally, it is observed that the RF decreases and the RR increases with each performed (sub)cycle. In comparing a triple SMM to a quadruple SMM, Dolog and Weiss reported a decrease of RR and an increase of RF when adding an extra shape to their cycle.55 More research is needed on the repetition of cycles and subcycles of multiple shape memory materials in order to assess the lifetime and quality of the behavior of these materials upon extended use.
2.3. Electro- and Magnetoactivation
Up to now, the notions and definitions given in this section are common to thermally triggered SMMs, regardless of the type of activation. In electroactivated thermally triggered SMMs, the temperature is a result of Joule resistive heating, which requires the flow of an electric current I through a material with a certain electrical resistance R. The power dissipated P is the product P = RI2 = V2/R, where V is the voltage drop across the material. From the formula of the dissipated power, it is straightforward that, at finite voltages and currents, P = 0 either for a perfectly conducting (R = 0) or perfectly insulating (R = ∞) material. For resistive heating to take place, the material should ideally have a “medium” electrical resistance R, taking into account that reasonable values of currents (e.g., <100 mA) and voltages (e.g., <10 V) need to be used. In practice, reaching a dissipated power leading to a measurable temperature increase, e.g. P ≈ 0.1 W, requires resistance values usually in the rage of ∼100–10000 Ω, which is significantly lower than the resistance of the polymer. Such resistance values are achieved by incorporating electrically conductive fillers in the insulating SMP matrix.
Another alternative way of triggering the heating of the material and, thus, the activation and shape recovery of SMCs is by the application of alternating magnetic fields. In order to obtain magnetosensitive SMCs, ferromagnetic fillers need to be incorporated inside the SMP. There are three magnetic heating mechanisms that can arise in magnetosensitive SMCs: (i) eddy currents, (ii) hysteresis, and (iii) losses related to the rotation of the magnetic spin due to Néel–Brown relaxation.70 The heating efficiency is highly dependent on the size of the magnetic fillers. The composites with embedded fillers in the microscale can be heated through eddy current and hysteresis loss by applying magnetic fields at moderate frequencies. In contrast, when magnetic fillers in the nanoscale are introduced in the SMP, these two heating mechanisms become less effective. Hysteresis loss decreases and becomes negligible for particle diameters below a certain threshold. For instance, iron oxide (Fe3O4) turns superparamagnetic, i.e. the magnetization curve shows no hysteresis, for diameters smaller than 20 nm. Furthermore, a multidomain structure is theoretically estimated to happen for diameters higher than 62.9–128 nm.71,72 These characteristic sizes depend on the nature and shape of the magnetic nanoparticles.
The main magnetic heating mechanism for magnetic nanoparticles is related to Néel–Brown relaxations. High frequencies are needed for this type of magnetic heating to generate a considerable temperature increase. The Néel relaxation consists of the rotation of the nanoparticle magnetic moment to align with the direction of the applied magnetic field while the nanoparticle itself stays motionless. The Brown relaxation leads to the physical rotation of the magnetic nanoparticles. Both relaxation mechanisms exist simultaneously when the particles are subjected to AC magnetic fields. Nevertheless, the Néel relaxation tends to dominate for smaller nanoparticles (<10 nm in diameter) in viscous media and Brownian motion tends to dominate for larger particles in media with low viscosities.70,73
3. Electro- and Magnetoactive SMCs: Fillers and Activation
SMPs are preferred over SMAs for various applications due to their low cost and their ability to sustain large deformations. Nevertheless, they also present some drawbacks or limitations in comparison to SMAs: SMPs have lower strength, stiffness, and recovery force and are electrically and thermally insulating. In order to overcome these limitations, different fillers have been incorporated within SMPs: hence, the terminology shape memory polymer composite. The addition of some of these fillers has been proven to improve some properties and give extra characteristics to the pristine polymers themselves. This includes the possibility of different types of stimuli for triggering the necessary temperature difference in order to make use of the shape memory properties. The properties of the resulting SMCs can be affected by several factors such as filler distribution, filler–polymer interface, filler size and aspect ratio, the nature of the SMP, and even the processing techniques.74
Regarding the size of the fillers, continuous fibers such as carbon, glass, or aramid fibers have been used to reinforce SMPs.75−77 An alternative to continuous fibers is to use short (chopped) fibers78,79 of a few millimeters in length. SMA wires or fibers have also been reported in the literature.80 The fibers are generally incorporated within the SMP with the aim of improving the mechanical properties of the material. Even though not common, a few references have been found where carbon fibers, thanks to their high electrical conductivity, were also used to heat the SMC by means of electric Joule heating.81−83 This review focuses on shape memory composites with dispersed nanofillers that are used in order to give the resulting nanocomposite the ability to be heated with an electric current or a magnetic field. A schematic diagram showing an overview of the main existing nanofillers is shown in Figure 3.
Figure 3.
Main existing conductive and magnetic nanofillers reported in the literature to conform electro- and magnetoactive shape memory polymer composites.
3.1. Carbon Nanofillers
In the past couple of decades, shape memory polymer nanocomposites with embedded carbon nanofillers have been intensely investigated to electrically induce the shape memory effect. Thanks to their excellent electrical conductivity, these nanofillers facilitate resistive heating and permit the electrical activation of the shape memory characteristics of SMCs. Moreover, the thermal conductivity of the resulting composite is improved. This feature is particularly interesting for applications where direct heat cannot be applied in order to trigger the shape memory effects. Such applications include, for instance, biomedical devices or self-deployable structures.84 Within the group of carbon nanofillers, investigations focus on SMCs with embedded carbon black nanoparticles (CBs), carbon nanofibers (CNFs), single-walled and multiwalled carbon nanotubes (SWCNTs and MWCNTs, respectively), and graphene.
3.1.1. Carbon Black Nanoparticles
CBs are effective fillers that can be used for the reinforcement of shape memory polymers. Due to their spherical geometry, a high concentration of CBs is needed in order to surpass the percolation threshold. At low concentrations, the addition of CBs does not create an electrically conductive network that has the ability to heat due to an electric current. Nevertheless, concentrations of 0.5–1.0 vol % CBs have been shown to improve the actuation ratio of the 2W-SM effect and the elastic modulus at high temperature of a polyethylene-based SMC.39 The same work showed that the 2W-SM effect of the composite was lost when the concentration of CB was increased to 20 vol %. At the required high concentrations for resistive heating at moderate voltages (>15 wt %), this family of SMCs was reported to have lower recovery ratios,85 more brittle behavior leading to failure of the material at shorter deformations,86,87 and, in some cases, severely worsened shape fixity.88 The last effect is found in semicrystalline SMP composites, since the addition of CBs decreases the amount of crystalline regions within the composite,39 which are in charge of fixing the temporary shape.
In order to achieve an electrical conducting network without a severe worsening of the shape memory characteristics of SMCs, other conductive fillers may be considered. Carbon nanotubes (CNTs) are an excellent alternative.
3.1.2. Carbon Nanotubes
CNTs are conductive fillers with other advantageous characteristics that enable the formation of SMCs with low electrical resistivity at low filler concentrations. This is due to the high aspect ratio of CNTs, which facilitates the formation of an electrically conductive network within the composite. In other words, the percolation threshold for nanocomposites using CNTs is attained at lower concentrations than for other conductive fillers such as CBs.89 The resistivity once past this threshold for a given concentration of CNTs is also found to be lower than for other fillers at the same concentration. This can be advantageous, because a higher concentration of fillers may result in more brittle materials, high cost, high weight, or difficulty in processing due to an enormous increase in viscosity.
It also has been shown that, at the same concentrations, CNT-filled SMCs exhibit higher shape fixity than other fillers, such as CBs,87 which may be attributed to the alignment of the CNTs in the direction of the deformation. In the same work, the authors also showed that their shape memory elastomer filled with CNTs displays a better shape recovery ratio and better overall shape memory characteristics, as can be seen in Figure 4. The shape recovery was electrically triggered in their SMC with 15 parts per hundred rubber (phr) of CNTs at 50 V within 15 min. Conversely, the nature of the polymer matrix itself also influences the shape memory characteristics of the resulting composite. In a recent work, Tekay studied polymer blend networks of polycaprolactone (PCL) and a maleic anhydride grafted block copolymer (denominated SEBS-g-MA) with and without dispersed MWCNTs.90 The author showed that varying the composition of the polymer network toward a 50/50 blend increases the recovery and fixity ratios. Changing the relative concentration to 30/70 of PCL/SEBS-g-MA decreased these ratios by 35% and 3%, respectively. The SMC with 10 phr MWCNTs has a resistivity of 0.0382 Ω m that is able to undergo electrically triggered shape recovery (91.17%) when it is subjected to a constant voltage of 40 V in 56 s.
Figure 4.
Relationship between the shape recovery ratio and the shape fixity ratio of a SMC with different concentrations (in phr) of CBs or CNTs (reproduced with permission from ref (87); published by MDPI, 2022.)
The method of dispersion of the CNTs also plays an important role in the electrical and mechanical properties of the SMCs: investigations have focused on how to avoid CNT aggregation due to van der Waals forces and achieve a random but uniform dispersion within the composite. It has been shown that mini twin-screw melt mixing,91 cross-linking MWCNTs onto semicrystalline networks,92in situ polymerization,93 or surface modification94 can decrease the electrical resistivity of the SMC. The last three methods also enhance the interfacial cohesiveness between the nanotubes and the polymer matrix, which is directly related to a better stress distribution and strength of the resulting composite. Nevertheless, surface modification should be used with caution since it may detrimentally affect the shape recovery ratio during the lifetime of the composite.94 Furthermore, a decrease in the electrical resistivity with an aggressive functionalization of CNTs is due to the generation of defects on the surface of the nanotubes.
Another solution for creating a conductive network with carbon nanotubes is by incorporating them in films or yarns onto the polymer. These are commonly referred to as MWCNT nanopapers or buckypapers, which have been shown, for example in the work of Lu and Gou,95 to facilitate the electrical activation of the shape memory composite when it is subjected to a constant current of 0.6 A or by Lu et al.96 to trigger the shape recovery of a composite at a constant voltage of 30 V while driving up a weight of 5 g by 30 mm. In another investigation, a shape memory epoxy matrix with MWCNT nanopaper was quickly cured by heating with an electric field at constant voltage, reaching 105 °C at 4.5 V.97 The electrical properties of the resulting composite were used once again to electrically trigger the shape recovery. A high concentration of 30 wt % MWCNTs could be embedded in the form of buckypaper in an epoxy shape memory polymer.98 This SMC has an elastic modulus 52% and 514% higher than the pristine material below and above the Ttrans, respectively. It shows shape recovery when it is subjected to 12 V under 22 s.
Wang et al. reported on a novel way of fabricating CNT-filled polyurethane SMC by spraying-evaporation modeling.99 They printed CNT layers at different desired locations of the SMC. Several CNT layers can be printed in order to tailor the electrical resistivity of certain regions. More layers signify a lower electrical resistivity and higher temperature due to resistive heating whenever the material is subjected to a constant electrical current. They show a temperature difference of 12 °C along a single stripe of SMC between the regions with 10 and 50 CNT layers, as shown in Figure 5. This way, sequential and selectively triggered SMCs can be produced. They report complete shape recovery under 30 s when the SMCs are subjected to 40 V.
Figure 5.
Stripe of a shape memory nanocomposite of polyurethane with different numbers of printed carbon nanotube layers along its length and the resulting temperature distribution due to resistive heating (reprinted from ref (99) with permission of Elsevier.)
In the resistive heating phenomenon, the heat generated within the material directly depends on the injected electrical current and the electrical resistivity of the material. A thorough investigation on the phenomenon of resistive heating of an electroactive SMC was recently carried out by our group, including surface temperature measurements on a polycaprolactone SMP with 3 wt % MWCNTs used to validate analytical formulas and a 3D thermoelectric numerical model.100 This investigation shows that the electrical resistivity of the SMC has a nonlinear nonmonotonic dependency on temperature. Other material properties such as the heat capacity or the thermal conductivity are also shown to affect the resulting temperature of the SMC. We also used a bespoke tensile test bench with integrated controllers for resistive heating in order to investigate the evolution of the electrical resistivity of the same SMC during 1W-SM cycles performed with uniaxial tensile deformation.101 Using a proposed dedicated PI controller, the resulting temperature on the surface of the SMC due to resistive heating can be efficiently and accurately controlled in order to follow a certain heating/cooling ramp or in order to maintain a constant temperature even though the electrical resistivity varies with time. The results presented in ref (101) give also indications of the interplay among the electrical, thermal, and mechanical properties of the electroactive composite within shape memory cycles. An example of this interplay is shown in Figure 6, where the strain in the loading direction is shown as a function of a thermal cycle performed at a constant stress of 600 kPa in order to investigate the 2W-SM characteristics of the SMC. The electrical resistivity during the process is also depicted. Further information on these measurements can be found in ref (101).
Figure 6.
Evolution of the Green–Lagrange strain in the loading direction (εyy in red) and of the instantaneous electrical resistivity (ρe in black) with temperature of an SMP of polycaprolactone with 3 wt % MWCNTs. The test was performed at a constant stress of 600 kPa in order to study the 2W-SM cycle.
The energetic efficiency of the resistive heating process has been studied by Cortés et al.102 for a thermosetting epoxy SMP with 0.2 wt % MWCNTs. Furthermore, because the heating can be generated locally in the surfaces of interest, this last work also illustrates the ability of sequentially activating the shape recovery of different regions within a SMC plate, as shown in Figure 7, under a voltage in the range 126–265 V. The same research group studied the IR-activated shape recovery of printed shape memory acrylic resins with 0.1 wt % MWCNTs.103
Figure 7.
Sequential shape recovery due to selective resistive heating of a shape memory composite stripe. Three electrodes are painted on the surface of the SMC and are located on the right and left tips and another in between. The regions that heat up depend on which electrodes are used for injecting the electric current. During the first 10 min the current is injected between the central and left electrodes and during the last 10 min between the central and right electrodes (reprinted from ref (102), with permission of Elsevier.)
SWCNTs have also been used within SMPs in order to achieve a multiresponse. In their work, Xiao et al. showed the beneficial characteristics of a pyrene-based SMP when 1–3 wt % SWCNTs was dispersed in a PCL copolymer in order to obtain a SMC that can be thermo-, electro-, or photoactivated.104 Upon the application of a constant voltage of 50 V, the SMC heated to 65 °C in 20 s, which proved satisfactory during shape recovery.
Electrically assisted CNT dispersion has been used in order to align CNTs within polymer matrices. Electric fields have successfully been applied during curing of polymer networks in order to successfully orient the conductive fillers in the direction of the current.105 Martin et al. showed that AC electrical fields achieve more linearly oriented conductive fillers in comparison to DC electric fields, which results in a more inhomogeneous and branched dispersion.106 These highly oriented nanocomposites overcome the percolation threshold at lower concentrations and show anisotropic electrical properties as well as decent optical transparency. Yu et al. created a SMC with embedded CB where the chained CNT dispersion was shown to decrease the electrical resistivity by over 1 order of magnitude in comparison to the SMC with a randomly oriented distribution.107
3.1.3. Carbon Nanofiber
Carbon nanofibers (CNFs) have been reported as a cheaper alternative to CNTs to improve the electrical conductivity of a PCL shape memory polymer matrix.108 The resulting electrical resistivity for a concentration of 10 wt % is reported to be double that of the composites with CNTs at the same concentration. In an attempt to further decrease the electrical resistivity with carbon nanofibers, another PCL-based composite using carbon fiber felt (CFF) was reported by Gong et al. to have an electrical resistivity of 0.78 Ω m at a lower concentration of 3.6 wt % CFF.109
3.1.4. Graphene
Graphene is another carbonaceous filler that is used in SMCs due to its excellent electrical and thermal conductivities. To facilitate a homogeneous dispersion within the SMP, graphene oxide (GO) and reduced graphene oxide (rGO) are usually utilized. Wang et al. investigated an epoxy nanocomposite incorporating rGO nanopaper that underwent complete shape recovery within 5 s by applying a low electric voltage of 6 V.110 The 2W-SM effect has also been shown in a shape memory polyurethane filled with graphene nanosheets.111 The work demonstrates the 2W-SM effect that can be electrically triggered for higher concentrations of 4 and 8 wt %, which resulted in SMC samples with electrical resistivities of 5.49 and 1.17 Ω m, respectively. During the conventional 1W-SM effect, this SMC shows improved shape memory characteristics, amounting to 93% shape fixity and 95% shape recovery at concentrations of 2 wt %. The investigation reported in this work also shows a better thermal stability of the resulting composite, with higher melting and crystallization temperatures after the addition of the conductive filler. A similar trend can be found in amorphous shape memory nanocomposites: an increased glass transition temperature is observed in amorphous shape memory nanocomposites by the addition of graphene112 with respect to the pristine polymer. After a certain concentration of graphene within the SMP, the glass transition temperature and the shape memory properties start to deteriorate due to poor dispersion and aggregation of the fillers.113 This investigation reported an optimized concentration of 1.5 phr of GO inside the polyurethane SMP. The same research group also examined the electroactive shape memory performance of the SMCs.114 The composites with 1.5 phr rGO could not be heated with their available setup due to a relatively high electrical resistivity on the order of 15 Ω m. However, the electrical resistivity of their SMC with 2.5 phr decreased to 0.4 Ω m. The values of the electrical resistivity reported here are calculated from the data presented in ref (114). Shape recovery of this SMC was obtained by applying 50 V and reaching a temperature of 64 °C after 90 s. A much faster electric shape recovery was reported in a poly(vinyl acetate) SMC with 4.5 wt % rGO.115 The resulting SMC has an electrical resistivity 0.037 Ω m. By applying 70 V, full shape recovery was demonstrated to happen within 2.5 s. Similarly, graphene-filled PCL has shown to achieve satisfactory electrical resistivities for slightly higher concentrations: i.e. 0.09 Ω m for 7 wt % reduced graphene oxide.116 Graphene–MWCNT hybrids have also been reported in a PCL matrix.117,118 Although the electrical resistivity using such hybrid conductive fillers may be somewhat smaller than those reported for MWCNT-filled SMCs, as the price of graphene (or SWCNTs) is much greater than that of MWCNTs for a limited improvement in the electrical characteristics, limits graphene competitiveness.119
Concerning the effect of graphene in the mechanical properties of the resulting SMPs, Chen et al. reported increases of 64% and 71% in the tensile strength and elastic modulus by adding 3 wt % of rGO in a shape memory epoxy matrix.120 It has been widely observed in the literature that the addition of high concentrations of graphene-based fillers increases the tensile strength of the resulting composite. The improved load-bearing capabilities usually come at the cost of worse shape memory characteristics due to the agglomeration of graphene at high concentrations. Interestingly, Zhang et al. proposed a method of functionalizing GO in order to produce a SMC that has both high load-bearing capabilities with an elastic modulus of 456.7 MPa and an excellent shape recovery of 100%.121
For further information on graphene-filled shape memory polymers, the reader can consult a recently written comprehensive review on this topic written by Kausar122 and the references therein.
3.2. Ferromagnetic Nanofillers
One of the most used magnetic fillers in SMCs is Fe3O4 because of its biocompatibility and nontoxicity, which enables their utilization for medical purposes. Other magnetic fillers that have been reported in the literature are particles of Fe2O3, Ni, NiZn, NdFeB, or NiMnGa. Examples of materials with these and other fillers can be found in a recent review written by van Vilsteren, Yarmand, and Ghodrat.123 Even though some SMCs contain nickel-based magnetic nanoparticles,124−126 no reference was found where these particles were used for the magnetic heating of the smart composites, with the exception of NiZn ferrite nanoparticles. For instance, Buckley et al. used 10 vol % NiZn ferrite nanoparticles dispersed in their SMP and achieved a temperature increase of 11 °C after 40 s in a 545 A m–1 magnetic field at 12.2 MHz.127 The fact that nickel nanoparticles are not or are rarely used in magnetoactive SMCs may be due to the lower heating efficiency of the polymer with Ni than with other nanoparticles such as iron or magnetite.128 Furthermore, the use of nickel nanoparticles may be limited in magnetic heating due to the complex synthesis,129 carcinogenic effects,130 and the unavoidable formation of oxide layers on the nickel nanoparticles, which is shown to detrimentally affect the magnetic heating ability.131 By adding polymer coatings to fight these drawbacks, the heating efficiency of nickel nanoparticles is dramatically decreased.129
Conventionally, magnetic particles are incorporated in the SMP matrix by physical blending. Due to van der Waals and magnetic forces, the unmodified magnetic particles tend to aggregate. This, together with the weak compatibility between the particles and the matrix, results in unsatisfactory mechanical and shape memory properties.132 In order to reduce the agglomeration of particles and improve the properties of the composite, several techniques have proven satisfactory, such as filler surface modification,132 chemical modification of the polymer matrix,133 or mechanical ultrasonication.134
The shape memory characteristics of magnetosensitive SMCs has been investigated in terms of the heating mechanism.135 It was found that the shape fixity ratio was higher in magnetosensitive SMCs when they were activated with magnetic heating rather than activated with conventional heating. The shape recovery ratio at a given magnetic field intensity, on the other hand, depends on the concentration of the magnetic fillers. It is higher for conventional heating when a low amount of magnetic fillers is dispersed. Nevertheless, after a certain threshold is attained, magnetic heating also results in a higher shape recovery ratio than conventional heating. Moreover, a few years later, the same research group investigated different parameters that have an influence on the temperature that can be reached with inductive heating on a biodegradable multiblock copolymer filled with Fe3O4.136 The magnetic field intensity H was varied between 7 and 30 kA m–1 (corresponding to μ0H = 8.8–37.7 mT, where μ0 is the magnetic permeability of free space) with frequencies between 253 and 732 kHz. They found that increasing the nanoparticle concentration and the magnetic field strength resulted in higher temperatures. They also studied the influence of the surrounding environment, leading to a lower steady-state temperature in distilled water than in air, and a considerably slower heating phenomenon when the material was submerged in a saline solution. Furthermore, the composition of the magnetic nanoparticles and their distribution within the polymer matrix were investigated. They found that, with a homogeneous distribution of 10 wt % of Fe3O4 nanoparticles, the mechanical properties of the composite were not dramatically different from those of the pristine SMP, while observing and improvement in its shape fixity by 7% and the shape recovery ratio by >2% on the third shape memory cycle.
An SMC with multiple fillers was fabricated by He et al.137 The conformation of the composite structure allowed for remotely and selectively triggering the shape recovery in specific regions in order to obtain different shapes. Their composite structure was divided into three regions with different compositions, as illustrated in Figure 8: on the right an epoxy SMP filled with 5 wt % Fe3O4 nanoparticles, on the center neat epoxy SMP, and on the left the epoxy SMP filled with 0.4 wt % CNTs. The selective triggering of the shape memory effect is due to a frequency effect: the region with Fe3O4 heats up by a magnetic field at 296 kHz, and the region with CNTs heats up by a magnetic field at 13.56 MHz. Among the plausible mechanisms responsible for the heating of the CNT-filled region when the material is subjected to a high-frequency magnetic field, one could be the presence of ferromagnetic metal impurities in the CNTs as a result of their manufacturing.138−140 After the shape programming process where the permanent (straight) shape is heated past the Ttrans and deformed into the temporary shape (temporary shape #1), the recovery back toward the permanent shape can follow five different recovery routes with five total intermediate temporary shapes. Figure 8 illustrates each recovery path with a different arrow color (in gray values) together with a sketch of the required frequency of activation. The last step of each route toward the flat permanent shape needs to be achieved with conventional heating of the composite since the neat SMP central area does not contain any filler.
Figure 8.
Illustration of the five shape recovery routes that can be followed from temporary Shape #1 to the permanent flat shape of a SMC with three different regions: CNT-SMP that can be activated with inductive heating at 13.56 MHz, neat SMP that can be activated with conventional heating, and Fe3O4-SMP that can be activated with inductive heating at 256 kHz. (Reproduced from ref (137) with permission of John Wiley and Sons. Copyright 2011 Wiley-VCH.)
SMCs with multiple ferromagnetic particles have also been produced in order to achieve a functional 2W-SM behavior.141 In this investigation, Fe3O4 and magnetized NdFeB particles were dispersed in an acrylate-based SMP. The Fe3O4 particles, with an average size of 30 μm, were heated by hysteresis loss with a high-frequency magnetic field of 10 mT at 60 kHz. The NdFeB particles, characterized by having huge coercivity, do not contribute to the heating due the small magnitude of the magnetic field. Instead, the NdFeB particles are used to generate a magnetic torque and induce a shape change when they are subjected to a DC magnetic field of 30 mT. As depicted in Figure 9, first the AC magnetic field Bh is applied onto the structure in order to heat up the material. When the Ttrans is surpassed, the DC magnetic field Ba will cause the cantilever beam to bend. The new temporary shape can be fixed by removing the AC magnetic field while keeping the DC field. Otherwise, the material can be used as an actuator by keeping the AC magnetic field and varying the sign of the DC magnetic field at a low frequency (0.25 Hz), hence creating a reversible bending transformation of the beam. Besides the interesting potential applications and versatility of shape morphing and actuation of this material, the authors also proved the excellent shape memory characteristics, amounting to a shape fixity ratio of 95% and complete shape recovery.
Figure 9.
Principle of operation of a multishape magnetosensitive SMC with embedded NdFeB and Fe3O4 particles. (a) Illustration of magnetization and composition. (b) The SMC is heated by the application of a high-frequency AC magnetic field Bh above the glass transition of the SMP. Deformation and actuation can be achieved by a DC (or low-frequency AC) magnetic field Ba. (c) Shape fixity can be achieved by removing Bh while keeping the desired Ba. (Reproduced from ref (141), with permission of John Wiley and Sons. Copyright 2019 Wiley-VCH.)
Even though magnetic particles are normally embedded inside SMPs to enable magnetic heating of the resulting composite, some research groups have used them as means of increasing the electrical conductivity of the polymer matrix and heated the resulting composite with the Joule effect by injecting an electric current. Leng and co-workers have produced a polyurethane SMP with dispersed Ni powder. They produced aligned Ni chains by using a magnetic field during the curing of the SMP.142 The aided alignment of the magnetic nanoparticles helped to reduce the electric resistivity to <0.1 Ω m for 20 vol % of Ni powder. Resistive heating of up to 55 °C was shown when applying 6 V.
3.3. Other Fillers
Noble metals in the form of nanoparticles have attracted interest as fillers of SMPs in order to achieve multistimuli SMCs. This is due to their photothermal effect, where they absorb light at certain wavelengths from UV to near-IR and convert it into thermal energy. In their investigation, Mishra and Tracy created two similar SMCs filled with either gold nanospheres or gold nanorods, and they were able to selectively trigger the shape memory effect depending on the wavelength of light used: 530 nm for the nanospheres and 860 nm for the nanorods.143 Other noble metals that have been used to photothermally activate SMCs are platinum144 and silver nanoparticles.145 Besides nanoparticles of noble metals, metallic ions have been incorporated as fillers in SMCs to activate the shape memory effect using light in the near-IR. Bai et al. reported a composite made by cross-linking poly(acrylic acid) and poly(vinyl alcohol) with different metallic ions.146 They looked into Cu2+, Cd2+, Cr2+, Al3+, and ions of ferromagnetic materials (Fe3+, Co2+ and Ni2+).
The SMCs containing polypyrrole (Ppy) can take advantage of the electrical conductivity of the particles for electrically triggering the shape memory effect.147,148 Moreover, Ppy-SMP composites are also shown to exhibit an excellent photothermal performance. Thus, these SMCs are multistimuli. There are many other fillers that have been used in shape memory polymer composites in order to improve mechanical, thermal, and electrical properties or to obtain multistimuli materials. For example, shape memory alloys have been embedded in SMPs in the form of wires149 or particules.80 Silicon carbide has been used due to its temperature stability and good microwave absorption that allows the SMM to be activated through multiple stimuli.150 Oxides of tungsten, silicon, or aluminum have been incorporated in order to improve the mechanical properties of SMC foams for medical devices.151 Other electroactive polymers have been blended with SMPs,152 liquid metals have been used to fill the SMP and achieve electroactivation,153 and fillers such as cellulose nanocrystals154 or nanoparticules of mineral silicates151 have been used to reinforce the SMP but do not enable electro- or magnetoactivation.
After this overview of fillers in SMCs, the reader can proceed to the Appendix, which lists some of the many existing references on electro- and magnetosensitive SMCs, together with the activation mechanism and general characteristics of the materials reported.
4. Applications
In view of the highly tailorable material properties of SMCs together with the excellent shape memory properties and possible and versatile multistimuli activation, SMCs are used (or intend to be used) in many areas of research and industry. In this section, a few applications of shape memory composites are described, paying special attention to those that can be electrically activated.
4.1. Biomedical Industry
Within the biomedical industry, shape memory polymers have attracted a great deal of attention. Most of the applications in this field use conventional heating of SMPs4 or heating with magnetic fields of SMCs with magnetic nanofillers. A composite made from a poly(lactic acid) SMP with dispersed 10 wt % Fe3O4 nanoparticles was used in 4D printing in order to fabricate prototypes of shape memory heart occluders for the treatment of congenital heart diseases.155 A similar SMC was used for conceiving a tracheal scaffold.156,157 The composite, with 15 wt % Fe3O4, is to be implanted as a flat surface and, under the influence of an AC magnetic field of 4 kA m–1 at 30 kHz, it curls to form a tube within 35 s, as shown in Figure 10a. Moreover, the same research group also reported on the feasibility of this SMC to create different designs of bioinspired scaffolds to be applied for bone repair and regeneration.158 One of the proposed structural designs is shown in Figure 10b, where the magnetic activation (4 kA m–1 at 30 kHz) expands the SMC scaffolds within 15 s. As an additional example, magnetoactive SMCs have also been conceived in the biomedical field to apply in drug delivery159 or as intravascular stents.160,161
Figure 10.
4D printed poly(lactic acid) SMP with embedded Fe3O4 composites actuated by a magnetic field for the conception of (a) tracheal scaffolds (reprinted from ref (156), with permission from Elsevier) and (b) porous bone tissue scaffolds (reprinted from ref (158). with permission from Elsevier.)
Electroactive SMCs have also been found to be the material of choice for biomedical applications in the literature. For example, a porous SMP foam with homogeneously dispersed CNTs has been fabricated for the potential treatment of intracranial aneurysms by implantation through a catheter (Figure 11a) and subsequent expansion of the foam in order to fill the aneurysm, as illustrated in Figure 11b.162 Thanks to the incorporation of the CNTs, X-ray imaging can be used during the implantation procedure to visualize the path of the SMP foam. The researchers used a foam with 1 wt % CNTs and investigated the steady-state temperature reached due to the injection of a constant current. After compressing the foam to 60% of its initial volume, the authors triggered shape recovery to the expanded foam in less than 1 min by injecting 1 A (Figure 11c–e), demonstrating in this way the potential of the SMC in the treatment of aneurysms. A similar investigation was reported in their previous work with 0.005 g mol–1 CNTs.163 The applied deformation toward the temporary shape is a compression of more than 50%. Complete electrically triggered shape recovery toward the expanded permanent shape happened within 2 min at 0.2 A and reached a maximum temperature of 46.8 °C, preventing neighboring tissue damage.
Figure 11.
Illustration of the treatment of aneurysms by filling them with a SMC. (a) Compressed SMC foam being placed in a saccular aneurysm via a catheter. (b) Recovered expanded shape of the SMC filling the aneurysm space. Demonstration of the prototype for the treatment of aneurysms of a SMC with 1 wt % CNTs at (c) 0 s, (d) 30 s, and (e) 60 s after the injection of an electric current (1 A). (Reproduced from ref (162) with permission of John Wiley and Sons. Copyright 2021 Wiley-VCH).
4.2. Aerospace Industry
Within the aerospace field, self-deployable structures have been reported using magnetoactive SMCs. A reconfigurable morphing antenna was reported to be heated in an AC magnetic field of 40 mT at 60 kHz to be used as a self-deployable structure or as a structure with variable resonant frequencies.141 Regarding electroactive SMCs, Paik et al. fabricated a SMC made of polyurethane with 7 wt % MWCNTs by in situ polymerization. The electrical resistivity of the SMC was around 0.8 Ω m. The SMC was applied in a microaerial vehicle to actuate the control surface (rudder).164 When an electric current was injected in the SMC, it shrunk and produced a 30° angle deflection. Unfortunately, because this was a 1W-SM effect, subsequent shape memory programming was be needed for further deflection of the control surface, which is not compatible with realistic flight control. More recently, a hybrid shape memory composite actuator has been developed as a laminate composed of SMA wires and SMP filled with nichrome for morphing flap flight.165 Resistive heating was induced in the composite structure by injecting a current of 0.7 A in the SMA wires and of 0.4 A in the SMP–nichrome composite. The authors used this composite for the morphing of the trailing edge of the wings of an unmanned aerial vehicle, resulting in deformation angles between 32 and 39°, as shown in Figure 12. The flapping of the trailing edge proved to create an extra lift force on the wing while decreasing the drag.
Figure 12.
(a) Wing of an unmanned aerial vehicle with a morphing flap at the trailing edge capable of changing the deformation angle between (b) 39° and (c) 32°. The actuation mechanism is through resistive heating of both the SMA wires and the SMP–nichrome composite. (Reprinted from ref (165), with permission from Elsevier.)
4.3. Soft Robotics
Using a magnetoactive SMC made of poly(aryl ether ketone) with dispersed Fe3O4 nanoparticles, Yang et al. showed the diversity of potential applications of their composite exhibiting a remotely and magnetically triggered 1W-SM effect.166 Their layered-film nanocomposite with 15 wt % magnetic nanoparticles was subjected to an AC magnetic field of 27.9 kHz for 50 s. They fabricated a ball launcher that propelled a ball forward by triggering the shape recovery with the AC magnetic field. Moreover, they showed the morphing abilities of their composite material with flower-shaped and plane-shaped geometries that unfolded back to a flat position with the application of the magnetic field. Other soft robotic applications were envisioned by, for instance, Cohn et al., using their SMC of polycaprolactone dimethacrylate with 5 wt % of Fe3O4 nanoparticles.167 In their published work, they showed the morphing of several structures when the material was subjected to a magnetic field (4 kA m–1 at 375 kHz) such as a self-deployable honeycomb cylinder or a spider web and the folding of a spiral. They investigated rolling and gripping movements using their SMC.
A crawling robot was reported by Peng et al.168 In this work, they used a double layer of a 3D porous CNT sponge SMP to create a nanocomposite that was able to undergo a 2W-SM effect when it was subjected to resistive heating. The CNT loading inside the composite was of 1.08 wt %, which resulted in an electrical resistivity of about 0.007 Ω m. They recreated an inchworm locomotion by applying voltages between 2 and 8 V. The inchworm-like robot is able to travel 1.2 cm in 10 min. Xu et al. produced another SMC made of poly(ethylene-co-octene) filled with CNTs to achieve a composite was is able to undergo a stress-free 2W-SM effect with a fast response when it was subjected to either low voltages (<36 V) or infrared light (250 mW cm–2).48 In this work, they built an electroactive gripper and a photoactive crawling robot. The gripper, shown in Figure 13, was formed in a C shape. At room temperature, the opening of both tips of the C shape was 0.4 cm. Upon the application of 36 V, the temperature of the composite increased to 55 °C and the shape changed by increasing the aperture to 0.8 cm in 18 s. When the material was cooled once again, the original opening of 0.4 cm was recovered after 168 s. They showed the lifting capabilities of this gripper by lifting 15 nuts (69.09 g). In their most recent work, Xu et al. also reported a high-temperature warning robot that was made possible by the addition of paraffin wax into the composite.169 They hot-pressed the SMC layer with sandpaper to make the surface uneven, and they stacked two uneven-surfaced layers together. At high temperature, the stiffness of the SMC decreased and the surface roughness disappeared, leaning to a better contact between both layers that could be monitored thanks to a change of electrical resistivity across the layers of more than 2 orders of magnitude.
Figure 13.
(a) Photographs and (b) IR thermal images of the SMC electroactive gripper. (Adapted with permission from ref (48). Copyright 2019 American Chemical Society.)
Other works have also exploited the weight-lifting and 2W-SM and multiple shape capabilities of SMCs. Xu et al. used a poly(ethylene-co-vinyl acetate) filled with CNTs that was activated with a voltage.170 Instead of using the SMC as the grip, they mounted it in a setup that permitted the opening and closing of a mechanical grip due to the elongation and contraction of the SMC during its 2W-SM effect. Furthermore, Hu et al. built a device with a magnetoresistive triple SMC that was capable of displaying text in Braille and self-refresh in order to potentially apply it as a Braille e-book.171
4.4. 4D Printing
Within 4D printing, Dong et al. investigated several applications in poly(lactic acid) filled with 8 wt % CNTs.172 They demonstrated the fast shape recovery (<44 s) of pyramid-, diamond-, and crown-shaped printed structures when they were subjected to voltages in the range 20–35 V. Interestingly, they also developed a multisegment self-deployable rectangular structure that can be selectively deployed in three sections. Each section was actuated with 25 V in order to reach a temperature >70 °C. Moreover, they also developed an external stent that can be implanted in an extended unfolded shape outside of the body part. By applying 20 V, the stent shrinks and bends within 60 s.
4.5. Self-Healing
Self-healing of structures is another popular application that has been investigated in the scope of electrically and magnetically triggered shape memory composites. This property can be used in a plethora of applications in order to self-repair surfaces and structures that have suffered from impact or fatigue, thus prolonging the life of the structure and delaying disposal. An example of this kind of composite was reported by Ren et al.173 It consisted of a polycaprolactone/thermoplastic polyurethane SMP with 2–6 wt % CNTs that exhibited fixity and shape recovery ratios of 96% and 94%, respectively. They reported excellent electrically triggered healing efficiencies of the composites ranging from 94.79% to 96.15%. Furthermore, comprehensive reviews on electroactive174 and magnetoactive175 self-healable nanocomposites have been recently published, where other examples of shape memory self-healable materials can be found.
Besides the applications that have already been reported in the literature, the interesting properties of electro- and magnetoactive SMCs make these materials excellent candidates for other prospective applications such as artificial muscles, self-deploying antennae and solar arrays, dental fixtures, self-foldable devices, and many more.176,177
5. Conclusions and Future Horizons
Bearing in mind all the information and the numerous references cited in this review, it is not surprising to say that SMPs and SMCs have attracted a great deal of attention and interest from researchers and the industry. Even though shape memory materials have been investigated for years, there are probably many things to discover, many properties to keep enhancing, and many applications to be conceived.
Considering the extensive literature review performed, some prospective future directions in the research field of shape memory polymer and their composites are listed here below.
-
(i)
High importance has been given in recent research to minimize the time during shape recovery of SMCs for their application as actuators. This trend will likely continue in order to obtain faster SMC actuators. Besides the ability to achieve a fast recovery time, a variable time scale or even a fast interruption of the shape recovery may be desirable for some applications, notably in the biomedical field.
-
(ii)
Materials with a tunable transition temperature are highly desirable, but further investigation is needed in order to do so accurately and repetitively. Furthermore, broad transitions may lead to multiple intermediate temporary shapes during a slow shape recovery. The accurate characterization, description, and prediction of these intermediate shapes could widen the application of these materials.
-
(iii)
Due to the specific requirements of the biomedical industry, surely more investigations will follow on biocompatible SMPs and SMCs that can be remotely actuated. Besides, if they are also biodegradable, some implants would not need a secondary surgical intervention to be removed.
-
(iv)
On the same green trend as biodegradability, dynamic covalent bonds on the SMP matrices will surely be kept under investigation. These allow for recyclability and reprocessability of the material. Similarly, research on self-healable SMCs is on the rise so as to increase the lifetime of the structures.
-
(v)
Especially within the area of electrically activated SMCs, there has been recent interest in selectively and sequentially triggering the shape recovery. By improving the localization of heat generation, sequential shape recovery may lead to more applications of electroactive SMCs in self-deployable and self-foldable devices.
-
(vi)
There are some cases where direct heating cannot be applied in order to trigger the shape recovery, such as when the material is surrounded by sensitive equipment or when it is inside of the human body. In general, the current trend in the investigation of SMCs is in multistimuli materials where, besides a conventional thermal treatment, alternative heating can be applied in order to trigger the shape memory effects remotely. These mechanisms include resistive heating, electromagnetic induction, and photothermal effects. An in-depth understanding of the energy conversion into heat of these alternative heating mechanisms may allow for complex and accurately actuated devices.
-
(vii)
Because of the direct link between the material and the mechanical and shape memory properties of SMCs, a predictable and controllable temperature through alternative heating mechanisms (resistive Joule heating, electromagnetic induction, etc.) is desired.
-
(viii)
Due to the one-time actuation or the required time-consuming shape reprogramming of conventional 1W-SM polymers and composites, materials exhibiting 2W-SM effects will likely be the object of future investigations in order to meet the challenging needs of future applications. Undoubtedly, more research is needed in this area to evaluate the fatigue and wearing of the 2W-SM characteristics in order to estimate the lifetime expectancy of these materials when in use. Similarly, the cyclic (or subcyclic) lifetime of multiple shape memory composites should be further investigated.
The outstanding and polyvalent characteristics of these materials offer an incredible platform to investigate and create applications that are, as of now, unimaginable.
Acknowledgments
This research was founded through the “Actions de recherche concertées 2017 – Synthesis, Characterization, and Multiscale Model of Smart Composite Materials (S3CM3) 17/21-07”, financed by the “Direction Générale de l’Enseignement non obligatoire de la Recherche scientifique, Direction de la Recherche scientifique, Communauté française de Belgique et octroyées par l’Académie Universitaire Wallonie-Europe”.
Glossary
Abbreviations
- BMI
bismaleimide
- BR
butadiene rubber
- Cp
heat capacity
- CB
carbon black
- CE
cyanate ester
- CF
carbon fiber
- CNF
carbon nanofiber
- CNT
carbon nanotube
- CSF
carbon short fiber
- DMA
dimethacrylate
- EOC
ethylene-1-octene copolymer
- EP
epoxy
- G
graphene
- Gr
graphite
- GO
graphene oxide
- GNP
graphene nanoplatelets
- LLDPE
linear low-density polyethylene
- MA
methacrylate
- MWCNT
multiwalled carbon nanotube
- NGDE
neopentyl glycol diglycidyl ether
- phr
parts per hundred rubber
- PBS
poly(butylene succinate)
- PCL
polycaprolactone
- PCLA
poly(l-lactide-co-ε-caprolactone)
- PCO
polycyclooctene
- PDL
ω-pentadecalactone
- PE
polyethylene
- PEG
poly(ethylene glycol)
- PEN
polyethylene naphthalate
- PEVA
poly(ethylene-vinyl acetate)
- PK
polyketone
- PMMA
poly(methyl methacrylate)
- PLA
poly(lactic acid)
- PPC
poly(propylene carbonate)
- PPDO
polydioxanone
- PPy
polypyrrole
- PS
polystyrene
- PU
polyurethane
- PVA
polyvinyl acetate
- PVAl
poly(vinyl alcohol)
- PVDF
polyvinylidene fluoride
- Py
pyrene
- RF
shape fixity ratio
- rGO
reduced graphene oxide
- RR
shape recovery ratio
- SBS
poly(styrene-butadiene-styrene)
- SMA
shape memory alloy
- SMP
shape memory polymer
- T
temperature
- Tc
crystallization temperature
- Tg
glass transition temperature
- Tm
melting temperature
- tR
recovery time
- ε
strain
- κ
thermal conductivity
- ρe
electrical resistivity
- ς
concentration
- χc
crystallinity
- 2W-SM
two-way shape memory
Appendix
This appendix provides Table 1, giving some of the many references available in the literature of shape memory composites with certain fillers that allows them to be activated with either an electric field or a magnetic field. Some other characteristics of the materials or the experiments are given as additional information. For a better understanding of the information listed in the table, please refer to the list of abbreviations given below.
Table 1. Compilation of References of Shape Memory Composites Activated with an Electric or Magnetic Field.
SMP | filler | ς | ρe (Ω m) | activation | activation magnitude | T reached (°C) | tR (s) | comments | ref |
---|---|---|---|---|---|---|---|---|---|
PCL | MWCNT | 3 wt % | 0.15–0.6 | electrical | controlled resistive heating (0.6–15.7 mA) | 60 | 130 | 1W-SM and 2W-SM effects; RF > 96.7% and RR > 88.9%; variable electric current to achieve constant temperature and constant heating and cooling rates; nonlinear evolution of ρe within shape memory cycle | (101) |
PU | MWCNT + Fe3O4 | 30 wt % Fe3O4 +0.25–1 wt % MWCNT | magnetic | 29.7 kA/m at 45 kHz | >45 | 60 | RR >95% | (178) | |
PU/SM-PU | CB | 1–5 wt % | 5 × 10–2 Ω m for 5 wt %; 10–3 Ω m for 3 wt % | electrical | 100 V | 25 | 2W-SM; RR 99.8% | (179) | |
PU | MWCNT | 5 wt % | ∼1 Ω m | electrical | 60 V | 100 | 10 | (180) | |
EP | MWCNT | 0.2 wt % | 1–10 Ω m | electrical | 126–265 V | 200–250 | <300 | sequential recovery | (102) |
EP | MWCNT | 30 wt % | 47 Ω m | electrical | 12 V | >55 | 22 | (98) | |
EP | rGO | 0.5 wt % | 0.025 Ω m | electrical | 10 V | >100 | 10 | RF = 95% and RR = 98%; electrical recovery | (181) |
PLA/PU | MWCNT and CF | 6 wt % | electrical | 10 V | >70 | 25 | RR = 94%; negative Poisson ratio | (182) | |
PVAl | MWCNT | 5–30 wt % | 0.52 Ω m for 30 wt % | electrical | 45–120 V | >74 | 35 | (183) | |
PCL | MWCNT coated Fe3O4 | magnetic | 6.8 kA/m at 20 kHz | T increase of 17.2 | 120 | alignment of Fe3O4-coated MWCNT due to magnetic field 0.2 T; increase of elastic modulus by almost 100% but reduction of max elongation | (184) | ||
PCL | MWCNT | 3 wt % | 0.05 Ω m | electrical | 0.1 W | ∼50 | analytical, numerical, and experimental investigation | (100) | |
BR | MWCNT and CB | 0–15 phr MWCNT or 0–30 phr CB | 10–2 Ω m for 15 phr CNT or 30 phr CB | electrical | 50 V | 110 after 15 min at 50 V | 900 | lower ρe and higher RF and RR for CNT than CB | (87) |
PU | Fe3O4 | 0–9 wt % | magnetic | 250 A/m at 20 kHz | >52 | 35 | RR decrease with ς, slight RF increase with ς | (185) | |
PBS–PEG | CNT | 0.2–1 wt % | 7.05 Ω m for 1 wt % | electrical | 20–120 V | 85 at 70 V for 1 wt % | 55 | CNT reduced ductility | (186) |
PU | CNF | 1–7 wt % | 560 Ω m for 7 wt % | electrical | 100–300 V | 25–55 | χc, RF decrease with ς of CNF; elastic modulus increase with CNF ς | (187) | |
PU | CNT | 3 wt % | 0.037 Ω m | electrical | 50 V | >40 | 40 | different preparation techniques and effect of dispersion and electrical properties | (93) |
PU | G sheet | 1–8 wt % | 1.17 Ω m for 8 wt % | electrical | 65–80 V | >70 | 8 | 2W-SM; study mechanical properties; RF = 95%; RR = 93% | (111) |
PU | CNT | 4 wt % | electrical | 40 V | 100 cross-linked CNT | 50 | CNT covalently bonded to PU; increase RR and RF for cross-linked CNT (>10%) | (92) | |
PU | rGO | 0–2.5 phr | 0.4 Ω m for 2.5 phr | electrical | 50 V | 64 for 2.5 phr | 120 | Tg and mechanical properties increase with ς; ductility increase with cycles; RR decreases with ς | (114) |
EOC | CB | 0–38 wt % | 0.08 Ω m for 17 wt % | electrical | 5–15 V for 17 wt % | 40–160 | 60 | two different CB fillers; lower RR for higher ς | (85) |
PU | rGO + Fe3O4 and γ-Fe2O3 | 3 wt % GO + 7 wt % Fe3O4 | magnetic | 287 kHz with 300A in coil | 70 for mixture, 60 for hybrid | 60 | Young modulus doubled from hybrid to only Fe3O4 | (188) | |
PU | Ni powder chains | 0–20 vol% | 0.005–0.01 Ω m for 20%, 104 for 4% | electrical | 6 V | 55 | 90 | alignment magnetic field; decrease Tg with ς | (142) |
PU | CB with and without 0.5 vol% Ni | 4–10% CB | 0.1 Ω m for 10% CB + chained 0.5% Ni | electrical | 30 V | 80 | 120 | ρe increases with number of cycles | (189) |
PVAl | CNT and G | 20 phr GO and 4–16 phr CNT | 0.1 Ω m for 20 GO + 16 CNT | electrical | 60 V | 100 max (normally 60) | 120 | ρe decreases with bending angle | (190) |
PS | CNT | 1.47–7.02 wt % | 0.01 Ω m for 7.02 wt % | electrical | 35 V | >85 | 80 | RR ≈ 100% | (96) |
PU | GNP | 1–3% | 4 Ω m for 3% GNP | electrical | 75 V | 100 for 3% | 60 | χc, elastic modulus, RF and RR increase with ς; strength, Tg, and Tm decrease with ς | (191) |
EP | Ag decorated rGO and CF mat | 2–4 wt % rGO and 11.6% CF | 2.3 × 103 Ω m | electrical | 8.6 V | 100 | 36 | RR = 99% | (192) |
PS | MWCNT nanopaper | 1.47–7.02 wt % | 0.008 Ω m for 7.02 wt % | electrical | 0.6 A | >62 | 300 | RR = 98% | (95) |
EP | CNF | 9.18 wt %? | 0.03 Ω m | electrical | 10–20 V | >50 | 2.1 | increase elastic modulus above Tg, increase κ | (193) |
PU | MWCNT | 3–7 wt % | 12 Ω m for 3 wt % | electrical | 40 V | >70 in 10 s | 9 | Young modulus increase by 50%; RR > 98%; Tc increase with ς | (194) |
PU | Fe3O4 | 0–10 wt % | magnetic | 30 kA/m at 258 kHz | 70 at 150 s 12.5 kA/m 258 kHz | 22 | magnetic heating depends on geometry; for high ς, RF and RR higher for magnetic actuation | (135) | |
PK BMI | MWCNT | 8 wt % | ∼ 0.06 Ω m | electrical | 20–50 V | 40–140 | RR = 90% by 40 V; self-healing | (195) | |
PU | MWCNT | 3–10% | 0.9 Ω m | electrical | 9.5 mA | 32 | (164) | ||
PU | GO | 0–20 phr | >7 × 102 Ω m | electrical | 100–200 V | <60 | (196) | ||
EP | Fe3O4 coated with oelic acid | 1.5–8 wt % | magnetic | 30 mT at 293 kHz | T increase of +25 | 60 | (133) | ||
PPC/PLA | MWCNT | 1–3 phr | 0.1 Ω m for 3 phr | electrical | 20–30 V | 130 max for PPC/PLA (50/50) 3 phr at 30 V | 30 | max tensile strength and strain at break increased with ς at room temperature; At 50 °C strength decrease with concentration; RR = 97% | (197) |
PLA/PU (70/30) | CB | 0–8 phr | ∼0.4 Ω m for 8 phr | electrical | 30 V | 98 for 8 phr | 80 | RF = 90%; RR increase from 59% to 85.9% for 8 phr CB; conductivity ∼1 order of magnitude higher for PLA/PU(70/30) than for PLA | (198) |
PU | CF/PU yarn fabric | 10.9% CF | 1 × 10–3 Ω m | electrical | 6 V | 87 | 80 | RR = 99.3%; applications of deployable structure | (82) |
PEVA | CNF | 0–15 wt % | 0.204 Ω m for 15 wt %; 5 × 109 Ω m for 3 wt % | electrical | 60 V | 100 after 25 s | 2W-SM; RR decreases with ς | (199) | |
PU/PCL | modified MWCNT | 2–10 wt % | 1 Ω m for 10 wt %; 102 Ω m for 4 wt % | electrical | 40 V | >50 | 15 | Tg of PLA increase with ς; Tgof PU decrease with ς; Tcdecrease with ς | (200) |
PU/PVDF (90/10) | modified MWCNT | 0–10 wt % | 50 Ω m for 10 wt % pristine CNT; 4 Ω m for 10 wt % modified CNT | electrical | 40 V | T rise of +40 | 15 | κ × 4 by adding 10 wt % MWCNT; increase 10% tensile strength and decrease 100% elongation at break for surface-modified MWCNT than pristine; RR decrease with cycles faster for modified than pristine CNT | (94) |
PU | Fe3O4 | 0–40 vol% | 106 Ω m for 40 vol% | magnetic | 4.4 kA/m at 50 Hz | 50 | 1200 | electric percolation at 30 wt %; Cp decrease with ς; κ increased by 0.4 for 40 vol % | (201) |
PEVA and PU | FE2O3 | 0–13.6 wt % | magnetic | 15–23 kA/m | 100 for PEVA; 108 for PU | (202) | |||
PDL | Fe3O4 | 5–11 wt % | magnetic | 7–30 kA/m at 258 kHz | 100 | 180 | decreased RF (95% to 90%) and increased RR (95% to 97%) for increased ς | (203) | |
EP | CB | 0–10 wt % | 200 Ω m for 10 wt % | electrical | 3.5 W constant power | 65 | <200 | ρe vs ε during elongation and recovery | (204) |
PVA | GNP | 1.5–4.5 wt % | 0.04 Ω m for 4.5 wt % | electrical | 50–70 V | 2.5 | elastic modulus increases by 1 order of magnitude from 0 to 4.5 wt %; Tg decrease with ς | (115) | |
PU | PPy | 8–21% | 0.105 Ω m | electrical | 40 V | 70 | <30 | Tm decrease with ς | (205) |
PCL DMA | Fe3O4 | 2–12 wt % | magnetic | 300 kHz, 5 W | 43 | 20 | (206) | ||
PLA/PU (50/50) | CNT | 1–5 wt % | 0.33 Ω m for 5 wt % | electrical | 50 V | 50–70 | <150 | decrease RR (13% smaller) in 5 wt %; faster shape recovery with electrical heating than conventional; higher RR with electrical heating than conventional | (207) |
EP | MWCNT nanopaper | 40 wt % | 3 × 10–3 Ω m | electrical | 0–6 W | 20–170 | 20 | curing with resistive heating 105 °C after 40 s at 4.6 V; recovery at 76 °C | (97) |
PE | CNF | 0–11.6 vol% | 0.1 Ω m for 11.6 vol% | electrical | 1–40 V | 100–110 | 100 | Tc and elastic modulus increase with ς; Tm decrease with ς | (208) |
PCL/MA (50/50) | MWCNT | 0–10 phr | 0.0384 Ω m | electrical | 40 V | >70 | 56 | (90) | |
PU | GNP | electrical | 10–30 V | 60 at 10 V | 20 | (209) | |||
PCO/PE | MWCNT | 8–15 vol% | 0.0105 Ω m for 10 vol% | electrical | 150 V | >120 after 2 min | 90 | triple SM effect; MWCNT dispersed in PCO; chemical cross-linking increases ρe due to agglomeration of fillers | (210) |
SBS/LLDPE | CB | 0–25 wt % | 1.08 Ω m for 1 wt % CB; 7.4 × 10–2 Ω m for 25 wt % CB | electrical | 0–120 V | 60 | elongation at break decrease by 50% from 0 to 25 wt % CB; tensile strength increase by >50%; recovery at 40 V | (211) | |
EP | rGO | 0.0027 Ω m | electrical | 1–6 V | 35–240 | 5 | bending | (110) | |
PEVA/PCL | MWCNT | 5 wt % | 0.0205 Ω m | electrical | 20 V | 97 | 12 | recovery at 30 V | (212) |
PU | CNT | 10 to 50 layers | 22.96 Ω m | electrical | 40 V | 90 | 30 | CNT layers at different locations; RR = 100% at 40 V; localized and selective actuation | (99) |
EP/CE | GNP | 0.8–2.4% | 11.3 × 10–2 Ω m for 2.4% | electrical | 20–120 V | 98 | strength increased by 25% for 1.6 wt %; recovery at 60 V | (213) | |
EP | Ag-CF | 5.4 wt %, 1.8 wt % | 95.1 Ω m for 5.4 wt %; 1 × 106 Ω m for 1.8 wt % | electrical | 60–120 V | 60 | RF increase with ς; RR decrease with ς | (81) | |
EP | MWCNT and CB | 1.9 wt % total | 2 × 105 Ω m for 0.8 wt % MWCNT | electrical | 225 V | >80 | 570 | RR = 99% | (214) |
PLA/PU (70/30) | MWCNT and CB | 3–5 phr CB and 0–2 phr MWCNT | <0.1 Ω m for CB 5 phr and >1 phr MWCNT | electrical | 30 V | 70 °C | 100 | 70 °C after 100 s for 3 phr CB and 1 phr CNT; RR = 0 | (215) |
PPDO–PCL/PU | Fe3O4 | 0–10 wt % | magnetic | 7–30 kA/m at 256–732 kHz | 65 | (136) | |||
PMMA–PEG | Modified Fe3O4 | 1–5 wt % | magnetic | 100 kHz, 8 kW | 40.4 | 59 | T rise 4 °C for-surface modified Fe3O4 than for pristine; RF = 90%; RR = 95% | (216) | |
PCL | modified MWCNT | 0–9 wt % | 20 Ω m for >7 wt % | electrical | 50 V for 5 wt % | 70 | 120 | decrease RR with ς | (217) |
PCL-Py | SWCNT | 2 wt % | 3.6 Ω m | electrical | 50 V | 65 | 20 | increase RR with ς | (104) |
MA-based thermoset | Fe3O4 | 0–2.5 wt % | magnetic | 0.33 mT at 297 kHz | 50 | 40 | (218) | ||
PS | MWCNT and CB | 1 wt % MWCNT and 10–15 wt % CB | 0.03 Ω m for CB. 0.025 Ω m for CB+CNT randomly oriented. 0.0075 Ω m for CB+CNT chained | electrical | 25 V | >95 | 75 | MWCNT in chains, reduction of ρe by more than 1 order of magnitude; ρe increases with number of shape memory cycles | (107) |
acrylate | NdFeB and Fe3O4 | 15 vol % of each | magnetic | 40 mT at 60 kHz (to heat up Fe3O4) + 30 mT at 0.25 Hz to actuate beam | 100 (and above) in 40 s | dual magnetic field activation; elastic modulus increases linearly with ς; RF = 95%; RR = 100% | (141) | ||
PLA | Fe3O4 | 10–40 wt % | magnetic | 30 kA/m at 268 kHz | >60 | >200 | nonmonotonic evolution of RR, modulus, tensile strength, and elongation at break with ς | (132) | |
Nafion | Fe3O4 | 15–25% | magnetic | Power 50–100% | internal T 50–110, surface T < 40 | 16 | surface modification of Fe3O4 to improve distribution | (219) | |
PCL/PU | GNP | 0–10 wt % | ∼1 Ω m for 10 wt % | electrical | 10–40 V | 60 for 10 wt % at 10 V; 100 for 10 wt % at 30 V | 60 | 5ecovery within 1 min for 7 wt % at 30 V; faster heating and recovery for higher ς; Tg, χc, viscosity increase with ς; Tmand Tc decrease with ς | (220) |
PLA | PPy | 0.028 Ω m | electrical | 15–40 V | 120 for 40 V | 2 | microfiber membrane PPy coated; worse mechanical properties with Ppy ρe depends on polymerization time and temperature | (221) | |
PCLA | GO | 4.86–13.29 vol % | electrical | 2.5 | RR = 100% | (222) | |||
PU/EP | G and CNT | 0.8–3 wt % CNT and 2–12 wt % graphene | 0.31 Ω m for 12 wt % graphene | electrical | 100 V for 8% graphene and 3%CNT | 60 | 150 | compressive strain; densification at ε > 70%; 2% permanent deformation after 100 cycles | (223) |
The authors declare no competing financial interest.
References
- Parameswaranpillai J.; Siengchin S.; George J. J.; Jose S.. Shape Memory Polymers, Blends and Composites: Advances and Applications; Advanced Structured Materials; Springer Singapore: 2020; Vol. 115. 10.1007/978-981-13-8574-2. [DOI] [Google Scholar]
- Behl M.; Lendlein A. Actively Moving Polymers. Soft Matter 2007, 3 (1), 58–67. 10.1039/B610611K. [DOI] [PubMed] [Google Scholar]
- Xin X.; Liu L.; Liu Y.; Leng J. Mechanical Models, Structures, and Applications of Shape-Memory Polymers and Their Composites. Acta Mech. Solida Sin. 2019, 32 (5), 535–565. 10.1007/s10338-019-00103-9. [DOI] [Google Scholar]
- Lendlein A.; Behl M.; Hiebl B.; Wischke C. Shape-Memory Polymers as a Technology Platform for Biomedical Applications. Expert Rev. Med. Devices 2010, 7 (3), 357–379. 10.1586/erd.10.8. [DOI] [PubMed] [Google Scholar]
- Mazurek-Budzyńska M.; Razzaq M. Y.; Behl M.; Lendlein A.. Shape-Memory Polymers. In Functional Polymers; Jafar Mazumder M. A., Sheardown H., Al-Ahmed A., Eds.; Springer International: 2019; Polymers and Polymeric Composites: A Reference Series, pp 605–663. 10.1007/978-3-319-95987-0_18. [DOI] [Google Scholar]
- Liu C.; Qin H.; Mather P. T. Review of Progress in Shape-Memory Polymers. J. Mater. Chem. 2007, 17 (16), 1543. 10.1039/b615954k. [DOI] [Google Scholar]
- Rousseau I. A. Challenges of Shape Memory Polymers: A Review of the Progress toward Overcoming SMP’s Limitations. Polym. Eng. Sci. 2008, 48 (11), 2075–2089. 10.1002/pen.21213. [DOI] [Google Scholar]
- Leng J.; Lu H.; Liu Y.; Huang W. M.; Du S. Shape-Memory Polymers—A Class of Novel Smart Materials. MRS Bull. 2009, 34 (11), 848–855. 10.1557/mrs2009.235. [DOI] [Google Scholar]
- Hu J.Introduction to Shape Memory Polymers. In Advances in Shape Memory Polymers; Elsevier: 2013; pp 1–22. 10.1533/9780857098542.1. [DOI] [Google Scholar]
- Safranski D. L.Introduction to Shape-Memory Polymers. In Shape-Memory Polymer Device Design; Elsevier: 2017; pp 1–22. 10.1016/B978-0-323-37797-3.00001-4. [DOI] [Google Scholar]
- Yakacki C. M.; Shandas R.; Safranski D.; Ortega A. M.; Sassaman K.; Gall K. Strong, Tailored, Biocompatible Shape-Memory Polymer Networks. Adv. Funct. Mater. 2008, 18 (16), 2428–2435. 10.1002/adfm.200701049. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hu J.Shape Memory Polymers: Fundamentals, Advances and Applications; Smithers: 2014. [Google Scholar]
- Hager M. D.; Bode S.; Weber C.; Schubert U. S. Shape Memory Polymers: Past, Present and Future Developments. Prog. Polym. Sci. 2015, 49 (50), 3–33. 10.1016/j.progpolymsci.2015.04.002. [DOI] [Google Scholar]
- Huang S.; Kong X.; Xiong Y.; Zhang X.; Chen H.; Jiang W.; Niu Y.; Xu W.; Ren C. An Overview of Dynamic Covalent Bonds in Polymer Material and Their Applications. Eur. Polym. J. 2020, 141, 110094. 10.1016/j.eurpolymj.2020.110094. [DOI] [Google Scholar]
- García F.; Smulders M. M. J. Dynamic Covalent Polymers. J. Polym. Sci., Part A: Polym. Chem. 2016, 54 (22), 3551–3577. 10.1002/pola.28260. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hammer L.; Van Zee N. J.; Nicolaÿ R. Dually Crosslinked Polymer Networks Incorporating Dynamic Covalent Bonds. Polymers 2021, 13 (3), 396. 10.3390/polym13030396. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li Z.; Yu R.; Guo B. Shape-Memory and Self-Healing Polymers Based on Dynamic Covalent Bonds and Dynamic Noncovalent Interactions: Synthesis, Mechanism, and Application. ACS Appl. Bio Mater. 2021, 4 (8), 5926–5943. 10.1021/acsabm.1c00606. [DOI] [PubMed] [Google Scholar]
- Peng S.; Sun Y.; Ma C.; Duan G.; Liu Z.; Ma C. Recent Advances in Dynamic Covalent Bond-Based Shape Memory Polymers. e-Polym. 2022, 22 (1), 285–300. 10.1515/epoly-2022-0032. [DOI] [Google Scholar]
- Koltzenburg S.; Maskos M.; Nuyken O.. Elastomers. In Polymer Chemistry; Springer Berlin Heidelberg: 2017; pp 477–491. 10.1007/978-3-662-49279-6_18. [DOI] [Google Scholar]
- Tobushi H.; Hara H.; Yamada E.; Hayashi S. Thermomechanical Properties in a Thin Film of Shape Memory Polymer of Polyurethane Series. Smart Mater. Struct. 1996, 5 (4), 483–491. 10.1088/0964-1726/5/4/012. [DOI] [Google Scholar]
- Lendlein A.; Langer R. Biodegradable, Elastic Shape-Memory Polymers for Potential Biomedical Applications. Science 2002, 296 (5573), 1673–1676. 10.1126/science.1066102. [DOI] [PubMed] [Google Scholar]
- Miaudet P.; Derré A.; Maugey M.; Zakri C.; Piccione P. M.; Inoubli R.; Poulin P. Shape and Temperature Memory of Nanocomposites with Broadened Glass Transition. Science 2007, 318 (5854), 1294–1296. 10.1126/science.1145593. [DOI] [PubMed] [Google Scholar]
- Ni Q.-Q.; Zhang C.; Fu Y.; Dai G.; Kimura T. Shape Memory Effect and Mechanical Properties of Carbon Nanotube/Shape Memory Polymer Nanocomposites. Compos. Struct. 2007, 81 (2), 176–184. 10.1016/j.compstruct.2006.08.017. [DOI] [Google Scholar]
- Behl M.; Zotzmann J.; Lendlein A. One-Way and Reversible Dual-Shape Effect of Polymer Networks Based on Polypentadecalactone Segments. Int. J. Artif Organs 2011, 34 (2), 231–237. 10.5301/IJAO.2011.6424. [DOI] [PubMed] [Google Scholar]
- Gopinath S.; Adarsh N. N.; Radhakrishnan Nair P.; Mathew S. One-Way Thermo-Responsive Shape Memory Polymer Nanocomposite Derived from Polycaprolactone and Polystyrene-Block-Polybutadiene-Block-Polystyrene Packed with Carbon Nanofiber. Mater. Today Commun. 2020, 22, 100802. 10.1016/j.mtcomm.2019.100802. [DOI] [Google Scholar]
- Leng J.; Lan X.; Liu Y.; Du S. Shape-Memory Polymers and Their Composites: Stimulus Methods and Applications. Prog. Mater. Sci. 2011, 56 (7), 1077–1135. 10.1016/j.pmatsci.2011.03.001. [DOI] [Google Scholar]
- Bothe M.Shape Memory and Actuation Behavior of Semicrystalline Polymer Networks; BAM-Dissertationsreihe; BAM: 2014. [Google Scholar]
- Zare M.; Prabhakaran M. P.; Parvin N.; Ramakrishna S. Thermally-Induced Two-Way Shape Memory Polymers: Mechanisms, Structures, and Applications. Chem. Eng. J. 2019, 374, 706–720. 10.1016/j.cej.2019.05.167. [DOI] [Google Scholar]
- Chung T.; Romo-Uribe A.; Mather P. T. Two-Way Reversible Shape Memory in a Semicrystalline Network. Macromolecules 2008, 41 (1), 184–192. 10.1021/ma071517z. [DOI] [Google Scholar]
- Murcia A. P.; Gomez J. M. U.; Sommer J.-U.; Ionov L. Two-Way Shape Memory Polymers: Evolution of Stress vs Evolution of Elongation. Macromolecules 2021, 54 (12), 5838–5847. 10.1021/acs.macromol.1c00568. [DOI] [Google Scholar]
- Pilate F.; Toncheva A.; Dubois P.; Raquez J.-M. Shape-Memory Polymers for Multiple Applications in the Materials World. Eur. Polym. J. 2016, 80, 268–294. 10.1016/j.eurpolymj.2016.05.004. [DOI] [Google Scholar]
- Hong S. J.; Yu W.-R.; Youk J. H. Two-Way Shape Memory Behavior of Shape Memory Polyurethanes with a Bias Load. Smart Mater. Struct. 2010, 19 (3), 035022. 10.1088/0964-1726/19/3/035022. [DOI] [Google Scholar]
- Westbrook K. K.; Parakh V.; Chung T.; Mather P. T.; Wan L. C.; Dunn M. L.; Qi H. J. Constitutive Modeling of Shape Memory Effects in Semicrystalline Polymers With Stretch Induced Crystallization. J. Eng. Mater. Technol. 2010, 132 (4), 041010. 10.1115/1.4001964. [DOI] [Google Scholar]
- Li J.; Rodgers W. R.; Xie T. Semi-Crystalline Two-Way Shape Memory Elastomer. Polymer 2011, 52 (23), 5320–5325. 10.1016/j.polymer.2011.09.030. [DOI] [Google Scholar]
- Pandini S.; Passera S.; Messori M.; Paderni K.; Toselli M.; Gianoncelli A.; Bontempi E.; Riccò T. Two-Way Reversible Shape Memory Behaviour of Crosslinked Poly(ε-Caprolactone). Polymer 2012, 53 (9), 1915–1924. 10.1016/j.polymer.2012.02.053. [DOI] [Google Scholar]
- Ge Q.; Westbrook K. K.; Mather P. T.; Dunn M. L.; Jerry Qi H. Thermomechanical Behavior of a Two-Way Shape Memory Composite Actuator. Smart Mater. Struct. 2013, 22 (5), 055009. 10.1088/0964-1726/22/5/055009. [DOI] [Google Scholar]
- Huang M.; Dong X.; Wang L.; Zhao J.; Liu G.; Wang D. Two-Way Shape Memory Property and Its Structural Origin of Cross-Linked Poly(ε-Caprolactone). RSC Adv. 2014, 4 (98), 55483–55494. 10.1039/C4RA09385B. [DOI] [Google Scholar]
- Kolesov I.; Dolynchuk O.; Jehnichen D.; Reuter U.; Stamm M.; Radusch H.-J. Changes of Crystal Structure and Morphology during Two-Way Shape-Memory Cycles in Cross-Linked Linear and Short-Chain Branched Polyethylenes. Macromolecules 2015, 48 (13), 4438–4450. 10.1021/acs.macromol.5b00097. [DOI] [Google Scholar]
- Ma L.; Zhao J.; Wang X.; Chen M.; Liang Y.; Wang Z.; Yu Z.; Hedden R. C. Effects of Carbon Black Nanoparticles on Two-Way Reversible Shape Memory in Crosslinked Polyethylene. Polymer 2015, 56, 490–497. 10.1016/j.polymer.2014.11.036. [DOI] [Google Scholar]
- Yang Q.; Fan J.; Li G. Artificial Muscles Made of Chiral Two-Way Shape Memory Polymer Fibers. Appl. Phys. Lett. 2016, 109 (18), 183701. 10.1063/1.4966231. [DOI] [Google Scholar]
- Xie H.; Cheng C.-Y.; Deng X.-Y.; Fan C.-J.; Du L.; Yang K.-K.; Wang Y.-Z. Creating Poly(Tetramethylene Oxide) Glycol-Based Networks with Tunable Two-Way Shape Memory Effects via Temperature-Switched Netpoints. Macromolecules 2017, 50 (13), 5155–5164. 10.1021/acs.macromol.6b02773. [DOI] [Google Scholar]
- Hui J.; Xia H.; Chen H.; Qiu Y.; Fu Y.; Ni Q.-Q. Two-Way Reversible Shape Memory Polymer: Synthesis and Characterization of Benzoyl Peroxide-Crosslinked Poly(Ethylene-Co-Vinyl Acetate). Mater. Lett. 2020, 258, 126762. 10.1016/j.matlet.2019.126762. [DOI] [Google Scholar]
- Raquez J.-M.; Vanderstappen S.; Meyer F.; Verge P.; Alexandre M.; Thomassin J.-M.; Jérôme C.; Dubois P. Design of Cross-Linked Semicrystalline Poly(ε-Caprolactone)-Based Networks with One-Way and Two-Way Shape-Memory Properties through Diels-Alder Reactions. Chem. - Eur. J. 2011, 17 (36), 10135–10143. 10.1002/chem.201100496. [DOI] [PubMed] [Google Scholar]
- Behl M.; Kratz K.; Zotzmann J.; Nöchel U.; Lendlein A. Reversible Bidirectional Shape-Memory Polymers. Adv. Mater. 2013, 25 (32), 4466–4469. 10.1002/adma.201300880. [DOI] [PubMed] [Google Scholar]
- Bothe M.; Pretsch T. Bidirectional Actuation of a Thermoplastic Polyurethane Elastomer. J. Mater. Chem. A 2013, 1 (46), 14491. 10.1039/c3ta13414h. [DOI] [Google Scholar]
- Biswas A.; Aswal V. K.; Sastry P. U.; Rana D.; Maiti P. Reversible Bidirectional Shape Memory Effect in Polyurethanes through Molecular Flipping. Macromolecules 2016, 49 (13), 4889–4897. 10.1021/acs.macromol.6b00536. [DOI] [Google Scholar]
- Wang K.; Jia Y.-G.; Zhu X. X. Two-Way Reversible Shape Memory Polymers Made of Cross-Linked Cocrystallizable Random Copolymers with Tunable Actuation Temperatures. Macromolecules 2017, 50 (21), 8570–8579. 10.1021/acs.macromol.7b01815. [DOI] [Google Scholar]
- Xu Z.; Ding C.; Wei D.-W.; Bao R.-Y.; Ke K.; Liu Z.; Yang M.-B.; Yang W. Electro and Light-Active Actuators Based on Reversible Shape-Memory Polymer Composites with Segregated Conductive Networks. ACS Appl. Mater. Interfaces 2019, 11 (33), 30332–30340. 10.1021/acsami.9b10386. [DOI] [PubMed] [Google Scholar]
- Burke K. A.; Mather P. T. Soft Shape Memory in Main-Chain Liquid Crystalline Elastomers. J. Mater. Chem. 2010, 20 (17), 3449. 10.1039/b924050k. [DOI] [Google Scholar]
- Wang Y.; Huang X.; Zhang J.; Bi M.; Zhang J.; Niu H.; Li C.; Yu H.; Wang B.; Jiang H. Two-Step Crosslinked Liquid-Crystalline Elastomer with Reversible Two-Way Shape Memory Characteristics. Mol. Cryst. Liq. Cryst. 2017, 650 (1), 13–22. 10.1080/15421406.2017.1318025. [DOI] [Google Scholar]
- Chen S.; Hu J.; Zhuo H.; Zhu Y. Two-Way Shape Memory Effect in Polymer Laminates. Mater. Lett. 2008, 62 (25), 4088–4090. 10.1016/j.matlet.2008.05.073. [DOI] [Google Scholar]
- Tobushi H.; Hayashi S.; Sugimoto Y.; Date K. Two-Way Bending Properties of Shape Memory Composite with SMA and SMP. Materials 2009, 2 (3), 1180–1192. 10.3390/ma2031180. [DOI] [Google Scholar]
- Wang K.; Jia Y.-G.; Zhao C.; Zhu X. X. Multiple and Two-Way Reversible Shape Memory Polymers: Design Strategies and Applications. Prog. Mater. Sci. 2019, 105, 100572. 10.1016/j.pmatsci.2019.100572. [DOI] [Google Scholar]
- Xia Y.; He Y.; Zhang F.; Liu Y.; Leng J. A Review of Shape Memory Polymers and Composites: Mechanisms, Materials, and Applications. Adv. Mater. 2021, 33 (6), 2000713. 10.1002/adma.202000713. [DOI] [PubMed] [Google Scholar]
- Dolog R.; Weiss R. A. Shape Memory Behavior of a Polyethylene-Based Carboxylate Ionomer. Macromolecules 2013, 46 (19), 7845–7852. 10.1021/ma401631j. [DOI] [Google Scholar]
- Wang K.; Jia Y.-G.; Zhu X. X. Biocompound-Based Multiple Shape Memory Polymers Reinforced by Photo-Cross-Linking. ACS Biomater. Sci. Eng. 2015, 1 (9), 855–863. 10.1021/acsbiomaterials.5b00226. [DOI] [PubMed] [Google Scholar]
- Dai L.; Song J.; Qu S.; Xiao R. Triple-Shape Memory Effect in 3D-Printed Polymers. Express Polym. Lett. 2020, 14 (12), 1116–1126. 10.3144/expresspolymlett.2020.91. [DOI] [Google Scholar]
- Pretsch T. Triple-Shape Properties of a Thermoresponsive Poly(Ester Urethane). Smart Mater. Struct. 2010, 19 (1), 015006. 10.1088/0964-1726/19/1/015006. [DOI] [Google Scholar]
- Zotzmann J.; Behl M.; Feng Y.; Lendlein A. Copolymer Networks Based on Poly(ω-Pentadecalactone) and Poly(ϵ-Caprolactone)Segments as a Versatile Triple-Shape Polymer System. Adv. Funct. Mater. 2010, 20 (20), 3583–3594. 10.1002/adfm.201000478. [DOI] [Google Scholar]
- Chatani S.; Wang C.; Podgórski M.; Bowman C. N. Triple Shape Memory Materials Incorporating Two Distinct Polymer Networks Formed by Selective Thiol–Michael Addition Reactions. Macromolecules 2014, 47 (15), 4949–4954. 10.1021/ma501028a. [DOI] [Google Scholar]
- Xie T.; Xiao X.; Cheng Y.-T. Revealing Triple-Shape Memory Effect by Polymer Bilayers: Revealing Triple-Shape Memory Effect by Polymer Bilayers. Macromol. Rapid Commun. 2009, 30 (21), 1823–1827. 10.1002/marc.200900409. [DOI] [PubMed] [Google Scholar]
- Tobushi H.; Hayashi S.; Pieczyska E.; Date K.; Nishimura Y. Three-Way Shape Memory Composite Actuator. MSF 2011, 674, 225–230. 10.4028/www.scientific.net/MSF.674.225. [DOI] [Google Scholar]
- Scalet G. Two-Way and Multiple-Way Shape Memory Polymers for Soft Robotics: An Overview. Actuators 2020, 9 (1), 10. 10.3390/act9010010. [DOI] [Google Scholar]
- Huang Y. N.; Fan L. F.; Rong M. Z.; Zhang M. Q.; Gao Y. M. External Stress-Free Reversible Multiple Shape Memory Polymers. ACS Appl. Mater. Interfaces 2019, 11 (34), 31346–31355. 10.1021/acsami.9b10052. [DOI] [PubMed] [Google Scholar]
- Li J.; Liu T.; Xia S.; Pan Y.; Zheng Z.; Ding X.; Peng Y. A Versatile Approach to Achieve Quintuple-Shape Memory Effect by Semi-Interpenetrating Polymer Networks Containing Broadened Glass Transition and Crystalline Segments. J. Mater. Chem. 2011, 21 (33), 12213. 10.1039/c1jm12496j. [DOI] [Google Scholar]
- Shao Y.; Lavigueur C.; Zhu X. X. Multishape Memory Effect of Norbornene-Based Copolymers with Cholic Acid Pendant Groups. Macromolecules 2012, 45 (4), 1924–1930. 10.1021/ma202506b. [DOI] [Google Scholar]
- Kong D.; Xiao X. High Cycle-Life Shape Memory Polymer at High Temperature. Sci. Rep 2016, 6 (1), 33610. 10.1038/srep33610. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rodinò S.; Curcio E. M.; Renzo D. A.; Sgambitterra E.; Magarò P.; Furgiuele F.; Brandizzi M.; Maletta C. Shape Memory Alloy—Polymer Composites: Static and Fatigue Pullout Strength under Thermo-Mechanical Loading. Materials 2022, 15 (9), 3216. 10.3390/ma15093216. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lu L.; Li G. One-Way Multishape-Memory Effect and Tunable Two-Way Shape Memory Effect of Ionomer Poly(Ethylene- Co -Methacrylic Acid). ACS Appl. Mater. Interfaces 2016, 8 (23), 14812–14823. 10.1021/acsami.6b04105. [DOI] [PubMed] [Google Scholar]
- Cotin G.; Perton F.; Blanco-Andujar C.; Pichon B.; Mertz D.; Bégin-Colin S.. Design of Anisotropic Iron-Oxide-Based Nanoparticles for Magnetic Hyperthermia. In Nanomaterials for Magnetic and Optical Hyperthermia Applications; Elsevier: 2019; pp 41–60. 10.1016/B978-0-12-813928-8.00002-8. [DOI] [Google Scholar]
- Li Q.; Kartikowati C. W.; Horie S.; Ogi T.; Iwaki T.; Okuyama K. Correlation between Particle Size/Domain Structure and Magnetic Properties of Highly Crystalline Fe3O4 Nanoparticles. Sci. Rep 2017, 7 (1), 9894. 10.1038/s41598-017-09897-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yani A.; Kurniawan C.; Djuhana D.. Investigation of the Ground State Domain Structure Transition on Magnetite (Fe3O4); Bali: 2018; p 020020. 10.1063/1.5064017. [DOI] [Google Scholar]
- Soto G. D.; Meiorin C.; Actis D. G.; Mendoza Zélis P.; Moscoso Londoño O.; Muraca D.; Mosiewicki M. A.; Marcovich N. E. Magnetic Nanocomposites Based on Shape Memory Polyurethanes. Eur. Polym. J. 2018, 109, 8–15. 10.1016/j.eurpolymj.2018.08.046. [DOI] [Google Scholar]
- Meng Q.; Hu J. A Review of Shape Memory Polymer Composites and Blends. Composites, Part A 2009, 40 (11), 1661–1672. 10.1016/j.compositesa.2009.08.011. [DOI] [Google Scholar]
- Cheng X.; Chen Y.; Dai S.; Bilek M. M. M.; Bao S.; Ye L. Bending Shape Memory Behaviours of Carbon Fibre Reinforced Polyurethane-Type Shape Memory Polymer Composites under Relatively Small Deformation: Characterisation and Computational Simulation. J. Mech. Behav. Biomed. Mater. 2019, 100, 103372. 10.1016/j.jmbbm.2019.103372. [DOI] [PubMed] [Google Scholar]
- Cuevas J. M.; Rubio R.; Laza J. M.; Vilas J. L.; Rodriguez M.; M León L. Shape Memory Composites Based on Glass-Fibre-Reinforced Poly(Ethylene)-like Polymers. Smart Mater. Struct. 2012, 21 (3), 035004. 10.1088/0964-1726/21/3/035004. [DOI] [Google Scholar]
- Korotkov R.; Vedernikov A.; Gusev S.; Alajarmeh O.; Akhatov I.; Safonov A. Shape Memory Behavior of Unidirectional Pultruded Laminate. Composites, Part A 2021, 150, 106609. 10.1016/j.compositesa.2021.106609. [DOI] [Google Scholar]
- Guo J.; Lv H.; Wang Z.; Tang X.; Liang W.; Zhang S. Thermo-Mechanical Property of Shape Memory Polymer/Carbon Fibre Composites. Mater. Res. Innovations 2015, 19 (sup8), S8-566–S8-572. 10.1179/1432891715Z.0000000001750. [DOI] [Google Scholar]
- Wang Z.; Liu J.; Guo J.; Sun X.; Xu L. The Study of Thermal, Mechanical and Shape Memory Properties of Chopped Carbon Fiber-Reinforced TPI Shape Memory Polymer Composites. Polymers 2017, 9 (11), 594. 10.3390/polym9110594. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pulla S. S.Development of Multifunctional Shape Memory Polymer and Shape Memory Polymer Composites; University of Kentucky: 2015. [Google Scholar]
- Wang Y.; Chen Z.; Niu J.; Shi Y.; Zhao J.; Ye J.; Tian W. Electrically Responsive Shape Memory Composites Using Silver Plated Chopped Carbon Fiber. Front. Chem. 2020, 8, 322. 10.3389/fchem.2020.00322. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Qi Y.; Sun B.; Gu B.; Zhang W. Electrothermally Actuated Properties of Fabric-Reinforced Shape Memory Polymer Composites Based on Core–Shell Yarn. Compos. Struct. 2022, 292, 115681. 10.1016/j.compstruct.2022.115681. [DOI] [Google Scholar]
- Margoy D.; Gouzman I.; Grossman E.; Bolker A.; Eliaz N.; Verker R. Epoxy-Based Shape Memory Composite for Space Applications. Acta Astronautica 2021, 178, 908–919. 10.1016/j.actaastro.2020.08.026. [DOI] [Google Scholar]
- Meng H.; Li G. A Review of Stimuli-Responsive Shape Memory Polymer Composites. Polymer 2013, 54 (9), 2199–2221. 10.1016/j.polymer.2013.02.023. [DOI] [Google Scholar]
- Le H. H.; Osazuwa O.; Kolesov I.; Ilisch S.; Radusch H.-J. Influence of Carbon Black Properties on the Joule Heating Stimulated Shape-Memory Behavior of Filledethylene-1-Octene Copolymer. Polym. Eng. Sci. 2011, 51 (3), 500–508. 10.1002/pen.21818. [DOI] [Google Scholar]
- Arun D. I.; Santhosh Kumar K. S.; Satheesh Kumar B.; Chakravarthy P.; Dona M.; Santhosh B. High Glass-Transition Polyurethane-Carbon Black Electro-Active Shape Memory Nanocomposite for Aerospace Systems. Mater. Sci. Technol. 2019, 35 (5), 596–605. 10.1080/02670836.2019.1575054. [DOI] [Google Scholar]
- González-Jiménez A.; Bernal-Ortega P.; Salamanca F. M.; Valentin J. L. Shape-Memory Composites Based on Ionic Elastomers. Polymers 2022, 14 (6), 1230. 10.3390/polym14061230. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang X.; Zhao J.; Chen M.; Ma L.; Zhao X.; Dang Z.-M.; Wang Z. Improved Self-Healing of Polyethylene/Carbon Black Nanocomposites by Their Shape Memory Effect. J. Phys. Chem. B 2013, 117 (5), 1467–1474. 10.1021/jp3098796. [DOI] [PubMed] [Google Scholar]
- Gunes I. S.; Jimenez G. A.; Jana S. C. Carbonaceous Fillers for Shape Memory Actuation of Polyurethane Composites by Resistive Heating. Carbon 2009, 47 (4), 981–997. 10.1016/j.carbon.2008.11.053. [DOI] [Google Scholar]
- Tekay E. Preparation and Characterization of Electro-Active Shape Memory PCL/SEBS-g-MA/MWCNT Nanocomposites. Polymer 2020, 209, 122989. 10.1016/j.polymer.2020.122989. [DOI] [Google Scholar]
- Pötschke P.; Villmow T.; Krause B. Melt Mixed PCL/MWCNT Composites Prepared at Different Rotation Speeds: Characterization of Rheological, Thermal, and Electrical Properties, Molecular Weight, MWCNT Macrodispersion, and MWCNT Length Distribution. Polymer 2013, 54 (12), 3071–3078. 10.1016/j.polymer.2013.04.012. [DOI] [Google Scholar]
- Jung Y. C.; Yoo H. J.; Kim Y. A.; Cho J. W.; Endo M. Electroactive Shape Memory Performance of Polyurethane Composite Having Homogeneously Dispersed and Covalently Crosslinked Carbon Nanotubes. Carbon 2010, 48 (5), 1598–1603. 10.1016/j.carbon.2009.12.058. [DOI] [Google Scholar]
- Jin Yoo H.; Chae Jung Y.; Gopal Sahoo N.; Whan Cho J. Polyurethane-Carbon Nanotube Nanocomposites Prepared by In-Situ Polymerization with Electroactive Shape Memory. J. Macromol. Sci., Part B: Phys. 2006, 45 (4), 441–451. 10.1080/00222340600767471. [DOI] [Google Scholar]
- Raja M.; Ryu S. H.; Shanmugharaj A. M. Influence of Surface Modified Multiwalled Carbon Nanotubes on the Mechanical and Electroactive Shape Memory Properties of Polyurethane (PU)/Poly(Vinylidene Diflouride) (PVDF) Composites. Colloids Surf., A 2014, 450, 59–66. 10.1016/j.colsurfa.2014.03.008. [DOI] [Google Scholar]
- Lu H.; Gou J. Fabrication and Electroactive Responsive Behavior of Shape-Memory Nanocomposite Incorporated with Self-Assembled Multiwalled Carbon Nanotube Nanopaper: Shape-Memory Nanocomposite. Polym. Adv. Technol. 2012, 23 (12), 1529–1535. 10.1002/pat.2074. [DOI] [Google Scholar]
- Lu H.; Liu Y.; Gou J.; Leng J.; Du S. Surface Coating of Multi-Walled Carbon Nanotube Nanopaper on Shape-Memory Polymer for Multifunctionalization. Compos. Sci. Technol. 2011, 71 (11), 1427–1434. 10.1016/j.compscitech.2011.05.016. [DOI] [Google Scholar]
- Slobodian P.; Riha P.; Olejnik R.; Matyas J. Accelerated Shape Forming and Recovering, Induction, and Release of Adhesiveness of Conductive Carbon Nanotube/Epoxy Composites by Joule Heating. Polymers 2020, 12 (5), 1030. 10.3390/polym12051030. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Datta S.; Henry T. C.; Sliozberg Y. R.; Lawrence B. D.; Chattopadhyay A.; Hall A. J. Carbon Nanotube Enhanced Shape Memory Epoxy for Improved Mechanical Properties and Electroactive Shape Recovery. Polymer 2021, 212, 123158. 10.1016/j.polymer.2020.123158. [DOI] [Google Scholar]
- Wang X.; Sparkman J.; Gou J. Electrical Actuation and Shape Memory Behavior of Polyurethane Composites Incorporated with Printed Carbon Nanotube Layers. Compos. Sci. Technol. 2017, 141, 8–15. 10.1016/j.compscitech.2017.01.002. [DOI] [Google Scholar]
- Pereira Sánchez C.; Houbben M.; Fagnard J.-F.; Laurent P.; Jérôme C.; Noels L.; Vanderbemden P. Resistive Heating of a Shape Memory Composite: Analytical, Numerical and Experimental Study. Smart Mater. Struct. 2022, 31 (2), 025003. 10.1088/1361-665X/ac3ebd. [DOI] [Google Scholar]
- Pereira Sánchez C.; Houbben M.; Fagnard J.-F.; Harmeling P.; Jérôme C.; Noels L.; Vanderbemden P. Experimental Characterization of the Thermo-Electro-Mechanical Properties of a Shape Memory Composite during Electric Activation. Smart Mater. Struct. 2022, 31 (9), 095029. 10.1088/1361-665X/ac8297. [DOI] [Google Scholar]
- Cortés A.; Pérez-Chao N.; Jiménez-Suárez A.; Campo M.; Prolongo S. G. Sequential and Selective Shape Memory by Remote Electrical Control. Eur. Polym. J. 2022, 164, 110888. 10.1016/j.eurpolymj.2021.110888. [DOI] [Google Scholar]
- Cortés A.; Aguilar J. L.; Cosola A.; Fernández Sanchez-Romate X. X.; Jiménez-Suárez A.; Sangermano M.; Campo M.; Prolongo S. G. 4D-Printed Resins and Nanocomposites Thermally Stimulated by Conventional Heating and IR Radiation. ACS Appl. Polym. Mater. 2021, 3 (10), 5207–5215. 10.1021/acsapm.1c00970. [DOI] [Google Scholar]
- Xiao W.-X.; Fan C.-J.; Li B.; Liu W.-X.; Yang K.-K.; Wang Y.-Z. Single-Walled Carbon Nanotubes as Adaptable One-Dimensional Crosslinker to Bridge Multi-Responsive Shape Memory Network via π–π Stacking. Compos. Commun. 2019, 14, 48–54. 10.1016/j.coco.2019.05.006. [DOI] [Google Scholar]
- Park C.; Wilkinson J.; Banda S.; Ounaies Z.; Wise K. E.; Sauti G.; Lillehei P. T.; Harrison J. S. Aligned Single-Wall Carbon Nanotube Polymer Composites Using an Electric Field. J. Polym. Sci. B Polym. Phys. 2006, 44 (12), 1751–1762. 10.1002/polb.20823. [DOI] [Google Scholar]
- Martin C. A.; Sandler J. K. W.; Windle A. H.; Schwarz M.-K.; Bauhofer W.; Schulte K.; Shaffer M. S. P. Electric Field-Induced Aligned Multi-Wall Carbon Nanotube Networks in Epoxy Composites. Polymer 2005, 46 (3), 877–886. 10.1016/j.polymer.2004.11.081. [DOI] [Google Scholar]
- Yu K.; Zhang Z.; Liu Y.; Leng J. Carbon Nanotube Chains in a Shape Memory Polymer/Carbon Black Composite: To Significantly Reduce the Electrical Resistivity. Appl. Phys. Lett. 2011, 98 (7), 074102. 10.1063/1.3556621. [DOI] [Google Scholar]
- Sanchez-Garcia M. D.; Lagaron J. M.; Hoa S. V. Effect of Addition of Carbon Nanofibers and Carbon Nanotubes on Properties of Thermoplastic Biopolymers. Compos. Sci. Technol. 2010, 70 (7), 1095–1105. 10.1016/j.compscitech.2010.02.015. [DOI] [Google Scholar]
- Gong X.; Liu L.; Liu Y.; Leng J. An Electrical-Heating and Self-Sensing Shape Memory Polymer Composite Incorporated with Carbon Fiber Felt. Smart Mater. Struct. 2016, 25 (3), 035036. 10.1088/0964-1726/25/3/035036. [DOI] [Google Scholar]
- Wang W.; Liu D.; Liu Y.; Leng J.; Bhattacharyya D. Electrical Actuation Properties of Reduced Graphene Oxide Paper/Epoxy-Based Shape Memory Composites. Compos. Sci. Technol. 2015, 106, 20–24. 10.1016/j.compscitech.2014.10.016. [DOI] [Google Scholar]
- Jiu H.; Jiao H.; Zhang L.; Zhang S.; Zhao Y. Graphene-Crosslinked Two-Way Reversible Shape Memory Polyurethane Nanocomposites with Enhanced Mechanical and Electrical Properties. J. Mater. Sci: Mater. Electron 2016, 27 (10), 10720–10728. 10.1007/s10854-016-5173-2. [DOI] [Google Scholar]
- Martin-Gallego M.; Verdejo R.; Lopez-Manchado M. A.; Sangermano M. Epoxy-Graphene UV-Cured Nanocomposites. Polymer 2011, 52 (21), 4664–4669. 10.1016/j.polymer.2011.08.039. [DOI] [Google Scholar]
- Kim J. T.; Kim B. K.; Kim E. Y.; Park H. C.; Jeong H. M. Synthesis and Shape Memory Performance of Polyurethane/Graphene Nanocomposites. React. Funct. Polym. 2014, 74, 16–21. 10.1016/j.reactfunctpolym.2013.10.004. [DOI] [Google Scholar]
- Kim J. T.; Jeong H. J.; Park H. C.; Jeong H. M.; Bae S. Y.; Kim B. K. Electroactive Shape Memory Performance of Polyurethane/Graphene Nanocomposites. React. Funct. Polym. 2015, 88, 1–7. 10.1016/j.reactfunctpolym.2015.01.004. [DOI] [Google Scholar]
- Sabzi M.; Babaahmadi M.; Samadi N.; Mahdavinia G. R.; Keramati M.; Nikfarjam N. Graphene Network Enabled High Speed Electrical Actuation of Shape Memory Nanocomposite Based on Poly(Vinyl Acetate): Shape Memory Nanocomposites of Poly(Vinyl Acetate). Polym. Int. 2017, 66 (5), 665–671. 10.1002/pi.5303. [DOI] [Google Scholar]
- Thinh P. X.; Basavajara C.; Kim J. K.; Huh D. S. Characterization and Electrical Properties of Honeycomb-Patterned Poly(ε-Caprolactone)/Reduced Graphene Oxide Composite Film. Polym. Compos 2012, 33 (12), 2159–2168. 10.1002/pc.22357. [DOI] [Google Scholar]
- Chen Y.-F.; Tan Y.-J.; Li J.; Hao Y.-B.; Shi Y.-D.; Wang M. Graphene Oxide-Assisted Dispersion of Multi-Walled Carbon Nanotubes in Biodegradable Poly(ε-Caprolactone) for Mechanical and Electrically Conductive Enhancement. Polym. Test. 2018, 65, 387–397. 10.1016/j.polymertesting.2017.12.019. [DOI] [Google Scholar]
- Kang S.; Kang T.-H.; Kim B. S.; Oh J.; Park S.; Choi I. S.; Lee J.; Son J. G. 2D Reentrant Micro-Honeycomb Structure of Graphene-CNT in Polyurethane: High Stretchability, Superior Electrical/Thermal Conductivity, and Improved Shape Memory Properties. Composites, Part B 2019, 162, 580–588. 10.1016/j.compositesb.2019.01.004. [DOI] [Google Scholar]
- Su X.; Wang R.; Li X.; Araby S.; Kuan H.-C.; Naeem M.; Ma J. A Comparative Study of Polymer Nanocomposites Containing Multi-Walled Carbon Nanotubes and Graphene Nanoplatelets. Nano Mater. Sci. 2022, 4, 185–204. 10.1016/j.nanoms.2021.08.003. [DOI] [Google Scholar]
- Chen L.; Li W.; Liu Y.; Leng J. Nanocomposites of Epoxy-Based Shape Memory Polymer and Thermally Reduced Graphite Oxide: Mechanical, Thermal and Shape Memory Characterizations. Composites, Part B 2016, 91, 75–82. 10.1016/j.compositesb.2016.01.019. [DOI] [Google Scholar]
- Zhang Y.; Hu J.; Zhu S.; Qin T.; Ji F. A “Trampoline” Nanocomposite: Tuning the Interlayer Spacing in Graphene Oxide/Polyurethane to Achieve Coalesced Mechanical and Memory Properties. Compos. Sci. Technol. 2019, 180, 14–22. 10.1016/j.compscitech.2019.04.033. [DOI] [Google Scholar]
- Kausar A. Shape Memory Polymer/Graphene Nanocomposites: State-of-the-Art. e-Polym. 2022, 22 (1), 165–181. 10.1515/epoly-2022-0024. [DOI] [Google Scholar]
- van Vilsteren S. J. M.; Yarmand H.; Ghodrat S. Review of Magnetic Shape Memory Polymers and Magnetic Soft Materials. Magnetochemistry 2021, 7 (9), 123. 10.3390/magnetochemistry7090123. [DOI] [Google Scholar]
- Gopinath S.; Adarsh N. N.; Nair P.; Mathew S. Nano-metal Oxide Fillers in Thermo-responsive Polycaprolactone-based Polymer Nanocomposites Smart Materials: Impact on Thermo-mechanical, and Shape Memory Properties. J. Vinyl Addit Technol. 2021, 27 (4), 768–780. 10.1002/vnl.21848. [DOI] [Google Scholar]
- Aaltio I.; Nilsén F.; Lehtonen J.; Ge Y. L.; Spoljaric S.; Seppälä J.; Hannula S. P. Magnetic Shape Memory - Polymer Hybrids. MSF 2016, 879, 133–138. 10.4028/www.scientific.net/MSF.879.133. [DOI] [Google Scholar]
- Zhang D. W.; Liu Y. J.; Leng J. S. Magnetic Field Activation of Thermoresponsive Shape-Memory Polymer with Embedded Micron Sized Ni Powder. AMR 2010, 123–125, 995–998. 10.4028/www.scientific.net/AMR.123-125.995. [DOI] [Google Scholar]
- Buckley P. R.; McKinley G. H.; Wilson T. S.; Small W.; Benett W. J.; Bearinger J. P.; McElfresh M. W.; Maitland D. J. Inductively Heated Shape Memory Polymer for the Magnetic Actuation of Medical Devices. IEEE Trans. Biomed. Eng. 2006, 53 (10), 2075–2083. 10.1109/TBME.2006.877113. [DOI] [PubMed] [Google Scholar]
- Bayerl T.Application of Particulate Susceptors for the Inductive Heating of Temperature Sensitive Polymer-Polymer Composites, Als Ms. gedr.; IVW-Schriftenreihe; Univ. Inst. für Verbundwerkstoffe: 2012. [Google Scholar]
- Deatsch A. E.; Evans B. A. Heating Efficiency in Magnetic Nanoparticle Hyperthermia. J. Magn. Magn. Mater. 2014, 354, 163–172. 10.1016/j.jmmm.2013.11.006. [DOI] [Google Scholar]
- Magaye R.; Zhao J.; Bowman L.; Ding M. Genotoxicity and Carcinogenicity of Cobalt-, Nickel- and Copper-Based Nanoparticles. Exp. Ther. Med. 2012, 4 (4), 551–561. 10.3892/etm.2012.656. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Suwanwatana W.; Yarlagadda S.; Gillespie J. W. Influence of Particle Size on Hysteresis Heating Behavior of Nickel Particulate Polymer Films. Compos. Sci. Technol. 2006, 66 (15), 2825–2836. 10.1016/j.compscitech.2006.02.033. [DOI] [Google Scholar]
- Zhang X.; Lu X.; Wang Z.; Wang J.; Sun Z. Biodegradable Shape Memory Nanocomposites with Thermal and Magnetic Field Responsiveness. J. Biomater. Sci., Polym. Ed. 2013, 24 (9), 1057–1070. 10.1080/09205063.2012.735098. [DOI] [PubMed] [Google Scholar]
- Puig J.; Hoppe C. E.; Fasce L. A.; Pérez C. J.; Piñeiro-Redondo Y.; Bañobre-López M.; López-Quintela M. A.; Rivas J.; Williams R. J. J. Superparamagnetic Nanocomposites Based on the Dispersion of Oleic Acid-Stabilized Magnetite Nanoparticles in a Diglycidylether of Bisphenol A-Based Epoxy Matrix: Magnetic Hyperthermia and Shape Memory. J. Phys. Chem. C 2012, 116 (24), 13421–13428. 10.1021/jp3026754. [DOI] [Google Scholar]
- Shahdan D.; Flaifel M. H.; Ahmad S. H.; Chen R. S.; Razak J. A. Enhanced Magnetic Nanoparticles Dispersion Effect on the Behaviour of Ultrasonication-Assisted Compounding Processing of PLA/LNR/NiZn Nanocomposites. Journal of Materials Research and Technology 2021, 15, 5988–6000. 10.1016/j.jmrt.2021.11.046. [DOI] [Google Scholar]
- Mohr R.; Kratz K.; Weigel T.; Lucka-Gabor M.; Moneke M.; Lendlein A. Initiation of Shape-Memory Effect by Inductive Heating of Magnetic Nanoparticles in Thermoplastic Polymers. Proc. Natl. Acad. Sci. U.S.A. 2006, 103 (10), 3540–3545. 10.1073/pnas.0600079103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Weigel T.; Mohr R.; Lendlein A. Investigation of Parameters to Achieve Temperatures Required to Initiate the Shape-Memory Effect of Magnetic Nanocomposites by Inductive Heating. Smart Mater. Struct. 2009, 18 (2), 025011. 10.1088/0964-1726/18/2/025011. [DOI] [Google Scholar]
- He Z.; Satarkar N.; Xie T.; Cheng Y.-T.; Hilt J. Z. Remote Controlled Multishape Polymer Nanocomposites with Selective Radiofrequency Actuations. Adv. Mater. 2011, 23 (28), 3192–3196. 10.1002/adma.201100646. [DOI] [PubMed] [Google Scholar]
- Lipert K.; Ritschel M.; Leonhardt A.; Krupskaya Y.; Büchner B.; Klingeler R. Magnetic Properties of Carbon Nanotubes with and without Catalyst. J. Phys.: Conf. Ser. 2010, 200 (7), 072061. 10.1088/1742-6596/200/7/072061. [DOI] [Google Scholar]
- Kletetschka G.; Inoue Y.; Lindauer J.; Hůlka Z. Magnetic Tunneling with CNT-Based Metamaterial. Sci. Rep 2019, 9 (1), 2551. 10.1038/s41598-019-39325-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Han M.; Deng L.. High Frequency Properties of Carbon Nanotubes and Their Electromagnetic Wave Absorption Properties. In Carbon Nanotubes Applications on Electron Devices; Marulanda J. M., Ed.; InTech: 2011. 10.5772/16629. [DOI] [Google Scholar]
- Ze Q.; Kuang X.; Wu S.; Wong J.; Montgomery S. M.; Zhang R.; Kovitz J. M.; Yang F.; Qi H. J.; Zhao R. Magnetic Shape Memory Polymers with Integrated Multifunctional Shape Manipulation. Adv. Mater. 2020, 32 (4), 1906657. 10.1002/adma.201906657. [DOI] [PubMed] [Google Scholar]
- Leng J. S.; Lan X.; Liu Y. J.; Du S. Y.; Huang W. M.; Liu N.; Phee S. J.; Yuan Q. Electrical Conductivity of Thermoresponsive Shape-Memory Polymer with Embedded Micron Sized Ni Powder Chains. Appl. Phys. Lett. 2008, 92 (1), 014104. 10.1063/1.2829388. [DOI] [Google Scholar]
- Mishra S. R.; Tracy J. B. Sequential Actuation of Shape-Memory Polymers through Wavelength-Selective Photothermal Heating of Gold Nanospheres and Nanorods. ACS Appl. Nano Mater. 2018, 1 (7), 3063–3067. 10.1021/acsanm.8b00394. [DOI] [Google Scholar]
- Manikandan M.; Hasan N.; Wu H.-F. Platinum Nanoparticles for the Photothermal Treatment of Neuro 2A Cancer Cells. Biomaterials 2013, 34 (23), 5833–5842. 10.1016/j.biomaterials.2013.03.077. [DOI] [PubMed] [Google Scholar]
- Toncheva A.; Khelifa F.; Paint Y.; Voué M.; Lambert P.; Dubois P.; Raquez J.-M. Fast IR-Actuated Shape-Memory Polymers Using in Situ Silver Nanoparticle-Grafted Cellulose Nanocrystals. ACS Appl. Mater. Interfaces 2018, 10 (35), 29933–29942. 10.1021/acsami.8b10159. [DOI] [PubMed] [Google Scholar]
- Bai Y.; Zhang J.; Wen D.; Yuan B.; Gong P.; Liu J.; Chen X. Fabrication of Remote Controllable Devices with Multistage Responsiveness Based on a NIR Light-Induced Shape Memory Ionomer Containing Various Bridge Ions. J. Mater. Chem. A 2019, 7 (36), 20723–20732. 10.1039/C9TA05329H. [DOI] [Google Scholar]
- Sahoo N.; Jung Y.; Yoo H.; Cho J. Influence of Carbon Nanotubes and Polypyrrole on the Thermal, Mechanical and Electroactive Shape-Memory Properties of Polyurethane Nanocomposites. Compos. Sci. Technol. 2007, 67 (9), 1920–1929. 10.1016/j.compscitech.2006.10.013. [DOI] [Google Scholar]
- Chen J.; Sun D.; Gu T.; Qi X.; Yang J.; Lei Y.; Wang Y. Photo-Induced Shape Memory Blend Composites with Remote Selective Self-Healing Performance Enabled by Polypyrrole Nanoparticles. Compos. Sci. Technol. 2022, 217, 109123. 10.1016/j.compscitech.2021.109123. [DOI] [Google Scholar]
- Cortes P.; Terzak J.; Kubas G.; Phillips D.; Baur J. W. The Morphing Properties of a Vascular Shape Memory Composite. Smart Mater. Struct. 2014, 23 (1), 015018. 10.1088/0964-1726/23/1/015018. [DOI] [Google Scholar]
- Du H.; Song Z.; Wang J.; Liang Z.; Shen Y.; You F. Microwave-Induced Shape-Memory Effect of Silicon Carbide/Poly(Vinyl Alcohol) Composite. Sensors and Actuators A: Physical 2015, 228, 1–8. 10.1016/j.sna.2015.01.012. [DOI] [Google Scholar]
- Hasan S. M.; Thompson R. S.; Emery H.; Nathan A. L.; Weems A. C.; Zhou F.; Monroe M. B. B.; Maitland D. J. Modification of Shape Memory Polymer Foams Using Tungsten, Aluminum Oxide, and Silicon Dioxide Nanoparticles. RSC Adv. 2016, 6 (2), 918–927. 10.1039/C5RA22633C. [DOI] [PMC free article] [PubMed] [Google Scholar]
- An Y.; Okuzaki H. Novel Electro-Active Shape Memory Polymers for Soft Actuators. Jpn. J. Appl. Phys. 2020, 59 (6), 061002. 10.35848/1347-4065/ab8e08. [DOI] [Google Scholar]
- Wang X.; Lan J.; Wu P.; Zhang J. Liquid Metal Based Electrical Driven Shape Memory Polymers. Polymer 2021, 212, 123174. 10.1016/j.polymer.2020.123174. [DOI] [Google Scholar]
- Garces I. T.; Aslanzadeh S.; Boluk Y.; Ayranci C. Cellulose Nanocrystals (CNC) Reinforced Shape Memory Polyurethane Ribbons for Future Biomedical Applications and Design. Journal of Thermoplastic Composite Materials 2020, 33 (3), 377–392. 10.1177/0892705718806334. [DOI] [Google Scholar]
- Lin C.; Lv J.; Li Y.; Zhang F.; Li J.; Liu Y.; Liu L.; Leng J. 4D-Printed Biodegradable and Remotely Controllable Shape Memory Occlusion Devices. Adv. Funct. Mater. 2019, 29 (51), 1906569. 10.1002/adfm.201906569. [DOI] [Google Scholar]
- Zhao W.; Zhang F.; Leng J.; Liu Y. Personalized 4D Printing of Bioinspired Tracheal Scaffold Concept Based on Magnetic Stimulated Shape Memory Composites. Compos. Sci. Technol. 2019, 184, 107866. 10.1016/j.compscitech.2019.107866. [DOI] [Google Scholar]
- Zhang F.; Wen N.; Wang L.; Bai Y.; Leng J. Design of 4D Printed Shape-Changing Tracheal Stent and Remote Controlling Actuation. International Journal of Smart and Nano Materials 2021, 12 (4), 375–389. 10.1080/19475411.2021.1974972. [DOI] [Google Scholar]
- Zhao W.; Huang Z.; Liu L.; Wang W.; Leng J.; Liu Y. Porous Bone Tissue Scaffold Concept Based on Shape Memory PLA/Fe3O4. Compos. Sci. Technol. 2021, 203, 108563. 10.1016/j.compscitech.2020.108563. [DOI] [Google Scholar]
- Yamagishi K.; Nojiri A.; Iwase E.; Hashimoto M. Syringe-Injectable, Self-Expandable, and Ultraconformable Magnetic Ultrathin Films. ACS Appl. Mater. Interfaces 2019, 11 (44), 41770–41779. 10.1021/acsami.9b17567. [DOI] [PubMed] [Google Scholar]
- Babaie A.; Rezaei M.; Razzaghi D.; Roghani-Mamaqani H. Synthesis of dual-stimuli-responsive Polyurethane Shape Memory Nanocomposites Incorporating isocyanate-functionalized Fe 3 O 4 Nanoparticles. J. Appl. Polym. Sci. 2022, 139 (33), e52790. 10.1002/app.52790. [DOI] [Google Scholar]
- Wei H.; Zhang Q.; Yao Y.; Liu L.; Liu Y.; Leng J. Direct-Write Fabrication of 4D Active Shape-Changing Structures Based on a Shape Memory Polymer and Its Nanocomposite. ACS Appl. Mater. Interfaces 2017, 9 (1), 876–883. 10.1021/acsami.6b12824. [DOI] [PubMed] [Google Scholar]
- Pineda-Castillo S. A.; Luo J.; Lee H.; Bohnstedt B. N.; Liu Y.; Lee C.-H. Effects of Carbon Nanotube Infiltration on a Shape Memory Polymer-Based Device for Brain Aneurysm Therapeutics: Design and Characterization of a Joule-Heating Triggering Mechanism. Adv. Eng. Mater. 2021, 23 (6), 2100322. 10.1002/adem.202100322. [DOI] [Google Scholar]
- Wang J.; Luo J.; Kunkel R.; Saha M.; Bohnstedt B. N.; Lee C.-H.; Liu Y. Development of Shape Memory Polymer Nanocomposite Foam for Treatment of Intracranial Aneurysms. Mater. Lett. 2019, 250, 38–41. 10.1016/j.matlet.2019.04.112. [DOI] [Google Scholar]
- Paik I. H.; Goo N. S.; Jung Y. C.; Cho J. W. Development and Application of Conducting Shape Memory Polyurethane Actuators. Smart Mater. Struct. 2006, 15 (5), 1476–1482. 10.1088/0964-1726/15/5/037. [DOI] [Google Scholar]
- Kim N.-G.; Han M.-W.; Iakovleva A.; Park H.-B.; Chu W.-S.; Ahn S.-H. Hybrid Composite Actuator with Shape Retention Capability for Morphing Flap of Unmanned Aerial Vehicle (UAV). Compos. Struct. 2020, 243, 112227. 10.1016/j.compstruct.2020.112227. [DOI] [Google Scholar]
- Yang S.; He Y.; Liu Y.; Leng J. Non-Contact Magnetic Actuated Shape-Programmable Poly(Aryl Ether Ketone)s and Their Structural Variation during the Deformation Process. Smart Mater. Struct. 2022, 31 (3), 035035. 10.1088/1361-665X/ac4ff7. [DOI] [Google Scholar]
- Cohn D.; Zarek M.; Elyashiv A.; Sbitan M. A.; Sharma V.; Ramanujan R. V. Remotely Triggered Morphing Behavior of Additively Manufactured Thermoset Polymer-Magnetic Nanoparticle Composite Structures. Smart Mater. Struct. 2021, 30 (4), 045022. 10.1088/1361-665X/abeaeb. [DOI] [Google Scholar]
- Peng Q.; Wei H.; Qin Y.; Lin Z.; Zhao X.; Xu F.; Leng J.; He X.; Cao A.; Li Y. Shape-Memory Polymer Nanocomposites with a 3D Conductive Network for Bidirectional Actuation and Locomotion Application. Nanoscale 2016, 8 (42), 18042–18049. 10.1039/C6NR06515E. [DOI] [PubMed] [Google Scholar]
- Xu Z.; Wei D.-W.; Bao R.-Y.; Wang Y.; Ke K.; Yang M.-B.; Yang W. Self-Sensing Actuators Based on a Stiffness Variable Reversible Shape Memory Polymer Enabled by a Phase Change Material. ACS Appl. Mater. Interfaces 2022, 14 (19), 22521–22530. 10.1021/acsami.2c07119. [DOI] [PubMed] [Google Scholar]
- Xu L.; Li Z.; Lu H.; Qi X.; Dong Y.; Dai H.; Islam Md Z.; Fu Y.; Ni Q. Electrothermally-Driven Elongating-Contracting Film Actuators Based on Two-Way Shape Memory Carbon Nanotube/Ethylene-Vinyl Acetate Composites. Adv. Materials Technologies 2022, 7 (7), 2101229. 10.1002/admt.202101229. [DOI] [Google Scholar]
- Hu T.; Xuan S.; Gao Y.; Shu Q.; Xu Z.; Sun S.; Li J.; Gong X. Smart Refreshable Braille Display Device Based on Magneto-Resistive Composite with Triple Shape Memory. Adv. Materials Technologies 2022, 7 (1), 2100777. 10.1002/admt.202100777. [DOI] [Google Scholar]
- Dong X.; Zhang F.; Wang L.; Liu Y.; Leng J. 4D Printing of Electroactive Shape-Changing Composite Structures and Their Programmable Behaviors. Composites, Part A 2022, 157, 106925. 10.1016/j.compositesa.2022.106925. [DOI] [Google Scholar]
- Ren D.; Chen Y.; Li H.; Rehman H. U.; Cai Y.; Liu H. High-Efficiency Dual-Responsive Shape Memory Assisted Self-Healing of Carbon Nanotubes Enhanced Polycaprolactone/Thermoplastic Polyurethane Composites. Colloids Surf., A 2019, 580, 123731. 10.1016/j.colsurfa.2019.123731. [DOI] [Google Scholar]
- Orellana J.; Moreno-Villoslada I.; Bose R. K.; Picchioni F.; Flores M. E.; Araya-Hermosilla R. Self-Healing Polymer Nanocomposite Materials by Joule Effect. Polymers 2021, 13 (4), 649. 10.3390/polym13040649. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cerdan K.; Moya C.; Van Puyvelde P.; Bruylants G.; Brancart J. Magnetic Self-Healing Composites: Synthesis and Applications. Molecules 2022, 27 (12), 3796. 10.3390/molecules27123796. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Melly S. K.; Liu L.; Liu Y.; Leng J. Active Composites Based on Shape Memory Polymers: Overview, Fabrication Methods, Applications, and Future Prospects. J. Mater. Sci. 2020, 55 (25), 10975–11051. 10.1007/s10853-020-04761-w. [DOI] [Google Scholar]
- Chakraborty D. D.; Chakraborty P.. Shape-Memory Polymer Composites and Their Applications. In Smart Polymer Nanocomposites; Elsevier: 2021; pp 103–115. 10.1016/B978-0-12-819961-9.00016-5. [DOI] [Google Scholar]
- Cai Y.; Jiang J.-S.; Liu Z.-W.; Zeng Y.; Zhang W.-G. Magnetically-Sensitive Shape Memory Polyurethane Composites Crosslinked with Multi-Walled Carbon Nanotubes. Composites, Part A 2013, 53, 16–23. 10.1016/j.compositesa.2013.05.016. [DOI] [Google Scholar]
- Chen S.; Yang S.; Li Z.; Xu S.; Yuan H.; Chen S.; Ge Z. Electroactive Two-Way Shape Memory Polymer Laminates. Polym. Compos. 2015, 36 (3), 439–444. 10.1002/pc.22958. [DOI] [Google Scholar]
- Cho J. W.; Kim J. W.; Jung Y. C.; Goo N. S. Electroactive Shape-Memory Polyurethane Composites Incorporating Carbon Nanotubes. Macromol. Rapid Commun. 2005, 26 (5), 412–416. 10.1002/marc.200400492. [DOI] [Google Scholar]
- D’Elia E.; Ahmed H. S.; Feilden E.; Saiz E. Electrically-Responsive Graphene-Based Shape-Memory Composites. Applied Materials Today 2019, 15, 185–191. 10.1016/j.apmt.2018.12.018. [DOI] [Google Scholar]
- Dong K.; Panahi-Sarmad M.; Cui Z.; Huang X.; Xiao X. Electro-Induced Shape Memory Effect of 4D Printed Auxetic Composite Using PLA/TPU/CNT Filament Embedded Synergistically with Continuous Carbon Fiber: A Theoretical & Experimental Analysis. Composites, Part B 2021, 220, 108994. 10.1016/j.compositesb.2021.108994. [DOI] [Google Scholar]
- Du F.-P.; Ye E.-Z.; Yang W.; Shen T.-H.; Tang C.-Y.; Xie X.-L.; Zhou X.-P.; Law W.-C. Electroactive Shape Memory Polymer Based on Optimized Multi-Walled Carbon Nanotubes/Polyvinyl Alcohol Nanocomposites. Composites, Part B 2015, 68, 170–175. 10.1016/j.compositesb.2014.08.043. [DOI] [Google Scholar]
- Gong T.; Li W.; Chen H.; Wang L.; Shao S.; Zhou S. Remotely Actuated Shape Memory Effect of Electrospun Composite Nanofibers. Acta Biomaterialia 2012, 8 (3), 1248–1259. 10.1016/j.actbio.2011.12.006. [DOI] [PubMed] [Google Scholar]
- Gu S.-Y.; Jin S.-P.; Gao X.-F.; Mu J. Polylactide-Based Polyurethane Shape Memory Nanocomposites (Fe 3 O 4 /PLAUs) with Fast Magnetic Responsiveness. Smart Mater. Struct. 2016, 25 (5), 055036. 10.1088/0964-1726/25/5/055036. [DOI] [Google Scholar]
- Huang C.-L.; He M.-J.; Huo M.; Du L.; Zhan C.; Fan C.-J.; Yang K.-K.; Chin I.-J.; Wang Y.-Z. A Facile Method to Produce PBS-PEG/CNTs Nanocomposites with Controllable Electro-Induced Shape Memory Effect. Polym. Chem. 2013, 4 (14), 3987. 10.1039/c3py00461a. [DOI] [Google Scholar]
- Jimenez G. A.; Jana S. C. Composites of Carbon Nanofibers and Thermoplastic Polyurethanes with Shape-Memory Properties Prepared by Chaotic Mixing. Polym. Eng. Sci. 2009, 49 (10), 2020–2030. 10.1002/pen.21442. [DOI] [Google Scholar]
- Lee S.-H.; Jung J.-H.; Oh I.-K. 3D Networked Graphene-Ferromagnetic Hybrids for Fast Shape Memory Polymers with Enhanced Mechanical Stiffness and Thermal Conductivity. Small 2014, 10 (19), 3880–3886. 10.1002/smll.201400624. [DOI] [PubMed] [Google Scholar]
- Leng J. S.; Huang W. M.; Lan X.; Liu Y. J.; Du S. Y. Significantly Reducing Electrical Resistivity by Forming Conductive Ni Chains in a Polyurethane Shape-Memory Polymer/Carbon-Black Composite. Appl. Phys. Lett. 2008, 92 (20), 204101. 10.1063/1.2931049. [DOI] [Google Scholar]
- Liu X.; Li H.; Zeng Q.; Zhang Y.; Kang H.; Duan H.; Guo Y.; Liu H. Electro-Active Shape Memory Composites Enhanced by Flexible Carbon Nanotube/Graphene Aerogels. J. Mater. Chem. A 2015, 3 (21), 11641–11649. 10.1039/C5TA02490K. [DOI] [Google Scholar]
- Lotfi Mayan Sofla R.; Rezaei M.; Babaie A.; Nasiri M. Preparation of Electroactive Shape Memory Polyurethane/Graphene Nanocomposites and Investigation of Relationship between Rheology, Morphology and Electrical Properties. Composites, Part B 2019, 175, 107090. 10.1016/j.compositesb.2019.107090. [DOI] [Google Scholar]
- Lu H.; Liang F.; Gou J.; Leng J.; Du S. Synergistic Effect of Ag Nanoparticle-Decorated Graphene Oxide and Carbon Fiber on Electrical Actuation of Polymeric Shape Memory Nanocomposites. Smart Mater. Struct. 2014, 23 (8), 085034. 10.1088/0964-1726/23/8/085034. [DOI] [Google Scholar]
- Luo X.; Mather P. T. Conductive Shape Memory Nanocomposites for High Speed Electrical Actuation. Soft Matter 2010, 6 (10), 2146. 10.1039/c001295e. [DOI] [Google Scholar]
- Mahapatra S. S.; Yadav S. K.; Yoo H. J.; Ramasamy M. S.; Cho J. W. Tailored and Strong Electro-Responsive Shape Memory Actuation in Carbon Nanotube-Reinforced Hyperbranched Polyurethane Composites. Sens. Actuators, B 2014, 193, 384–390. 10.1016/j.snb.2013.12.006. [DOI] [Google Scholar]
- Orozco F.; Kaveh M.; Santosa D. S.; Lima G. M. R.; Gomes D. R.; Pei Y.; Araya-Hermosilla R.; Moreno-Villoslada I.; Picchioni F.; Bose R. K. Electroactive Self-Healing Shape Memory Polymer Composites Based on Diels–Alder Chemistry. ACS Appl. Polym. Mater. 2021, 3 (12), 6147–6156. 10.1021/acsapm.1c00999. [DOI] [Google Scholar]
- Park J.; Dao T.; Lee H.; Jeong H.; Kim B. Properties of Graphene/Shape Memory Thermoplastic Polyurethane Composites Actuating by Various Methods. Materials 2014, 7 (3), 1520–1538. 10.3390/ma7031520. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Qi X.; Dong P.; Liu Z.; Liu T.; Fu Q. Selective Localization of Multi-Walled Carbon Nanotubes in Bi-Component Biodegradable Polyester Blend for Rapid Electroactive Shape Memory Performance. Compos. Sci. Technol. 2016, 125, 38–46. 10.1016/j.compscitech.2016.01.023. [DOI] [Google Scholar]
- Qi X.; Xiu H.; Wei Y.; Zhou Y.; Guo Y.; Huang R.; Bai H.; Fu Q. Enhanced Shape Memory Property of Polylactide/Thermoplastic Poly(Ether)Urethane Composites via Carbon Black Self-Networking Induced Co-Continuous Structure. Compos. Sci. Technol. 2017, 139, 8–16. 10.1016/j.compscitech.2016.12.007. [DOI] [Google Scholar]
- Qian C.; Zhu Y.; Dong Y.; Fu Y. Vapor-Grown Carbon Nanofiber/Poly(Ethylene-Co-Vinyl Acetate) Composites with Electrical-Active Two-Way Shape Memory Behavior. J. Intell. Mater. Syst. Struct. 2017, 28 (19), 2749–2756. 10.1177/1045389X17698246. [DOI] [Google Scholar]
- Raja M.; Ryu S. H.; Shanmugharaj A. M. Thermal, Mechanical and Electroactive Shape Memory Properties of Polyurethane (PU)/Poly (Lactic Acid) (PLA)/CNT Nanocomposites. Eur. Polym. J. 2013, 49 (11), 3492–3500. 10.1016/j.eurpolymj.2013.08.009. [DOI] [Google Scholar]
- Razzaq M. Y.; Anhalt M.; Frormann L.; Weidenfeller B. Thermal, Electrical and Magnetic Studies of Magnetite Filled Polyurethane Shape Memory Polymers. Materials Science and Engineering: A 2007, 444 (1–2), 227–235. 10.1016/j.msea.2006.08.083. [DOI] [Google Scholar]
- Razzaq M. Y.; Behl M.; Lendlein A. Magnetic Memory Effect of Nanocomposites. Adv. Funct. Mater. 2012, 22 (1), 184–191. 10.1002/adfm.201101590. [DOI] [Google Scholar]
- Razzaq M. Y.; Behl M.; Nöchel U.; Lendlein A. Magnetically Controlled Shape-Memory Effects of Hybrid Nanocomposites from Oligo(ω-Pentadecalactone) and Covalently Integrated Magnetite Nanoparticles. Polymer 2014, 55 (23), 5953–5960. 10.1016/j.polymer.2014.07.025. [DOI] [Google Scholar]
- Rogers N.; Khan F. Characterization of Deformation Induced Changes to Conductivity in an Electrically Triggered Shape Memory Polymer. Polym. Test. 2013, 32 (1), 71–77. 10.1016/j.polymertesting.2012.10.001. [DOI] [Google Scholar]
- Sahoo N. G.; Jung Y. C.; Goo N. S.; Cho J. W. Conducting Shape Memory Polyurethane-Polypyrrole Composites for an Electroactive Actuator. Macromol. Mater. Eng. 2005, 290 (11), 1049–1055. 10.1002/mame.200500211. [DOI] [Google Scholar]
- Schmidt A. M. Electromagnetic Activation of Shape Memory Polymer Networks Containing Magnetic Nanoparticles. Macromol. Rapid Commun. 2006, 27 (14), 1168–1172. 10.1002/marc.200600225. [DOI] [Google Scholar]
- Shao L.; Dai J.; Zhang Z.; Yang J.; Zhang N.; Huang T.; Wang Y. Thermal and Electroactive Shape Memory Behaviors of Poly(l -Lactide)/Thermoplastic Polyurethane Blend Induced by Carbon Nanotubes. RSC Adv. 2015, 5 (123), 101455–101465. 10.1039/C5RA20632D. [DOI] [Google Scholar]
- Tang Z.; Sun D.; Yang D.; Guo B.; Zhang L.; Jia D. Vapor Grown Carbon Nanofiber Reinforced Bio-Based Polyester for Electroactive Shape Memory Performance. Compos. Sci. Technol. 2013, 75, 15–21. 10.1016/j.compscitech.2012.11.019. [DOI] [Google Scholar]
- Valentini L.; Cardinali M.; Kenny J. Hot Press Transferring of Graphene Nanoplatelets on Polyurethane Block Copolymers Film for Electroactive Shape Memory Devices. J. Polym. Sci., Part B: Polym. Phys. 2014, 52 (16), 1100–1106. 10.1002/polb.23539. [DOI] [Google Scholar]
- Wang Z.; Zhao J.; Chen M.; Yang M.; Tang L.; Dang Z.-M.; Chen F.; Huang M.; Dong X. Dually Actuated Triple Shape Memory Polymers of Cross-Linked Polycyclooctene–Carbon Nanotube/Polyethylene Nanocomposites. ACS Appl. Mater. Interfaces 2014, 6 (22), 20051–20059. 10.1021/am5056307. [DOI] [PubMed] [Google Scholar]
- Wang Y.; Zhu G.; Cui X.; Liu T.; Liu Z.; Wang K. Electroactive Shape Memory Effect of Radiation Cross-Linked SBS/LLDPE Composites Filled with Carbon Black. Colloid Polym. Sci. 2014, 292 (9), 2311–2317. 10.1007/s00396-014-3266-0. [DOI] [Google Scholar]
- Wang W.; Liu Y.; Leng J. Recent Developments in Shape Memory Polymer Nanocomposites: Actuation Methods and Mechanisms. Coord. Chem. Rev. 2016, 320–321, 38–52. 10.1016/j.ccr.2016.03.007. [DOI] [Google Scholar]
- Wang Y.; Ma T.; Tian W.; Ye J.; Wang X.; Jiang X. Electroactive Shape Memory Properties of Graphene/Epoxy-Cyanate Ester Nanocomposites. PRT 2018, 47 (1), 72–78. 10.1108/PRT-04-2017-0037. [DOI] [Google Scholar]
- Wei K.; Zhu G.; Tang Y.; Li X.; Liu T. The Effects of Carbon Nanotubes on Electroactive Shape-Memory Behaviors of Hydro-Epoxy/Carbon Black Composite. Smart Mater. Struct. 2012, 21 (8), 085016. 10.1088/0964-1726/21/8/085016. [DOI] [Google Scholar]
- Wei Y.; Huang R.; Dong P.; Qi X.-D.; Fu Q. Preparation of Polylactide/Poly(Ether)Urethane Blends with Excellent Electro-Actuated Shape Memory via Incorporating Carbon Black and Carbon Nanotubes Hybrids Fillers. Chin J. Polym. Sci. 2018, 36 (10), 1175–1186. 10.1007/s10118-018-2138-3. [DOI] [Google Scholar]
- Xia S.; Li X.; Wang Y.; Pan Y.; Zheng Z.; Ding X.; Peng Y. A Remote-Activated Shape Memory Polymer Network Employing Vinyl-Capped Fe 3 O 4 Nanoparticles as Netpoints for Durable Performance. Smart Mater. Struct. 2014, 23 (8), 085005. 10.1088/0964-1726/23/8/085005. [DOI] [Google Scholar]
- Xiao Y.; Zhou S.; Wang L.; Gong T. Electro-Active Shape Memory Properties of Poly(ε-Caprolactone)/Functionalized Multiwalled Carbon Nanotube Nanocomposite. ACS Appl. Mater. Interfaces 2010, 2 (12), 3506–3514. 10.1021/am100692n. [DOI] [PubMed] [Google Scholar]
- Yakacki C. M.; Satarkar N. S.; Gall K.; Likos R.; Hilt J. Z. Shape-Memory Polymer Networks with Fe 3 O 4 Nanoparticles for Remote Activation. J. Appl. Polym. Sci. 2009, 112 (5), 3166–3176. 10.1002/app.29845. [DOI] [Google Scholar]
- Zhang F. H.; Zhang Z. C.; Luo C. J.; Lin I.-T.; Liu Y.; Leng J.; Smoukov S. K. Remote, Fast Actuation of Programmable Multiple Shape Memory Composites by Magnetic Fields. J. Mater. Chem. C 2015, 3 (43), 11290–11293. 10.1039/C5TC02464A. [DOI] [Google Scholar]
- Zhang Z.; Dou J.; He J.; Xiao C.; Shen L.; Yang J.; Wang Y.; Zhou Z. Electrically/Infrared Actuated Shape Memory Composites Based on a Bio-Based Polyester Blend and Graphene Nanoplatelets and Their Excellent Self-Driven Ability. J. Mater. Chem. C 2017, 5 (17), 4145–4158. 10.1039/C7TC00828G. [DOI] [Google Scholar]
- Zhang F.; Xia Y.; Wang L.; Liu L.; Liu Y.; Leng J. Conductive Shape Memory Microfiber Membranes with Core–Shell Structures and Electroactive Performance. ACS Appl. Mater. Interfaces 2018, 10 (41), 35526–35532. 10.1021/acsami.8b12743. [DOI] [PubMed] [Google Scholar]
- Zhang X.-J.; Yang Q.-S.; Shang J.-J.; Liu X.; Leng J. The Shape Memory Properties of Multi-Layer Graphene Reinforced Poly(L-Lactide-Co-ϵ-Caprolactone) by an Atomistic Investigation. Smart Mater. Struct. 2021, 30 (5), 055005. 10.1088/1361-665X/abeef9. [DOI] [Google Scholar]
- Zhou J.; Li H.; Liu W.; Dugnani R.; Tian R.; Xue W.; Chen Y.; Guo Y.; Duan H.; Liu H. A Facile Method to Fabricate Polyurethane Based Graphene Foams/Epoxy/Carbon Nanotubes Composite for Electro-Active Shape Memory Application. Composites, Part A 2016, 91, 292–300. 10.1016/j.compositesa.2016.10.021. [DOI] [Google Scholar]