Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Mechanical study of perovskite solar cells: opportunities and challenges for wearable power source

Open Access Open Access

Abstract

We provide a review of current understandings of mechanical properties and fracture behaviors of perovskites that are essential for flexible and stretchable solar cell (SC) applications. We first review the mechanical failure modes in perovskites. We further discuss the underlying mechanisms of mechanical failure and its impact on device degradation in flexible perovskite solar cells (PSCs). Then, we examine the strategies to mitigate these mechanical issues in flexible PSCs. Lastly, we assess the elevated challenges and present recommendations for future research directions to advance the technology towards a fully stretchable and wearable energy source.

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

Corrections

Letian Dou, Xiwen Gong, Sergey Makarov, Barry Rand, and Yongsheng Zhao, "Halide Perovskites for Photonics and Optoelectronics feature issue: publisher’s note," Opt. Mater. Express 12, 1009-1009 (2022)
https://opg.optica.org/ome/abstract.cfm?uri=ome-12-3-1009

3 February 2022: A typographical correction was made to the article title.

1. Introduction

Metal halide perovskites (MHPs) have emerged as promising photovoltaic (PV) materials due to their excellent optical and electrical properties and the ease of manufacturing. Within the MHP community, improving power conversion efficiency (PCE), operational stability, and the replacement of lead (Pb) have been the main topics of interest. Recently, mechanical stability and flexibility of perovskites have attracted increasing attentions. Compared to rigid solar cells (SCs), flexible solar cells (fSCs) can minimize the damage caused by the vibration and turbulence during the handling and transportation process, which accounts for 6% of the device damage in the conventional rigid SCs [1]. The lightweight nature of perovskite solar cells (PSCs) enables their installation to remote areas by reducing the cost of transportation and installation compared to the conventional Si-based SCs. Moreover, enhancing mechanical conformability of SCs will generate new pathways to harvest solar energy beyond solar panels on the roof or in the dessert. For instance, fSCs can harvest solar energy on non-planar and dynamic structures by serving as power source for wearable electronics, vehicles, and small personal gadgets. In addition, fSCs also enable fast and large-area roll-to-roll manufacturing, which reduces the cost and allows the scaling-up on an industry relevant level.

There have been significant efforts in understanding the mechanical properties of perovskite thin films. In this article, we first review the mechanical properties of MHPs and present the most commonly observed mechanical failure modes in PSCs: crack and delamination. We further review solutions to address these issues: 1) reducing the grain boundaries, 2) grain boundary crosslinking 2) interfacial engineering, 4) pre-strained substrate, and 5) neutral plane strategy.

Moving towards the era of wearable technologies, there are increasing needs to further improve the mechanical deformability in electronics, to achieve versatility and compatibility with existing technology in energy sources [2,3]. Beyond flexibility, stretchability needs to be introduced in perovskites to develop next-generation wearable SCs.

2. Mechanical properties, common failures modes in MHPs and solutions to improve their flexibility

2.1 Fundamental mechanical properties of MHPs

MHPs feature low exciton binding energy (37 - 50 meV) [4,5], shallow trap states [6], high charge carrier mobility [7], and long charge diffusion length [8]. These superior optoelectronic properties make MHPs ideal candidates for PV applications. In addition, the solution processability of perovskites promises exciting applications in flexible and portable power sources for self-powered electronics.

To achieve flexibility in PSCs, significant efforts have been devoted to understanding their mechanical properties. Young’s modulus(E) is one of the most commonly reported indicators to evaluate the stiffness and mechanical compatibility of materials for flexible and wearable applications. Several researchers reported E value of MHPs based on experimental and theoretical studies. For instance, Rakita et al. reported an E of 14.3 GPa and 19.6 GPa for MAPbI3 and MAPbBr3, respectively, using the nano-indentation method [9]. Other studies on the mechanical properties of MAPbI3 also measured the E value with nano-indentation giving a range of values (Table 1) depending on dwell time and lattice plane [10,11,12,13]. Feng investigated the elastic properties and anisotropy of MABX3 (B = Pb, Sn; X = Br, I) using first-principles calculations based on density functional theory (DFT) [14]. Their research suggests that the elastic properties of perovskite depend on the chemical bonds between B and X. The calculated E of cubic phase MHPs ranges from 15 to 37 GPa (with E of MAPbI3 = 22.2 GPa). Faghihnasiri et al. also reported similar results (E of MAPbI3 = 22.8 GPa) [15] taking into account the interactions between MA and PBX3. The E of MHPs is in between that of rigid inorganic semiconductors (> 50 GPa) and flexible organic semiconductors (< 5 GPa) (Table 1).

Tables Icon

Table 1. Mechanical Properties of Metal Halide Perovskite and other Photovoltaic Materials

Critical strain energy release rate (GC) is another important mechanical property which evaluates the fracture toughness of materials. Strain energy release rate (G) is defined as the amount of strain energy released by generating one unit of crack. Once the strain energy release rate reaches the critical value (GC), crack will propagate. Hence, the higher value of GC indicate that it is more energetically costly for a crack to propagate, i.e. more resistive to crack propagation. MHP thin films show comparable GC to organic semiconductors, which is significantly lower than that of silicon or CIGS thin films (Table 1) [16]. Therefore, it is more energetically favorable for MHP and organic semiconductors to form cracks than other inorganic semiconductors. In part 2.2, we will discuss in detail that using E or ${G_C}$ alone is insufficient to evaluate the flexibility of materials. Instead, the value of (E/GC)1/2 governs the critical (maximum possible) bending radius of materials, which better reflects the intrinsic flexibility of materials. Thus, the value of (E/GC)1/2 should be considered more often in the future research of flexible and stretchable electronics.

Overall, MHPs are regarded as promising candidates for fSCs applications in view of their reduced rigidity compared to inorganic semiconductors, excellent optoelectronic properties, solution processability, and low cost. However, in order to achieve the same level of flexibility such as in organic fSCs while maintaining the outstanding optoelectronic properties of perovskites, new material design and device engineering strategies are needed.

2.2 Effect of mechanical deformation in perovskite thin films

2.2.1 Strain-induced structural and optoelectronic property changes in perovskites

Flexible PSCs often operate under external strain, and thus the effect of strain needs to be considered carefully to develop high-performance and stable PSCs. Recent work has revealed that the strain in MHP can modulate their optoelectronic properties. For example, tensile strain increases the bandgap of MHP, while compressive strain decreases it [32]. Such change of the bandgap is ascribed to the different sensitivity of valence band and conduction band to the change of strength of Pb-X bond induced by the lattice expansion. In addition, strain also modulates the carrier mobility: compressive strain leads to the increased hole mobility due to the decrease of the hole effective mass [33].

Strain also negatively impacts the stability of MHP films. For example, tensile strain decreases the activation energy for ion migration, which accelerates the degradation of perovskite into PbI2. Conversely, compressive strain increases the activation energy thus suppressing such degradation [34]. Strain also affects the phase stability of perovskites. For example, residual tensile strain in FAPbI3 drives the photo-active α phase into photo-inactive δ phase at room temperature [33]. In the case of CsPbI3, tensile strain stabilizes the photoactive black γ-phase, preventing its transition to the photo-inactive yellow δ-phase [35].

Due to the importance of strain on MHPs, several groups have conducted thorough review on strain analysis and engineering in MHPs [3638]. In the following discussion, we will instead focus on an important yet currently underexplored area: the fracture mechanics of PSCs and its impact on the photovoltaic performance.

2.2.2 Mechanical failure modes in perovskites

Due to the rigidity of MHP materials, mechanical failures are often present when rendering PSCs in flexible forms. Crack formation is one of the most commonly encountered mechanical failures in flexible PSCs (Fig. 1(b)) [3942]. In general, cracks tend to initiate at 1) the grain boundaries of polycrystalline MHP, 2) within the grains, and 3) the interfaces between MHP/carrier transporting layers.

 figure: Fig. 1.

Fig. 1. Schematic illustration of mechanical failure modes in MHP thin films. (a) MHP thin film on PET without external stress(bending). (b) Crack formed on MHP thin film upon bending. (c) Crack sites make MHP film susceptible to moisture-induced degradation. (d) Diffused moisture degrades MHP into PbI2. (e) Hole transport layer materials penetrating through crack sites, leading to undesirable charge carrier recombination as indicated by the red dashed circle. The blue dots and white dots represent electrons and holes respectively. (f) Delamination results in the loss of electrical contact. Charge carriers fail to transport between the layers, leading to increased series resistance and device failure. Delamination also makes the film susceptible to degradation by generating pathways for moisture.

Download Full Size | PDF

Dai et al. reported that grain boundaries are the primary locations of crack formation [41]. They found that MHPs with smaller grain size have grain boundaries parallel to the substrate. Using the double cantilever beam test on MHP film, they observed that cracks form along the grain boundaries, i.e. horizontally. In comparison, in MHP films with larger grain sizes, the horizontal grain boundaries are mostly absent. In this case, cracks tend to form at the next weakest sites: the MHP/electron transport layer interface. In addition, cracks can also form within the grain of MHP. Yadavalli et al. reported that upon e-beam exposure, α-FAPbI3 films experience crack formation primarily within each grain [39]. Lee et al. also observed cracks formation within the grain in MAPbI3 film during the bending test [40].

Interfacial delamination is another major cause of mechanical failures in flexible PSCs. Dong et al. observed interfacial delamination between MHP film and ITO/PET substrates post cyclic bend tests, in addition to the vertical crack formation [42]. They further confirmed that the gap caused by delamination and crack allows water and oxygen molecules to permeate and interact with perovskites, which accelerates the degradation of the MHP films (Fig. 1(c)).

2.2.3 Mechanistic studies of crack formation and interfacial delamination of perovskite thin films

Several in-depth studies have been carried out to unveil the mechanism of the mechanical failure in flexible PSCs. Dong et al. elucidated the effect of bending on crack formation in MHP films based on fracture mechanics studies [42]. Upon bending, thin film undergoes a combination of tensile stress at the outer layer and compressive stress at the inner layer of the film. When the film is under convex bending, the maximum tensile stress (σT) is given by Eq. (1) [39]:

$${\sigma _T} = \frac{{Eh}}{{2R}}$$
where E is Young’s modulus of the MHP, h is the thickness of the substrate (when thickness of MHP is relatively small it is neglected), and R is the bending radius. The tensile stress drives the formation of fractures in MHP films. Using Eq. (2), one can predict the critical tensile stress when cracks start to form [43]:
$$K = \psi {\sigma _T}c_0^{0.5} \ge {K_{IC}} = \sqrt {{G_C}E}$$
where K is the stress intensity factor, ψ is the geometrical constant (∼π0.5), c0 is the half length of the incipient crack (in the case of MHP films, grain-boundary grooves and facets are considered as incipient crack), KIC (MPa m0.5) is the critical stress intensity factor of MHP for mode I fracture, Gc (J/m2) is the critical strain energy release rate. When this criterion is met, i.e., σT ≥ (GCE)0.5/ψc00.5, an array of vertical channel cracks will form parallel to the bending line.

In addition to the bending stress, residual stress(σR) that accumulated during the MHP crystallization also leads to the crack formation [44]. Specifically, the large difference of the thermal expansion coefficients between MHP and substrate generates significant tensile residual stress post annealing [10]. Furthermore, the lattice mismatch between the MHP film and substrate materials also results in the concentrated stress in the MHP film, which leads to crack formation [36]. Therefore, the net tensile stress should be considered in Eq. (2) and can be written as σnet = σT + σR (σR > 0 for tensile stress, σR < 0 for compressive stress).

Substituting ${\sigma _T}$ in Eq. (2) with Eq. (1), we can obtain the critical bending radius (Rc) at which the cracks start to propagate with Eq. (3). It should be noted that Rc is the most commonly reported parameter to quantify the flexibility of thin films. Smaller Rc corresponds to greater flexibility.

$$R \le \sqrt {\frac{E}{{{G_C}}}} \frac{{h\psi c_0^{0.5}}}{2} = {R_C}$$
From Eq. (3), it is clear that the critical bending radius of thin film is governed by the term (E/GC)1/2: as the value of (E/GC)1/2 gets smaller, thin films can be bent to a smaller radius without forming cracks. It should be noted that we assumed certain thickness and initial crack size in Eq. (3), thus the value of (E/GC)1/2 reflects the intrinsic properties of material and is independent to material thickness or initial crack size.

From Table 1, we can see that organic semiconductors exhibit lower values of (E/GC)1/2, while brittle inorganic materials such as silicon, CIGS, and ITO show higher values. MHP is comparable to those of brittle materials ((E/GC)1/2 value of 2.98 × 104.5 m-0.5). However, PSCs have demonstrated enhanced flexibility compared to the conventional SCs. The increased flexibility (smaller RC) of PSCs can be attributed to their thinner film thickness (lower h), due to their high absorption coefficient. I.e., PSCs can realize the comparable power conversion efficiencies with thinner absorber layers than Si SCs [45,46].

Since we assume the brittle fracture condition, the model described by Eq. (3) does not precisely predict the critical bending radius of materials that go through ductile fracture, such as elastomers, metals, and organic semiconductors. However, the value of (E/GC)1/2 reflects the intrinsic flexibility of a broad range of materials of interest. In Fig. 2, we presented the value of the inverse of (E/GC)1/2 of photovoltaic relevant materials: semiconductors, conductors, and substrate materials that are commonly employed in SCs. The increasing value of (GC/E)1/2 generally agrees with the higher flexibility and mechanical ductility of corresponding materials. MAPbI3 is at the higher end of the region of inorganic crystals, presenting both exciting opportunities and challenges for its applications in fSCs.

 figure: Fig. 2.

Fig. 2. Chart of (GC/E)1/2 value for photovoltaic materials (semiconductors, metals, and elastomers). Upper right direction indicates higher (Gc/E)1/2 and thus lower Rc and higher flexibility, which applies to elastomers and metals. The bottom left area features materials with lower (GC/E)1/2 and thus higher Rc and rigidity, which applies to inorganic crystals. See Supplementary Material 1 for numerical values.

Download Full Size | PDF

In general, the value of (E/GC)1/2 can be considered as an empirical guide in search for new flexible and stretchable electronic materials. In the past, E has been the most commonly used indicator to predict the mechanical flexibility and stretchability in the wearable electronic community [47]. However, the tensile modulus (E) alone does not fully reflect the intrinsic flexibility of a material. For a given strain, materials with high E will have high stress building up. Should the material exhibit a high GC value meanwhile, it may as well sustain the high stress without fracture. Compared to E, the parameter of (GC/E)1/2 better reflects the resistance to deformation and to fracture of a material.

During the mechanical deformation of electronic device, interfacial slippage and delamination can also take place at the interface between different layers. While there are reports on interlayer slippage at the weak van der Waals interaction interfaces within the stack of 2D MHP layers [48,49], there has been yet study on the interfacial slippage between the MHP and adjacent layers of PSCs. The reports on the interfacial slippage have been so far scarce in PSCs, as it is not as noticeable as crack formation or delamination. However, interfacial slippage in inorganic flexible electronics has been investigated [5053]. It has been shown that slippage always precedes delamination as the applied loading increases [51]. Clearly, the understanding of the interfacial slippage behavior is essential to analyze the failure behavior of flexible PSCs, more attentions should be paid from our community in the future.

The process of mechanical delamination of PSCs, on the other hand, has been well studied. The condition of delamination is described by Eq. (4) [10,42,54]:

$$G = \frac{{(1 - {v^2})t{\sigma _{net}}}}{{2E}} \ge {G_C}$$
where G is the driving strain energy release rate, v is the Poisson’s ratio (∼0.33 for MAPbI3), t is the thickness of the thin film, and GC is the critical strain energy release rate of the interface between photovoltaic active layer and its adjacent layers. Based on Eq. (4), there are three main directions to avoid delamination in PSCs: 1) decreasing the MHP film thickness, 2) reducing residual stress, and 3) increasing the interface adhesion between the MHP layer and the adjacent layer.

2.2.4 Impact of mechanical failures on the photovoltaic performance of perovskites

It is widely accepted that the mechanical failures are detrimental to the functionality of PSCs, and the underlying mechanism of how mechanical deformation translates into decreased PV performance is being explored.

Hu et al. suggested that the cracks formed at the grain boundaries serve as the pathway for moisture to diffuse in and decompose MHPs (Fig. 2(d)) [55]. In addition to accelerated perovskite degradation, mechanical failures can negatively impact the PCE of PSCs. Figure 2(e) demonstrates that cracks lead to physical contact between the charge carrier transport layers, which leads to the undesirable recombination of charge carriers and the decrease in shunt resistance [56,57]. The shunting of PSCs leads to the leakage current and low fill factor (FF).

Furthermore, interfacial delamination leads to physical disconnection between the charge carrier transport layers and the MHP layer (Fig. 2(f)). This leads to the increased series resistance and decreased fill factor (FF) and PCE [58]. The effect of series resistance and shunting resistance on the fill factor (FFsh+s) can be described by Eq. (5) [59],

$$F{F_{sh + s}} = F{F_0}(1 - {r_s})[1 - \frac{{({v_{OC}} + 0.7)}}{{{v_{OC}}}}\frac{{F{F_0}(1 - {r_s})}}{{{r_{sh}}}}]$$
where ${v_{OC}}$ is the normalized open circuit voltage, ${r_{sh}}$ is the normalized shunt resistance, ${r_s}$ is the normalized series resistance, and FF0 is as expressed in Eq. (6),
$$F{F_0} = \frac{{{v_{OC}} - \ln ({v_{OC}} + 0.72)}}{{{v_{OC}} + 1}}$$

2.3 Material and device design strategies to improve mechanical flexibility of MHPs

To mitigate the mechanical failures in perovskites, several strategies have been developed to improve the mechanical stability of PSCs and performance in flexible devices (Fig. 3). From the previous discussion, it is established that grain boundaries not only facilitate mechanical failures but also accelerate the degradation of PSCs. Therefore, minimizing grain boundaries in MHP is one of the main focuses in the PSCs community.

 figure: Fig. 3.

Fig. 3. Mechanical Failure managing strategies. (a) Reducing grain boundaries leads to the improved electronic and mechanical stability of perovskites. (b) Crosslinking grain boundaries enhances the fracture resistance of MHPs. (c) Adhesive and low-modulus interfacial layer changes the strain distribution and reduces the overall strain in the MHP layer and the substrate. It also provides strong adhesion between the rigid layers, which increases the resistance to delamination. (d) Deposition of MHP film on the pre-strained substrate. After coating on the pre-stretched elastomer substrate, the thin film shows wrinkled structure that improves the mechanical flexibility and stretchability. (e) Reducing stress on MHP layer by placing it closer to stress-free neutral plane. The closed-up image shows the stress distribution in the layers, x-axis indicates the stress (tensile stress for positive direction) and y-axis indicates the distance from the neutral plane. Initially, the maximum tensile stress occurs at the surface of MHP film(left). However, by adjusting the neutral plane (red dashed line, right) in the PSC and placing it close to the MHP film, the stress in MHP film is reduced.

Download Full Size | PDF

Deceleration of crystallization in MHP through additive engineering is one of the most effective approaches. Feng et al. reported that the addition of dimethyl sulfide led to the decreased Gibbs free energy and crystallization rate [60]. They attribute the prolonged crystallization process to the formation of a stable chelated intermediate $-{-}$ a product of strong interactions between PbI2 and dimethyl sulfide molecules. The resultant MHP film exhibited increased grain size and crystallinity, which led to an enhanced PCE value of 18.40%. In addition, the devices demonstrated an improved mechanical flexibility at 4 mm bending radius while maintaining 82% of the initial PCE after 5000 bending cycles. Other additives including DIO/H2O [61], NH4Cl [62], polycaprolactone [63] have also been explored to increase the grain size and reduce the grain boundaries of perovskite thin films.

Another approach is to crosslink the grains of perovskite to enhance its fracture resistance. Li et al. employed photo-crosslinked [6,6]-phenylC61-butyric oxetane dendron ester(C-PCBOD) to form the organic networks between grain boundaries of perovskites [64]. Not only the C-PCBOD network improves the mechanical stability of PSC, but it also reduces the compressive stress during thermal expansion in the film and leads to the reduced crack formation. The device maintained 62% of its initial PCE of 18.1% under 1% strain. Hu et al. used sulfonated graphene oxide(s-GO) to form cementitious grain boundaries in the film [55]. The sulfonic group of s-GO strongly interacts with [PbI6]4- at grain boundaries of perovskite film. The s-GO-[PbI6]4- complex enhances the bending resistance of perovskite film. The devices maintain 80% of their initial PCE after 10,000 bending cycles at bending radius of 3 mm. Polymers such as ethoxylated trimethylolpropane triacrylate (ETPTA) [65] or trimethyltrivinyl-cyclotrisiloxane (V3D3) [66] have also been introduced to form cross-linked network between grain boundaries and enhance the mechanical stability of PSCs.

The introduction of the adhesive and low-modulus interfacial layer is another important strategy to prevent the crack formation and delamination. It is known that continuous bending leads to the accumulation of highly localized stress in the MHP film. The concentrated stress often initiates mechanical failures. When a low-modulus interfacial layer is placed in between two rigid layers, shear deformation can be accommodated within the soft interfacial layer [67]. Thus, strain no longer continuously builds up linearly throughout the multilayer structure. This effect leads to the decoupling of strain in adjacent rigid layers and results in a decrease in overall strain in each rigid layer [68] as demonstrated in the right panel of Fig. 3(b).

Moreover, the strong adhesion provided by the interfacial layer between rigid films also increases the resistance to delamination. For example, Meng et al. introduced a polymer interfacial layer (Poly(3,4-ethylenedioxythiophene): poly(ethylene-co-vinyl acetate), PEDOT:EVA) in between the ITO and MHPs [69]. The soft polymer layer (E = 139 MPa) provides strong adhesion between the ITO and MHP films, and effectively reduces stress in the MHP layer and PET/ITO. The MHP films with PEDOT:EVA interfacial layer showed an enhanced bending resistance: no crack formed after 7000 cycles of bending with a bending radius of 3 mm. Xue et al. introduced a viscous PEDOT:graphene oxide(PEDOT:GO) gel (E = 347 MPa) as interfacial layer between the MHP and PET/ITO layers [70]. PEDOT:GO strongly adheres the adjacent layers and reduced the strain accordingly. In addition to improving the mechanical stability, PEDOT:EVA and PEDOT:GO interfacial layers also facilitate hole transport in the PSCs.

Depositing MHP films on pre-strained substrates can further enhance their flexibility and stretchability. Kaltenbrunner et al. fabricated PSC thin films (3 µm) and transferred them on pre-stretched elastomers to enhance the device flexibility [71]. When the strain of the substrate is released after the transfer, the device forms sinusoidal wave-like wrinkled geometry. This initial wrinkled structure allows the PSC to be repeatedly stretched and compressed without significant decrease in performance. Tavakoli et al. fabricated flexible PSC on substrate with inverted nanocone structures [72]. In the bending test, their device maintained 95% of the initial PCE after 200 cycles of bending with 6 mm bending radius, showing the improved flexibility over device on the flat substrate. The inverted cone structure makes the film foldable and leads to the effective stress release during the stretching and thermal expansion.

Adjusting the location of MHP layer to a specific plane in the PSC is another effective method to minimize stress on MHP. During the bending of a thin film, there exists a plane in which the net strain and stress is zero, and such location is defined as the neutral plane. By positioning the neutral plane closer to or within the MHP, researchers managed to reduce the strain on MHP. Specifically, the position of the neutral plane can be controlled by modifying the film thickness and/or modulus of the adjacent layers of MHPs [73], which is described by Eq. (7),

$$y = \frac{{\sum\nolimits_{i = 1}^n {E_i^{\prime}{h_i}{y_i}} }}{{\sum\nolimits_{i = 1}^n {E_i^{\prime}{h_i}} }}$$
where y is the position of the neutral plane, Ei′ = Ei / (1 - vi2) is the plane strain modulus, Ei, vi, hi represents the Young’s modulus, Poisson’s ratio, and the thickness of the ith layer, respectively. yi stands for the distance from the bottom of the film to the center of the ith layer.

Lee at al. decreased the strain on the MHP layer by reducing the thickness of the flexible PET substrate from 100 µm to 2.5 µm [40]. More importantly, they added a protective parylene layer on top of their full device to adjust the position of neutral plane of PSC onto the MHP layer. By minimizing the strain applied on the MHP film, they achieved an improved cyclic durability from 20 cycles to 100 cycles of crumpling.

In the work by Lei et al., an innovative strategy that simultaneously reduces the grain boundaries and places MHP at the neutral plane was reported [74]. The authors fabricated PSCs using single crystal MHPs to eliminate the grain boundaries within MHP. The single crystal PSCs was placed in between PET(top) and SU-8/PDMS (bottom) layers. By precisely controlling the thickness of each layer, they managed to place the single crystal MHP film on the neutral plane. The thin film composite features great flexibility: no crack formation until the bending radius reaches 2.5 mm. They further demonstrated that single crystal MHP solar cells show improved mechanical and electronic stability over polycrystalline MHP solar cells, due to the absence of grain boundaries. By fabricating single crystal MHP SCs on the neutral plane, the authors achieved efficient flexible solar cells with the PCE of 20.04%. Their findings further validate that increasing the grain size of MPHs slows down the mechanical degradation of PSCs.

3. From flexible to stretchable solar cells

Looking forward, further advancing the mechanical deformability and stability of PSCs will be one of the key areas of interest. For example, introducing stretchability to PSCs can enable applications that are previously unimaginable, including wearable electronic textile, biomedical sensing devices, and solar harvesting on irregular areas (Fig. 4). However, the promise of stretchable PSCs entails the improved mechanical deformability, which are not yet attainable from the current flexible PV technology.

 figure: Fig. 4.

Fig. 4. Application of Stretchable Solar Cell. (a) textile solar cell. (b) power source for wearable electronics. (c) self-powered electronics for in-situ health monitoring. (d) light-weighted wearable power source for solar backpack.

Download Full Size | PDF

For example, to achieve the desired stretchability for wearable applications, E of the electronic material needs to be reduced to the range of 200-500 MPa [3,75]. Few materials reported in today’s PSCs meet this requirement (Table 1). Crack on-set strain (CoS) is another commonly used parameter to quantify the stretchability of a material. CoS is defined as the strain at which crack occurs [76]. CoS of stretchable materials and structures is typically above 70%, while CoS of α-FAPbI3 is only 2.4% [36].

To improve the mechanical deformability of PSCs, innovative material design and device engineering strategies are urgently needed. Building heterostructures between perovskites and other low modulus materials is one potential pathway to achieve intrinsic stretchability [77]. Due to their ability to reconfigure the polymeric chain conformations, polymeric semiconductors feature superior mechanical deformability. Therefore, semiconducting polymers are considered as the promising matrix material to enable stretchable MHPs. Upon deformation, the entangled polymeric chains get straightened and aligned; once the stress is released, polymeric chains can resume their original state. The reversible process is driven by the change of entropy and the Gibbs free energy of the system. As a result, the CoS of semiconducting polymers have been reported beyond 120% [30]. Other innovative material syntheses have also reported semiconducting polymer stretchability over 100% [78,79]. In comparison, the covalent bonds in inorganic semiconducting materials can only withstand the deformation at a fraction of its length, leading to their poor stretchability (e.g., 1% for a-Si).

4. Challenges and outlook

The intrinsic mechanical and optoelectronic properties of perovskites allow ample opportunities in innovating high performance flexible and wearable photovoltaics. Considerable progress has been made to improve the mechanical stability of PSCs, as highlighted in this Review article. With the research on flexible and wearable PSCs continues to expand, multiple considerations, including lead toxicity, large-area scalability, and long-term stability, need to be advanced.

4.1 Lead toxicity

The best performing PSCs today still rely on Pb as the main element [80]. Lead toxicity gives rise to concerns in the commercialization of PSCs and will particularly restrict their applications in wearable technology [81]. Significant efforts have been made in overcoming the lead toxicity issue in PSCs: lead-free perovskite innovation, encapsulation, and recycling of PSCs have been the three main focuses [82,–85]. For wearable applications of PSCs, the devices often operate under unfavorable moist or straining conditions, which leads to the increased chance of lead leakage [81]. Advances of novel encapsulation techniques for wearable PSCs are urgently needed.

4.2 Device upscaling

PSCs with larger device area often suffer from lower PCEs than their smaller area counterparts [85]. The decreased performance is ascribed to the nonuniformity caused by pinholes and particulates and the difficulty in maintaining a thin yet robust film on a large scale. Significant efforts have been devoted to improving the deposition methods from spin coating (small-scale) to roll-to-roll techniques [86]. Yet, to successfully scale up PSCs, better control of the experimental conditions and further advances in the large area deposition of MHPs, electrodes, and carrier transporting layers are needed [87,88]. For instance, the deposition of brittle ITO onto flexible substrates often yields high surface roughness and low wettability, posing challenges in scaling up flexible PSCs.

4.3 Long-term stability

Improving the long-term stability of PSCs has been of paramount significance to realize their commercialization. MHPs are known to be sensitive to the external environment: humidity, temperature, and mechanical stimuli, etc. This should be of elevated concern for device with large area [88]. For flexible and wearable PSCs, the mechanical stimuli further accelerate the degradation of PSC through crack formation and delamination. In addition, the adhesion between MHPs and adjacent layers is weaker in flexible or stretchable PSCs due to the deformation, swelling, and poor surface roughness of the substrates, which will further limit the long-term stability of PSCs [89].

5. Summary

In this article, we reviewed the mechanical properties and fracture behaviors of MHPs and their impact on the photovoltaic performance of PSCs. We introduced a new term of (E/GC)1/2 that can be used to evaluate the intrinsic flexibility of electronic materials. We analyzed several strategies to address the crack formation and delamination in PSCs: 1) reducing grain boundaries, 2) crosslinking grain boundaries, 3) interfacial engineering, 4) pre-strained substrate, and 5) neutral plane strategy. Looking forward, we anticipate perovskite PVs to continuously advance, evolving from its current flexible form towards its stretchable and wearable form. With the strong growth in wearable technology, stretchable PSCs represent one of the most exciting opportunities for wearable energy sources. To accommodate unprecedented demand for the high mechanical deformability, we need to reimagine the material design and device engineering strategy. We recommend stretchable heterostructures between perovskites and polymer semiconductors as one possible pathway for future wearable PSCs design.

Acknowledgments

The authors acknowledge support from the Department of Chemical Engineering, and the College of Engineering at the University of Michigan. The authors thank Dr. Teng Cui and Dr. F Pelayo García de Arquer for helpful discussions. The authors also thank Hazel Mohamedali for the helpful comments on the manuscript.

Contributions. Sijun Seong and Yanmeng Liu have contributed equally to this work. Sijun Seong, Yanmeng Liu, and Xiwen Gong wrote the manuscript.

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are available in the Supplementary Material and Refs. [9––31,90,91].

Supplemental document

See Supplement 1 for supporting content.

References

1. L. Li, S. Zhang, Z. Yang, E. E. S. Berthold, and W. Chen, “Recent advances of flexible perovskite solar cells,” J. Energy Chem. 27(3), 673–689 (2018). [CrossRef]  

2. A. Novikov, J. Goding, C. Chapman, E. Cuttaz, and R. A. Green, “Stretchable bioelectronics: Mitigating the challenges of the percolation threshold in conductive elastomers,” APL Mater. 8(10), 101105 (2020). [CrossRef]  

3. H. S. Jung, G. S. Han, N. G. Park, and M. J. Ko, “Flexible perovskite solar cells,” Joule 3(8), 1850–1880 (2019). [CrossRef]  

4. K. Tanaka, T. Takahashi, T. Ban, T. Kondo, K. Uchida, and N. Miura, “Comparative study on the excitons in lead-halide-based perovskite-type crystals CH3NH3PbBr3 CH3NH3PbI3,” Solid State Commun. 127(9-10), 619–623 (2003). [CrossRef]  

5. M. Hirasawa, T. Ishihara, T. Goto, K. Uchida, and N. Miura, “Magnetoabsorption of the lowest exciton in perovskite-type compound (CH3NH3)PbI3,” Phys. B 201(C), 427–430 (1994). [CrossRef]  

6. W. J. Yin, T. Shi, and Y. Yan, “Unusual defect physics in CH3NH3PbI3 perovskite solar cell absorber,” Appl. Phys. Lett. 104(6), 063903 (2014). [CrossRef]  

7. C. S. Ponseca, T. J. Savenije, M. Abdellah, K. Zheng, A. Yartsev, T. Pascher, T. Harlang, P. Chabera, T. Pullerits, A. Stepanov, J. P. Wolf, and V. Sundström, “Organometal halide perovskite solar cell materials rationalized: ultrafast charge generation, high and microsecond-long balanced mobilities, and slow recombination,” J. Am. Chem. Soc. 136(14), 5189–5192 (2014). [CrossRef]  

8. G. Xing, N. Mathews, S. Sun, S. S. Lim, Y. M. Lam, M. Grätzel, S. Mhaisalkar, and T. C. Sum, “Long-range balanced electron- and hole-transport lengths in organic-inorganic CH 3 NH 3 PbI 3,” Science 342(6156), 344–347 (2013). [CrossRef]  

9. Y. Rakita, S. R. Cohen, N. K. Kedem, G. Hodes, and D. Cahen, “Mechanical properties of APbX3 (A = Cs or CH3NH3; X = i or Br) perovskite single crystals,” MRS Commun. 5(4), 623–629 (2015). [CrossRef]  

10. C. Ramirez, S. K. Yadavalli, H. F. Garces, Y. Zhou, and N. P. Padture, “Thermo-mechanical behavior of organic-inorganic halide perovskites for solar cells,” Scr. Mater. 150, 36–41 (2018). [CrossRef]  

11. M. A. Reyes-Martinez, A. L. Abdelhady, M. I. Saidaminov, D. Y. Chung, O. M. Bakr, M. G. Kanatzidis, W. O. Soboyejo, and Y. L. Loo, “Time-dependent mechanical response of APbX3 (A = Cs, CH3NH3; X = I, Br) single crystals,” Adv. Mater. 29(24), 1606556 (2017). [CrossRef]  

12. M. Spina, A. Karimi, W. Andreoni, C. A. Pignedoli, B. Náfrádi, L. Forró, and E. Horváth, “Mechanical signatures of degradation of the photovoltaic perovskite CH3NH3PbI3 upon water vapor exposure,” Appl. Phys. Lett. 110(12), 121903 (2017). [CrossRef]  

13. S. Sun, Y. Fang, G. Kieslich, T. J. White, and A. K. Cheetham, “Mechanical properties of organic-inorganic halide perovskites, CH3NH3PbX3 (X = I, Br and Cl), by nanoindentation,” J. Mater. Chem. A 3(36), 18450–18455 (2015). [CrossRef]  

14. J. Feng, “Mechanical properties of hybrid organic-inorganic CH3NH3BX3 (B = Sn, Pb; X = Br, I) perovskites for solar cell absorbers,” APL Mater. 2(8), 081801 (2014). [CrossRef]  

15. M. Faghihnasiri, M. Izadifard, and M. E. Ghazi, “DFT study of mechanical properties and stability of cubic methylammonium lead halide perovskites (CH3NH3PbX3, X = I, Br, Cl),” J. Phys. Chem. C 121(48), 27059–27070 (2017). [CrossRef]  

16. N. Rolston, B. L. Watson, C. D. Bailie, M. D. McGehee, J. P. Bastos, R. Gehlhaar, J. E. Kim, D. Vak, A. T. Mallajosyula, G. Gupta, A. D. Mohite, and R. H. Dauskardt, “Mechanical integrity of solution-processed perovskite solar cells,” Extreme Mech. Lett. 9, 353–358 (2016). [CrossRef]  

17. S. E. Grillo, M. Ducarroir, M. Nadal, E. Tourni, and J.-P. Faurie, “Nanoindentation of Si, GaP, GaAs and ZnSe single crystals,” J. Phys. D: Appl. Phys. 36(1), L5–L9 (2003). [CrossRef]  

18. R. Ballarini, R. L. Mullen, Y. Yin, H. Kahn, S. Stemmer, and A. H. Heuer, “The fracture toughness of polysilicon microdevices: a first report,” J. Mater. Res. 12(4), 915–922 (1997). [CrossRef]  

19. E. Mazhnik and A. R. Oganov, “A model of hardness and fracture toughness of solids,” J. Appl. Phys. 126(12), 125109 (2019). [CrossRef]  

20. S. W. Freiman and J. J. Mecholsky, “The fracture of brittle materials: testing and analysis,” in The Fracture of Brittle Materials: Testing and Analysis (Wiley, 2012).

21. S. Luo, J. H. Lee, C. W. Liu, J. M. Shieh, C. H. Shen, T. T. Wu, D. Jang, and J. R. Greer, “Strength, stiffness, and microstructure of Cu(In,Ga)Se2 thin films deposited via sputtering and co-evaporation,” Appl. Phys. Lett. 105(1), 011907 (2014). [CrossRef]  

22. D. G. Neerinck and T. J. Vink, “Depth profiling of Thin IT0 films by grazing incidence x-ray diffraction,” Thin Solid Films 278, 12 (1996). [CrossRef]  

23. Z. Chen, B. Cotterell, and W. Wang, “The fracture of brittle thin films on compliant substrates in flexible displays,” Eng. Fracture Mechanics 69, 597 (2002). [CrossRef]  

24. C. Lee, X. Wei, J. W. Kysar, and J. Hone, “Measurement of the elastic properties and intrinsic strength of monolayer graphene,” Science 321(5887), 385–388 (2008). [CrossRef]  

25. P. Zhang, L. Ma, F. Fan, Z. Zeng, C. Peng, P. E. Loya, Z. Liu, Y. Gong, J. Zhang, X. Zhang, P. M. Ajayan, T. Zhu, and J. Lou, “Fracture toughness of graphene,” Nat. Commun. 5(1), 1–7 (2014).

26. U. Lang, N. Naujoks, and J. Dual, “Mechanical characterization of PEDOT:PSS thin films,” Synth. Met. 159(5-6), 473–479 (2009). [CrossRef]  

27. I. Lee, G. W. Kim, M. Yang, and T. S. Kim, “Simultaneously enhancing the cohesion and electrical conductivity of PEDOT:PSS conductive polymer films using DMSO additives,” ACS Appl. Mater. Interfaces 8(1), 302–310 (2016). [CrossRef]  

28. S. Savagatrup, A. S. Makaram, D. J. Burke, and D. J. Lipomi, “Mechanical properties of conjugated polymers and polymer-fullerene composites as a function of molecular structure,” Adv. Funct. Mater. 24(8), 1169–1181 (2014). [CrossRef]  

29. V. Brand, C. Bruner, and R. H. Dauskardt, “Cohesion and device reliability in organic bulk heterojunction photovoltaic cells,” Sol. Energy Mater. Sol. Cells 99, 182–189 (2012). [CrossRef]  

30. N. Balar and B. T. O’Connor, “Correlating crack onset strain and cohesive fracture energy in polymer semiconductor films,” Macromolecules 50(21), 8611–8618 (2017). [CrossRef]  

31. K. Gonzalez, J. Xue, A. Chu, and K. Kirane, “Fracture and energetic strength scaling of soft, brittle, and weakly nonlinear elastomers,” J. Appl. Mech. 87(4), 041009 (2020). [CrossRef]  

32. C. Zhu, X. Niu, Y. Fu, N. Li, C. Hu, Y. Chen, X. He, G. Na, P. Liu, H. Zai, Y. Ge, Y. Lu, X. Ke, Y. Bai, S. Yang, P. Chen, Y. Li, M. Sui, L. Zhang, H. Zhou, and Q. Chen, “Strain engineering in perovskite solar cells and its impacts on carrier dynamics,” Nat. Commun. 10(1), 815 (2019). [CrossRef]  

33. Y. Chen, Y. Lei, Y. Li, Y. Yu, J. Cai, M. H. Chiu, R. Rao, Y. Gu, C. Wang, W. Choi, H. Hu, C. Wang, Y. Li, J. Song, J. Zhang, B. Qi, M. Lin, Z. Zhang, A. E. Islam, B. Maruyama, S. Dayeh, L. J. Li, K. Yang, Y. H. Lo, and S. Xu, “Strain engineering and epitaxial stabilization of halide perovskites,” Nature 577(7789), 209–215 (2020). [CrossRef]  

34. J. Zhao, Y. Deng, H. Wei, X. Zheng, Z. Yu, Y. Shao, J. E. Shield, and J. Huang, “Strained hybrid perovskite thin films and their impact on the intrinsic stability of perovskite solar cells,” Sci. Adv. 3(11), eaao5616 (2017). [CrossRef]  

35. J. A. Steele, H. Jin, I. Dovgaliuk, R. F. Berger, T. Braeckevelt, H. Yuan, C. Martin, E. Solano, K. Lejaeghere, S. M. J. Rogge, C. Notebaert, W. Vandezande, K. P. F. Janssen, B. Goderis, E. Debroye, Y. K. Wang, Y. Dong, D. Ma, M. Saidaminov, H. Tan, Z. Lu, V. Dyadkin, D. Chernyshov, V. van Speybroeck, E. H. Sargent, J. Hofkens, and M. B. J. Roeffaers, “Thermal unequilibrium of strained black CsPbI3 thin films,” Science 365(6454), 679–684 (2019). [CrossRef]  

36. D. Liu, D. Luo, A. N. Iqbal, K. W. P. Orr, T. A. S. Doherty, Z. H. Lu, S. D. Stranks, and W. Zhang, “Strain analysis and engineering in halide perovskite photovoltaics,” Nat. Mater. 20(10), 1337–1346 (2021). [CrossRef]  

37. E. G. Moloney, V. Yeddu, and M. I. Saidaminov, “Strain engineering in halide perovskites,” ACS Mater. Lett. 2(11), 1495–1508 (2020). [CrossRef]  

38. L. Gu, D. Li, L. Chao, H. Dong, W. Hui, T. Niu, C. Ran, Y. Xia, L. Song, Y. Chen, and W. Huang, “Strain engineering of metal–halide perovskites toward efficient photovoltaics: advances and perspectives,” Sol. RRL 5(3), 2000672 (2021). [CrossRef]  

39. S. K. Yadavalli, Z. Dai, H. Zhou, Y. Zhou, and N. P. Padture, “Facile healing of cracks in organic–inorganic halide perovskite thin films,” Acta Mater. 187, 112–121 (2020). [CrossRef]  

40. G. Lee, M. C. Kim, Y. W. Choi, N. Ahn, J. Jang, J. Yoon, S. M. Kim, J. G. Lee, D. Kang, H. S. Jung, and M. Choi, “Ultra-flexible perovskite solar cells with crumpling durability: Toward a wearable power source,” Energy Environ. Sci. 12(10), 3182–3191 (2019). [CrossRef]  

41. Z. Dai, S. K. Yadavalli, M. Hu, M. Chen, Y. Zhou, and N. P. Padture, “Effect of grain size on the fracture behavior of organic-inorganic halide perovskite thin films for solar cells,” Scr. Mater. 185, 47–50 (2020). [CrossRef]  

42. Q. Dong, M. Chen, Y. Liu, F. T. Eickemeyer, W. Zhao, Z. Dai, Y. Yin, C. Jiang, J. Feng, S. Jin, S. Liu, S. M. Zakeeruddin, M. Grätzel, N. P. Padture, and Y. Shi, “Flexible perovskite solar cells with simultaneously improved efficiency, operational stability, and mechanical reliability,” Joule 5(6), 1587–1601 (2021). [CrossRef]  

43. B. Lawn, Fracture of Brittle Solids (Cambridge University, 1993).

44. N. Rolston, K. A. Bush, A. D. Printz, A. Gold-Parker, Y. Ding, M. F. Toney, M. D. McGehee, and R. H. Dauskardt, “Engineering stress in perovskite solar cells to improve stability,” Adv. Energy Mater. 8, 1802139 (2018). [CrossRef]  

45. W. J. Yin, T. Shi, and Y. Yan, “Unique properties of halide perovskites as possible origins of the superior solar cell performance,” Adv. Mater. 26(27), 4653–4658 (2014). [CrossRef]  

46. A. Shah, P. Torres, R. Tscharner, N. Wyrsch, and H. Keppner, “Photovoltaic technology: the case for thin-film solar cells,” Science 285(5428), 692–698 (1999). [CrossRef]  

47. S. E. Root, S. Savagatrup, A. D. Printz, D. Rodriquez, and D. J. Lipomi, “Mechanical properties of organic semiconductors for stretchable, highly flexible, and mechanically robust electronics,” Chem. Rev. 117(9), 6467–6499 (2017). [CrossRef]  

48. Q. Tu, I. Spanopoulos, P. Yasaei, C. C. Stoumpos, M. G. Kanatzidis, G. S. Shekhawat, and V. P. Dravid, “Stretching and breaking of ultrathin 2D hybrid organic-inorganic perovskites,” ACS Nano 12(10), 10347–10354 (2018). [CrossRef]  

49. Q. Tu, I. Spanopoulos, S. Hao, C. Wolverton, M. G. Kanatzidis, G. S. Shekhawat, and V. P. Dravid, “Out-of-plane mechanical properties of 2D hybrid organic-inorganic perovskites by nanoindentation,” ACS Appl. Mater. Interfaces 10(26), 22167–22173 (2018). [CrossRef]  

50. Y. Huang, X. Feng, and B. Qu, “Slippage toughness measurement of soft interface between stiff thin films and elastomeric substrate,” Rev. Sci. Instrum. 82(10), 104704 (2011). [CrossRef]  

51. H. Chen, B. W. Lu, Y. Lin, and X. Feng, “Interfacial failure in flexible electronic devices,” IEEE Electron Device Lett. 35(1), 132–134 (2014). [CrossRef]  

52. S.-I. Park, J.-H. Ahn, X. Feng, S. Wang, Y. Huang, and J. A. Rogers, “Theoretical and experimental studies of bending of inorganic electronic materials on plastic substrates**,” (n.d.).

53. L. Dai, X. Feng, B. Liu, and D. Fang, “Interfacial slippage of inorganic electronic materials on plastic substrates,” Appl. Phys. Lett. 97(22), 221903 (2010). [CrossRef]  

54. L. ben Freund and S. Suresh, Thin Film Materials: Stress, Defect Formation and Surface Evolution (Cambridge University, 2004).

55. X. Hu, X. Meng, X. Yang, Z. Huang, Z. Xing, P. Li, L. Tan, M. Su, F. Li, Y. Chen, and Y. Song, “Cementitious grain-boundary passivation for flexible perovskite solar cells with superior environmental stability and mechanical robustness,” Sci. Bull. 66(6), 527–535 (2021). [CrossRef]  

56. E. Ochoa-Martinez, M. Ochoa, R. D. Ortuso, P. Ferdowsi, R. Carron, A. N. Tiwari, U. Steiner, and M. Saliba, “Physical passivation of grain boundaries and defects in perovskite solar cells by an isolating thin polymer,” ACS Energy Lett. 6(7), 2626–2634 (2021). [CrossRef]  

57. J. Ye, G. Liu, L. Jiang, H. Zheng, L. Zhu, X. Zhang, H. Wang, X. Pan, and S. Dai, “Crack-free perovskite layers for high performance and reproducible devices via improved control of ambient conditions during fabrication,” Appl. Surf. Sci. 407, 427–433 (2017). [CrossRef]  

58. B. Qi and J. Wang, “Fill factor in organic solar cells,” Phys. Chem. Chem. Phys. 15(23), 8972–8982 (2013). [CrossRef]  

59. M. A. Green, “Solar cell fill factors: General graph and empirical expressions,” Solid-State Electron. 24(8), 788–789 (1981). [CrossRef]  

60. J. Feng, X. Zhu, Z. Yang, X. Zhang, J. Niu, Z. Wang, S. Zuo, S. Priya, S. Liu, and D. Yang, “Record efficiency stable flexible perovskite solar cell using effective additive assistant strategy,” Adv. Mater. 30(35), 1801418 (2018). [CrossRef]  

61. L. Hou, Y. Wang, X. Liu, J. Wang, L. Wang, X. Li, G. Fu, and S. Yang, “18.0% efficiency flexible perovskite solar cells based on double hole transport layers and CH3NH3PbI3-xClx with dual additives,” J. Mater. Chem. C 6(32), 8770–8777 (2018). [CrossRef]  

62. X. Dai, Y. Deng, C. H. van Brackle, S. Chen, P. N. Rudd, X. Xiao, Y. Lin, B. Chen, and J. Huang, “Scalable fabrication of efficient perovskite solar modules on flexible glass substrates,” Adv. Energy Mater. 10(1), 1903108 (2020). [CrossRef]  

63. Y. Lan, Y. Wang, and Y. Song, “Efficient flexible perovskite solar cells based on a polymer additive,” Flex. Print. Electron. 5(1), 014001 (2020). [CrossRef]  

64. M. Li, Y. G. Yang, Z. K. Wang, T. Kang, Q. Wang, S. H. Turren-Cruz, X. Y. Gao, C. S. Hsu, L. S. Liao, and A. Abate, “Perovskite grains embraced in a soft fullerene network make highly efficient flexible solar cells with superior mechanical stability,” Adv. Mater. 31(25), 1901519 (2019). [CrossRef]  

65. Y. Guo, J. Cheng, L. Liu, and Z. Tan, “Grains boundary networking interconnection design for robust flexible perovskite solar cell,” Mater. Lett. 292, 129559 (2021). [CrossRef]  

66. C. Ge, Z. Yang, X. Liu, Y. Song, A. Wang, and Q. Dong, “Stable and highly flexible perovskite solar cells with power conversion efficiency approaching 20% by elastic grain boundary encapsulation,” CCS Chem. 3(7), 2035–2044 (2021). [CrossRef]  

67. S. Li, X. Liu, R. Li, and Y. Su, “Shear deformation dominates in the soft adhesive layers of the laminated structure of flexible electronics,” Int. J. Solids Struct. 110-111, 305–314 (2017). [CrossRef]  

68. J. Hu, L. Li, H. Lin, P. Zhang, W. Zhou, and Z. Ma, “Flexible integrated photonics: where materials, mechanics and optics meet [Invited],” Opt. Mater. Express 3(9), 1313 (2013). [CrossRef]  

69. X. Meng, Z. Cai, Y. Zhang, X. Hu, Z. Xing, Z. Huang, Z. Huang, Y. Cui, T. Hu, M. Su, X. Liao, L. Zhang, F. Wang, Y. Song, and Y. Chen, “Bio-inspired vertebral design for scalable and flexible perovskite solar cells,” Nat. Commun. 11(1), 3016 (2020). [CrossRef]  

70. T. Xue, G. Chen, X. Hu, M. Su, Z. Huang, X. Meng, Z. Jin, J. Ma, Y. Zhang, and Y. Song, “Mechanically robust and flexible perovskite solar cells via a printable and gelatinous interface,” ACS Appl. Mater. Interfaces 13(17), 19959–19969 (2021). [CrossRef]  

71. M. Kaltenbrunner, G. Adam, E. D. Głowacki, M. Drack, R. Schwödiauer, L. Leonat, D. H. Apaydin, H. Groiss, M. C. Scharber, M. S. White, N. S. Sariciftci, and S. Bauer, “Flexible high power-per-weight perovskite solar cells with chromium oxide-metal contacts for improved stability in air,” Nat. Mater. 14(10), 1032–1039 (2015). [CrossRef]  

72. M. M. Tavakoli, Q. Lin, S. F. Leung, G. C. Lui, H. Lu, L. Li, B. Xiang, and Z. Fan, “Efficient, flexible and mechanically robust perovskite solar cells on inverted nanocone plastic substrates,” Nanoscale 8(7), 4276–4283 (2016). [CrossRef]  

73. W. Jo, H. Lee, Y. Lee, B. S. Bae, and T. S. Kim, “Controlling neutral plane of flexible substrates by asymmetric impregnation of glass fabric for protecting brittle films on foldable electronics,” Adv. Eng. Mater. 23(6), 2001280 (2021). [CrossRef]  

74. Y. Lei, Y. Chen, R. Zhang, Y. Li, Q. Yan, S. Lee, Y. Yu, H. Tsai, W. Choi, K. Wang, Y. Luo, Y. Gu, X. Zheng, C. Wang, C. Wang, H. Hu, Y. Li, B. Qi, M. Lin, Z. Zhang, S. A. Dayeh, M. Pharr, D. P. Fenning, Y. H. Lo, J. Luo, K. Yang, J. Yoo, W. Nie, and S. Xu, “A fabrication process for flexible single-crystal perovskite devices,” Nature 583(7818), 790–795 (2020). [CrossRef]  

75. J. Yu, M. Wang, and S. Lin, “Probing the soft and nanoductile mechanical nature of single and polycrystalline organic-inorganic hybrid perovskites for flexible functional devices,” ACS Nano 10(12), 11044–11057 (2016). [CrossRef]  

76. K. Kim, X. Jia, C. Fuentes-Hernandez, B. Kippelen, S. Graham, and O. N. Pierron, “Optimizing crack onset strain for silicon nitride/fluoropolymer nanolaminate barrier films,” ACS Appl. Nano Mater. 2(4), 2525–2532 (2019). [CrossRef]  

77. P. Jia, M. Lu, S. Sun, Y. Gao, R. Wang, X. Zhao, G. Sun, V. L. Colvin, and W. W. Yu, “Recent advances in flexible perovskite light-emitting diodes,” Adv. Mater. Interfaces 8(17), 2100441 (2021). [CrossRef]  

78. J. Mun, G. J. N. Wang, J. Y. Oh, T. Katsumata, F. L. Lee, J. Kang, H. C. Wu, F. Lissel, S. Rondeau-Gagné, J. B. H. Tok, and Z. Bao, “Effect of nonconjugated spacers on mechanical properties of semiconducting polymers for stretchable transistors,” Adv. Funct. Mater. 28(43), 1804222 (2018). [CrossRef]  

79. J. Y. Oh, S. Rondeau-Gagné, Y. C. Chiu, A. Chortos, F. Lissel, G. J. N. Wang, B. C. Schroeder, T. Kurosawa, J. Lopez, T. Katsumata, J. Xu, C. Zhu, X. Gu, W. G. Bae, Y. Kim, L. Jin, J. W. Chung, J. B. H. Tok, and Z. Bao, “Intrinsically stretchable and healable semiconducting polymer for organic transistors,” Nature 539(7629), 411–415 (2016). [CrossRef]  

80. H. Min, D. Y. Lee, J. Kim, G. Kim, K. S. Lee, J. Kim, M. J. Paik, Y. K. Kim, K. S. Kim, M. G. Kim, T. J. Shin, and S. il Seok, “Perovskite solar cells with atomically coherent interlayers on SnO2 electrodes,” Nature 598(7881), 444–450 (2021). [CrossRef]  

81. J. L. Stauber, T. M. Florence, B. L. Gulson, and L. S. Dale, “Percutaneous absorption of inorganic lead compounds 0:a a,” (1994).

82. X. Li, F. Zhang, J. Wang, J. Tong, T. Xu, and K. Zhu, “On-device lead-absorbing tapes for sustainable perovskite solar cells,” Nat Sustain 4(12), 1038–1041 (2021). [CrossRef]  

83. E. Horváth, M. Kollár, P. Andričević, L. Rossi, X. Mettan, and L. Forró, “Fighting health hazards in lead halide perovskite optoelectronic devices with transparent phosphate salts,” ACS Appl. Mater. Interfaces 13(29), 33995–34002 (2021). [CrossRef]  

84. B. Chen, C. Fei, S. Chen, H. Gu, X. Xiao, and J. Huang, “Recycling lead and transparent conductors from perovskite solar modules,” Nat. Commun. 12(1), 5859 (2021). [CrossRef]  

85. M. A. Green, E. D. Dunlop, J. Hohl-Ebinger, M. Yoshita, N. Kopidakis, and A. W. Y. Ho-Baillie, “Solar cell efficiency tables (Version 55),” Prog. Photovoltaics 28(1), 3–15 (2020). [CrossRef]  

86. Z. Yang, S. Zhang, L. Li, and W. Chen, “Research progress on large-area perovskite thin films and solar modules,” J. Materiomics 3(4), 231–244 (2017). [CrossRef]  

87. M. Davis and Z. Yu, “A review of flexible halide perovskite solar cells towards scalable manufacturing and environmental sustainability,” J. Semicond. 41(4), 041603 (2020). [CrossRef]  

88. S. W. Lee, S. Bae, D. Kim, and H. S. Lee, “Historical analysis of high-efficiency, large-area solar cells: toward upscaling of perovskite solar cells,” Adv. Mater. 32(51), 2002202 (2020). [CrossRef]  

89. Z. Wang, M. Xu, Z. Li, Y. Gao, L. Yang, D. Zhang, and M. Shao, “Intrinsically stretchable organic solar cells beyond 10% power conversion efficiency enabled by transfer printing method,” Adv. Funct. Mater. 31(35), 2103534 (2021). [CrossRef]  

90. A. H. Tsou, J. S. Hord, G. D. Smith, and R. W. Schrader, “Strength analysis of polymeric films,” Polymer 33(14), 2970–2974 (1992). [CrossRef]  

91. G. Scetta, N. Selles, P. Heuillet, M. Ciccotti, and C. Creton, “Cyclic fatigue failure of TPU using a crack propagation approach,” Polym. Test. 97, 107140 (2021). [CrossRef]  

Supplementary Material (1)

NameDescription
Supplement 1       Numerical values of plot in Fig.2

Data availability

Data underlying the results presented in this paper are available in the Supplementary Material and Refs. [931,90,91].

9. Y. Rakita, S. R. Cohen, N. K. Kedem, G. Hodes, and D. Cahen, “Mechanical properties of APbX3 (A = Cs or CH3NH3; X = i or Br) perovskite single crystals,” MRS Commun. 5(4), 623–629 (2015). [CrossRef]  

31. K. Gonzalez, J. Xue, A. Chu, and K. Kirane, “Fracture and energetic strength scaling of soft, brittle, and weakly nonlinear elastomers,” J. Appl. Mech. 87(4), 041009 (2020). [CrossRef]  

90. A. H. Tsou, J. S. Hord, G. D. Smith, and R. W. Schrader, “Strength analysis of polymeric films,” Polymer 33(14), 2970–2974 (1992). [CrossRef]  

91. G. Scetta, N. Selles, P. Heuillet, M. Ciccotti, and C. Creton, “Cyclic fatigue failure of TPU using a crack propagation approach,” Polym. Test. 97, 107140 (2021). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (4)

Fig. 1.
Fig. 1. Schematic illustration of mechanical failure modes in MHP thin films. (a) MHP thin film on PET without external stress(bending). (b) Crack formed on MHP thin film upon bending. (c) Crack sites make MHP film susceptible to moisture-induced degradation. (d) Diffused moisture degrades MHP into PbI2. (e) Hole transport layer materials penetrating through crack sites, leading to undesirable charge carrier recombination as indicated by the red dashed circle. The blue dots and white dots represent electrons and holes respectively. (f) Delamination results in the loss of electrical contact. Charge carriers fail to transport between the layers, leading to increased series resistance and device failure. Delamination also makes the film susceptible to degradation by generating pathways for moisture.
Fig. 2.
Fig. 2. Chart of (GC/E)1/2 value for photovoltaic materials (semiconductors, metals, and elastomers). Upper right direction indicates higher (Gc/E)1/2 and thus lower Rc and higher flexibility, which applies to elastomers and metals. The bottom left area features materials with lower (GC/E)1/2 and thus higher Rc and rigidity, which applies to inorganic crystals. See Supplementary Material 1 for numerical values.
Fig. 3.
Fig. 3. Mechanical Failure managing strategies. (a) Reducing grain boundaries leads to the improved electronic and mechanical stability of perovskites. (b) Crosslinking grain boundaries enhances the fracture resistance of MHPs. (c) Adhesive and low-modulus interfacial layer changes the strain distribution and reduces the overall strain in the MHP layer and the substrate. It also provides strong adhesion between the rigid layers, which increases the resistance to delamination. (d) Deposition of MHP film on the pre-strained substrate. After coating on the pre-stretched elastomer substrate, the thin film shows wrinkled structure that improves the mechanical flexibility and stretchability. (e) Reducing stress on MHP layer by placing it closer to stress-free neutral plane. The closed-up image shows the stress distribution in the layers, x-axis indicates the stress (tensile stress for positive direction) and y-axis indicates the distance from the neutral plane. Initially, the maximum tensile stress occurs at the surface of MHP film(left). However, by adjusting the neutral plane (red dashed line, right) in the PSC and placing it close to the MHP film, the stress in MHP film is reduced.
Fig. 4.
Fig. 4. Application of Stretchable Solar Cell. (a) textile solar cell. (b) power source for wearable electronics. (c) self-powered electronics for in-situ health monitoring. (d) light-weighted wearable power source for solar backpack.

Tables (1)

Tables Icon

Table 1. Mechanical Properties of Metal Halide Perovskite and other Photovoltaic Materials

Equations (7)

Equations on this page are rendered with MathJax. Learn more.

σ T = E h 2 R
K = ψ σ T c 0 0.5 K I C = G C E
R E G C h ψ c 0 0.5 2 = R C
G = ( 1 v 2 ) t σ n e t 2 E G C
F F s h + s = F F 0 ( 1 r s ) [ 1 ( v O C + 0.7 ) v O C F F 0 ( 1 r s ) r s h ]
F F 0 = v O C ln ( v O C + 0.72 ) v O C + 1
y = i = 1 n E i h i y i i = 1 n E i h i
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.