Hakjun Kim‡
a,
Cheong Beom Lee‡bc,
Bum Ho Jeonga,
Jongmin Leea,
Jia Choia,
Kyeounghak Kim
*c and
Hui Joon Park
*ad
aDepartment of Organic and Nano Engineering & Human-Tech Convergence Program, Hanyang University, Seoul 04763, Republic of Korea. E-mail: huijoon@hanyang.ac.kr
bDepartment of Chemistry, Hanyang University, Seoul 04763, Republic of Korea
cDepartment of Chemical Engineering, Hanyang University, Seoul 04763, Republic of Korea. E-mail: chemekim@hanyang.ac.kr
dDepartment of Semiconductor Engineering, Hanyang University, Seoul 04763, Republic of Korea
First published on 14th February 2025
Tin (Sn) halide perovskites have shown significant potential as channels for field-effect transistors (FETs) due to their low effective mass, reduced Fröhlich interaction, as well as lead-free composition, a requirement for electronic components. However, their inherent instability has limited their practical application. Here, we reveal that alkyl ammonium additives of appropriate size can efficiently passivate A-site defects in two-dimensional (2D) Sn halide perovskites, thereby promoting ideal octahedral formation and enhancing hydrogen bonding between A-site and X-site components. These effects lead to improved structural stability, as evidenced by enhanced crystallinity, reduced non-radiative recombination, and decreased Sn oxidation. FETs incorporating perovskites with alkylammonium cations of optimal chain length and multiple functional groups–specifically, propane-1,3-diammonium iodide–exhibit superior performance metrics, including a maximum field-effect mobility of 2.6 cm2 V−1 s−1, an on/off current ratio exceeding 106, and a threshold voltage approaching 0 V.
Early research on FETs utilized three-dimensional (3D) structures of lead (Pb) halide perovskites as the channel layer.5 However, the toxicity of Pb-based perovskites has prompted studies to explore Pb-free perovskite material, a crucial prerequisite for electronic components. Sn halide perovskites, in particular, have gained attention due to their similar ionic radius (115 pm for Sn2+, 119 pm for Pb2+) and analogous atomic orbitals in the valence shell. More importantly, their lower effective mass and reduced Fröhlich interaction6 contribute to enhanced charge carrier mobility, making them even more promising for high-performance FETs.
However, in 3D metal halide perovskites, ion migration poses a significant challenge, leading to hysteresis in transistor operation. To address this issue, two-dimensional (2D) perovskite systems with reduced ion migration have been proposed by replacing small organic cations, such as methylammonium (MA+) and formaidinium (FA+), with bulky organic cations like phenethylammonium (PEA+) and butylammonium (BA+), forming a layered structure with a metal halide octahedral configuration.7 2D perovskites offer additional advantages, including increased stability in ambient conditions due to the hydrophobic nature of bulky organic cations. Furthermore, quantum confinement effects, induced by the bulky organic cations, allow for precise control over charge carrier movement, thereby reducing scattering.8,9 These characteristics make 2D Sn perovskites suitable for FET applications. One commonly studied composition of 2D Sn perovskite FETs is (PEA)2SnI4, adopting a Ruddlesden–Popper (RP) structure. Despite the numerous advantages of this 2D Sn perovskite, the facile oxidation of Sn2+ to Sn4+, an intrinsic difficulty of the Sn halide perovskites, leads to unintended p-doping and induces defects in the crystal lattice structure, which remains a critical issue.10 Moreover, the rapid crystallization kinetics of Sn halide perovskites, resulting in poor perovskite films with high defect density, must be resolved to suppress ion migration for stable FET operation with improved carrier mobility.
To overcome these challenges and achieve high-performance and stable Sn halide perovskite FETs, a range of strategies have been explored. These include engineering the precursor solution,11 modifying the interface between the dielectric and channel layer,12 and passivating the top and bottom layers of the perovskite.13,14 Particularly, the introduction of additives into the perovskite film has proven to be a simple and effective method for defect passivation, crystallinity enhancement, and improvement of charge transport.15 In this work, we ascertain the crucial role of alkyl ammonium chain length in the additives for efficient defect passivation in Sn halide perovskite thin films, aiming for high performance and stable FETs. We investigate this aspect by employing various alkyl ammonium additives with different chain lengths, including methylammonium iodide (MAI), propylammonium iodide (PAI), and butylammonium iodide (BAI), on (PEA)2SnI4 perovskite thin films for FETs. Moreover, we confirm that an increased number of ammonium functional groups (propane-1,3-diammonium iodide, PDAI2) further maximizes the passivation effect. This results in a notable enhancement in the mobility of the p-type channel FET, increasing from 0.42 cm2 V−1 s−1 to 2.41 cm2 V−1 s−1, and an effective reduction of the hysteresis, demonstrating increased stability under operation conditions.
While the crystal structure remains unaffected by the small portion of alkyl ammonium additives, we observed a slightly higher XRD intensity in (PEA)2SnI4 films with PAI additives (Fig. 1b), indicative of enhanced crystallinity. This improved perovskite film quality with the PAI additive was further confirmed by steady state PL (Fig. 1c). In contrast to the PL spectra of the perovskite films with MAI and BAI, which showed similar intensity to that of the pristine perovskite film, the PL intensity of the perovskite film with PAI noticeably increased compared to the pristine film. This indicates that PAI, unlike other additives, effectively reduces non-radiative recombination within the film.
The improved film quality with PAI significantly enhances the performance of the FET devices. Fig. 2a illustrates the FET device structure, which features a (PEA)2SnI4 perovskite thin film as the channel layer, grown on a p+-Si/silicon dioxide (SiO2, 300 nm) substrate, with a bottom-gate-top-contact configuration. The transfer curves of the devices (at Vd = −40 V in the saturation regime, Fig. 2b–e) demonstrate a substantial performance enhancement in the device with PAI, compared to the pristine device. Specifically, the field-effect mobility (μFE) increased from 0.42 cm2 V−1 s−1 to 1.11 cm2 V−1 s−1, and the on–off current ratio (Ion/off) improved from 1.05 × 105 to 3.76 × 105. Additionally, the FET device with PAI exhibits a threshold voltage (Vth) closer to 0 V (from −6.17 V to −1.93 V) and a minimized subthreshold swing (SS) (reduced from 1.83 V dec−1 to 1.09 V dec−1). Conversely, the devices with MAI and BAI additives show decreased μFE and Ion/off, to 0.21 cm2 V−1 s−1 and 1.01 × 104 for MAI, and to 0.18 cm2 V−1 s−1 and 3.88 × 104 for BAI, respectively. Furthermore, their Vth and SS values are also degraded, to −14.3 V and 3.78 V dec−1 for MAI, and −18.4 V and 1.79 V dec−1 for BAI. This degradation is attributed to the induced perovskite lattice instability caused by these additives. All these data are based on average values from 30 devices, and these average values, including maximum values, are summarized in Table 1. The positive effects of additives on the 2D (PEA)2SnI4 perovskite structure will be further evaluated by density functional theory (DFT) calculations later.
Pristine | MAI | PAI | BAI | PDAI2 | ||
---|---|---|---|---|---|---|
a Avg. values represent average ± standard deviation from forward scan curves of 30 devices. | ||||||
μFE (cm2 V−1 s−1) | aAvg. | 0.42 ± 0.12 | 0.21 ± 0.12 | 1.11 ± 0.13 | 0.18 ± 0.11 | 2.41 ± 0.12 |
Max. | 0.62 | 0.35 | 1.37 | 0.31 | 2.60 | |
Ion/Ioff | aAvg. | 1.05 × 105 | 1.01 × 104 | 3.76 × 105 | 3.88 × 104 | 2.43 × 106 |
Max. | 2.34 × 105 | 1.59 × 104 | 8.13 × 105 | 4.04 × 104 | 5.08 × 106 | |
Vth (V) | aAvg. | −6.17 ± 1.06 | −14.3 ± 1.05 | −1.93 ± 1.09 | −18.4 ± 1.03 | −1.12 ± 0.89 |
SS (V dec−1) | aAvg. | 1.83 ± 0.14 | 3.78 ± 0.23 | 1.09 ± 0.11 | 1.79 ± 0.15 | 0.88 ± 0.12 |
Min. | 1.68 | 2.54 | 0.97 | 1.58 | 0.75 |
Inspired by the favorable impact of PAI on the (PEA)2SnI4 perovskite thin film, leading to improved FET performance, we introduced PDAI2, which contained an additional ammonium functional group, to the film (1 mol%). Similar to PAI, the introduction of PDAI2 resulted in minimal changes in the position of XRD peaks [(002): 5.5°, interlayer spacing 16.1 Å] (Fig. 3a and b) and no alterations in bandgap, as estimated by absorbance (Fig. 3d), Tauc plot (Fig. S1, ESI†), and PL spectra (Fig. 3e). However, with the addition of PDAI2, a slight increase in XRD peak intensity was observed compared to PAI (Fig. 3b), accompanied by a significant increase in PL intensity (Fig. 3e), suggesting more efficient suppression of non-radiative recombination compared to PAI. Scanning electron microscopy (SEM) analysis revealed a reduction in sharp grain boundaries seen in pristine films with the addition of PAI, which were further diminished with the addition of PDAI2 (Fig. 3f–h). To clearly show the morphological improvements in (PEA)2SnI4 induced by the additives, low-magnification SEM images (Fig. S2, ESI†) and non-contact mode AFM analyses (Fig. S3, ESI†) are provided. In particular, the AFM line profiles reveal variations in grain boundary depth: the pristine (PEA)2SnI4 film exhibits a depth of 39.47 nm, whereas the PAI and PDAI2 added films show reduced depths of 27.26 and 12.02 nm, respectively. This reduction indicates that PAI and PDAI2 effectively passivate the charged defects at the grain boundaries, thereby improving the overall film quality.17 As a result, the contact resistance at the device interface of the pristine film, estimated through the transmission line method, decreases with PAI and is further reduced with PDAI2 (Fig. 3c).
The improved film quality with additives such as PAI and PDAI2 is also confirmed by X-ray photoelectron spectroscopy (XPS) analysis. Fig. 4a–c show the core level spectra of Sn 3d5/2 and Sn 3d3/2, representing peaks at 486 and 494 eV, respectively. These peaks are further resolved into main peaks (485.5 eV and 494 eV) and two shoulder peaks at lower (484.5 eV and 493 eV) and higher binding energies (486.5 eV and 495 eV). The main peaks correspond to Sn2+ species, while the shoulder peaks at higher binding energies (486.5 eV and 495 eV) are associated with Sn4+ species in an oxidized state. These peaks are diminished with PAI and further reduced with PDAI2, indicating prevention of Sn oxidation from Sn2+ to Sn4+ by effective passivation. Additionally, the shoulder peaks at lower binding energies (484.5 eV and 493 eV) corresponding to undercoordinated Sn in an oxidized state (denoted as Snδ<2+)22 are also decreased with PAI and markedly reduced with PDAI2. The oxidation of Sn2+ to Sn4+ has been reported to generate Sn vacancies in the lattice, which act as hole traps and these vacancies contribute to charge trapping, negatively impacting hole charge transport.23 To further investigate the oxidation induced defect density within Sn perovskite bulk, we conducted space-charge limited current (SCLC) measurements using a hole-only device with the structure of [ITO/PEDOT:PSS/(PEA)2SnI4/PTAA/Ag] (Fig. S4, ESI†). The trap density (Ntrap) was calculated using the equation Ntrap = 2εεoVTFL(eL2)−1,24 where the trap-filled limited voltage (VTFL) was determined to be 0.778 V for pristine (PEA)2SnI4 films and decreased to 0.636 V and 0.581 V with the incorporation of PAI and PDAI2, respectively. Consequently, the trap density of pristine (PEA)2SnI4, initially 3.11 × 1016 cm−3 was reduced to 2.54 × 1016 cm−3 with PAI and further decreased to 2.32 × 1016 cm−3 with PDAI2. For reference, the energy band structures of perovskite films with and without additives (PAI and PDAI2) were investigated using ultra-violet photoelectron spectroscopy (UPS) (Fig. S5, ESI†). The work function (φ) was calculated using the equation: φ = hν − |Ecut-off − EF|, where Ecut-off and EF are the secondary electron cut-off energy and Fermi energy level, and the valence band maximum (VBM) is estimated from the onset energy.25 The conduction band minimum (CBM) was then estimated by incorporating the bandgap value obtained from the Tauc plot (Fig. S1, ESI†). The results indicate that the variations are minimal, demonstrating that the addition of a small portion of additives has only a marginal effect on the energy band structure.
Consequently, the μFE and Ion/off of the FET device were further enhanced with PDAI2 compared to PAI, as demonstrated in the transfer curves (Fig. 5a–c). Specifically, μFE (in saturation regime) increased from 1.11 cm2 V−1 s−1 to 2.41 cm2 V−1 s−1, and Ion/off improved from 3.76 × 105 to 2.43 × 106. The corresponding output curves are presented in Fig. 5d–f. Additionally, compared to PAI, Vth further shifted with PDAI2 (from −1.93 V to −1.12 V), and SS values were further reduced (from 1.09 to 0.88). These data are based on average values from 30 devices, and the variations in μFE, Ion/off, Vth, and SS of the devices with and without additives (PAI and PDAI2) are summarized in Fig. 5g–j and Table 1. To accurately assess the mobility values of FET devices, we estimated the effective mobility (μeffective) by multiplying the claimed μFE by the reliability factor in the saturation regime (rsat).26 Detailed information regarding the mobility calculations and the obtained data is provided in Note 1 and Table S1 of the (ESI†). Additionally, the transfer characterization , including fitting lines used for mobility calculations, is presented in Fig. S6 (ESI†).
The transfer curves of 30 FET devices with PDAI2-added perovskite are represented in Fig. S7 (ESI†). The maximum interface trap density (Nmax) within FET devices was also calculated from SS values using the equation27 Nmax = [(SSlog
e)/(kT/q) − 1](Ci/q), where k is the Boltzmann constant, T is the absolute temperature, q is the elementary charge, and Ci is the capacitance of the insulator. Nmax is minimized in the FETs with PAI and PDAI2, being lowest with PDAI2 (Fig. S8, ESI†). For reference, the performance variation of FETs depending on the portion of additive (PAI and PDAI2) is investigated in Fig. S9 (ESI†), with 1 mol% showing the best performance.
To evaluate the operational stability of the FET devices, bias-stress stability tests were conducted by monitoring ID under constant gate and drain voltages (VG = −40 V and VD = −40 V). As shown in Fig. S10 (ESI†), the FET device with pristine perovskite retained less than 20% of its initial ID value after 2000 s, indicating noticeable carrier trapping within the device. However, the devices with PAI and PDAI2 exhibited much lower ID decay, retaining over 40% and 60% of their initial ID values, respectively, even after 2000 s. This suggests a reduction in trap states in the devices with these additives.28
To elucidate the advantageous effects of additive molecules on the perovskite structures, DFT calculations were conducted. In (PEA)2SnI4, PEA vacancies can be easily formed, and ammonium cations can effectively passivate these cation sites.29 Given that the defect passivation enhances device performance by mitigating non-radiative recombination centers within the perovskite,30 we calculated the formation energy (Eform) of various organic molecules that could fill a PEA vacancy, thereby improving the stability of (PEA)2SnI4 (Fig. 6a and b). The formation energy decreases as the length of carbon chain in the additive increases (methylammonium (MA+) → ethylammonium (EA+) → propylammonium (PA+)), which stabilizes the perovskite (Fig. 6a). However, the formation energy increases when the carbon chain length exceeds that of PA+, indicating an optimal chain length for perovskite stabilization. We also examined the effect of diammonium additives on the stability of the 2D (PEA)2SnI4 perovskite. Similarly, the formation energy decreased with increasing carbon chain length (methyldiammonium (MDA+) → ethane-1,2-diammonium (EDA+) → propane-1,3-diammonium (PDA+)), reaching its lowest at (PEA)1.5(PDA)0.5SnI4 (Eform = −2.65 eV), but increased when butane-1,4-diammonium (BDA+) was used, akin to the single ammonium group additives. Overall, diammonium additives exhibited lower formation energies than single ammonium group additives, regardless of carbon chain (Fig. 6a and b).
To explore the origin of the formation energy differences influenced by carbon chain length, we initially examined the equatorial I–Sn–axial I angle (Fig. 6c). Greater deviation from the ideal octahedral structure (90°) increases distortion and instability in perovskite materials.31 The equatorial I–Sn–axial I angle varied with the number of carbon atoms, with angles in propylammonium (PA+) and propyldiammonium (PDA+) approaching 90° more closely than in other cases, thus promoting an ideal octahedron formation (Fig. 6d). This suggests that steric hindrance, dependent on the number of carbon atoms, affects the stability of the inorganic framework.32,33
We also evaluated the impact of increased ammonium groups by quantifying the number of H⋯I bonds shorter than 3.0 Å, which serve as indicators of effective hydrogen bonding and contribute to the overall structural stability of organic–inorganic halide perovskites. Our results showed a higher number of hydrogen bonds in diammonium additives (MDA+, EDA+, PDA+, and BDA+) compared to monoammonium additives (MA+, EA+, PA+, and BA+) (Fig. 6e). The increased hydrogen bonding due to diammonium groups further stabilizes the 2D perovskite structures, enhancing their structural robustness (e.g., raising their melting temperature), as previously reported.34 Additionally, the average Sn–axial I bond length increased with diammonium additives (Fig. S11, ESI†), which can be attributed to the H⋯I hydrogen bonding interaction. Therefore, the incorporation of organic molecules with diammonium groups results in a higher number of hydrogen bonds, providing greater stabilization to the system compared to the additives with a single ammonium group.
Considering that strong van der Waals interactions between organic ligands stabilize the 2D perovskite structure, longer chain lengths seem more appropriate as additives.35 However, structural mismatches, such as octahedral tilting and size mismatches between the cavity size of PEA vacancies and the length of addictive molecules, could destabilize the perovskite structure. Therefore, an optimal carbon chain length likely exists. Consequently, the PDA additive enhances device efficiency and stability by effectively passivating the PEA vacancies.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4tc05307a |
‡ These authors contribute equally to this work |
This journal is © The Royal Society of Chemistry 2025 |