Introduction

Direct and tunable bandgap, easy fabrication, low cost, and color purity are the significant properties of metal halide perovskite. The metal halide perovskites have been widely used in different applications, including light-emitting diodes (LEDs), solar cells, lasers, and sensors1,2,3,4,5,6,7,8,9. The quasi-two-dimensional perovskites with Ruddlesden–Popper (RP) phase have the chemical formula of L2An−1BnX3n+1, where L is a spacer cation, A is a small cation, B is a divalent metal cation, and X is a halide anion. The value of n shows the number of perovskite layers sandwiched between L cations. The quasi-2D perovskites are similar to multiple quantum wells (MQWs), and the width of the quantum well can be adjusted by the amount of n. The chemical formula of 2D perovskite is L2BX4 (n = 1). In this structure, L2 and BX4 represent the quantum barrier and the quantum well, respectively10.

The perovskite light emitting diode (PeLED) consists of a perovskite active layer sandwiched between the electron transport layer (ETL), and the hole transport layer (HTL). By applying a forward bias to PeLED, electrons and holes are injected into the active layer from the cathode side through the ETL and the anode side through the HTL, respectively. Subsequently, the electrons and holes recombine in the active region, leading the photon generation by the phenomenon of spontaneous emission11. The main challenge of PeLEDs is their low stability. The most critical factor of its short lifetime is the instability of the active material. 2D perovskites are more stable than 3D perovskites12. Also, the 2D perovskites with aromatic cations, such as phenylethylammonium (PEA), are more stable than aliphatic cations, such as butylammonium (BA)13. The fluorophenethylamine (FPEA) (FC6H4(CH2)2NH3) aromatic cation is formed by replacing one hydrogen atom in the aromatic ring of the PEA (C6H5(CH2)2NH3) cation with a fluorine (F) atom. The 2D perovskite with FPEA aromatic cation is more stable than PEA because the ionic radius of F is larger than that of hydrogen and the length of the cation in FPEA is larger than that of PEA13. Therefore, regarding stability, FPEA2PbI4 is a suitable option for use in the active layer of PeLED. To the best of our knowledge, the simulation of PeLED with 2D perovskites has not been done so far.

Some studies investigated the properties of 2D perovskite materials. The optical and electronic properties of BA2PbX4 are investigated with density functional theory (DFT) calculations by Silver et al.14. C. Song, et al.15 studied the electronic and optical properties of p-FPEA2PbI4 with DFT calculations. In another work, the mechanical properties of BZA2PbI4 perovskite were computed by DFT calculations16. In our previous work, we demonstrated that among the RP 2D perovskites with L2PbI₄ formula (L = PEA, FPEA, BA, and BZA), FPEA2PbI₄ exhibits greater thermodynamic, mechanical, and moisture stabilities, attributed to the presence of a fluorine atom in its structure12. Consequently, altering the position of the fluorine atom can influence various properties of this material. Therefore, there is a need to thoroughly investigate the properties of X-FPEA2PbI4 (X = p, m, and o) perovskites to understand more about this material for use in different applications.

In this study, we investigate the optical, mechanical, and electronic properties, as well as the thermodynamic, thermal, and moisture stability of 2D Ruddlesden popper (RP) perovskites with the chemical formula of X-FPEA2PbI4 (X = p, m, and o). We utilize density functional theory calculations to assess their potential use in PeLED. To our knowledge, this is the first comprehensive investigation into the effect of the displacement of the fluorine (F) atom in the aromatic cation on the properties and stability of 2D perovskites X-FPEA2PbI4 (X = p, m, and o). Our results indicate that the p-FPEA2PbI4 perovskite exhibits high stability and is proposed as the active layer in PeLEDs. Furthermore, we simulate PeLEDs based on p-FPEA2PbI4, PEA2PbI4, and MAPbI3 using the finite element method (FEM) and compare their internal quantum efficiency, external quantum efficiency, current–light intensity (L–I) curve, emission spectrum, and stability.

Simulation methods

Kohn–Sham equation

The foundation of DFT calculations is the Kohn–Sham (KS) equation, as expressed in Eq. (1), which is derived from the multi-particle Schrödinger equation. The Hamiltonian in this equation encompasses both kinetic energy (T) and potential energy (V). The potential energy consists of the nuclear potential (Vn), Hartree potential (VH), and exchange–correlation potential (Vxc) (see section S1)17.

$$\begin{gathered} (T + V)\phi_{i} (r) = \varepsilon_{i} (r)\phi_{i} (r) \hfill \\ V = V_{n} + V_{H} + V_{xc} \hfill \\ \end{gathered}$$
(1)

The KS equation is solved using the iterative self-consistent field (SCF) method, as illustrated in Fig. 1.

Fig. 1
figure 1

Algorithm of SCF for solving the KS equation.

The process begins with an initial guess for the electron density (n(r)). The Hartree and exchange potentials are calculated based on this electron density. Subsequently, the KS equation is solved and the electron density is recalculated. This new value is then compared to the initial guess, and the process continues until the two densities converge closely. By solving the KS equation, we can obtain the eigenvalues, eigenfunctions, total energy, and electron density17.

Electronic properties

Bloch’s theorem is employed for crystal structures characterized by periodic potential. The Bloch wave function, expressed in Eq. (2), consists of a periodic component (uik(r)) and a plane wave component (exp(ikr)). By substituting the Bloch wave function into the Kohn–Sham equation, we derive the crystalline version of the equation, as illustrated in Eq. (3). This refined equation is instrumental in obtaining the band structure of the material. In this equation, i represents the eigenstate index and k denotes the wave vector17,18.

$$\psi_{nk} (r) = u_{nk} (r)\exp (jkr)$$
(2)
$$\begin{gathered} [ - \frac{1}{2}(\nabla + ik)^{2} + V_{tot} ]u_{ik} = \varepsilon_{ik} u_{ik} (r) \hfill \\ V_{tot} (r) = V_{n} (r) + V_{H} (r) + V_{xc} (r) \hfill \\ \end{gathered}$$
(3)

The effective mass of charge carriers (i.e., electrons and holes) is derived from the band structure as19:

$$m^{*} = \frac{{\hbar^{2} }}{{\frac{{\partial^{2} E}}{{\partial k^{2} }}}}$$
(4)

Optical properties

To calculate the optical properties with DFT calculations, the imaginary part of the dielectric function (ε2(ω)) is first determined from Fermi’s golden rule as20:

$$\varepsilon_{2} (\hbar \omega ) = \frac{{2q^{2} \pi }}{{V\varepsilon_{0} }}\sum\limits_{k,v,c} {| < \Psi_{k}^{c} |u.r|\Psi_{k}^{v} > |}^{2} \delta (E_{k}^{c} - E_{k}^{v} + \hbar \omega )$$
(5)

In this equation, the parameters q, V, ħω, u, and ε0 are the electronic charge, the volume of a primitive cell, the photon energy, the polarization of the incident electric field, and the dielectric constant in the vacuum, respectively. In addition, ψkv and ψkc denote the wave functions of electrons in the valence and conduction bands, respectively. The term includes the delta function for energy conservation. Next, the real part of the dielectric function (ε1(ω)) is calculated from ε2(ω) using the Kramer-Kronig relation as:

$$\varepsilon_{1} (\hbar \omega ) = 1 + \frac{2}{\pi }\int\limits_{0}^{\infty } {\frac{{\omega ^{\prime}\varepsilon_{2} (\omega ^{\prime})}}{{\omega ^{{\prime}{2}} - \omega^{2} }}}$$
(6)

Finally, the complex refractive index (N = nr + ik) can be calculated from the complex dielectric function (ε(ω)) as follows:

$$\varepsilon = \varepsilon_{1} + i\varepsilon_{2} = N^{2}$$
(7)
$$\varepsilon_{1} = n_{r}^{2} - k^{2} ,\varepsilon_{2} = 2n_{r} k$$
(8)

The exciton binding energy (Ebx) is computed using the reduced effective mass (mr) and the static dielectric constant (εs) as19:

$$\begin{gathered} E_{bex} = 13.6\frac{{m_{r} }}{{\varepsilon_{s}^{2} }},\frac{1}{{m_{r} }} = \frac{1}{{m_{e} }} + \frac{1}{{m_{h} }} \hfill \\ \hfill \\ \end{gathered}$$
(9)

Mechanical properties

To investigate the mechanical properties of perovskite materials, we first calculate the elastic constants (Cij). The elastic constants, according to Hooke’s law in Voigt notation, are expressed as21:

$$\sigma_{i} = C_{ij} \varepsilon_{j}$$
(10)

where σi is stress and εj is strain.

Using DFT calculations, the elastic constants can be determined as:

$$\Delta U = \frac{\Omega }{2}\sum\nolimits_{i = 1}^{6} {\sum\nolimits_{j = 1}^{6} {c_{ij} e_{i} e_{j} } }$$
(11)

ΔU is the total energy change under external force, Ω is primitive cell volume, and eij is strain-related deformation vectors. The conditions for the mechanical stability of the monoclinic crystal structure are as follows22:

$$\begin{gathered} C_{11} > \, 0, \, C_{22} > 0, \, C_{33} > 0, \, C_{44} > \, 0, \, C_{55} > 0, \, C_{66} > 0 \hfill \\ C_{11} + C_{22} + C_{33} + 2C_{12} + 2C_{13} + 2C_{23} > 0 \hfill \\ C_{33} C_{55} - 2C^{2}_{35} > \, 0, \, C_{44} C_{66} - 2C^{2}_{46} > \, 0, \, C_{22} + C_{33} - 2C_{23} > \, 0 \hfill \\ \end{gathered}$$
(12)

The mechanical properties, including bulk modulus (B), Young’s modulus (E), Shear’s modulus (G), and Poisson’s ratio (υ), are expressed from the Cij and approximations of Voigt (upper limit of values), Reuss (lower limit of values) and Hill (average values) with the following equations23:

$$B_{V} = \frac{1}{9}[(C_{11} + C_{22} + C_{33} ) + 2(C_{12} + C_{13} + C_{23} )]$$
(13)
$$G_{V} = \frac{1}{15}[(C_{11} + C_{22} + C_{33} ) - (C_{12} + C_{13} + C_{23} ) + 3(C_{44} + C_{55} + C_{66} )]$$
(14)
$$E_{V} = \frac{{9B_{V} G_{V} }}{{3B_{V} + G_{V} }}$$
(15)
$$\upsilon_{V} = \frac{{3B_{V} - 2G_{V} }}{{2(3B_{V} + G_{V} )}}$$
(16)
$$B_{R} = \frac{1}{{[(S_{11} + S_{22} + S_{33} ) + 2(S_{12} + S_{13} + S_{23} )]}}$$
(17)
$$G_{R} = \frac{15}{{[4(S_{11} + S_{22} + S_{33} ) + 3(S_{44} + S_{55} + S_{66} ) - 4(S_{12} + S_{13} + S_{23} )]}}$$
(18)
$$E_{R} = \frac{{9B_{R} G_{R} }}{{3B_{R} + G_{R} }}$$
(19)
$$\upsilon_{R} = \frac{{3B_{R} - 2G_{R} }}{{2(3B_{R} + G_{R} )}}$$
(20)
$$B_{H} = \frac{{B_{V} + B_{R} }}{2}$$
(21)
$$G_{H} = \frac{{G_{V} + G_{R} }}{2}$$
(22)
$$E_{H} = \frac{{E_{V} + E_{R} }}{2}$$
(23)
$$\upsilon_{H} = \frac{{\upsilon_{V} + \upsilon_{R} }}{2}$$
(24)

The subscripts V, R, and H in these equations represent the Voigt, Reuss, and Hill approximations, respectively. A Poisson’s ratio value of 0.26 serves as a threshold between ductility and brittleness; thus, for values greater than 0.26, perovskite exhibits ductile behavior. Additionally, lower values of mechanical moduli—such as the bulk modulus, Young’s modulus, and shear modulus—indicate greater flexibility of the material.

Stability

The formation energy (FE) of the 2D perovskite with the general formula L2BX4 is calculated using Eq. (25), where E expresses the total energy of the material and its components. A higher absolute value of FE indicates greater thermodynamic stability of the material24.

$$FE = E(L_{2} BX_{4} ) - (2E(LX) + E(BX_{2} ))$$
(25)

To assess moisture stability, the adsorption energy (Eads) of water on the surface of perovskite is calculated as25:

$$E_{ads} = E_{water/surface} - E_{water} - E_{surface}$$
(26)

A negative value of Eads indicates that the water molecule (H₂O) is more strongly attracted to the perovskite surface. Conversely, a positive Eads value suggests that the water molecule is repelled from the surface of the perovskite.

Efficiency of LEDs

To calculate the LED efficiency, the carrier transport model is first solved using Eq. (27). In this equation, ϕ represents the electrostatic potential, μ is the mobility of carriers, D is the diffusion coefficient, q is the electron charge, n is the electron density, p is the hole density, ND is the donor concentration, and NA is the acceptor concentration. In addition, Gn and Gp denote the generation rates, while Rn, and Rp are the recombination rates. By solving the equations of the carrier transport model, the values of n, p, and ϕ can be obtained26.

$$\begin{aligned} \frac{\partial n(t)}{{\partial t}} & = G_{n} - R_{n} + \frac{1}{q}\nabla \left[ { - q\mu_{n} n\nabla \phi + qD_{n} \nabla_{n} } \right] \\ \frac{\partial p(t)}{{\partial t}} & = G_{p} - R_{p} + \frac{1}{q}\nabla \left[ { - q\mu_{p} p\nabla \phi - qD_{p} \nabla_{p} } \right] \\ \nabla .(\varepsilon \nabla \phi ) & = - q(p - n + N_{D} - N_{A} ) \\ \end{aligned}$$
(27)

Next, we calculate the carrier rate equation using Eq. (28), where R = Rn-Gn or R = Rp-Gp represents the net recombination rate. Moreover, A denotes the nonradiative Shockley–Read–Hall (SRH) recombination, B is the radiative recombination, and C is the Auger recombination.

$$\frac{dn}{{dt}} = - R = - (An + Bn^{2} + Cn^{3} )$$
(28)

The internal quantum efficiency (IQE) is calculated as the ratio of radiative recombination to nonradiative recombination using Eq. (29).

$$IQE = \frac{Bn}{{A + Bn + Cn^{2} }}$$
(29)

Considering excitonic recombination, the IQE for 2D perovskites is modified as27:

$$IQE = \frac{{Bn + k_{exciton} }}{{A + k_{exciton} + Bn + Cn^{2} }}$$
(30)

The EQE is calculated from the IQE using Eq. (31), assuming that the light emission occurs from the edge. In this equation, nr is the refractive index of the active layer28.

$$EQE = \frac{IQE}{{2n_{r}^{2} }}$$
(31)

The optical output power (L) is related to total photon flux (ϕt) and external quantum efficiency (EQE) as:

$$L = \phi_{t} .h\nu = EQE(\frac{I}{q}h\nu )$$
(32)

Here, the total photon flux (ϕt) is defined as:

$$\phi_{t} = R_{spon} .V_{0}$$
(33)

where V0 is the active layer volume. The total spontaneous emission rate (Rspon) per unit volume is expressed by29:

$$R_{spon} = \int\limits_{{E_{g} }}^{\infty } {\frac{1}{{\tau_{r} }}} f_{c} (1 - f_{v} )g(E)dE$$
(34)

where fc is the Fermi –Dirac distribution in the conduction band, fv is the Fermi –Dirac distribution in the valence band, g(E) is the joint density of states and τr is the radiative recombination lifetime.

Results and discussions

Structural properties

The crystal structure and lattice parameters of X-FPEA2PbI4 (X = p, m, and o) perovskites are shown in Fig. 2. At room temperature, these perovskites exhibit a monoclinic crystal phase30.

Fig. 2
figure 2

Crystal structure and lattice parameters of X-FPEA2PbI4.

The spacer cation FPEA (FC6H4(CH2)2NH3) is characterized as an aromatic cation. In its structure, the ammonium group (NH₃) is linked to the aromatic ring via a carbon chain. The position of the fluorine (F) atom on the aromatic ring leads to three distinct configurations: para (p)-FPEA, meso (m)-FPEA, and ortho (o)-FPEA, as depicted in Fig. 3. Importantly, it has been observed that the experimental lattice parameters remain consistent regardless of the position of the F atom on the aromatic ring.

Fig. 3
figure 3

The aromatic spacer cation of (a) p-FPEA, (b) m-FPEA, and (c) o-FPEA.

The optimized lattice parameters and primitive cell volumes of X-FPEA2PbI4 (X = p, m, and o) are obtained through DFT calculations and used to determine the stability and various properties of the perovskites. These parameters are presented in Table 1 (see section S1).

Table 1 Optimized lattice parameters and volumes of X-FPEA2PbI4 (X = p, m, and o).

Material stability

To investigate thermodynamic stability, the formation energy (FE) values for X-FPEA2PbI4 (X = p, m, and o) perovskites are calculated using DFT calculation and presented in Table 2. The p-FPEA2PbI4 perovskite is more stable than m-FPEA2PbI4 and o-FPEA2PbI4 due to its larger absolute value of FE. By examining the hydrogen bonding in these perovskites (see section S3), it can be concluded that pFPEA has a stronger hydrogen bond due to the shorter N…X distance. So, it is more stable.

Table 2 Formation energy of X-FPEA2PbI4 (X = p, m, and o).

Ab initio molecular dynamics (AIMD) combines molecular dynamics with quantum mechanical calculations, using first principles to derive atomic forces. This method provides accurate insights into the thermal stability and dynamics of materials. The NVT ensemble maintains a constant particle number, volume, and temperature, allowing the system to exchange energy with a heat bath. This is essential for simulating realistic thermal conditions in AIMD studies. The thermal stability of X-FPEA2PbI4 (X = p, m, and o) perovskites is explored using AIMD analysis with the NVT ensemble, with results shown in Fig. 4. In all three materials, the energy changes due to temperature variations around room temperature are minimal, indicating their thermal stability. Among these perovskites, p-FPEA2PbI4 exhibits the lowest energy changes, suggesting it has better stability compared to the other materials.

Fig. 4
figure 4

Thermal stability of X-FPEA2PbI4 (X = p, m, and o) using AIMD simulation.

The values of adsorption energy for X-FPEA2PbI4 (X = p, m, and o) perovskites are presented in Table 3. The results show that p-FPEA2PbI4 exhibits greater stability against water exposure (indicated by a lower absolute value of Eads), suggesting it has superior moisture stability compared to the other two perovskites. These findings are consistent with experimental results obtained using the contact angle method31. To calculate Eads, the average from several cases involving the placement of water molecules on the perovskite surface is computed using DFT, as illustrated in Figs. S2S4.

Table 3 Adsorption energy of X-FPEA2PbI4 (X = p, m, and o).

The results presented in Section S2 show that the formation energy of X-FPEA2PbI4 (X = p, m, o) is nearly 25 times larger than that of MAPbI3. Thus, the thermodynamic stability of the proposed 2D perovskites is significantly higher than that of 3D perovskites. Comparing the adsorption energy values between MAPbI3 and the proposed materials reveals that the moisture stability of X-FPEA2PbI4 (X = p, m, o) is approximately four times greater than that of MAPbI3. Moreover, p-FPEA2PbI4 is more stable than PEA2PbI4, which has an FE value of − 4.689 eV13.

Electrical and optical properties

The values of the complex refractive index and dielectric function for X-FPEA2PbI4 (X = p, m, and o) perovskites are shown in Fig. 5. The refractive index at zero frequency (ns) for p-FPEA2PbI4, m-FPEA2PbI4, and o-FPEA2PbI4 perovskite is 2.102, 2.067, and 2.059, respectively. In addition, the dielectric constant values at zero frequency (εs) for these perovskites are 4.420, 4.272, and 4.239, respectively. Variations in the refractive index and dielectric constant are proportional to changes in the position of the F atom on the aromatic ring. Reducing the distance between the negative charge (fluorine atom) and the positive charge (NH3 part) in the organic cation decreases the dipole moment, which in turn reduces the polarization and dielectric constant31.

Fig. 5
figure 5

Complex refractive index for X-FPEA2PbI4 (a) X = p, (b) X = m, and (c) X = o and dielectric function for X-FPEA2PbI4 (d) X = p, (e) X = m, and (f) X = o.

Figure 6a–c show the band structure and bandgap energy (Eg) for X-FPEA2PbI4 (X = p, m, and o) perovskites, respectively. The value of Eg varies with the position of the F atom. Specifically, the Eg values for p-FPEA2PbI4, m-FPEA2PbI4, and o-FPEA2PbI4 are 2.097, 2.133, and 2.177 eV, respectively. The bandgap inversely correlates with refractive index values.

Fig. 6
figure 6

Band structure and bandgap energy of X-FPEA2PbI4 (a) X = p, (b) X = m, and (c) X = o without spin–orbit coupling effect.

The band structure and Eg for X-FPEA2PbI4 (X = p, m, and o) perovskites, considering the spin–orbit coupling (SOC) effect, are indicated in Fig. 7. The obtained values of Eg for p-FPEA2PbI4, m-FPEA2PbI4, and o-FPEA2PbI4 are 1.323, 1.329, and 1.357 eV, respectively.

Fig. 7
figure 7

Band structure and bandgap energy of X-FPEA2PbI4 (a) X = p, (b) X = m, and (c) X = o with SOC effect.

The band structure (with and without the SOC effect) exhibits a flat band in the Z–Γ region. This flat band region (Γ–Z) in the band structure corresponds to the x-axis in the real space which represents the quantum barrier effect in the multiple quantum well structure of 2D perovskites. This means that large effective mass and the flat band dispersion have no considerable effect on charge transport and the device’s efficiency. Therefore, to analyze the charge transport inside the device, we calculate the effective mass in the Γ–X dispersive region corresponding to the z-axis of the real space (along with the charge transport direction in the device). The calculated effective masses, and the obtained values of Ebx using Eq. (9) for X-FPEA2PbI4 (X = p, m, and o) perovskites are presented in Table 4. The decrease in Ebx with the change in the F position in p-FPEA2PbI4 compared to o-FPEA2PbI4 is attributed to the increase in dipole moment and dielectric constant.

Table 4 Effective masses, dielectric constants, bandgap energies, dielectric, and exciton binding energies of X-FPEA2PbI4 (X = p, m, and o).

Mechanical properties

Mechanical stability refers to the ability of a material or structure to maintain its shape and resist deformation under applied forces or loads. It indicates how well a material can withstand mechanical stress without undergoing failure, such as fracture, yielding, or excessive deformation. Mechanical stability is crucial for ensuring that materials can perform effectively in their intended applications, especially in structural components where safety and reliability are paramount. To investigate the mechanical stability conditions for X-FPEA2PbI4 (X = p, m, and o) perovskites, the elastic constants (Cij) are obtained using DFT calculations and are presented in Table 5. By examining the conditions in Eq. (12), it is evident that all three perovskites are mechanically stable.

Table 5 The elastic constants of X-FPEA2PbI4 (X = p, m, and o).

The mechanical properties of X-FPEA2PbI4 (X = p, m, and o) perovskites, computed using VRH approximations, are presented in Table 6. From the obtained values, it can be concluded that o-FPEA2PbI4 is more flexible than the other two materials (i.e., m-FPEA2PbI4 and p-FPEA2PbI4). The position of the fluorine atom affects the steric hindrance around the cation. In o-FPEA, the proximity of the fluorine to the ammonium group creates a more favorable configuration that allows for greater movement and flexibility in the crystal structure. The specific bonding angles and distances in o-FPEA2PbI4 facilitate easier deformation under stress compared to the more rigid structures of m-FPEA2PbI4 and p-FPEA2PbI4. It should be noted that all three materials are ductile due to Poisson’s ratio being higher than 0.26.

Table 6 Mechanical properties of X-FPEA2PbI4 (X = p, m, and o).

Results of PeLED simulation

Here, a 2D perovskite-based PeLED is presented, as shown in Fig. 8a. Since p-FPEA2PbI4 possesses the most stability among the proposed materials (explained in Section “Results and discussions”–“Simulation methods”), it is chosen as the active layer of the PeLED. In addition, another 2D perovskite (PEA2PbI4) and a 3D perovskite (MAPbI3) are investigated as active layers for comparison.

Fig. 8
figure 8

(a) Schematic of PeLED and (b) The different structures of PeLEDs and energy levels of ETL, HTL, and active layer.

The proposed PeLED is examined using the following layer structure, as shown in Fig. S5 (a-m):

  • Structure 1: HTL (poly[N, N′-bis(4-butyl phenyl)-N, N′-bis(phenyl)-benzidine] (poly-TPD))/Active layer (MAPbI3)/ETL (1,3,5-Tris(1-phenyl-1h benzimidazole-2-yl)benzene (TPBi))

  • Structure 2: HTL (Poly-TPD)/Active layer (PEA2PbI4)/ETL (TPBi)

  • Structure 3: HTL (Poly-TPD) /Active layer (pFPEA2PbI4)/ETL (TPBi)

  • Structure 4: HTL (poly (9-vinyl carbazole) (PVK))/Active layer (MAPbI3)/ETL (1,3,5-Tris(3-pyridyl-3-phenyl)benzene (TmPyPB))

  • Structure 5: HTL (PVK)/Active layer (PEA2PbI4)—ETL (TmPyPB)

  • Structure 6: HTL (PVK)/Active layer (pFPEA2PbI4)/ETL (TmPyPB)

  • Structure 7: HTL (Poly-TPD)/Active layer (PEA2PbI4)/ETL (TmPyPB)

  • Structure 8: HTL (Poly-TPD)/Active layer (pFPEA2PbI4)/ETL (TmPyPB)

  • Structure 9: HTL (Poly-TPD) /Active layer (PEA2PbI4)/ETL (TmPyPB)

  • Structure 10: HTL (PVK)/Active layer (pFPEA2PbI4)/ETL (TPBi)

  • Structure 11: HTL (PVK)/Active layer (PEA2PbI4)/ETL (TPBi)

  • Structure 12: HTL (PVK)/Active layer (pFPEA2PbI4)/ETL (TPBi)

The ETL, HTL, and active layer thicknesses are set at 30, 60, and 80 nm, respectively, and the device area is 0.1 cm2. The simulation parameters for the PeLED with different materials at room temperature include electron affinity (χe), bandgap energy (Eg), dielectric constant (εr), density of states in the conduction band (Nc), density of states in the valence band (Nv), mobility (μ), SRH recombination (A), radiative recombination (B), and Auger recombination (C), as listed in Tables S7 and S8. The energy levels of the materials used in the PeLED are illustrated in Fig. 8b.

IQE is calculated using Eqs. (2730). The results for the proposed PeLED structures are shown in Fig. 9. From these results, it can be seen that the PeLED with MAPbI3 exhibits the lowest efficiency, while PEA2PbI4 has the highest efficiency. The p-FPEA2PbI4-based PeLED achieves lower efficiency than PEA2PbI4; however, according to Table 3, it offers greater stability. The reduced efficiency of p-FPEA2PbI4 can be attributed to its lower exciton binding energy compared to PEA2PbI4. Conversely, the enhanced stability of FPEA2PbI4 arises from the substitution of an F atom for H in the aromatic ring, which has a larger atomic radius, thereby increasing the length of the cation. In addition, changes in the ETL and HTL layers impact efficiency more significantly in 3D perovskites than in 2D perovskites. Therefore, PeLEDs utilizing 2D perovskites exhibit better stability. Based on the efficiency values for all three active layer materials, the combination of TPBi for the ETL layer and the Poly-TPD for the HTL layer yields the highest efficiency values. Consequently, we choose these materials for the ETL and HTL layers of the PeLED structure.

Fig. 9
figure 9

The IQE values of PeLED for different ETLs and HTLs for (a) MAPbI3, (b) PEA2PbI4, and (c) p-FPEA2PbI4.

The IQE values under temperature fluctuations (from T = 290 K to T = 300 K) are presented in Fig. 10. The results indicate that the efficiency fluctuations for PeLEDs with 2D perovskite active layers are significantly lower than those for 3D perovskites. Consequently, PeLEDs utilizing 2D perovskites exhibit better stability. The changes in efficiency due to temperature fluctuations for p-FPEA2PbI4, PEA2PbI4, and MAPbI3 are 0.01 (1/K), 0.0282 (1/K), and 0.038 (1/K), respectively. These findings align with the thermodynamic stability of the perovskite materials. Thus, the more stable the perovskite material used in the active layer, the more stable the efficiency of the PeLED component under temperature fluctuations.

Fig. 10
figure 10

The IQE of PeLED for different temperatures for (a) MAPbI3, (b) PEA2PbI4, and (c) p-FPEA2PbI4.

The EQE value is calculated according to Eq. (31). Figure 11 displays the EQE curves with respect to temperature fluctuations. The variations in EQE with temperature fluctuations for p-FPEA2PbI4, PEA2PbI4, and MAPbI3 are 0.0112 1/K, 0.032 1/K, and 0.043 1/K, respectively. Consequently, the PeLEDs with an active layer of p-FPEA2PbI4 demonstrate better EQE stability in response to temperature fluctuations.

Fig. 11
figure 11

The EQE of PeLED with different active layers (a) MAPbI3, (b) PEA2PbI4, and (c) p-FPEA2PbI4.

Figure 12 presents the light output (L)- current (I) for PeLEDs with different active layers. From these curves, it can be concluded that p-FPEA2PbI4 exhibits better L–I stability against temperature fluctuations, while PEA2PbI4 has a higher light output value than the other materials. The variations of L with temperature fluctuations at 100 mA for p-FPEA2PbI4, PEA2PbI4, and MAPbI3 are 0.00073 1/K, 0.0012 1/K, and 0.0043 1/K, respectively.

Fig. 12
figure 12

The L–I curve of PeLED for different active layers (a) MAPbI3, (b) PEA2PbI4, and (c) p-FPEA2PbI4.

The spontaneous emission spectrum of the PeLED is examined at I = 100 mA, with temperature fluctuations from T = 290 K to T = 300 K. The results for the different active layer materials (i.e., MAPbI3, PEA2PbI4, and p-FPEA2PbI4) are shown in Fig. 13. The full width at half maximum (FWHM) values in the emission spectrum, considering temperature fluctuations, are 0.0516 nm/K for MAPbI3, 0.0245 nm/K for PEA2PbI4 and 0.0156 nm/K for p-FPEA2PbI4. Thus, the change in FWHM for p-FPEA2PbI4 is smaller under temperature fluctuations, indicating that it has better spectral stability.

Fig. 13
figure 13

The spontaneous emission of PeLED with different active layers (a) MAPbI3, (b) PEA2PbI4, and (c) FPEA2PbI4.

In Fig. 14, the changes in the FWHM of the emission spectrum with current variations for MAPbI3, PEA2PbI4, and p-FPEA2PbI4 are measured as 0.00282 nm/mA, 0.00169 nm/mA, and 0.00138 nm/mA, respectively. This data indicates that p-FPEA2PbI4 exhibits the smallest change in FWHM with current variation, demonstrating superior spectral stability compared to MAPbI3 and PEA2PbI4. This enhanced stability suggests that p-FPEA2PbI4 can be more reliable for applications requiring stable optical performance under varying operational conditions.

Fig. 14
figure 14

The spontaneous emission of PeLED for different currents for (a) MAPbI3, (b) PEA2PbI4, and (c) FPEA2PbI4.

The chromaticity coordinates of the LED emission are plotted on the Commission Internationale de L’Eclairage (CIE) 1931 color space diagram, as shown in Fig. 15. The position of the chromaticity point, labeled “B”, indicates that the LED emits light concentrated in this portion of the visible spectrum. The proximity of the chromaticity point to the outer boundary of the color space triangle suggests that the LED has an excellent degree of color purity at (0.7923, 0.2706), (0.400, 0.596), and (0.2015, 0.75512) for MAPbI3, PEA2PbI4, FPEA2PbI4, respectively.

Fig. 15
figure 15

The CIE 1931 chromatic coordinates for the PeLEDs with (a) MAPbI3, (b) PEA2PbI4, and (c) FPEA2PbI4.

By examining these results, we observe that the stability of PeLEDs with an active layer of 2D perovskites (PEA2PbI4 and p-FPEA2PbI4) against temperature fluctuations is superior to that of the 3D perovskite (MAPbI3). On the other hand, among the two 2D perovskites, PEA2PbI4 exhibits greater efficiency than p-FPEA2PbI4, although it is less stable. Therefore, we conclude that there is a trade-off between the stability and efficiency of PeLEDs. Table 7 indicates that p-FPEA2PbI4 is more stable than the other materials, while PEA2PbI4 exhibits higher efficiency compared to the others. Overcoming the trade-off between stability and efficiency in 2D perovskite-based PeLEDs is achievable through a combination of chemical modifications, structural optimization, and interface engineering. By carefully adjusting material properties, optimizing device structure, and improving interfaces, we can design PeLEDs that exhibit both high efficiency and long-term stability, making them suitable for practical applications.

Table 7 The tradeoff between properties of PeLED.

Conclusion

In this study, we conducted a comprehensive analysis of the stability and opto-electro-mechanical properties of RP 2D perovskites, focusing specifically on X-FPEA2PbI4, where X denotes different configurations: para (p), meso (m), and ortho (o). Our findings provide critical insights into how these variations influence both stability and efficiency, which are essential factors for their application in perovskite light-emitting diodes (PeLEDs). The comparative analysis of the three configurations indicated that p-FPEA2PbI4 stands out with a formation energy of − 4.825 eV, demonstrating superior thermodynamic stability compared to m-FPEA2PbI4 (− 4.647 eV) and o-FPEA2PbI4 (− 4.580 eV). This enhanced stability is particularly significant in practical applications, where environmental factors can lead to material degradation. Furthermore, p-FPEA2PbI4 exhibited lower efficiency variations under temperature fluctuations, quantified as 0.01 1/K, in contrast to 0.0282 1/K and 0.0415 1/K for PEA2PbI4 and MAPbI3, respectively. This stability under varying real conditions positions p-FPEA2PbI4 as a highly reliable candidate for PeLED applications. With temperature fluctuations, the values of FWHM in the emission spectrum are 0.0516 nm/K for MAPbI3, 0.0245 nm/K for PEA2PbI4, and 0.0156 nm/K for p-FPEA2PbI4. While PEA2PbI4 achieves the highest IQE of 6.289%, the trade-off between efficiency and stability becomes evident when considering the fluctuations associated with each material. The IQE for p-FPEA2PbI4 is 2.285%, which is lower than that of PEA2PbI4, however, it is compensated by its robust stability profile. This balance is crucial for the longevity and performance of PeLEDs, especially in commercial applications where consistent output is required.