This study theoretically explores the spin-polarized current ratio, conductivity, and spin moments in multilayered polymer field-effect transistors. Respective layers are formed by dimerized backbones with repeat units coupled to magnetic side-groups. We show that within a small interval of the gate voltage, a considerable formation of the non-zero spin-polarized current ratio is formed together with a considerable decrease of the current. We demonstrate that this phenomenon is correlated with the formation of an alternating (ferrimagnetic) spin alignment, resulting in a net spin moment, controlled by interaction couplings, such as electron–phonon, magnetic coupling, and intra-site Coulomb (Hubbard) interaction. We demonstrated that by tuning the Fermi energy of respective chains with an applied gate voltage, the spin polarization of the current can change its sign, with no change of the orientation of the total spin. These findings provide insights for optimizing the spin-polarized current ratio through gate voltage modulation and interaction couplings, offering potential applications in spintronic device design.

Organic electronics (OE)1–3 has attracted increasing attention due to several outstanding advantages compared to inorganic electronics, such as flexibility, lower cost, and reduced environmental impact.4–6 However, OE faces challenges in terms of conductivity, durability, and thermal stability,7–10 particularly in high-power or high-frequency applications. Similarly, organic spintronics (OS)11–18 is advancing at a pace comparable to inorganic spintronics (IS),19–21 offering benefits similar to OE. One notable advantage of OS is the significantly longer spin relaxation time (SRT),22,23 attributed to the weak spin–orbit interactions (SOIs) in organic materials because the strength of spin–orbital interaction scales with the atomic number Z as Z 4. The OS materials, thus, commonly provide the SRT at least 1 μs.15 Nevertheless, organic materials suffer from lower spin injection efficiency, which limits their potential in specific spintronic applications.24,25

Most methods for producing magnetic organic materials differ significantly from those used for magnetic inorganics. Due to the weak SOI, especially 1D organic materials are required to generate spin signals for OS applications, while inorganic spintronics can rely in this regard on magnetism or the strong inherent SOI within inorganic materials or phenomena like, e.g., inverse spin Hall or Rashba–Edelstein effects.19–21 However, due to the long SRT, also PEDOT:PSS composite polymer reveals measurable inverse spin Hall voltage.26 Besides, Hu et al. theoretically show that in quasi-1D organic ladder polymers, a polaron moving under a static electric field can produce a significant, rapidly oscillating spin Hall signal.27 

In the case of organics, magnetism can be achieved either in chiral molecules,14,17,18 when the spin formation is believed to be coupled with the helical potential, or by linking organic molecules with magnetic functional groups (MFG)28–30 through exchange coupling between the magnetic ions of the MFG and the electrons of the organic molecules.31–34 This magnetic exchange coupling can be either ferromagnetic or antiferromagnetic (ferrimagnetic). However, in both cases, the resulting spin moments induced in the organic material may exhibit either ferromagnetic or antiferromagnetic (ferrimagnetic) coupling, depending on the specific interactions within the system. In addition to the above methods, magnetic signals induced by photoexcitation of an organic-based semiconductor were also observed.35 

The viability of OS is assessed by whether a device can generate a significant magnetic signal through electric induction. The spin-polarized current ratio (SPR), which measures the ratio of different spin-charge currents, effectively indicates magnetic materials with a magnetic order. It has been evaluated to 100% when the organic ferromagnet is connected to metallic electrodes as a good spin filter.36 Even by changing the electrodes to ferromagnetic metallic ones, significant magnetoresistance can be produced.37 In this study, we propose a magnetic polymer multilayer field-effect transistor (FET),38–41 as depicted in Fig. 1, left. In this design, we note that polymer chains are horizontally oriented and coupled to metallic leads so that the current is fully controlled only by the on-chain charge transfer process. To stabilize the spin states, the length of chains is below 10 nm, i.e., below the threshold of the spin diffusion length, which is commonly several tens of nanometers.15 However, when the source-to-drain voltage V SD is applied, the length L of the chain can be even longer due to the following reasons. The travel time along the chain equals to τ = L 2 μ V SD ( μ is the charge mobility). Taking just the lowest common experimental value for the source-to-drain voltage V SD = 0.1 V in FETs in the Ohmic regime and possibly the lowest values of the bulk mobility in organic systems μ = 10 4 c m 2 V 1 s 1, we find for the travel time τ at the length of 10 nm the value τ = 10 7 s, which is by an order of magnitude lower than the SRT being at least 1 μ s.15 More, for high on-chain mobilities of ca. μ = 1 c m 2 V 1 s 1, the length of the chain L could be almost by two order of magnitudes longer and no spin relaxation would occur.

FIG. 1.

Left, the schematic structure of a magnetic polymer multilayered field-effect transistor. The sandwiched polymer layers are deposited on the oxide insulator (OX). The polymer chains within layers are oriented in the source-to-drain (S–D) direction. The distance between repeat unit sites and the thickness of the polymer layer are a u and u z, respectively. t ox is the thickness of the insulator, which is placed above the gate electrode G. The source electrode S is grounded. Right, the respective polymer layers are formed by chains with a backbone composed of repeat units with alteration of single and double bonds. The repeat units are linked to electron-rich functional groups containing metals, which enables formation of induced magnetic moments { S i } inside these groups, coupled due to the spin–spin interaction with local spins { σ i } residing on respective repeat units.

FIG. 1.

Left, the schematic structure of a magnetic polymer multilayered field-effect transistor. The sandwiched polymer layers are deposited on the oxide insulator (OX). The polymer chains within layers are oriented in the source-to-drain (S–D) direction. The distance between repeat unit sites and the thickness of the polymer layer are a u and u z, respectively. t ox is the thickness of the insulator, which is placed above the gate electrode G. The source electrode S is grounded. Right, the respective polymer layers are formed by chains with a backbone composed of repeat units with alteration of single and double bonds. The repeat units are linked to electron-rich functional groups containing metals, which enables formation of induced magnetic moments { S i } inside these groups, coupled due to the spin–spin interaction with local spins { σ i } residing on respective repeat units.

Close modal

For respective polymer chains, we assume that MFG are attached to the polymer backbone (Fig. 1, right). However, since the polymer host is inherently non-magnetic, the magnetic coupling between the polymer and the MFG is expected to be relatively weak, resulting in a weak Zeeman effect and small spin moments within the polymer. Depending on the structure of the polymer backbone, the polymer can exhibit either ferromagnetic or antiferromagnetic (ferrimagnetic) behavior. Due to the long SRT in both ferromagnetic and antiferromagnetic (ferrimagnetic) polymers, a substantial SPR is expected when a source-to-drain voltage ( V SD ) is applied. Contrary to OS materials exhibiting the non-zero SPR due to their chirality, we show that the SPR is formed intrinsically as a consequence of the antiferromagnetic (ferrimagnetic) order. A pronounced SPR is observed within a narrow range of gate-source voltages ( V GS ) and can reverse signs with changing of the V GS polarities, highlighting its strong potential for applications in spintronics.

In addition to the introduction, the theoretical section outlines the self-consistent approach for solving the quantum theory-based distribution of induced charges, polymer unit displacements, and the electric potentials along each polymer layer. Simultaneously, the coupling among layers and influence of the gate voltage is described by means of the classical physics in accordance with their macroscopic character and lower energy of mutual interaction. Furthermore, the Keldysh formalism was employed to compute the spin-dependent currents for the evaluation of the SPR.42,43 The results and discussion section provide a thorough interpretation of the calculation outcomes. Finally, a clear conclusion is presented in the closing section.

The nanoscaled multilayered field-effect transistor (FET) we will investigate is schematically drawn in Fig. 1. When such FET system is in an on-regime, the following phenomena take place:

(i) The applied source-to-drain voltage V SD > 0 switches-on the current of charge carries along the chain (the so-called x-direction). The local intra-unit Coulomb (Hubbard) interaction and the dimerization effect on the backbone induce a spin resolution of charges. This effect is amplified by the spin–spin interaction between itinerant spin-charges on repeat units and induced magnetic moments on functional side-groups. Due to the partial non-homogeneity of the spin-charge distribution near both ends, also the regular dimerization of bonds partially changes by the charge-lattice coupling. Further, as the length of chains is taken below their conjugation length, the on-chain kinetic should be treated fully by means of the quantum mechanics.

(ii) The applied gate-to-source voltage V GS < 0 brings two effects. First, it induces the charge carries (holes) inside the whole semiconducting sample. For the second, it forms the thin conducting channel near the polymer semiconductor—insulator boundary with the strongest gradient of both charge concentration and electric field in the vertical (so-called z−) direction, perpendicular to the insulated gate electrode. However, in the steady state, both diffusion current stemming from the charge concentration gradient and oppositely oriented drift current created by the gate voltage fully compensate each other in the vertical (z−) direction. Hence, we can restrict the charge motion just to the x−, i.e., source-to-drain, direction. Next, as both the diffusion coefficient D z and the mobility μ z in the vertical direction are given by the inter-chain charge transfer processes, they are controlled by the thermally activated hopping kinetics.

Since the on-regime of the FET assumes both drain and gate voltages applied simultaneously, the spin-resolved charge and current distribution should be solved self-consistently, together with the total charge density and the vertical charge distribution given by the gate voltage as well as with the on-chain spin-charge delocalization controlled by the drain voltage and coupled with the lattice displacement. Such theoretical description of the operating device is given in the following paragraphs.

Throughout the article, we assume that charges are generated both intrinsically (by dopants or presence of impurities in the FET structures) and induced by voltages applied to the contacts. For the first case, the sample is always macroscopically neutral; however, holes at the density δ h are created uniformly on the backbones of polymer chains. For the second case, i.e., charge induction by the gate voltage, the sample becomes charged with the induced density δ ρ ( z ). Then, on polymer chains, both contributions are summed up together. In the polymer multilayered field-effect transistor (FET) structure, as shown in Fig. 1, we propose for the distribution of the gate-induced charge density δ ρ ( z ) in the vertical ( z ) direction the following ansatz:
(1)
where δ ρ ( z ) = q δ n ( z ), δ n ( z ) and V(z) are the induced carrier density and the potential at height z, respectively, and q is the unit of the electric charge ( 1.6 × 10 19 C ). δ n, positive or negative, corresponds to the induced electron or hole density. In simulations below, we restrict ourselves to the hole conducting samples. Denoting Dz the hole diffusion coefficient and μz the hole mobility in the z-direction, we easily find that for ansatz (1), the total (diffusion + drift) current j z in the vertical direction is zero as j z = D z z δ ρ ( z ) μ z δ ρ ( z ) z V ( z ) = δ ρ ( z ) ( q k B T D z μ z ) z V ( z ) = 0 due to the Einstein relation. Hence, the ansatz implicitly satisfies the thermalized state condition. The strong gate field brings dominant divergences of the electric field in the vertical direction so that the Gauss law can be restricted just to the vertical direction as follows:
(2)
in which ϵ s is the dielectric constant of the polymer layers. When we substitute Eq. (1) into Eq. (2), the induced carrier density along the z-axis is found to follow the relation:29 
(3)
Taking into an account the shape of this dependence, the FET can be seen as a capacitor composed of a thin conducting channel and the gate metal separated by an insulator. The total gate-induced “surface” charge density, i.e., volume density integrated over thickness of the sample, is represented by the “capacity” relation as follows:
(4)
where zL is the thickness of the polymer film. The left-hand side of Eq. (4) approximates 2 ϵ s k B T δ n ( z = 0 ) at large values of z L. Then, from Eqs. (3) and (4), we find that the interface charge density δ n ( z = 0 ) can be written as
(5)
Comparing Eqs. (4) and (5), we see that while the gate-induced “surface” charge density scales linearly with ( V ( z = 0 ) V G ), the gate-induced “volume” charge density at the bottom of the sample scales quadratically with ( V ( z = 0 ) V G ) . Consequently, the “effective” thickness of the gate-induced conducting channel 0 z L δ n ( z ) d z δ n ( z = 0 ) 1 ( V ( z = 0 ) V G ), i.e., it decreases with the gate voltage. When the drain-to-source voltage V DS is applied, the induced charge density and the electric potential become both vertical height- and horizontal position-dependent. Therefore, Eq. (1) can be rewritten in the discrete form as
(6)
where δ n ( k , i ) is the induced charge density per the unit volume and k and i represent the kth layer and ith polymer repeat unit, counted from the bottom and the left side (source electrode) of FET, respectively. Notably, the layer number k = 1 and k = nL are corresponding to the first and top layers, respectively. For the FET structures, the vertical electric field on the top layer can be set as zero. Besides, for the Ohmic limit, when the source-to-drain-to-source voltage is much lower than that of the gate-to-source | V SD | | V GS |, we can approximate the profile of the potential along the most occupied bottom layer by a linear function. Thus, the potential on the first layer ( k = 1 ) along the source-to-drain direction follows the formula:
(7)
where V S and V D are source and drain and voltages; L is the channel length between the drain and source electrodes; and x ( i ) is the position of the ith polymer unit from the source electrode. For the grounded source contact, we set V S = 0. Expressing also Gauss law (2) in discrete space variables ( k , i ) similarly as in Eqs. (6) and (7), together with the boundary condition at the interface between the sample and the insulator, we find
(8)
(9)
where E I and E ( k , i ) are the projections of the vertical electric fields (facing to the gate) inside the insulator and the polymer semiconductor, respectively, at the points ( k , i ), while u z is the layer thickness in the vertical direction. From Eqs. (8) and (9), and assuming the electric field on the top layer is zero, hereafter, we have the following electric potential equations for calculating the electric potential on each position:
(10)
(11)
and consequently,
(12)

We can, thus, directly see that the discrete sets of induced charge densities δ n ( k , i ) and potentials V ( k , i ) inside the sample are unambiguously externally controlled by potentials at the source, drain, and gate contacts.

Each chain in the kth polyacetylene-like polymer chain (layer) is composed of N repeat units linked to the magnetic functional groups. Then, its Hamiltonian can be written as
(13)
The first term H T ( k ) is the kinetic energy of phonon-controlled π electrons (polarons) on a polymer chain formed by local transfer integrals. It is written in the mathematical form as
(14)

The equation means that the electrons hop between nearest-neighbor on-chain repeat units (sites), i.e., i i + 1, through the transfer integral t k i ; k i + 1 and simultaneously interact with phonons, represented in the adiabatic representation. Summation goes over all N sites and both spin orientations. The second term in the square brackets represents the phonon coupling with electrons, i.e., the electron–phonon interaction, where α and u k i are the phonon coupling constant and the displacement of the ith repeat unit, respectively. The third term in the square brackets, ( 1 ) i t e, is employed to break the energy’s degeneracy, which is also used to simulate the regular alteration of single and double bonds in the polymer chain. c ^ k i σ + and c ^ k i σ are the electron creation and annihilation operators acting on the ith site with the spin σ. It is worth mentioning that the electron–phonon Hamiltonian includes any possible phonon modes without explicitly including their eigenfrequency so that they simulate the effective reaction coordinate of phonons.

The second term in Eq. (13), H P ( k ), represents the electronic (potential) energy term of the kth polymer, which can be written as
(15)
where q V ( k , i ) is the orbital energy of an electron on the ith site. For an electrically unbiased polymer, q V ( k , i ) is constant along the chain. However, upon biasing the polymer with the source-to-drain voltage V SD, q V ( k , i ) changes along the polymer chain. The values of discrete sets of { V ( k , i ) } were determined by Eqs. (7), (10), and (11).
The third term in Eq. (13) represents the intra-site Coulomb (Hubbard) interaction between two electrons with mutually opposite spins residing at the same site, that is,
(16)
where U is the intra-site Coulomb (Hubbard) interaction constant; n ^ k i σ c ^ k i σ + c ^ k i σ is the operator of the electron density with the spin σ on the ith unit site; and σ ¯ means the opposite spin to σ. For simplicity, we can employ the mean-field approximation to Eq. (16) resulting in
(17)
in which n k i σ means the expectation value of n ^ k i σ, i.e., n k i σ = n ^ k i σ.
The fourth term in Eq. (13) determines the magnetic interaction between on-chain sites and the magnetic functional groups, which can be written as
(18)
where J s is a positive constant, corresponding to a ferromagnetic coupling between the magnetic spins in the functional groups S ^ k i and the on-chain spins σ ^ k i, in the site position ( k , i ). Formally, the interaction in Eq. (18) is a 3D problem. However, due to a notable geometrical anisotropy, we can restrict it to the 1D problem. Hence, such a dimensional restriction defines effectively an axis, onto which the spins can be projected. Thus, projections of spins on repeat units onto such axis can be taken formally as “up” ( ) and “down” ( ). From Eq. (18), we can realize that the spin moment of the functional groups can be seen as a magnetic field source to the on-chain spins. It leads to the Zeeman energy splitting for the conducting electrons in the polymer layers and results in a magnetization of the polymer backbone. The mean value of the spin moment of the ith polymer repeat unit is then σ ( k , i ) = 1 2 ( n k i n k i ). Because the functional groups are bulky with respect to the coupled repeat units of the polymer backbone, we can take the mean-field approximation to calculate both magnetic moments self-consistently. For the mean value of the projection of spins on respective side-groups, we then find
(19)
where Z m ( k , i ) is the partition function, Z m ( k , i ) = m = s m = s e J s m σ ( k , i ) / k B T, for s m s, and s is the spin quantum number of the magnetic functional group. Then, in Eq. (18), we can also employ the mean-field approximation S ^ k i S ( k , i ) for the spins of the MFG and we obtain
(20)
In this context, from Eq. (19), it stems that the sign of the coupling constant J s is not important for the amplification of the magnitude of the on-chain itinerant spins σ ^ k i. Upon a change J s J s, S ( k , i ) changes as S ( k , i ) S ( k , i ). Then, the mean-field Hamiltonian (20) is invariant against the symmetry J s J s. Due to the Zeeman energy splitting, the orbital energies become effectively σ-spin dependent, i.e., it undergoes the following change q V ( k , i ) q V ( k , i ) σ J s S ( k , i ). The sign of σ ( or ) selects the sign in the Zeeman energy split, while the “external” magnetic moment S ( k , i ), caused by linked functional groups, is self-consistently dependent on the mean value of the on-chain magnetic moments σ ( k , i ).
The final term of Eq. (13) represents the potential energy of the spring between the nearest-neighbor on-chain units, which has the classic form,
(21)
where k S is the spring constant.
We note that after introducing the mean-field approximations, the total Hamiltonian H in Eq. (13) can be expressed in the single-orbital representation, and the spin-dependent eigenfunction of an electron on the polymer chain can be constructed as a linear combination of states localized on single repeat units. We also note that from the algebraic structure of the Hamiltonian (13) in the site representation it follows that its off-diagonal matrix blocks H σ σ ¯ in spin variables are zero. Then, eigenstates for “up” and “down” spins can be separated and written in the form,
(22)
where | k , i , σ means the atomic orbital state with spin σ ( σ = or ) on the ith repeat unit of the kth chain and the index λ numerates both eigenvectors | Ψ λ σ ( k ) and corresponding eigenvalues E λ σ ( k ) while ϕ λ ; k i σ are expansion coefficients of the eigenvector | Ψ λ σ ( k ) . The spin-dependent total Hamiltonian H σ ( k ), spanned on the basis | k , i , σ and represented in the matrix form, can be written as
(23)
The matrix elements are defined as a i j ( k ) ( q V ( k , i ) + U n k i σ ¯ 1 2 σ J S S ( k , i ) ) δ i j t k i ; k j + 1 2 k S i ( u k i + 1 u k i ) 2. We note that values of V ( k , i ) were determined in Sec. II A. As the multilayer system is in contact with metallic leads and the insulator, the thermal equilibrium systems exchange heat and charges. Therefore, the grand thermodynamic (Landau) potential Ω H T S μ N , with the (room) temperature T and the chemical potential μ as thermodynamic constraints of the system, should take the minimum value for each layer. Therefore, for the eigenenergy E λ σ ( k ) = Ψ λ σ ( k ) | H σ ( k ) | Ψ λ σ ( k ) , we find
(24)
For the eigenenergies E λ σ = E λ σ ( k ) ( { u k i } ), we see that they are explicitly dependent on the instant values of the lattice displacements { u k i }. Because of the chain coupling to metallic contacts, the thermodynamical system is defined by the temperature T and the Fermi energy ε f ( k ). Then, the equilibrium takes place, when the thermodynamic grand (Landau) potential Ω H T S ε f N = λ σ ( E λ σ ( k ) ( { u k i } ) ε f ) f ( E λ σ ( k ) ) k B T λ σ { f ( E λ σ ( k ) ) ln ( f ( E λ σ ( k ) ) ) + ( 1 f ( E λ σ ( k ) ) ) ln ( 1 f ( E λ σ ( k ) ) ), with f ( E λ σ ( k ) ) being the Fermi–Dirac distribution, takes its minimum. It can be analytically shown that it only happens when the following relation takes place
(25)
Equation (25), thus, minimizes a sum of eigenenergies populated according to the Fermi–Dirac distribution concerning the vibrational displacements. It provides the following relations for static displacements:
(26)

We also note that the last “additive” term in Eq. (26) guarantees the conservation law of the spring length. Eigenvalues, eigenvectors, and ensemble of displacements { u k i } in Eqs. (22)–(26) are coupled together, and we must solve them self-consistently.

The numerical calculation proceeds as follows. First, from Eqs. (7)–(11), we obtain profiles of local potentials V ( k , i ), which will be fixed throughout the calculation. Then, the local charge densities δ n ( k , i ) will be estimated from Eq. (6). These values will be taken as an initial input for the mean densities δ n ( k , i ) = n k i + n k i at repeat units. We then assume that the local densities can partly delocalize along the chain due to the quantum processes. For these processes, we take an eigensolution to Eqs. (22) and (23). We also note that during the charge delocalization, the total sum of the on-chain charges δ n ( k , i ) = i ( n k i + n k i ) is conserved so that we can adjust, accordingly, the Fermi energy ε f ( k ) for each chain separately. Having determined the expansion coefficients ϕ λ ; k i σ, we also determine displacements { u k i }, as well as new mean values for σ ( k , i ), S ( k , i ) , and δ n ( k , i ). We procced this iterative procedure until we obtain self-consistency.

Since we have obtained distributions of the spin-polarized charges and repeat unit displacements from the above calculations for each polymer chain, respectively, we can then evaluate the spin-resolved charge current I σ ( k ) along the respective k-chain, with quantum-controlled delocalization of spin-charge distribution. We use the same method as in Refs. 42 and 43 based on Keldysh nonequilibrium Green’s function method. We get
(27)
Here, f L ( R ) is the Fermi–Dirac distribution function for the left (right) contact and Γ L ( R ) the coupling function between the chain and the left (right) contact is approximately proportional to the density of left (right) lead states. Here, we assume that both Γ coupling functions are identical and independent of the incident energy. Then, the current I σ ( k ) is proportional to the factor Γ R Γ L Γ R + Γ L. The results presented below are, thus, exactly up to this multiplicative factor. G σ r ( a ) ( k , ε ) is spin-dependent retarded (advanced) Green’s function of the respective k-chain defined as
(28)
which can be calculated utilizing the Hamiltonian H σ ( k ) in Eq. (23). In Eq. (28), “i” stands for the “imaginary unit.”
For evaluating the degree of spin-resolved polarization of the charge current, we define SPR,
(29)
where I is the total current, i.e., I = I + I .

We remark that due to the symmetry of the Hamiltonian with respect to spin orientations ( σ =↑ or ), the solution is ambiguous against the spin orientation ↑↔↓ inversion. Thus, the simulation below always corresponds to one selected branch of the bi-stable solution. More, in the model Hamiltonian, we neglected potential coupling of spins between respective chains. This coupling contributes to the formation of domains with an identical “choice” of the polarization in the bulk. Presented results below then show dependences on applied voltages V G S and V S D within such domains. During calculations, values of applied voltages are continuously changed and the selected spin orientation of previous calculation is taken as the initial input for the next set of voltages. In this way, the initially chosen sign of the bulk polarization is preserved for all calculations performed for different values of voltages. Thus, the obtained results correspond to common experiments, where a time window between respective measurements is faster than the relaxation of the spin-charge distribution.

For the numerical simulation, we set the following parameter values. The dimensions a u , u z, t o x , and S a in the field-effect transistor are 0.15 nm, 1.6 nm, 50 nm, and 0.24 nm2, respectively. The number of repeat units is 50, corresponding, thus, to the length of respective chains of ca. 7.5 nm. The number n L of stacked chains is 30. The permittivities of the insulating oxide ϵ I and the polymers ϵ s are 3.9 and 4.2 times the vacuum permittivity ϵ 0, respectively. Unless otherwise stated, the parameters α and k s in the electron–phonon interaction are 2 eV/Å and 20 eV/Å2, respectively. The temperature in all calculations is set to 300 K. We also assume that the conducting chain contains a small monomer intrinsic density of holes δh. Both values of the spectral function Γ R and Γ L are set to 1 × 10 2 eV. The values of the spin–spin Js and the intra-site Coulomb interaction U parameters used below are introduced in eV n m 6 units.

Spin-charge distribution in each polymer layer in the multilayer FET is formed through the following effects. We observed that the spin moments along the polymer layer alternate ferrimagnetically, resulting in a net spin moment. This alternating alignment is favored because ferrimagnetic states can evolve through the second-order perturbation in t / U coupling, thereby reducing the total energy. The edge effect is pronounced due to the finite length of the polymer, with spin moments being most prominent near the polymer edges. This pronounced edge effect is caused by the electron localization, which enhances the spin moments. Away from the edges, the spin-charges become more delocalized, the spin moments decrease due to finite intra-site Coulomb (Hubbard) interaction U within the polymer units, reducing the overall Coulombic energy. As expected, the spin moments are influenced by the on-chain interaction couplings within each polymer layer, such as electron–phonon and intra-site Coulomb couplings, as shown in Figs. 2(a) and 2(b). Additively to that, we also note a significant influence of the magnetic coupling of repeat units with magnetic functional groups [Fig. 2(c)] on the spin moments. Figure 2(a) demonstrates that the electron–phonon interactions diminish the amplitude of ferrimagnetic order of spin moments along the chain. Figure 2(b) shows that the intra-site Coulomb interaction U enhances both the magnitude of the spin moments and the net spin moment because the coupling U increases the energy split between the up- and down-spin electrons. Figure 2(c) shows that the magnetic coupling J s between repeat units and side-groups increases the spin moments both near and far from the edges, resulting in a significant enhancement in the magnitude of the alternating spin moments. This is expected as that the mean-field Hamiltonian of the magnetic coupling is invariant against a change of the sign of the coupling constant J s (see above). Thus, this coupling will amplify the local magnitude of the on-chain spins, but this cannot change the antiferromagnetic order between nearest-neighbor spins, where the antiferromagnetism arises from dimerization of the chain. The pronounced antiferromagnetic spin moments observed for a larger J s = 0.2 result, thus, from the competition between J s and U on one side and the electron–phonon coupling α on the other side [note its opposite effect in Fig. 2(a)]. Notably, the spin moments farther from the edges are primarily dominated by magnetic coupling J s rather than by the intra-site Coulomb interaction U. The electron density, closely linked to the spin moments, can be controlled either intrinsically or by the gate voltage V G S, with positive and negative V G S corresponding to electron and hole induction, respectively. Obviously, at finite U and J s, the amplitudes of spin moments increase with the gate-controlled electron density (lesser hole density), as shown in Fig. 2(d) for intrinsically generated hole density δ h = 0.02. The role of the on-chain couplings studied above can be also estimated in the current–voltage characteristics.

FIG. 2.

The spin moment σ z of the ith polymer repeat unit for different values of the on-chain interaction couplings (electron–phonon, Hubbard, magnetic coupling between repeat units and side-groups) and the gate voltage. (a) With V G S = 0 and V S D = 0 for various values of the electron–phonon coupling constant α, the intra-site Coulomb interaction is U = 0.5 and the magnetic coupling is J s = 0.1. (b) With V G S = 0 and V S D = 0 for various values of the intra-site Coulomb interaction constant U, the magnetic coupling and the electron–phonon coupling are J s = 0.1 and α = 2. (c) With V G S = 0 and V S D = 0 for various values of the magnetic coupling J s, the intra-site Coulomb interaction and the electron–phonon coupling are U = 0.5 and α = 2. (d) Of the first layer for various values of the gate voltage V G S with V S D = 0, the magnetic coupling, the intra-site Coulomb interaction, and the electron–phonon coupling are J s = 0.1, U = 0.5, and α = 2. The temperature and the intrinsic hole density are T = 300 K and δ h = 0.02. We note a ferrimagnetic order with increased amplitudes near both ends, which increases with the Hubbard and magnetic coupling. Oppositely, the ferrimagnetic order decreases with the electron–phonon interaction.

FIG. 2.

The spin moment σ z of the ith polymer repeat unit for different values of the on-chain interaction couplings (electron–phonon, Hubbard, magnetic coupling between repeat units and side-groups) and the gate voltage. (a) With V G S = 0 and V S D = 0 for various values of the electron–phonon coupling constant α, the intra-site Coulomb interaction is U = 0.5 and the magnetic coupling is J s = 0.1. (b) With V G S = 0 and V S D = 0 for various values of the intra-site Coulomb interaction constant U, the magnetic coupling and the electron–phonon coupling are J s = 0.1 and α = 2. (c) With V G S = 0 and V S D = 0 for various values of the magnetic coupling J s, the intra-site Coulomb interaction and the electron–phonon coupling are U = 0.5 and α = 2. (d) Of the first layer for various values of the gate voltage V G S with V S D = 0, the magnetic coupling, the intra-site Coulomb interaction, and the electron–phonon coupling are J s = 0.1, U = 0.5, and α = 2. The temperature and the intrinsic hole density are T = 300 K and δ h = 0.02. We note a ferrimagnetic order with increased amplitudes near both ends, which increases with the Hubbard and magnetic coupling. Oppositely, the ferrimagnetic order decreases with the electron–phonon interaction.

Close modal

For the output curves ( I vs V SD ) in Fig. 3(a), we see that the slope (conductivity) significantly decreases with the electron–phonon coupling α. We can attribute this to the formation of the dynamical energy disorder by the lattice interaction. For the intra-site Coulomb (Hubbard) coupling, we note that the conductivity only slightly decreases with it [Fig. 3(b)]. Here, two effects compete with each other. While increased ferrimagnetism promotes the alteration of the spin distribution, which promotes the on-chain transfer of spin-charges due to the Pauli-exclusion principle, the locally increased values of the Hubbard terms contribute to the local energy disorder, in turn. Increasing conductivity with the values of the magnetic coupling J s [Fig. 3(c)] may look as counter-intuitive, as it is the coupling of itinerant spin-resolved charges with the side-groups. However, we remind that this coupling promotes the formation of the on-chain ferrimagnetic state, which favors, due to the Pauliexclusion principle, the on-chain transfer of spin-resolved charges between occupied and unoccupied orbitals.

FIG. 3.

The current I as a function of the applied bias V S D for different values of the on-chain interaction couplings (electron–phonon, Hubbard, magnetic coupling between repeat units and side-groups): (a) various α with J s = 0.1 and U = 0.5, (b) various U with J s = 0.1 and α = 2, and (c) various J s with α = 2 and U = 0.5. The gate voltage, the temperature, and the intrinsic hole density are V G S = 1 V, T = 300 K, and δ h = 0.02, respectively, in all figures. We note that the increased electron–phonon coupling decreases the conductivity. The Hubbard coupling has almost no impact on the conductivity because, while it promotes the ferrimagnetic order (Fig. 2) and then also the charge transfer between adjacent repeat units due to the Pauli-exclusion principle, it also increases the on-chain energy disorder. These two effects outbalance each other. The magnetic coupling between polymer backbone and side-groups promotes the ferrimagnetism (Fig. 2) and hence the on-chain spin-charge mobility.

FIG. 3.

The current I as a function of the applied bias V S D for different values of the on-chain interaction couplings (electron–phonon, Hubbard, magnetic coupling between repeat units and side-groups): (a) various α with J s = 0.1 and U = 0.5, (b) various U with J s = 0.1 and α = 2, and (c) various J s with α = 2 and U = 0.5. The gate voltage, the temperature, and the intrinsic hole density are V G S = 1 V, T = 300 K, and δ h = 0.02, respectively, in all figures. We note that the increased electron–phonon coupling decreases the conductivity. The Hubbard coupling has almost no impact on the conductivity because, while it promotes the ferrimagnetic order (Fig. 2) and then also the charge transfer between adjacent repeat units due to the Pauli-exclusion principle, it also increases the on-chain energy disorder. These two effects outbalance each other. The magnetic coupling between polymer backbone and side-groups promotes the ferrimagnetism (Fig. 2) and hence the on-chain spin-charge mobility.

Close modal

In Fig. 4, we calculated the output curves for different intrinsic densities of holes or values of the gate voltage. While in the first case, the hole density changes uniformly in the volume, and in the second case, the sample could be macroscopically charged both positively and negatively; however, due to high intrinsic on-chain hole densities, the sample would be always hole conducting even for V G S > 0. In fact, for V G S > 0, the highest hole concentration is at the top, while for V G S < 0, it is at the bottom of the sample. In Fig. 4(a), we could easily see that the conductivity (slope) is notably non-linear with respect to the intrinsic hole densities. Namely, it superlinearly increases with the hole densities. This non-linearity can also be anticipated from Fig. 4(b). We see that the highest conductivity is for the highest absolute value of V G S, i.e., the highest cumulation of holes either at the top or bottom of the sample. Profiles are almost symmetric against V G S V G S inversion, i.e., induction of small charge densities by the gate voltage is less important (4) than the hole redistribution across the sample. Thus, this indicates that the conductivity is a convex function of the charge density.

FIG. 4.

The current I as a function of applied bias V S D for different values of the intrinsic hole density and gate voltage: (a) various intrinsic hole densities δ h with V G S = 1 V and (b) various gate voltages V G S with δ h = 0.02. The magnetic coupling, Hubbard coupling, electron–phonon coupling, and temperature are J s = 0.1, U = 0.5, α = 2 , and T = 300 K, respectively, in both figures. We note a strong non-linear dependence of the slope (conductivity) with respect to both intrinsic hole density and the gate voltage. This means that the charge mobility is dependent on the charge concentration.

FIG. 4.

The current I as a function of applied bias V S D for different values of the intrinsic hole density and gate voltage: (a) various intrinsic hole densities δ h with V G S = 1 V and (b) various gate voltages V G S with δ h = 0.02. The magnetic coupling, Hubbard coupling, electron–phonon coupling, and temperature are J s = 0.1, U = 0.5, α = 2 , and T = 300 K, respectively, in both figures. We note a strong non-linear dependence of the slope (conductivity) with respect to both intrinsic hole density and the gate voltage. This means that the charge mobility is dependent on the charge concentration.

Close modal

The above-mentioned non-linearity (superlinearity) of the conductivity on the hole concentration means that the on-chain charge carrier mobility becomes dependent on (increasing with) the charge carrier concentration. Namely, within a macroscopic approach, the total current I in thin film FETs becomes I = w L V SD 0 Z L μ ( ρ ( z ) ) ρ ( z ) d z. When the mobility μ becomes independent of the charge density ρ, the current I turns out to be directly proportional to the total surface charge density 0 Z L ρ ( z ) d z, i.e., the conductivity scales linearly with the “surface charge density.” Instead, we observed a notable superlinear dependence of the conductivity ( I V SD ) on the charge density. This indicates that μ ρ > 0. As the charge density in FETs is also controlled by the gate voltage U SG, we should have also μ U SG > 0. Such mobility dependences on the gate voltage in semiconducting polymers were also found experimentally in FETs,44 where the mobility of charges was explicitly controlled by the gate voltage. Increasing mobility dependences with the charge concentrations were explained theoretically45 by the following model. The energy band of highly delocalized on-chain orbitals is broadened due to the partial energy disorder in local potentials and transfer integrals. This broadening creates the so-called deep energy trap tail states. When polymer chains are coupled to metallic leads, the grand canonical ensemble is formed and holes, first, occupy these deep energy trap states with very low mobility. Upon increasing hole concentrations, higher energy states (less disordered) with higher mobility are then populated. Thus, with an increasing concentration of mean density of holes, mean mobility of holes will increase. In our case, the energy disorder states are controlled by the disorder in transfer integrals (vibration-controlled dimerization of bonds and the ferrimagnetic order) and in local potential by the Hubbard term with the ferrimagnetic order. Similarly, the correlation between conductivity and the spin order was recently shown by the dependence of the resistivity R ISHE (inverse spin-Hall effect) on the gate voltage.21 

In Fig. 5, we studied the SPR dependence on the applied source-to-drain voltage V S D in the Ohmic limit. We see that SPR generally slightly decreases with the applied voltage V S D, except some regions of the value of the coupling constant J s. The formation of the SPR for the antiferromagnetic systems was shown to be controlled by the local translation symmetry breaking.46 In our case, the latter is amplified by the finite length of polymer chain with the influence of both ends on the spin-charge distributions, the bonds dimerization and the electron–phonon coupling α within the transfer integral term, see Eq. (14). In Fig. 5(a), we see that the increasing value of the electron–phonon coupling increases the SPR, further we see that this dependence is fully anticorrelated with changes in conductivity in Fig. 3(a). As expected, we see the Hubbard coupling U amplifies the spin resolution of SPR in Fig. 5(b); however, it only slightly decreases the conductivity in Fig. 3(b), because increased local energy disorder due to the Hubbard coupling is partly compensated by increased amplitudes of the ferrimagnetic order in Fig. 2(b), which promotes the spin-charge transfer between occupied and unoccupied orbitals. Concerning the magnetic coupling J s between repeat units and the side-groups, the situation is more complicated. The coupling J s increases the SPR only up to some value. For higher J s, SPR decreases. Anyway, it increases magnitude of local ferrimagnetic order in Fig. 2(c) so that it promotes the spin-charge transfer between occupied and unoccupied orbitals, with a high superlinear increase of the conductivity in Fig. 3(c). Nontrivial dependences of SPR on V S D for different J s can be explained by competing effects within density-of-states (DOS) for almost half-filled orbitals. The formation of the magnetic order in low-dimensional carbon-based nanostructures was found to be controlled with the formation of so-called zero-energy in-bandgap states.47 The local static and dynamic energy disorder contribute to the formation of discrete in-bandgap states within the DOS. Such trap states have opposite impact on the formation of S P R and the charge carrier mobility.

FIG. 5.

The spin-polarization ratio S P R as a function of the applied bias V S D for different values of the on-chain interaction couplings (electron–phonon, Hubbard, magnetic coupling between repeat units and side-groups): (a) various α with J s = 0.1 and U = 0.5, (b) various U with J s = 0.1 and α = 2, and (c) various J s with α = 2 and U = 0.5. The gate voltage, the temperature, and the intrinsic hole density are V G S = 1 V, T = 300 K, and δ h = 0.02, respectively, in all figures. We note that SPR fully increases with both electron–phonon coupling and Hubbard coupling, but relation of the magnetic coupling between polymer backbone and side-groups to the SPR is nontrivial.

FIG. 5.

The spin-polarization ratio S P R as a function of the applied bias V S D for different values of the on-chain interaction couplings (electron–phonon, Hubbard, magnetic coupling between repeat units and side-groups): (a) various α with J s = 0.1 and U = 0.5, (b) various U with J s = 0.1 and α = 2, and (c) various J s with α = 2 and U = 0.5. The gate voltage, the temperature, and the intrinsic hole density are V G S = 1 V, T = 300 K, and δ h = 0.02, respectively, in all figures. We note that SPR fully increases with both electron–phonon coupling and Hubbard coupling, but relation of the magnetic coupling between polymer backbone and side-groups to the SPR is nontrivial.

Close modal

For almost half-filled system due to the relatively low hole concentration, the Fermi energy becomes close to these spin-polarized (zero-energy in-bandgap) states. By tuning the gate voltage, we will not only change the total mean on-chain hole density (intrinsic + gate induced) in the sample, but we can considerably change their vertical distribution between respective layers.

The so-called transfer curves (I vs VGS) are shown in Fig. 6. We see that when the absolute value of the gate voltage satisfies | V G S | 4 V, the smooth profile of the total current changes into the region with local resonant minima, which are sensitive to interaction parameters α, U , and J s. However, inside this region, the current rather increases with |VGS| so that sets of local resonant minima are close to V G S 0. We also note that the current I seems to be partly symmetric against the polarity inversion of the gate voltage. In this region, the current decreases with the electron–phonon coupling α, while it increases with the magnetic coupling J s for negative gate voltage V G S, but it decreases with the magnetic coupling J s for positive gate voltage V G S. On the other hand, for regions with | V G S | > 4 V, the current I significantly decreases with |VGS| for | V G S | > 4 V. In Fig. 7, we see that these changes in the current (Fig. 6) are strongly correlated with the formation of the S P R 0, which appears when | V G S | 4 V. Here, we note a remarkable phenomenon, when for low negative gate voltage, the S P R < 0, while for low positive gate voltage S P R > 0.

FIG. 6.

The current I as a function of the gate voltage V G S for different values of the on-chain interaction couplings (electron–phonon, Hubbard, magnetic coupling between repeat units and side-groups): (a) various α with J s = 0.1 and U = 0.5, (b) various U with J s = 0.1 and α = 2, and (c) various J s with α = 2 and U = 0.5. The source-to-drain voltage, the temperature, and the intrinsic hole density are V S D = 0.01 V, T = 300 K, and δ h = 0.02, respectively, in all figures. We note a considerable decrease in the transfer curves together with resonant features for | V G S | 4 V. The effect is associated with the formation of the SPR (Fig. 7) near the zero-energy in-bandgap states.

FIG. 6.

The current I as a function of the gate voltage V G S for different values of the on-chain interaction couplings (electron–phonon, Hubbard, magnetic coupling between repeat units and side-groups): (a) various α with J s = 0.1 and U = 0.5, (b) various U with J s = 0.1 and α = 2, and (c) various J s with α = 2 and U = 0.5. The source-to-drain voltage, the temperature, and the intrinsic hole density are V S D = 0.01 V, T = 300 K, and δ h = 0.02, respectively, in all figures. We note a considerable decrease in the transfer curves together with resonant features for | V G S | 4 V. The effect is associated with the formation of the SPR (Fig. 7) near the zero-energy in-bandgap states.

Close modal
FIG. 7.

The spin-polarization ratio S P R as a function of the gate voltage V G S for different values of the on-chain interaction couplings (electron–phonon, Hubbard, magnetic coupling between repeat units and side-groups): (a) various α with J s = 0.1 and U = 0.5, (b) various U with J s = 0.1 and α = 2, (c) and various J s with α = 2 and U = 0.5. The source-to-drain voltage, the temperature, and the intrinsic hole density are V S D = 0.01 V, T = 300 K, and δ h = 0.02, respectively, in all figures. A considerable formation of the SPR for | V G S | 4 V, i.e., when the Fermi energy is near the zero-energy in-bandgap states is observed. Note that the polarity of the SPR is changed upon changing the polarity of the gate voltage, however, without changes in the sign of the total net spin moment (Fig. 8). Note also a correlation of the SPR with the current decrease in transfer curves in Fig. 6.

FIG. 7.

The spin-polarization ratio S P R as a function of the gate voltage V G S for different values of the on-chain interaction couplings (electron–phonon, Hubbard, magnetic coupling between repeat units and side-groups): (a) various α with J s = 0.1 and U = 0.5, (b) various U with J s = 0.1 and α = 2, (c) and various J s with α = 2 and U = 0.5. The source-to-drain voltage, the temperature, and the intrinsic hole density are V S D = 0.01 V, T = 300 K, and δ h = 0.02, respectively, in all figures. A considerable formation of the SPR for | V G S | 4 V, i.e., when the Fermi energy is near the zero-energy in-bandgap states is observed. Note that the polarity of the SPR is changed upon changing the polarity of the gate voltage, however, without changes in the sign of the total net spin moment (Fig. 8). Note also a correlation of the SPR with the current decrease in transfer curves in Fig. 6.

Close modal

To avoid misinterpretation for the change of the sign of the S P R, we calculated in Fig. 8 the mean value of the total spin σ t k i σ ^ k i inside the sample. We see that during change in polarity of the gate voltage V G S, the total spin does not change its sign, but it increases continuously. More, near values of V G S 0, the change in σ t is lower. To understand, how it is possible, we remind the definition of S P R in Eq. (28).

FIG. 8.

The total spin σ t as a function of the gate voltage V G S for different values of the on-chain interaction couplings (electron–phonon, Hubbard, magnetic coupling between repeat units and side-groups): (a) various α with J s = 0.1 and U = 0.5, (b) various U with J s = 0.1 and α = 2, and (c) various J s with α = 2 and U = 0.5. The source-to-drain voltage, the temperature, and the intrinsic hole density are V S D = 0.01 V, T = 300 K, and δ h = 0.02, respectively, in all figures. We note a continuous change of the total spin σ t for | V G S | 4 V, even upon changing the polarity of the gate voltage, despite a remarkable change in the polarity of the SPR (Fig. 7).

FIG. 8.

The total spin σ t as a function of the gate voltage V G S for different values of the on-chain interaction couplings (electron–phonon, Hubbard, magnetic coupling between repeat units and side-groups): (a) various α with J s = 0.1 and U = 0.5, (b) various U with J s = 0.1 and α = 2, and (c) various J s with α = 2 and U = 0.5. The source-to-drain voltage, the temperature, and the intrinsic hole density are V S D = 0.01 V, T = 300 K, and δ h = 0.02, respectively, in all figures. We note a continuous change of the total spin σ t for | V G S | 4 V, even upon changing the polarity of the gate voltage, despite a remarkable change in the polarity of the SPR (Fig. 7).

Close modal

The S P R compares only currents, and in fact, only conductivities driven by opposite spins, but not their occupation density. The very critical factor is, thus, different mobilities of charges with oppositely oriented spins. In Fig. 2, we recall clearly resolved ferrimagnetic on-chain order, with very pronounced amplitudes near the chain ends. We also recall that tuning of parameters of interactions as α, U , and J s or the gate voltage V G S, changes amplitudes of opposite spin occupations on the nearest-neighbor sites of the ferrimagnetic state. This will also tune on-chain mobilities of spin-resolved charges in the ferrimagnetic state. In the region, where the gate voltage V G S 0, the Fermi energies of respective chains are almost identical and close (or in resonance) with the zero-energy in-bandgap states. This will amplify the ferrimagnetic order, giving rise to the formation of non-zero S P R. By tuning the gate voltage, the Fermi energies within respective chains cross the discrete set of states, where the S P R can change considerably, promoting, thus, mobility from one spin orientation to the opposite. Our simulation shows that while one side of the in-bandgap DOS promotes mobility with one spin orientation, the other side of the in-bandgap DOS promotes the mobility with the opposite spin orientation (see the scheme in Fig. 9). The discrete nature of such in-bandgap states is easily seen in transfer curves in Fig. 6 due to the resonant local minima. Decreased current in such states can be explained by decreased conductivity (mobility) for the states with increased on-chain energy disorder, which evolves together with enhanced both ferrimagnetic order and S P R. The formation of the zero-energy in-bandgap states was also proved for the interaction of the carbon-based linear chain interacting with metallic leads, where the charge transfer between the chain and leads occurs.48 While the authors in Ref. 48 assumed the preferred choice of the spin polarization in the Hubbard coupling, promoting, thus, a gap between both spin states, in our model on the contrary, in the Hubbard coupling [see Eqs. (16) and (17)], both spin states are energetically equivalent. Thus, their energy separation stems only from the bi-stable solution of the dimerized chain, which breaks the translational symmetry, induces the alteration of spin orientation (similarly as in Ref. 46) and is accompanied with the local charge and spin transfer. The effect is much more amplified near both ends, where the loss of the translational symmetry breaking is more amplified.

FIG. 9.

The schematic figure of the DOS of near-zero-energy in-bandgap states (blue). These states proved the formation of the ferrimagnetic order. Tuning the Fermi energy (red) by the gate voltage, we can map two different regions, where the charge mobility corresponding to one spin orientation is greater than for the opposite orientation. Thus, by tuning the gate voltage, we find not only formation of the current spin-polarization ratio (SPR), but that the sign of the SPR changes, without the change in sign of the net spin of the ferrimagnetic order on the backbone of the polymer chain.

FIG. 9.

The schematic figure of the DOS of near-zero-energy in-bandgap states (blue). These states proved the formation of the ferrimagnetic order. Tuning the Fermi energy (red) by the gate voltage, we can map two different regions, where the charge mobility corresponding to one spin orientation is greater than for the opposite orientation. Thus, by tuning the gate voltage, we find not only formation of the current spin-polarization ratio (SPR), but that the sign of the SPR changes, without the change in sign of the net spin of the ferrimagnetic order on the backbone of the polymer chain.

Close modal

Due to the relatively high intrinsically generated concentration of holes, the applied gate voltage mainly affects the distribution of holes (and the Fermi energy) among chains. In Fig. 10, we show dependences of both S P R ( k ) and current I ( k ) for the respective kth layer on the gate voltage V G S, for the interval | V G S | < 4, where the S P R is formed. We see that S P R and I are almost anti-correlated, but this anti-correlation is not ideal because the gate voltage also influences the density of holes. However, while the current is almost identical upon a change of the gate voltage V G S polarity, the S P R changes its sign.

FIG. 10.

The spin-polarized current ratio S P R (a) and the current I (b) of the kth layer for various values of the V G S at V S D = 0.01 V, with J s = 0.1 and U = 0.5. The intrinsic hole density and the temperature are δ h = 0.02 and T = 300 K, respectively. As the applied gate voltage is in the interval | V G S | 4 V, where significant non-zero values of the SPR occur, the vertical distribution of the source-to-drain current is far from standard FETs. Namely, near the gate contact, the current is smaller due to the formed spin polarization. We note that the gate voltage with the opposite sign changes the polarity of the SPR, but with negligible impact on the current.

FIG. 10.

The spin-polarized current ratio S P R (a) and the current I (b) of the kth layer for various values of the V G S at V S D = 0.01 V, with J s = 0.1 and U = 0.5. The intrinsic hole density and the temperature are δ h = 0.02 and T = 300 K, respectively. As the applied gate voltage is in the interval | V G S | 4 V, where significant non-zero values of the SPR occur, the vertical distribution of the source-to-drain current is far from standard FETs. Namely, near the gate contact, the current is smaller due to the formed spin polarization. We note that the gate voltage with the opposite sign changes the polarity of the SPR, but with negligible impact on the current.

Close modal

We can then conclude that the on-chain mobility of charges with a given spin orientation of one side of in-bandgap spectrum is equal to the on-chain mobility of charges with an opposite spin orientation of the other side of in-bandgap spectrum.

This work establishes a comprehensive theoretical picture of how spin interactions and charge-lattice couplings govern spin-polarized transport in multilayer magnetic polymer field-effect transistors (FETs). Dimerized transfer integrals and on-site Hubbard repulsion stabilize a ferrimagnetic ground state along each conjugated chain, yielding a finite net spin moment further reinforced by exchange with magnetic side-groups. Finite chain length naturally amplifies the spin amplitude at the chain ends.

Gate bias proves to be a powerful tuning knob. A modest positive V G S lowers the intrinsic hole density, strengthens the ferrimagnetic order, and produces a positive spin-polarized current ratio (SPR). Conversely, a comparable negative V G S injects holes, weakens the order, and reverses the SPR sign. The emergence of a pronounced SPR coincides with a drop in the source–drain current and a series of resonant minima, which we trace to the alignment of the Fermi level with spin-polarized zero-energy in-gap states.

Electron–phonon coupling, Hubbard interaction, and backbone–side-group exchange modulate magnetism and transport in distinct ways: increasing the electron–phonon coupling λ or the Hubbard parameter U enhances the SPR but generally reduces conductivity, while the exchange coupling J s shows a non-monotonic influence, boosting mobility at moderate values yet suppressing the SPR at large couplings.

Spatially, the largest SPR is found in lower polymer layers, where gate-induced carriers accumulate. However, excessive spin moments in the very first layer over-stabilize the ferrimagnetic pattern and curtail the SPR, underscoring the need for layer-by-layer optimization.

Overall, our results outline a clear strategy for gate-tunable control of spin-polarized currents in organic spintronic devices: engineer the interplay between lattice dimerization, Coulomb repulsion, and magnetic exchange, and then exploit gate bias to select both the polarity and magnitude of the SPR without reversing the net spin orientation. These insights can guide the design of low-power, electrically controllable organic spin transistors, and allied spin-logic elements, and they invite experimental validation in short-chain magnetic polymers.

We thank the National University of Kaohsiung and Institute of Macromolecular Chemistry in Prague for the research support. P.T. thanks the Ministry of Education, Youth and Sports of the Czech Republic, the program INTER-EXCELLENCE, LUAUS24032, for the financial support.

The authors have no conflicts to disclose.

Shih-Jye Sun: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Software (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). Miroslav Menšík: Conceptualization (equal); Investigation (equal); Methodology (equal); Validation (equal); Visualization (equal); Writing – review & editing (equal). Petr Toman: Investigation (equal); Validation (equal); Writing – review & editing (equal).

The data that support the findings of this study are available from the corresponding authors upon reasonable request.

1.
J. M.
Shaw
and
P. F.
Seidler
, “
Organic electronics: Introduction
,”
IBM J. Res. Dev.
45
,
3
9
(
2001
).
2.
S. R.
Forrest
and
M. E.
Thompson
, “
Introduction: Organic electronics and optoelectronics
,”
Chem. Rev.
107
(
4
),
923
925
(
2007
).
3.
M.
Fahlman
,
S.
Fabiano
,
V.
Gueskine
et al, “
Interfaces in organic electronics
,”
Nat. Rev. Mater.
4
,
627
650
(
2019
).
4.
H.
Ölmez
and
U.
Kretzschmar
, “
Potential alternative disinfection methods for organic fresh-cut industry for minimizing water consumption and environmental impact
,”
LWT Food Sci. Technol.
42
,
686
693
(
2009
).
5.
C.
Liu
,
C.
Xiao
,
C.
Xie
, and
W.
Li
, “
Flexible organic solar cells: Materials, large-area fabrication techniques and potential applications
,”
Nano Energy
89
,
106399
(
2021
).
6.
S.
Forrest
, “
The path to ubiquitous and low-cost organic electronic appliances on plastic
,”
Nature
428
,
911
918
(
2004
).
7.
L. S.
Xie
,
G.
Skorupskii
, and
M.
Dincă
, “
Electrically conductive metal–organic frameworks
,”
Chem. Rev.
120
(
16
),
8536
8580
(
2020
).
8.
K.
Mayandi
,
N.
Rajini
,
N.
Ayrilmis
,
M. P.
Indira Devi
,
S.
Siengchin
,
F.
Mohammad
, and
H. A.
Al-Lohedan
, “
An overview of endurance and ageing performance under various environmental conditions of hybrid polymer composites
,”
J. Mater. Res. Technol.
9
,
15962
15988
(
2020
).
9.
S.
Behzadi
and
M. M.
Farid
, “
Long term thermal stability of organic PCMs
,”
Appl. Energy
122
,
11
16
(
2014
).
10.
M.
Asim
,
M. T.
Paridah
,
M.
Chandrasekar
et al, “
Thermal stability of natural fibers and their polymer composites
,”
Iran Polym. J.
29
,
625
648
(
2020
).
11.
W. J. M.
Naber
et al, “
Organic spintronics
,”
J. Phys. D: Appl. Phys.
40
,
R205
(
2007
).
12.
C.
Boehme
and
J.
Lupton
, “
Challenges for organic spintronics
,”
Nat. Nanotechnol.
8
,
612
615
(
2013
).
13.
X.
Yang
,
A.
Guo
,
L.
Guo
,
Y.
Liu
,
X.
Sun
, and
Y.
Guo
, “
Organic semiconductors for room-temperature spin valves
,”
ACS Mater. Lett.
4
(
5
),
805
814
(
2022
).
14.
J. M.
Abendroth
,
D. M.
Sterner
,
B. P.
Bloom
et al, “
Spin selectivity in photoinduced charge-transfer mediated by chiral molecules
,”
ACS Nano
13
,
4928
4946
(
2019
).
15.
V.
Alek Dediu
,
L. E.
Hueso
,
I.
Bergenti
, and
C.
Taliani
, “
Spin routes in organic semiconductors
,”
Nat. Mater.
8
,
707
716
(
2009
).
16.
D.
Sun
,
E.
Ehrenfreund
, and
Z.
Valy Vardeny
, “
The first decade of organic spintronics research
,”
Chem. Commun.
50
,
1781
1793
(
2014
).
17.
N.
Goren
,
T. K.
Das
,
N.
Brown
et al, “
Metal organic spin transistor
,”
Nano Lett.
21
,
8657
8663
(
2021
).
18.
R.
Gupta
,
A.
Balo
,
R.
Garg
et al, “
The chirality-induced spin selectivity effect in asymmetric spin transport: From solution to device applications
,”
Chem. Sci.
15
,
18751
(
2024
).
19.
G.
Dastgeera
,
A. M.
Afzal
et al, “
Gate modulation of the spin current in graphene/WSe2 van der Waals heterostructure at room temperature
,”
J. Alloys Compd.
919
,
165815
(
2022
).
20.
G.
Dastgeera
,
M. A.
Shehzad
, and
J.
Eom
, “
Distinct detection of thermally induced spin voltage in Pt/WS2/Ni81Fe19 by the inverse spin Hall effect
,”
ACS Appl. Mater. Interfaces
11
,
48533
48539
(
2019
).
21.
G.
Dastgeer
,
M. F.
Khan
et al, “
Surface spin accumulation due to the inverse spin Hall effect in WS2 crystals
,”
2D Mater.
6
,
011007
(
2019
).
22.
S.
Pramanik
,
C. G.
Stefanita
,
S.
Bandyopadhyay
et al, “
Observation of extremely long spin relaxation times in an organic nanowire spin valve
,”
Nat. Nanotechnol.
2
,
216
219
(
2007
).
23.
J.
Rybicki
,
T. D.
Nguyen
,
Y.
Sheng
, and
M.
Wohlgenannt
, “
Spin–orbit coupling and spin relaxation rate in singly charged π-conjugated polymer chains
,”
Synth. Met.
160
,
280
284
(
2010
).
24.
Q.
Tian
and
S.
Xie
, “
Spin injection and transport in organic materials
,”
Micromachines
10
,
596
(
2019
).
25.
J. W.
Yoo
,
C. Y.
Chen
,
H. W.
Jang
et al, “
Spin injection/detection using an organic-based magnetic semiconductor
,”
Nat. Mater.
9
,
638
642
(
2010
).
26.
K.
Ando
,
S.
Watanabe
,
S.
Mooser
et al, “
Solution-processed organic spin–charge converter
,”
Nat. Mater.
12
,
622
627
(
2013
).
27.
G. C.
Hu
,
Y. Y.
Miao
, and
C.
Timm
, “
Oscillating spin Hall effect from polaron transport in organic chains
,”
Phys. Rev. B
106
(
14
),
144309
(
2022
).
28.
C.
Benelli
and
D.
Gatteschi
, “
Magnetism of lanthanides in molecular materials with transition-metal ions and organic radicals
,”
Chem. Rev.
102
(
6
),
2369
2388
(
2002
).
29.
J. S.
Miller
and
A. J.
Epstein
, “
Organic and organometallic molecular magnetic materials—Designer magnets
,”
Angew. Chem. Int. Ed. Engl.
33
,
385
415
(
1994
).
30.
J.
Vallejo
et al, “
Guest-dependent single-ion magnet behaviour in a cobalt(II) metal–organic framework
,”
Chem. Sci.
7
,
2286
2293
(
2016
).
31.
Y.-Z.
Zheng
,
W.
Xue
,
M.-L.
Tong
,
X.-M.
Chen
,
F.
Grandjean
, and
G. J.
Long
, “
A two-dimensional iron(II) carboxylate linear chain polymer that exhibits a metamagnetic spin-canted antiferromagnetic to single-chain magnetic transition
,”
Inorg. Chem.
47
(
10
),
4077
4087
(
2008
).
32.
M.
Ma
,
Q.
Zhang
,
J.
Dou
,
H.
Zhang
,
D.
Yin
,
W.
Geng
, and
Y.
Zhou
, “
Fabrication of one-dimensional Fe3O4/P(GMA–DVB) nanochains by magnetic-field-induced precipitation polymerization
,”
J. Colloid Interface Sci.
374
,
339
344
(
2012
).
33.
J. L.
Wilson
,
P.
Poddar
,
N. A.
Frey
,
H.
Srikanth
,
K.
Mohomed
,
J. P.
Harmon
,
S.
Kotha
, and
J.
Wachsmuth
, “
Synthesis and magnetic properties of polymer nanocomposites with embedded iron nanoparticles
,”
J. Appl. Phys.
95
(
3
),
1439
1443
(
2004
).
34.
S.
Ganguly
and
S.
Margel
, “
Fabrication and applications of magnetic polymer composites for soft robotics
,”
Micromachines
14
,
2173
(
2023
).
35.
Y.
Jung-Woo
et al, “
Multiple photonic responses in films of organic-based magnetic semiconductor V(TCNE)x, x ∼ 2
,”
Phys. Rev. Lett.
97
(
24
),
247205
(
2006
).
36.
G.
Hu
,
Y.
Guo
,
J.
Wei
, and
S.
Xie
, “
Spin filtering through a metal/organic-ferromagnet/metal structure
,”
Phys. Rev. B
75
(
16
),
165321
(
2007
).
37.
G. C.
Hu
,
M. Y.
Zuo
,
Y.
Li
,
J. F.
Ren
, and
S. J.
Xie
, “
Multi-state magnetoresistance in ferromagnet/organic-ferromagnet/ferromagnet junctions
,”
Appl. Phys. Lett.
104
(
3
),
033302
(
2014
).
38.
S.
Scheinert
and
G.
Paasch
, “
Fabrication and analysis of polymer field-effect transistors
,”
Phys. Status Solidi A
201
,
1263
1301
(
2004
).
39.
H.
Sirringhaus
,
M.
Bird
, and
N.
Zhao
, “
Charge transport physics of conjugated polymer field-effect transistors
,”
Adv. Mater.
22
,
3893
3898
(
2010
).
40.
Y. A. O.
Ze-Fan
,
W. A. N. G.
Jie-Yu
, and
P. E. I.
Jian
, “
High-performance polymer field-effect transistors: From the perspective of multi-level microstructures
,”
Chem. Sci.
12
,
1193
1205
(
2021
).
41.
S.
Cheng
,
Y.
Wang
,
R.
Zhang
,
H.
Wang
,
C.
Sun
, and
T.
Wang
, “
Recent progress in GaS sensors based on P3HT polymer field-effect transistors
,”
Sensors
23
,
8309
(
2023
).
42.
S. J.
Sun
,
M.
Menšík
, and
P.
Toman
, “
Spin-polarized currents induced in antiferromagnetic polymer multilayered field-effect transistors
,”
Phys. Chem. Chem. Phys.
26
,
13261
(
2024
).
43.
S. J.
Sun
,
M.
Menšík
,
C.
Ganzorig
,
P.
Toman
, and
J.
Pfleger
, “
Formation of spin-polarized current in antiferromagnetic polymer spintronic field-effect transistors
,”
Phys. Chem. Chem. Phys.
24
,
25999
(
2022
).
44.
M.
Menšík
,
P.
Toman
,
U.
Bielecka
,
W.
Bartkowiak
, and
J.
Pfleger
, “
On the methodology of the determination of charge concentration dependent mobility from organic field-effect transistor characteristics
,”
Phys. Chem. Chem. Phys.
20
,
2308
(
2018
).
45.
P.
Toman
,
M.
Menšík
,
W.
Bartkowiak
, and
J.
Pfleger
, “
Modelling of the charge carrier mobility in disordered linear polymer materials
,”
Phys. Chem. Chem. Phys.
19
,
7760
(
2017
).
46.
J.
Železný
,
Y.
Zhang
,
C.
Felser
, and
B.
Yan
, “
Spin-polarized current in noncollinear antiferromagnets
,”
Phys. Rev. Lett.
119
,
187204
(
2017
).
47.
D. G.
de Oteyza
and
T.
Frederiksen
, “
Carbon-based nanostructures as a versatile platform for tunable π-magnetism
,”
J. Condens. Matter Phys.
34
,
443001
(
2022
).
48.
Y.
Miao
,
S.
Qiu
,
G.
Zhang
,
J.
Ren
,
C.
Wang
, and
G.
Hu
, “
Ground-state properties of metal/organic-ferromagnet heterojunctions
,”
Phys. Rev. B
98
,
235415
(
2018
).