Controlling Phase Transition in Monolayer Metal Di

Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Controlling phase transition in monolayer metal

arXiv:2107.13287v1 [cond-mat.mtrl-sci] 28 Jul 2021

diiodides XI2 (X: Fe, Co, and Ni) by carrier doping


Teguh Budi Prayitno
Department of Physics, Faculty of Mathematics and Natural Science, Universitas
Negeri Jakarta, Kampus A Jl. Rawamangun Muka, Jakarta Timur 13220, Indonesia
E-mail: [email protected]

Abstract. We applied the generalized Bloch theorem to verify the ground state
(most stable state) in monolayer metal diiodides 1T-XI2 (X: Fe, Co, and Ni), a family
of metal dihalides, using the first-principles calculations. The ground state, which can
be ferromagnetic, antiferromagnetic, or spiral state, was specified by a wavevector in
the primitive unit cell. While the ground state of FeI2 is ferromagnetic, the spiral
state becomes the ground state for CoI2 and NiI2 . Since the multiferroic behavior
in the metal dihalide can be preserved by the spiral structure, we believe that CoI2
and NiI2 are promising multiferroic materials in the most stable state. When the
lattice parameter increases, we also show that the ground state of NiI2 changes to
a ferromagnetic state while others still keep their initial ground states. For the
last discussion, we revealed the phase transition manipulated by hole-electron doping
due to the spin-spin competition between the ferromagnetic superexchange and the
antiferromagnetic direct exchange. These results convince us that metal diiodides
have many benefits for future spintronic devices.

Keywords: metal diiodides, spin spiral, phase transition

1. Introduction

Recently, multiferroic properties in materials bring great attention from point of view of
both theoretical aspects and experimental researches [1, 2, 3]. These properties happen
due to a combination of the magnetic and dielectric orders appearing in the same phase.
It then leads to extraordinary applications in spintronics especially for the data storage
devices [4, 5, 6]. However, coupling between those two orders to generate multiferroic
behavior, in general, is very weak. A possible resolution to encounter this problem is
to introduce the magnetically-induced ferroelectricity in some frustrated helimagnets
which leads to the so-called giant magnetoelectric properties [7].
Previous authors reported that the multiferroicity in some layered metal dihalides
may appear in the spiral (SP) structure [8, 9]. This is triggered by the ferroelectric
polarization induced by the magnetic SP structure, thus combining the ferroelectricity
2

and magnetism in the same state. However, there are no sufficient reports if monolayer
two-dimensional metal dihalides may also possess the SP ground state. Former authors
only tabulated the ferromagnetic (FM) state or the antiferromagnetic (AFM) state as
the most stable state and excluded the SP ground state [10, 11]. It can be understood
since exploring the SP state needs a very large unit cell, especially for a long period,
thus yielding high computational cost. The SP state itself is a typical helimagnetic
state with a fixed cone angle [12]. Some applications of the SP structure can be seen in
constructing the domain wall [13, 14, 15].
Metal dihalides are layered compounds due to a weak van der Waals (vdW)
interaction between layers. In the two-dimensional system, the magnetic order in these
materials ignores the Mermin-Wagner theorem which prohibits magnetism due to a
strong suppression from the thermal fluctuation. This happens because the vdW type
yields the magnetic anisotropy to keep a long-range magnetic order that overcomes the
thermal fluctuations, thus leading to the intrinsic ferromagnetism [16, 17]. Therefore,
it is believed that the two-dimensional metal dihalides are more prominent than any
other two-dimensional materials such as graphene and MoS2 , which have no intrinsic
ferromagnetism. Besides, we are also allowed to exfoliate the layered structure into a
stable monolayer structure [18]. In the monolayer limit, the magnetism in the metal
dihalides can be tuned either by strain [19] or by Hubbard U [20]. It has also been
reported that the magnetism in the monolayer metal dihalides exhibits a topological
insulator [21] and a half-metal [22].
Our purpose is to explore the SP ground state in the monolayer metal diiodides XI2
(X: Fe, Co, and Ni), one of the families of metal dihalides, by utilizing the generalized
Bloch theorem (GBT), which is widely used to investigate the SP ground state in some
materials [23, 24, 25, 26], to minimize the high computational cost. The interesting
case of metal diiodides, CoI2 and NiI2 exhibit multiferroicity for the layered structures
[27, 28]. In this case, we use a primitive unit cell that consists of a 3d transition magnetic
atom and two non-magnetic I atoms. We set a flat spiral formation instead of a conical
spiral to include the FM, SP, and AFM states as the most stable states. These states
will be assigned by a spiral vector from the rotation of the magnetic moment of the
magnetic atom.
We confirm that the ground state of CoI2 and NiI2 is an SP ground state, in good
suitability with the layered systems [27, 28]. Meanwhile, we show an FM ground state
in FeI2 , which also confirms a half-metallic behavior as reported in Ref. [29]. When we
enhance the lattice parameter, the SP ground state of NiI2 changes to the FM ground
state, thus showing a sensitivity of state to the large lattice parameters. Meanwhile,
FeI2 and CoI2 still preserve their ground states, showing the robustness of state to the
lattice parameters. However, the magnetic moment for all the systems seems robust
to the lattice parameter. Based on these results, we assure that multiferroicity in the
metal dihalides should also occur in the monolayer systems.
For the last session, we display the dependence of ground state on hole-electron
doping, which reveals the phase transition of state in the certain interval of doping.
3

Since the phase transition depends on the competition between the FM superexchange
and AFM direct exchange, we show a simple mechanism of how the phase transition
occurs. The interesting case is that the FM ground state can be changed to the SP
ground state by applying hole doping. Therefore, we believe that introducing doping
may also achieve the multiferroic properties in the monolayer metal dihalides.

2. Computational Details

The crystal structure of monolayer 1T-XI2 (X: Fe, Co, and Ni) is depicted by Fig.
1(a). The primitive unit cell shown by a parallelogram contains a transition metal X
atom (cation) and two I atoms (anion). In the bulk layered structures, FeI2 and CoI2
crystallize into a CdI2 structure (P 3m1), but CoI2 does into a CdCl2 structure (R3m)
[30]. To avoid the layer-layer interactions in the z- axis, we assigned a vacuum distance
of more than 17 Å. Meanwhile, we specified the experimental lattice parameters of
4.04 Å for FeI2 , 3.96 Å for CoI2 , and 3.89 Å for NiI2 [10]. In the periodic lattices, we
established the lattice vectors
a a√
a = aêx , b = êx + 3êy , (1)
2 2
and the corresponding reciprocal lattice vectors
2π 2π 4π
A= êx − √ êy , B = √ êy , (2)
a a 3 a 3
where a defines the lattice parameter.
The computations were carried out in a 20 × 20 × 1 k-point sampling by a
linear combination of pseudo-atomic orbitals (LCPAO) approach [31, 32] combined
with the norm-conserving pseudopotential [33] as implemented in the OpenMX code
[34]. For treating the electron-electron interactions, we selected the generalized gradient
approximation (GGA) according to Perdew, Burke, and Ernzerhof [35] with an energy
cutoff of 200 Ryd. For the basis sets, we used an abbreviation snpmdafc, which means
that s, p, d, f are the primitive orbitals and n, m, a, c are the number of orbitals. We
set s3p3d3 for Fe atom, s3p3d3f2 for Co atom, s3p3d2f1 for Ni atom, and s2p2d1 for I
atom. In addition, the cutoff radii as the confinement at the atomic site, beyond which
the orbitals vanish, are 4.0 Bohr for X atom and 7.0 Bohr for I atom. The choices of
basis sets and cutoff radii are the minimum requirements to get converged results with
good accuracy.
To use the GBT via density functional theory (DFT), we introduced a wavevector
q into an LCPAO [36]
N
" !
1 X X ↑ 1
ψµk (r) = √ ei(k−q/2)·Rn Cµk,iα ξiα (r − τi − Rn )
N n iα 0
N
!#
X
i(k+q/2)·Rn
X ↓ 0
+ e Cµk,iα ξiα (r − τi − Rn ) , (3)
n iα 1
4

Figure 1. (Color online) Crystal structure of monolayer 1T-XI2 from top view (a).
Here, the FM and AFM ground states are achieved by assigning φ = 0 (b) and φ = 1
(c), respectively.

where ξiα describes the localized orbital. To create the spiral configurations, the
magnetic moment of metal atom is rotated from site to site obeying [37]
 
cos (ϕ0i + q · Ri ) sin θi
0
 
 sin (ϕi + q · Ri ) sin θi  .
Mi = Mi   (4)
cos θi
From Eq. 4, we see that the cone angle θ should be constant while the azimuthal angle
ϕ will be rotated along q direction. Since we want to tune the FM, SP, and AFM states
by q, we apply the flat spiral by setting the initial direction of magnetic moment of
metal atom to θ = π/2. Here, we concern the direction of q along x- axis, which is given
by

q = φ(A + 0.5B) = φ êx , (5)
a
where φ = 0 and φ = 1 give the FM state (Fig. 1(b)) and AFM state (Fig. 1(c)),
respectively. Then, the SP state will be generated by specifying φ between 0 and 1. We
also applied the penalty functional to keep θ = π/2.

3. Results and Discussions

For the convenience, we divide this section into two subsections, i.e., the non-doped and
doped cases
5

3.1. Non-doped Case


Figure 2 shows the ground states of XI2 (X: Fe, Co, and Ni) and magnetic moments for
the non-doped case. Here, the ground state is determined by the lowest total energy
difference ∆E as a function of φ. We observe that the ground state of FeI2 is an FM
state while the ground state of CoI2 and NiI2 is an SP state. As for CoI2 and NiI2 , our
SP ground states are not predicted by Refs. [10, 11], where they only considered the
FM and AFM states. By observing Fig. 2(a), we deduce that the AFM (FM) state is
more stable than the FM (AFM) state for CoI2 (NiI2 ), in good agreement with Refs.
[10, 11]. Detailed data can be found in table 1. Moreover, we also notice that all the
magnetic moment remains stable with respect to φ, starting from the higher magnetic
moment in FeI2 and then gradually reducing to NiI2 . This tendency is in nice agreement
with Hund’s rule for the free transition metal atoms.

100 4
FeI2
CoI2
80 NiI2 3.5

3
60
2.5
∆E (meV)

40
M (µB)

(a) 2 (b)
20
1.5
0
1
−20 FeI
0.5 CoI2
2
NiI2
−40 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
φ φ

Figure 2. (Color online) Ground states of some metal diiodides (a) and related
magnetic moments (b) for the non-doped case. Here, the ground state is associated
with the minimum energy ∆E = E(φ) − E(φ = 0).

The appearing SP state without any external treatments exhibits a frustrated spin
as the FM and AFM orders compete with each other. This competition of magnetic
orders will lead to one of stable ground states, namely, the FM, AFM, or SP state.
A similar stable SP state also occurs in the monoatomic transition metal chain and
square lattice, which is caused by the spin-spin interaction between the metal atom and
its nearest neighbor [38]. They also showed that the ground state is influenced by the
distance of nearest-neighbor atoms having different configurations of magnetic moment,
for example, the FM and AFM order. Similar discussions on the frustrated spin due to
spin-spin interaction can also be found in Refs. [39, 40]. In the other situations, some
materials require external treatments to produce the SP state, either using the doping
[41] or electric field [42], leading to the new stable form. Nevertheless, these treatments
usually reduce the magnetic moment, thus decreasing the magnetism in the magnetic
materials.
In general, the ground state in metal dihalides is influenced by either the AFM
6

Table 1. Predicted ground states and magnetic moments of metal diiodides from some
DFT calculations for the non-doped case. Here, GS and M denote the ground state
and magnetic moment in units of µB , respectively.
Metal (GS, M)1 (GS, M)2 (GS, M)3
FeI2 (FM, 3.49) (FM, 4.0) (FM, 3.45)
CoI2 (SP, 2.52) (AFM, 3.0) (AFM, 2.23)
NiI2 (SP, 1.38) (FM, 2.0) (FM, 1.53)
1
Present calculations.
2
Calculations from Ref. [10].
3
Calculations from Ref. [11].

direct exchange or the FM superexchange, the detailed mechanism of these interactions


can be found in Refs. [43, 44, 45, 46, 47]. These two interactions are determined by
the bonding cation-anion-cation where the electron hops from an X site to the nearest
X site following Hund’s rule and may lead to a spin frustration as explained previously.
As for FeI2 , since the ground state is an FM state, the FM superexchange overcomes
the AFM direct exchange. The different situation appears in CoI2 and NiI2 even though
they have the SP ground state. The interaction will be determined by the total energy
difference between the FM (φ = 0) and AFM (φ = 1) states. As shown in Fig. 2(a),
the interaction prefers the AFM direct exchange (FM superexchange) for CoI2 (NiI2 )
because the AFM state (FM state) is more stable than the FM state (AFM state).

Figure 3. (Color online) Schematic illustration of magnetic orders in FeI2 and NiI2
(a) and CoI2 (b).

Figure 3 shows a simple illustration of how the FM superexchange and AFM direct
exchange work due to electron hopping from an X atom to its nearest neighbor X atom
in appropriate with Hund’s rule. For the FM superexchange, the net magnetic moment
of the neighboring X atom (Fe or Ni atom) becomes parallel to reduce the kinetic
energy. This activates the inter-site virtual hopping and thus delocalizes the electrons
over X-I-X. Conversely, when the kinetic energy reduces because of antiferromagnetically
coupled magnetic moments of neighboring X atom (Co atom), it leads to the AFM direct
exchange. Furthermore, when a spin competition happens from these two interactions,
7

it may frustrate the magnetic moment to generate an SP ground state, as shown in CoI2
and NiI2 .
180 3.8
4.0 Å
160 4.04 Å
4.1 Å 3.7
140 4.2 Å
120 3.6
∆E (meV)

M (µB)
100
3.5
80
60 3.4
40 4.0 Å
3.3 4.04 Å
20 (a) (b) 4.1 Å
4.2 Å
0 3.2
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
φ φ

0 2.8
3.92 Å 3.92 Å
−5 3.96 Å 3.96 Å
4.1 Å 2.7 4.1 Å
−10 4.2 Å 4.2 Å
2.6
∆E (meV)

−15
M (µB)

−20 2.5

−25 2.4
−30
2.3
−35 (c) (d)
2.2
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
φ φ
1.8
30 3.87 Å 3.87 Å
3.89 Å 3.89 Å
4.1 Å 1.7 4.1 Å
20 4.2 Å 4.2 Å
1.6
∆E (meV)

10
M (µB)

1.5
0
1.4
−10
1.3
(e) (f)
−20 1.2
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
φ φ

Figure 4. Dependence of ground state on lattice parameter (a, c, e) and related


magnetic moments (b, d, e) for the non-doped case. Here, FeI2 , CoI2 and NiI2 are
depicted in (a, b), (c, d), and (e, f), respectively. In addition, the total energy difference
is defined as ∆E = E(φ) − E(φ = 0).

Regarding the metal dihalides, our previous studies reported that the ground state
of metal dichlorides can be tuned by varying the lattice parameter [48, 49]. So, our
next investigation is to check if the ground state of metal diiodides can also change
due to varying the lattice parameter. Figures 4(a, c, e) display the ground state of
metal diiodides with respect to the lattice parameter. As for FeI2 and CoI2 , the ground
states are robust to the lattice constant. However, we see a change of state from the
SP state to the FM state, which means that the ground state of NiI2 is sensitive to the
8

lattice parameter. Regarding the magnetic moment in Figs. 4(b, d, f), all the magnetic
moments remain stable for all the lattice parameters, indicating that all the systems
possess a high spin state. This state happens due to a weak ligand that the energy
difference between t2g and eg in the metal atoms is so weak. This implies that the
electron will occupy the empty orbital first by following Hund’s rule. Note that varying
lattice parameter may also lead to a low spin state, as occurred in the fcc iron [50, 51].

35

30

25
∆E (meV)

20
(a)
15

10

5 FeI2
CoI2
NiI2
0
1

0.8

0.6
φ

(b)

0.4

0.2

0
3.9 4.0 4.1 4.2 4.3
Lattice Parameter (Å)

Figure 5. Dependence of total energy difference ∆E = E(φ = φlowest ) − E(φ = 0)


(a) and related ground state φ = φlowest (b).

A more detailed dependence of the ground state of metal diiodides on the lattice
parameter can also be seen in Fig. 5. As for FeI2 and CoI2 , there is no transition
of ground state as the lattice parameter increases. This also implies that FeI2 is a
robust ferromagnet. A different situation occurs in NiI2 , namely, the FM ground state
is preferred at the large lattice parameter. However, for all the lattice parameters, our
calculations show that the FM state (AFM state) is more stable than the AFM state
(FM state) for FeI2 and NiI2 (CoI2 ). This indicates that FeI2 and NiI2 (CoI2 ) are still
controlled by the FM superexchange (AFM direct exchange). We will see in the next
subsection that this control can be changed by hole-electron doping.
9

3.2. Doped Case


Our final discussion is to consider the magnetic ground state on hole-electron doping.
We implement the Fermi level shift approach to run the self-consistent calculations for
the doped case [52]. In this approach, a background charge is inserted to make the
system neutral. We introduce x as the number of doped concentrations in units of
e/cell, where the positive (negative) e is addressed to hole (electron) doping. Figures
6(a, c, e) show the change of ground state as the doping is introduced. As for FeI2 , the
FM ground state changes to the SP state as x = 0.4 e/cell is applied. Contrarily, the
SP ground state in CoI2 shifts to the FM ground state when x = −0.4 e/cell is taken
into account. Lastly, the SP ground state in NiI2 is replaced by the FM ground state
as x = 0.4 e/cell or x = −0.4 e/cell is introduced. As shown in Figs. 6(b, d, f), the
magnetic moments tend to stable with respect to the doping, still remaining in the high
spin state.
For the detailed phase transition in the metal diiodides because of the doping, we
present Fig. 7 to see the role of the AFM direct exchange and FM superexchange for
creating the phase transition. As shown in Figs. 7(a, b), the phase transition arises in
the interval of doping x. As for FeI2 , the FM ground state still preserves in the interval
of x ≤ 0.1 e/cell while the SP ground state starts to emerge in the interval of x > 0.1
e/cell and no AFM ground state emerges. At the same time, the SP ground state in CoI2
still maintains in the range of −0.3 e/cell ≤ x ≤ 0.15 e/cell and x ≥ 0.25 e/cell, the FM
ground state takes the role in the range of x < −0.3 e/cell, while the AFM ground state
only happens in x = 0.2 e/cell. As for NiI2 , the SP ground state only appears in the
interval of 0 e/cell ≤ x ≤ 0.15 e/cell while the FM ground state emerges in the interval
of x < 0 e/cell and x > 0.15 e/cell and no AFM ground state emerges. From these
results, we conclude that introducing hole doping, in general, frustrates increasingly the
spins for all the systems. On the contrary, the frustrated spins due to increasing electron
doping only happen in CoI2 , while the FM ground state becomes robust in FeI2 and
NiI2 , thus decreasing the frustrated order to become the FM ground state.
As previously mentioned, the AFM direct exchange competes with the FM
superexchange to determine the magnetic order. To discuss the competition, we define
the exchange parameter Jij = (1/12)∆E/M 2 [53], which represents the interaction
between a magnetic X atom with its nearest neighbor X atom. Here, ∆E = E(φ =
1) − E(φ = 0) and M mean the total energy difference and the magnetic moment of X
atoms. Since an X atom has six nearest neighbor X atoms, we insert 1/12 to prevent
the double counting in the calculation. In this case, the positive and negative Jij are
addressed to the FM superexchange and AFM direct exchange. As shown in Fig. 7(c),
we see the different tendencies for all the systems.
In FeI2 , the strength of FM superexchange increases (decreases) as the electron
(hole) increases, leading to either the FM or SP ground state. In this case, the FM
superexchange (AFM direct exchange) works in the range of x < 0.35 e/cell (x ≥ 0.35
e/cell). As for CoI2 , the strength of AFM direct exchange becomes very strong (weak)
10
200 3.8
x = 0.0 (e/cell)
x = 0.4 (e/cell)
x = −0.4 (e/cell) 3.7
150
3.6
∆E (meV)

100

M (µB)
3.5
50
3.4
0
3.3 x = 0.0 (e/cell)
(a) (b) x = 0.4 (e/cell)
−50 x = −0.4 (e/cell)
3.2
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
φ φ

60 2.6
x = 0.0 (e/cell)
x = 0.4 (e/cell)
40 x = −0.4 (e/cell) 2.5
20
2.4
∆E (meV)

0 M (µB)
2.3
−20
−40 2.2

−60 2.1 x = 0.0 (e/cell)


(c) (d) x = 0.4 (e/cell)
−80 x = −0.4 (e/cell)
2
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
φ φ
100 1.5
x = 0.0 (e/cell)
x = 0.4 (e/cell)
80 x = −0.4 (e/cell) 1.4
60
∆E (meV)

1.3
M (µB)

40
1.2
20

0 1.1 x = 0.0 (e/cell)


(e) (f) x = 0.4 (e/cell)
x = −0.4 (e/cell)
−20 1
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
φ φ

Figure 6. Doping-induced magnetic ground state (a, c, e) and related magnetic


moments (b, d, f) for the experimental lattice parameters. We depict FeI2 , CoI2 and
NiI2 as in (a, b), (c, d), and (e, f), respectively. Here, positive and negative x are
addressed to hole and electron doping, respectively. We also define the total energy
difference as ∆E = E(φ) − E(φ = 0)

as the hole (electron) doping increases, which leads to either the SP or AFM ground
state. At this situation, the FM superexchange (AFM direct exchange) performs in the
range of x ≤ −0.1 e/cell (x > −0.1 e/cell). As for NiI2 , the FM superexchange wins
against the AFM direct exchange for both the non-doped and doped cases, which is
preferable to the SP or FM ground state. For this case, only the FM superexchange
proceeds for all doping.
As illustrated in Fig. 8, introducing the doping lets the Coulomb repulsion to
control the occupation of electron in each X site, where the electron concentration can
be either reduced by hole doping or increased by electron doping in the 3d orbital. When
11
90
FeI2
80 CoI2
NiI2
70

60
∆E (meV)

50
(a)
40

30

20

10

0
1

0.8

0.6
φ

(b)

0.4

0.2

0
1.5

1.0

0.5
Jij (meV)

0 (c)

−0.5

−1.0

−1.5
−0.4 −0.3 −0.2 −0.1 0 0.1 0.2 0.3 0.4
Doping x (e/cell)

Figure 7. Magnetic phase transition of XI2 (X: Fe, Co, and Ni) with respect to
hole-electron doping x. Here, we define the total energy difference ∆E = E(φ =
0) − E(φlowest ) (a) the ground state φ = φlowest (b), and the exchange parameter
Jij .

the FM superexchange wins against the AFM direct exchange (Jij > 0), the kinetic
energy reduces as the magnetic moments of nearest neighbor X atoms become parallel
and the electron over X-I-X becomes delocalized. This condition occurs by increasing
the electron doping for all the systems as shown in Fig. 7(c). On the contrary, when the
hole doping increases for FeI2 and CoI2 , the kinetic energy also reduces if the magnetic
moments of nearest neighbor X atoms become antiparallel, thus dominated by the AFM
direct exchange (Jij < 0). It also becomes the electron over X-I-X delocalized. As for
NiI2 , the FM superexchange still does exist as hole doping increases, thus indicating a
robust ferromagnet with respect to the doping.
12

Figure 8. (Color online) Schematic illustration of competition between AFM direct


exchange and FM superexchange.

4. Conclusions

In summary, we explore the magnetic ground state of two-dimensional monolayer metal


diiodides XI2 (X: Fe, Co, and Ni) by utilizing the generalized Bloch theorem (GBT)
within the self-consistent flat spin spiral calculations. FeI2 exhibits an FM ground state
while CoI2 and NiI2 show an SP ground state. Based on the spin-spin interaction, CoI2
prefers the AFM order while NiI2 tends to the FM order, even though having the same
SP ground state. So, the SP ground state is a manifestation of frustrated spin due to
competition between the FM and AFM orders. The emergence of SP state without any
external treatments shows that monolayer metal diiodides should be able to generate
the multiferroic property that is very useful for applicable spintronic instruments.
We also show that the new ground state can be tuned from the initial ground
state by applying hole-electron doping. This utilizes the spin-spin competition between
the FM and AFM exchange orders, in which the inter-site virtual electron hopping
determines the ground state. The available ground state due to the magnetic
configuration of nearest neighbor metal atoms will appear if the kinetic energy can be
reduced, thus delocalizing the electron. This mechanism also creates a frustrated spin
which leads to an SP ground state. Our findings highlight the importance of introducing
the doping to generate a multiferroic property in the metal dihalides for the general case.

Acknowledgments

This research was performed by using a personal high computer and available facilities
at Universitas Negeri Jakarta. We hereby state that this research is an independent
work without any fundings.
13

References

[1] Tokura Y 2006 Science 321 1481.


[2] Cheong S W and Mostovoy M 2007 Nat. Mater. 6 13.
[3] Tokura Y and Seki S 2010 Adv. Mater. 22 1554.
[4] Spaldin N A and Fiebig M 2005 Science 309 391.
[5] Ramesh R and Spaldin N A 2007 Nat. Mater. 6 21.
[6] Scott J F 2007 Nat. Mater. 6 256.
[7] Zhai K, Wu Y, Shen S, Tian W, Cao H, Chai Y, Chakournakos B C, Shang D, Yan L, Wang F
and Sun Y 2017 Nat. Commun. 8 519.
[8] Tokunaga Y, Okuyama D, Kurumaji T, Arima T, Nakao H, Murakami Y, Taguchi Y and Tokura
Y 2011 Phys. Rev. B 84 060406(R).
[9] Wu X, Cai Y, Xie Q, Weng H, Fan H and Hu J 2012 Phys. Rev. B 86 134413.
[10] Kulish V V and Huang W 2017 J. Mater. Chem. C 5 8734.
[11] Botana A S and Norman M R 2019 Phys. Rev. Materials 3 044001.
[12] Kurz Ph, Förster F, Nordström L, Bihlmayer G and Blügel S 2004 Phys. Rev. B 69 024415.
[13] Sabirianov R F, Solanki A K, Burton J D, Jaswal S S and Tsymbal E Y 2005 Phys. Rev. B 72
054443.
[14] Pyatakov A P, Zvezdin A K, Vlasov A M, Sergeev A S, Sechin D A, Nikolaeva E P, Nikolaev A
V, Chou H, Sun S J and Calvet L E 2012 Ferroelectrics 438 79.
[15] Chen G, Kang S P, Ophus C, N’Diaye A T, Kwon H Y, Qiu R T, Won C, Liu K, Wu Y and Schmid
K 2017 Nat. Commun. 8 15302.
[16] Wang M -C, Huang C -C, Cheung C -H, Chen C -Y, Tan S G, Huang T -W, Zhao Y, Zhao Y, Wu
G, Feng Y -P, Wu H -C and Chang C -R 2020 Ann. Phys. 532, 1900452.
[17] Han R, Jiang Z and Yan Y 2020 J. Phys. Chem. C 124 7956.
[18] McGuire M A 2020 J. Appl. Phys. 128, 110901.
[19] Mushtaq M, Zhou Y and Xiang X 2017 RSC Adv. 7, 22541.
[20] Li X, Zhang Z and Zhang H 2020 Nanoscale Adv. 2, 495.
[21] Chen P, Zou J -Y and Liu B -G 2017 Phys. Chem. Chem. Phys. 19, 13432.
[22] Ashton M, Gluhovic D, Sinnott S B, Guo J, Stewart D A and Hennig R G 2017 Nano Lett. 17,
5251.
[23] Knöpfle K, Sandratskii L M and Kübler J 2000 Phys. Rev. B 62 5564.
[24] Sandratskii L M and Bruno P 2003 Phys. Rev. B 67 214402.
[25] Garcı́a-Suárez V M, Newman C M, Lambert CJ, Pruneda J M and Ferrer J 2004 J. Phys.: Condens.
Matter 16 5453.
[26] Kunes̆ J and Laskowski R 2004 Phys. Rev. B 70 174415.
[27] Kurumaji T, Seki S, Ishiwata S, Murakawa H, Tokunaga Y, Kaneko Y and Tokura Y 2011 Phys.
Rev. Lett. 106 167206.
[28] Kurumaji T, Seki S, Ishiwata S, Murakawa H, Kaneko Y and Tokura Y 2013 Phys. Rev. B 87
014429.
[29] Kovaleva E A, Melchakova I, Mikhaleva N S, Tomilin F N, Ovchinnikov S G, Baek W, Pomogaev
V A, Avramov P and Kuzubov A A 2019 J. Phys. Chem. Solids 134 324.
[30] McGuire M A 2017 Crystals 7 121.
[31] Ozaki T and Kino H 2004 Phys. Rev. B 69 195113.
[32] Ozaki T 2003 Phys. Rev. B 67 155108.
[33] Troullier N and Martins J L 1991 Phys. Rev. B 43 1993.
[34] Ozaki T et al., OpenMX code (http://www.openmx-square.org).
[35] Perdew J P, Burke K and Ernzerhof M 1996 Phys. Rev. Lett. 77 3865.
[36] Prayitno T B and Budi E 2020 J. Phys.: Condens. Matter 32 105802.
[37] Sandratskii L M 1998 Adv. Phys. 47 91.
[38] Töws W and Pastor G M 2012 Phys. Rev. B 86 054443.
14

[39] Zelený M, Sǒb M and Hafner J 2009 Phys. Rev. B 80 144414.


[40] Saubanère M, Tanveer M. Ruiz-Dı́az P and Pastor G M 2010 Phys. Status Solidi B 247 2610.
[41] Inoue J and Maekawa S 1995 Phys. Rev. Lett. 74 3407.
[42] Prayitno T B 2021 J. Phys.: Condens. Matter 33 065805.
[43] Goodenough J B 1955 Phys. Rev. 100 564.
[44] Goodenough J B 1958 J. Phys. Chem. Solids 6 287.
[45] Kanamori J 1959 J. Phys. Chem. Solids 10 87.
[46] Kanamori J 1960 J. Appl. Phys. 31 S14.
[47] Anderson P W 1959 Phys. Rev. 115 2.
[48] Prayitno T B and Ishii F 2019 J. Phys. Soc. Jpn. 88 104705.
[49] Prayitno T B 2021 J. Magn. Magn. Mater. 517 167386.
[50] Uhl M, Sandratskii L M and Kübler J 1992 J. Magn. Magn. Mater. B 103 314.
[51] Mryasov O N, Gubanov V A and Liechtenstein A I 1992 Phys. Rev. B 45 12330.
[52] Sawada K, Ishii F, Saito M, Okada S and Kawai T 2009 Nano Lett. 9 269.
[53] Torun E, Sahin H, Singh S K and Peeters F M, 2015 Appl. Phys. Lett. 106, 192404.

You might also like