• No results found

Magnetizing lead-free halide double perovskites

N/A
N/A
Protected

Academic year: 2021

Share "Magnetizing lead-free halide double perovskites"

Copied!
9
0
0

Loading.... (view fulltext now)

Full text

(1)

S C I E N C E A D V A N C E S

|

R E S E A R C H A R T I C L E

M A T E R I A L S S C I E N C E

Magnetizing lead-free halide double perovskites

Weihua Ning1, Jinke Bao2, Yuttapoom Puttisong1, Fabrizo Moro1, Libor Kobera3, Seiya Shimono4,

Linqin Wang5, Fuxiang Ji1, Maria Cuartero5, Shogo Kawaguchi6, Sabina Abbrent3,

Hiroki Ishibashi7, Roland De Marco8, Irina A. Bouianova1, Gaston A. Crespo5,

Yoshiki Kubota7, Jiri Brus3, Duck Young Chung2, Licheng Sun5,9, Weimin M. Chen1,

Mercouri G. Kanatzidis2,10, Feng Gao1*

Spintronics holds great potential for next-generation high-speed and low–power consumption information technology. Recently, lead halide perovskites (LHPs), which have gained great success in optoelectronics, also show inter-esting magnetic properties. However, the spin-related properties in LHPs originate from the spin-orbit coupling of Pb, limiting further development of these materials in spintronics. Here, we demonstrate a new generation of halide perovskites, by alloying magnetic elements into optoelectronic double perovskites, which provide rich chemical and structural diversities to host different magnetic elements. In our iron-alloyed double perovskite,

Cs2Ag(Bi:Fe)Br6, Fe3+ replaces Bi3+ and forms FeBr6 clusters that homogenously distribute throughout the double

perovskite crystals. We observe a strong temperature-dependent magnetic response at temperatures below 30 K, which is tentatively attributed to a weak ferromagnetic or antiferromagnetic response from localized regions. We anticipate that this work will stimulate future efforts in exploring this simple yet efficient approach to develop new spintronic materials based on lead-free double perovskites.

INTRODUCTION

As a new generation of solution-processed semiconductors, lead halide perovskites (LHPs) have taken a dominant position within the portfolio of compounds due to their outstanding optoelectronic properties, including high carrier mobility, long carrier lifetime, low trap density, and high absorption coefficient. These properties result in high-performance optoelectronic devices, including solar cells (1, 2), light-emitting diodes (3, 4), lasers (5), and photodetectors (6, 7). Very recently, LHPs also show interesting spin-related prop-erties due to spin-orbit coupling (SOC) (8) of Pb. Thus, they exhibit spin-dependent optical selection rules (9, 10), large-scale Rashba and Dresselhaus splitting (11–13), magneto-optical effects (14, 15), and polarized light-related effects (16). These observations make LHPs interesting for spintronic devices (17). However, the magnetic properties resulting from the SOC are weak, therefore limiting their further development in the field.

In addition to SOC enhancement, a straightforward yet more efficient approach to achieving spintronic materials is magnetic doping/alloying into nonmagnetic semiconductors. This approach has played an increasingly prominent role to revolutionize electronics and spintronics in conventional semiconductors and has also been very recently examined in LHPs, showing optically switched mag-netism at 5 K (18). Since the magnetic properties in this approach

mainly originate from magnetic dopants, it is more desired to turn to other optoelectronic lead-free halide perovskites (e.g., double perovskites), which provide much more room for doping/alloying (19), in addition to being environmentally friendly and stable.

Here, we focus on the benchmark double perovskite Cs2AgBiBr6,

which is a promising lead-free and stable optoelectronic material. Through magnetic element iron alloying, we provide new strategies to develop a new generation of materials that can potentially couple optoelectronics with spintronics. Combining near edge x-ray absorp-tion fine structure (NEXAFS) and solid-state nuclear magnetic res-onance (ssNMR) measurements, we reveal that iron is in the form

of Fe3+, replacing Bi3+ and forming FeBr6 clusters, which

homoge-nously distribute throughout the double perovskite crystals. Our

double perovskite alloy, Cs2Ag(Bi:Fe)Br6, exhibits a structural phase

transition at ~120 K and a strongly temperature-dependent magnetic response at temperatures below 30 K.

RESULTS

Basic characterizations of double perovskite Cs2Ag(Bi:Fe)Br6

Double perovskites Cs2Ag(Bi:Fe)Br6 were prepared by using a

pro-cedure similar to that reported for Cs2AgBiBr6 (20, 21). CsBr, AgBr,

BiBr3, and FeBr3 precursors were mixed into an HBr solution in a

hydrothermal autoclave, which was heated at 120°C for 24 hours and then slowly cooled down, resulting in 2- to 5-mm black octa-hedral crystals as the final products. Figure 1A presents

photo-graphs of the Cs2AgBiBr6 and Cs2Ag(Bi:Fe)Br6 single crystals. The

crystal color of Cs2Ag(Bi:Fe)Br6 changes from red to black after

alloying, indicating enhanced absorption (fig. S1). Electron spin

resonance (ESR) measurements of the Cs2Ag(Bi:Fe)Br6 crystal show

a strong resonance located at g ~ 2.032 at room temperature, point-ing out the existence of paramagnetic Fe (Fig. 1B). The

concen-tration of Fe in Cs2Ag(Bi:Fe)Br6 is ~11.4%, as confirmed by both

energy dispersive spectroscopy (EDS) and inductively coupled plasma optical emission spectroscopy (ICP-OES) measurements (fig. S2 and table S1).

1Department of Physics, Chemistry and Biology (IFM), Linköping University,

Linköping SE-581 83, Sweden. 2Materials Science Division, Argonne National

Labo-ratory, Argonne, IL 60439, USA. 3Institute of Macromolecular Chemistry of the

Czech Academy of Sciences, Heyrovskeho nam. 2, 162 06 Prague 6, Czech Republic.

4Department of Materials Science and Engineering, National Defense Academy,

Yokosuka, Kanagawa 239-8686, Japan. 5Department of Chemistry, KTH Royal Institute

of Technology, SE-10044 Stockholm, Sweden. 6Japan Synchrotron Radiation Research

Institute (JASRI), SPring-8, Sayo, Hyogo 679-5198, Japan. 7Department of Physical

Science, Osaka Prefecture University, Sakai, Osaka 599-8531, Japan. 8Faculty of Science,

Health, Education and Engineering, University of the Sunshine Coast, 90 Sippy Downs Drive, Sippy Downs, QLD 4556, Australia. 9Center of Artificial Photosynthesis for

Solar Fuels, School of Science, Westlake University, Hangzhou 310024, China.

10Department of Chemistry, Northwestern University, Evanston, IL 60208, USA.

*Corresponding author. Email: feng.gao@liu.se

Copyright © 2020 The Authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original U.S. Government Works. Distributed under a Creative Commons Attribution NonCommercial License 4.0 (CC BY-NC). on January 4, 2021 http://advances.sciencemag.org/ Downloaded from

(2)

Both single-crystal x-ray diffraction (SC-XRD) and power XRD

(PXRD) measurements indicate that Cs2Ag(Bi:Fe)Br6 retains the

cubic phase of Cs2AgBiBr6. On the basis of the SC-XRD

measure-ments, Cs2AgBiBr6 crystallizes in the cubic Fm-3m space group

with the lattice parameter of a = 11.27 Å (table S2), manifesting a three-dimensional classic perovskite structure (Fig. 1A, bottom).

When 11.4% Fe is alloyed into Cs2AgBiBr6, the structure retains its

original cubic phase with lattice parameter slightly decreasing to

a = 11.25 Å (table S2). This lattice shrink phenomenon results from

smaller radius of Fe3+ (0.65 Å)/Fe2+ (0.78 Å) relative to Bi3+ (1.03 Å)/Ag+

(1.15 Å) (22), consistent with the Vegard’s law (23). The PXRD patterns further confirm the cubic double perovskite structure of

Cs2Ag(Bi:Fe)Br6, since all the diffraction peaks are almost the same

as the simulated and experimental Cs2AgBiBr6 (Fig. 1C). The highly

intense reflections (220) and (400) shift toward high 2 angles together with symmetric to asymmetric transformation after Fe alloying (Fig. 1, D and E), which is consistent with the lattice shrink observed from SC-XRD.

We additionally notice that Fe exhibits multiple valence states in

compounds (e.g., FeO, Fe2O3, and Fe3O4) and subsequently perform

NEXAFS, x-ray photoelectron spectroscopy (XPS), and synchrotron radiation–XPS (SR-XPS) measurements so as to understand the valence state and surface distribution of Fe in our alloyed double

perovskites. The NEXAFS Fe 2p3/2 edge spectrum for Cs2Ag(Bi:Fe)

Br6 (Fig. 1F) shows a double peak in the range of 705 to 713 eV. Peak

deconvolution indicates an intensity ratio of approximately 1:3

between the Fe 2p3/2 signal at 708.1 eV and that at 709.9 eV, pointing

out that the Fe is in the state of trivalence (24). The valence state of

Fe is further confirmed by the XPS Fe 2p3/2 core level spectrum

(fig. S3), as evidenced by the small atomic shake-up satellite as a

high binding energy shoulder at 712.5 eV on the major Fe 2p3/2 peak

at 707.5 eV that is characteristic of the trivalent state (25). It is important to note that SR-XPS of the Fe 2p level at a low photoelectron kinetic energy of 50 eV (electrons escaping from a shallow depth of one monolayer) does not reveal the presence of Fe in the sampled region, while laboratory XPS using deeper originating Fe 2p photo-electrons at a kinetic energy of 777 eV (compared with an Auger electron energy of 650 eV in the NEXAFS measurement, a compa-rable kinetic energy and photoelectron inelastic mean free path) gives a clearly discernable Fe 2p XPS signal. The laboratory XPS, SR-XPS, and NEXAFS results confirm that the alloyed Fe is not at the surface of the crystal but is instead buried beneath the surface monolayer.

Temperature-dependent structure of double perovskite Cs2Ag(Bi:Fe)Br6

We further investigate the effect of Fe alloying on the crystal structures at low temperatures. We perform specific heat capacity (Cp)

mea-surement to examine the phase transition behavior of Cs2Ag(Bi:Fe)

Br6 (Fig. 2A). A thermal anomaly is observed at ~120 K, which is

probably attributed to a weak first-order or second-order phase transition. To deeply understand the thermal anomaly observed from the Cp measurements, we characterize the crystal structures

of Cs2Ag(Bi:Fe)Br6 by using temperature-dependent synchrotron

powder diffraction (SPD). Figure 2B and fig. S4 show the SPD

patterns of Cs2Ag(Bi:Fe)Br6 in 300 to 30 K, where the cubic Bragg

reflection (4 0 0) split into two peaks at ~120 K, indicating that the phase transition at this temperature is a symmetric breaking structural Fig. 1. Basic characterizations of the double perovskite Cs2Ag(Bi:Fe)Br6. (A) Photographs and crystal structures of double perovskites Cs2AgBiBr6 and

Cs2Ag(Bi:Fe)Br6. (B) ESR signals from Cs2AgBiBr6 and Cs2Ag(Bi:Fe)Br6 at room temperature (RT). a.u., arbitrary units. (C) Powder XRD patterns for Cs2AgBiBr6 and

Cs2Ag(Bi:Fe)Br6, and the simulated XRD pattern of Cs2AgBiBr6 as reference, an expansion of highly intense reflections (220) (D) and (400) (E), illustrating shift toward

higher 2 angles and the symmetric to asymmetric transformation after Fe3+ alloying. (F) NEXAFS Fe 2p3/2 edge spectrum for Cs2Ag(Bi:Fe)Br6, in which the baseline

has been corrected.

on January 4, 2021

http://advances.sciencemag.org/

(3)

S C I E N C E A D V A N C E S

|

R E S E A R C H A R T I C L E

phase transition. In contrast, the reflection peaks of (4 4 4) only regularly shift toward the high-angle side, possibly due to lattice shrink during the cooling process.

To determine the low-temperature crystal structure of Cs2Ag(Bi:Fe)

Br6 accurately, we use the JANA2006 soft package (26) to do the Rietveld

refinements. The Rietveld refinement shows a perfect single-phased

material; thus, Cs2Ag(Bi:Fe)Br6 sample is free of CsBr, BiBr3, AgBr,

and Fe clusters or any other impurity (fig. S5). Table S3 lists the lattice parameters at different temperatures. A cubic to tetragonal (c > a) first-order structural transition is observed at 120 K. This temperature agrees well with thermal anomaly temperature (Tc) observed in the temperature-dependent Cp (Fig. 2A). With

decreas-ing temperature, Cs2Ag(Bi:Fe)Br6 experiences a symmetry

break-ing phase transition with the space group changbreak-ing from Fm-3m to I4/m. Meanwhile, an abnormal change is observed from the temperature-dependent lattice constant and rotation angle between

AgBr6 and (Bi:Fe)Br6 octahedron (Fig. 2, C and D), further

con-firming the structural phase transition at 120 K. As the temperature gradually decreases, the crystal phase retains I4/m with the lattice volume slightly decreasing at 30 K, which correlates well with the

temperature dependence of Cp. As for pristine Cs2AgBiBr6, the

structural behaviors are almost the same as Cs2Ag(Bi:Fe)Br6,

in-dicating that the structural phase transition is intrinsic, without

any influence from Fe3+ alloying (27). The lattice parameters of

Cs2Ag(Bi:Fe)Br6 are observed to be smaller than those of Cs2AgBiBr6,

consistent with SC-XRD results (22).

ssNMR of double perovskites Cs2AgBiBr6 and Cs2Ag(Bi:Fe)Br6

Next, we proceed to investigate how Fe ions distribute within the

Cs2AgBiBr6 matrix. We analyze miscellaneous 133Cs and 209Bi ssNMR

[magic angle spinning (MAS) ssNMR)] experiments to gain deep

insight in the structure for both Cs2AgBiBr6 and Cs2Ag(Bi:Fe)Br6

systems. As shown in Fig. 3A, only one symmetric peak located at

iso = 80.7 ± 0.5 parts per million (ppm) in 133Cs MAS NMR spectra

of Cs2AgBiBr6, confirming the existence of one crystallographic

position of the Cs+ ions in pristine Cs

2AgBiBr6. In contrast, besides

the peak at iso = 80.7 ± 0.5 ppm, an additional weak peak appears at

iso = 83.4 ± 0.5 ppm (Fig. 3B) in Cs2Ag(Bi:Fe)Br6, indicating the

presence of paramagnetic ions (Fe3+) in close proximity to Cs+ ions

(28–30). The presence of paramagnetic metal ions (e.g., Fe3+ and

Cu2+) usually causes extremely rapid longitudinal and transverse

relaxation of the nearby nuclei due to electron spin couplings. This behavior allows us to probe a more precise distribution of the

para-magnetic species in the perovskite matrix by using 133Cs NMR T

1

relaxation measurements (31).

The detected 133Cs NMR signal contains two partially overlapped

peaks (components), and determination of their relaxation

param-eters (T1 relaxation times and amounts of individual components)

is performed in two steps. First, the saturation recovery buildup

curves of the detected 133Cs NMR signal are analyzed using a double-

exponential saturation recovery function and preliminary values of

T1(133Cs) relaxation times, and the corresponding fractions of

indi-vidual components are obtained. Second, each 133Cs NMR spectrum

recorded in T1(133Cs) saturation recovery experiment is fitted. Then,

the buildup curve of individual components is separately analyzed by using the single-exponential saturation recovery functions (Fig. 3, C and D). This way, both peaks are analyzed, and the values

of T1 relaxation times and the amounts of individual phases are

re-fined. Figure 3 (C and D) shows the resulting 133Cs T1 relaxation

curves for Cs2AgBiBr6 and Cs2Ag(Bi:Fe)Br6. We obtain a slow T1

relaxation time of 306 s in Cs2AgBiBr6 and one slow- and one

fast-relaxing phase corresponding to signals at iso = 80.7 ± 0.5 ppm

Fig. 2. Temperature-dependent specific heat capacity and SPD. Specific heat capacity (A) and SPD patterns (B) of Cs2Ag(Bi:Fe)Br6 at different temperatures. The lattice

constant (C) and rotation angle (D) in 293 to 30 K for Cs2Ag(Bi:Fe)Br6. The crystal structures of Cs2Ag(Bi:Fe)Br6 at high temperature (HT, 300 K) (E) and low temperature (LT,

30 K) (F), red lines representing unit cells. Inset plot in (A) represents a magnification of the Cp peak at 120 K.

on January 4, 2021

http://advances.sciencemag.org/

(4)

(31 s) and iso = 83.4 ± 0.5 ppm (4 s) in Cs2Ag(Bi:Fe)Br6. The

shortened 133Cs T

1 relaxation times indicate that the

paramag-netic ions (Fe3+) are alloyed into the perovskite framework. Since

we now know the amount of Fe3+ and also the amount of Cs+, which

is in close proximity to Fe3+, simple calculations lead us to

con-clude that Fe3+ is homogenously distributed as small FeBr

6 clusters

throughout the lattice (more details in the Supplementary Ma-terials) (31).

The remaining doubt is whether Fe3+ ions replace Bi3+. Hence,

we perform the 209Bi ssNMR experiment on both Cs

2AgBiBr6 and

Cs2Ag(Bi:Fe)Br6 systems. As shown in Fig. 3E, a very broad signal

locates at about iso = 5500 ppm for both Cs2AgBiBr6 and Cs2Ag(Bi:Fe)

Br6. The 209Bi ssNMR spectrum of Cs2AgBiBr6 provides relatively

symmetric signal with evident J-coupling (JX−Bi = 2002 Hz, where X

corresponds to Br). In the case of Cs2Ag(Bi:Fe)Br6, the 209Bi ssNMR

spectrum shows broadened spectral line with J-coupling completely

missing. This phenomenon is caused by the fast T1 relaxation times

of 209Bi nuclei induced by the paramagnetic Fe3+, which is

homoge-nously distributed in the matrix, replacing Bi3+ ions.

Magnetic properties characterization of double perovskite Cs2Ag(Bi:Fe)Br6

Encouraged by the fact that Fe3+ ions are incorporated into the

lattice, we perform temperature-dependent magnetic susceptibility

for Cs2Ag(Bi:Fe)Br6 by superconducting quantum interference

device (SQUID). The results exhibit diamagnetism at room tempera-ture and weak temperatempera-ture dependence of magnetic susceptibility above 30 K (Fig. 4A), indicating that diamagnetism from core

elec-trons in this compound dominates over paramagnetic contribution

from the alloyed magnetic ion Fe3+. The magnetic susceptibility

in-creases rapidly below 20 K, indicating that a certain kind of magnetic response appears. The magnetic susceptibilities under zero-field– cooled (ZFC) and field-cooled (FC) procedures basically overlap with each other in the whole temperature range (Fig. 4A) and exhibit very tiny difference below 30 K (fig. S7). Field-dependent magnetization M(H) is linear with a negative slope above 50 K, which is consistent with temperature-dependent magnetic suscepti-bility (Fig. 4B). M(H) at 2 K has a positive component at low fields

and saturates to ~0.15 B per Fe after subtracting the diamagnetic

component. The saturated magnetic moment is much smaller than

both high spin (5 B per Fe) and low spin (1 B per Fe) state of Fe3+.

In addition, there is no obvious magnetic hysteresis in the M(H) loop in the low fields (see the inset of Fig. 4B).

To gain further insights into the observed temperature dependence of the magnetic response probed by SQUID, we perform ESR mea-surements. ESR is well-known to be capable of providing microscopic information associated with the chemical origins of the spin (mag-netic) species, being hence probed. In our case, this has enabled us to identify the paramagnetic centers or possible magnetic resonance signals arising from magnetic coupling such as ferromagnetic or antiferromagnetic resonance (FMR or AFMR) signals that are in-volved in the total ESR spectra as well as their contributions to the overall magnetic response evolving with temperature. Figure 4C

shows ESR spectra from Cs2Ag(Bi:Fe)Br6 crystal powder (the red

curves) obtained at 7.2 and 40 K, respectively. Strong multiline ESR signals are observed throughout the temperature range of 5 to 300 K: Fig. 3. ssNMR of Cs2AgBiBr6 and Cs2Ag(Bi:Fe)Br6. 133Cs-133Cs SD/MAS NMR and 133Cs Hahn-echo MAS NMR spectra of Cs2AgBiBr6 (A) and Cs2Ag(Bi:Fe)Br6 (B). 133Cs T1

saturation recovery buildup curves of Cs2AgBiBr6 (C) and Cs2Ag(Bi:Fe)Br6 (D). (E) 209Bi ssNMR of Cs2AgBiBr6 and Cs2Ag(Bi:Fe)Br6 conducted at static conditions. (F)

Sche-matic representation of possible scenarios for Fe3+ distribution inside the perovskite lattice: parent perovskite lattice, isolated Fe3+ ions, small Fe3+ clusters, and large Fe3+

clusters. The percentages are the ideally calculated values for Cs+ exhibiting fast relaxation in different scenarios.

on January 4, 2021

http://advances.sciencemag.org/

(5)

S C I E N C E A D V A N C E S

|

R E S E A R C H A R T I C L E

These are introduced upon the incorporation of Fe in Cs2AgBiBr6,

as they are absent in the ESR spectra from the pristine Cs2AgBiBr6

(shown as the gray curves in Fig. 4C). As shown in Fig. 4C (right),

the ESR spectrum at 40 K from the Cs2Ag(Bi:Fe)Br6 crystal powder

is best described by a sum of two paramagnetic species: (i) isolated

Fe3+ of tetragonal symmetry with S = 5/2 and g = 2.032 and (ii) a

defect center of unknown origin with S = 1/2 and g = 2.032. The assignment is supported by the good agreement between the simu-lated ESR spectra of these two centers by the spin Hamiltonian and the experimental spectrum (the spin Hamiltonian analysis is described in the Supplementary Materials). Upon cooling down below 35 K, in addition to the two aforementioned paramagnetic centers, an extra, broad component emerges and becomes dominant in the ESR spectrum at 7.2 K.

To find a correlation with the temperature-dependent magnetic response from the SQUID and additionally identify the relative importance of various spin species, we carefully investigate and analyze the ESR signal intensity (directly proportional to ) as a function of temperature. The results are summarized in Fig. 4D, where the temperature dependences of the ESR intensities of the individual ESR components and the total ESR signal intensity are displayed. Evidently, the total ESR signal intensity exhibits a similar temperature-dependent trend as the SQIUD data shown in Fig. 4A, confirming a direct correlation in measuring  by the two

tech-niques. Using the protocol of spectral deconvolution of the three

different ESR contributions, the intensities of the isolated Fe3+ and

the S = 1/2 center are found to gradually and monotonically increase with deceasing temperature, as expected according to the paramagnetic centers. Our analysis proves that the sharp rising of the magnetic response at T < 35 K is not related to these two paramagnetic species but rather due to the third and broad component that becomes dominant at T < 35 K, as shown in Fig. 4D.

Combining the results from the ESR and SQUID studies, the fol-lowing conclusions can be drawn. First, judging from the fact that

the Fe3+ ions remain magnetically isolated throughout the entire

temperature range of 5 to 300 K, the Cs2Ag(Bi:Fe)Br6 crystal is not

globally magnetic. Otherwise, the isolated Fe3+ ESR signal would

have decreased due to their strong magnetic coupling occurring below the magnetic phase transition temperature, which would have directly correlated with a concomitant increase of an FMR- or AFMR-like signal. Second, although the origin of the broad ESR component that is responsible for the steep rising of magnetic re-sponse at low temperatures is unknown, it is induced by the Fe in-corporation in the crystal and it resembles that of an FMR or AFMR signal. As no obvious shift in the resonance field of the broad ESR signal can be resolved, which should otherwise reflect the internal magnetic field induced by magnetization, it is not possible at pres-ent to single out whether or not the broad ESR signal corresponds to an FMR or AFMR signal. If it is induced by FMR or AFMR coupling, the corresponding magnetic response should then be (i) local, only representing a small part of the total sample volume, and (ii) rather

weak, to account for the negligible reduction of the isolated Fe3+ ESR

signal intensity and the small saturated magnetic moment detected in SQUID. A possible origin of the corresponding magnetic response

is ferromagnetic coupling from localized regions with high Fe3+

con-centrations due to a strong composition fluctuation. If this is true,

then it could hint that it may be possible to make the Cs2Ag(Bi:Fe)

Br6 crystal a magnetic semiconductor provided that a higher Fe

composition could be achieved. Thus, this work points toward some possible directions for further studies aiming to explore lead-free double perovskites for future spintronic applications. We cannot either rule out the possibility of magnetic inclusions of phase separated Fe clusters/composites beyond the detection limit of our characterization techniques in this work. The origin of this magnetic response needs further investigations on more samples with different iron concentrations, which is beyond the scope of this study at this stage.

DISCUSSION

In summary, we alloy magnetic element Fe3+ into optoelectronic

double perovskite Cs2AgBiBr6, providing a new class of halide double

perovskite alloys, which can be potentially used for spintronics.

The magnetic Fe3+ ions homogenously distribute with small FeBr

6

clusters in the crystal lattice. The temperature-dependent specific

heat and SPD of Cs2Ag(Bi:Fe)Br6 demonstrate a cubic-tetragonal

structural transition at ~120 K. Combined SQUID and ESR mea-surements reveal a strongly temperatudependent magnetic re-sponse at temperatures below 30 K, which is tentatively attributed a weak ferromagnetic or antiferromagnetic response from localized regions containing, e.g., phase separated magnetic clusters/composites or high Fe compositions due to alloy fluctuations. We believe that halide double perovskites hold great potential for a next generation Fig. 4. Magnetic characterizations of Cs2Ag(Bi:Fe)Br6. (A) Temperature-dependent

magnetic susceptibility of Cs2Ag(Bi:Fe)Br6 under zero-field-cooled (ZFC) and

field-cooled (FC) procedures with a magnetic field H = 1000 Oe. (B) Isothermal field-dependent magnetization of Cs2Ag(Bi:Fe)Br6. The inset plot presents the

low–magnetic field magnetization from −0.1 to 0.1 T. (C) ESR spectrum from Cs2AgBiBr6 (the gray curves) and Cs2Ag(Bi:Fe)Br6 (the red curves) crystal powder,

measured at 7.2 and 40 K, respectively. The simulated ESR spectra (the blue curves) are obtained from a spin Hamiltonian analysis, which consist of the contributions from Fe3+, with S = 5/2 and g = 2.032, and a defect center of unknown origin, with S = 1/2 and g = 2.032. The discrepancy between the simulated and experimental

ESR spectra indicates the existence of a third ESR signal—a broad background signal, which becomes more pronounced at low temperatures. (D) Temperature dependence of total microwave absorption (the black spheres), together with the contributions from the three individual ESR components revealed from the spin Hamiltonian analysis.

on January 4, 2021

http://advances.sciencemag.org/

(6)

of spintronic devices that have been widely studied in oxide-based perovskites and merit further study to realize their full potential. MATERIALS AND METHODS

Preparation of Cs2AgBiBr6 crystals

Solid CsBr (213 mg, 1.00 mmol) and BiBr3 (225 mg, 0.5 mmol) were

dissolved in 3 ml of 47% HBr. Solid AgBr (94 mg, 0.5 mmol) was then added to the solution, and the mixture was transferred into a

25 cm3 Teflon-lined digestion autoclave. The autoclave was sealed

and placed in the oven where it was heated at 120°C for 24 hours. After being slowly cooled to room temperature, red octahedral single crystals were achieved (yield ca. 79% based on Ag).

Preparation of Cs2Ag(Bi:Fe)Br6 crystals

Solid CsBr (213 mg, 1.00 mmol), BiBr3 (135 mg, 0.3 mmol), and

FeBr3 (59.1 mg, 0.2 mmol) were dissolved in 4 ml of 47% HBr. Solid

AgBr (94 mg, 0.5 mmol) was then added to the solution, and the

mixture was transferred into a 25-cm3 Teflon-lined digestion

auto-clave. The autoclave was sealed and placed in the oven where it was heated at 120°C for 24 hours. After being slowly cooled to room temperature, black octahedral single crystals were achieved (yield ca. 67% based on Ag).

Basic characterizations of Cs2AgBiBr6 and

Cs2Ag(Bi:Fe)Br6 crystals

The optical absorption measurements were carried out using a PE Lambda 950 ultraviolet-visible spectrophotometer. XRD patterns of the products were recorded with a X’Pert PRO x-ray diffractometer using Cu K1 irradiation ( = 1.5406 Å). Energy-dispersive x-ray spectroscopy analysis was performed using an LEO 1550 scan-ning electron microscope operated at 20-kV accelerating voltage,

with an Oxford Instruments X-Max 80-mm2 Silicon Drift Detector.

ICP-OES measurements were carried out using a PerkinElmer Avio

500 spectrometer with the Cs2Ag(Bi:Fe)Br6 crystal powder

dis-solved in aqua regia. XPS of Cs2Ag(Bi:Fe)Br6 crystals was performed

using a Scienta ESCA200 spectrometer with a base pressure of

2 × 10−10 mbar (ultrahigh vacuum) and monochromatized Al (K)

radiation (h = 1486 eV). NEXAFS spectroscopy and SR-XPS measurements were performed at the Elettra synchrotron radiation facility in Trieste, Italy. Specific heat (Cp) measurement was per-formed using physical property measurement system (PPMS-9 T)

Quantum Design Co. with Cs2Ag(Bi:Fe)Br6 crystals.

Temperature-dependent SPD

The SPD experiments were carried out at beamline BL02B2 at

SPring-8, Japan (32). The Cs2AgBiBr6 and Cs2Ag(Bi:Fe)Br6 crystal

samples were crushed into a fine powder and filled into a borosilicate glass capillary with a diameter of 0.1 mm. The incident x-ray beam was monochromatized to 24.8 keV (0.5 Å) using Si(111) double crystals. The detail of the wavelength was 0.50018 Å determined

by using standard reference materials CeO2.The temperature of the

powder sample was controlled using a nitrogen and helium gas blower. Rietveld refinements were performed using the JANA2006 software package (26).

Solid-state nuclear magnetic resonance

The ssNMR spectra were recorded at 11.7 T using a Bruker Avance III HD spectrometer. The 4-mm cross-polarization MAS (CP/

MAS) probe was used for 133Cs and 209Bi experiments at Larmor

frequency of (133Cs) = 65.611 MHz and (209Bi) = 80.858 MHz,

respectively. 133Cs MAS NMR experiments were collected at 10-kHz

spinning speed without 1H decoupling. The recycle delay was 4 s for

all ssNMR experiments. The 133Cs chemical shift was calibrated

using solid CsCl (133Cs, 228.1 ppm) (33). The pulse length was set to

2.4 s at 100 W for maximal signal intensity. The 133Cs rotor-

synchronized (1 loop) spin-echo MAS NMR experiments (90°-nR

-180°-nR-acq.) (34) were performed. The 133Cs-133Cs correlation

MAS NMR spectra were recorded using nuclear Overhauser effect spectroscopy–type three-pulse sequence. Duration of the spin- exchange period between the second and third pulse was 10 s. Spectral width in both frequency dimensions was rotor-synchronized

to be 10 kHz. The indirect detection period t1 consisted of 128

in-crements each made of 48 scans. The 209Bi chemical shift was

cali-brated using a saturated solution of Bi(NO3)3·5H2O in concentrated

HNO3 (209Bi, 0.0 ppm) (35). The pulse length was 9.0 s at 100 W at

maximal signal intensity. 209Bi NMR experiments were collected at

static conditions using spin-echo NMR experiments (90°--180°-acq.) (34), and the delay between pulses was 20 s. To compensate fric-tional heating of the spinning samples, all NMR experiments were measured under active cooling. The sample temperature was main-tained at 298 K, and the temperature calibration was performed on

Pb(NO3)2 using a calibration procedure described in the literature

(36). Dried sample was packed into ZrO2 rotors and subsequently

stored at room temperature. All NMR spectra were processed using the TopSpin 3.5 pl2 software package.

Magnetic characterization

Magnetic susceptibility measurements were performed in the com-mercial Quantum Design magnetic property measurement system

(MPMS3). Lots of small Cs2Ag(Bi:Fe)Br6 crystals with a total mass

of 29.9 mg were loaded into a polycrystalline sample holder, which was fixed in a brass sample rod. Both ZFC and FC procedures were used to check possible magnetic transition in this compound. ESR was performed with a Bruker Elexsys E500 spectrometer operating

at about 9.3 GHz. ESR spectra were recorded in dark. Cs2Ag(Bi:Fe)

Br6 and Cs2AgBiBr6 crystal powder samples were prepared via

col-lecting many randomly oriented small crystals into sealed and evac-uated quartz tubes and placed in an He-flow cryostat.

X-ray single crystallography

The SC-XRD data for Cs2Ag(Bi:Fe)Br6 and Cs2AgBiBr6 were

col-lected at 298 K with graphite monochromated Mo K ( = 0.71073 Å) on a charge-coupled device area detector (Bruker D8 VENTURE and Bruker-SMART), respectively. Data reductions and absorption corrections were performed with the APEX3 suite, SAINT and SADABS software packages, respectively. Structures were solved by a direct method using the SHELXL-97 software package. The non-hydrogen atoms were anisotropically refined using the full-matrix

least-squares method on F2. The details about data collection,

struc-ture refinement, and crystallography are summarized in table S2. SUPPLEMENTARY MATERIALS

Supplementary material for this article is available at http://advances.sciencemag.org/cgi/ content/full/6/45/eabb5381/DC1

REFERENCES AND NOTES

1. A. Kojima, K. Teshima, Y. Shirai, T. Miyasaka, Organometal halide perovskites as visible-light sensitizers for photovoltaic cells. J. Am. Chem. Soc. 131, 6050–6051 (2009).

on January 4, 2021

http://advances.sciencemag.org/

(7)

S C I E N C E A D V A N C E S

|

R E S E A R C H A R T I C L E

2. S. Bai, P. Da, C. Li, Z. Wang, Z. Yuan, F. Fu, M. Kawecki, X. Liu, N. Sakai, J. T.-W. Wang, S. Huettner, S. Buecheler, M. Fahlman, F. Gao, H. J. Snaith, Planar perovskite solar cells with long-term stability using ionic liquid additives. Nature 571, 245–250 (2019).

3. N. Wang, L. Cheng, R. Ge, S. Zhang, Y. Miao, W. Zou, C. Yi, Y. Sun, Y. Cao, R. Yang, Y. Wei, Q. Guo, Y. Ke, M. Yu, Y. Jin, Y. Liu, Q. Ding, D. Di, L. Yang, G. Xing, H. Tian, C. Jin, F. Gao, R. H. Friend, J. Wang, W. Huang, Perovskite light-emitting diodes based on solution-processed self-organized multiple quantum wells. Nat. Photonics 10, 699–704 (2016).

4. W. Xu, Q. Hu, S. Bai, C. Bao, Y. Miao, Z. Yuan, T. Borzda, A. J. Barker, E. Tyukalova, Z. Hu, M. Kawecki, H. Wang, Z. Yan, X. Liu, X. Shi, K. Uvdal, M. Fahlman, W. Zhang, M. Duchamp, J.-M. Liu, A. Petrozza, J. Wang, L.-M. Liu, W. Huang, F. Gao, Rational molecular passivation for high-performance perovskite light-emitting diodes. Nat. Photonics 13, 418–424 (2019).

5. H. Zhu, Y. Fu, F. Meng, X. Wu, Z. Gong, Q. Ding, M. V. Gustafsson, M. T. Trinh, S. Jin, X.-Y. Zhu, Lead halide perovskite nanowire lasers with low lasing thresholds and high quality factors. Nat. Mater. 14, 636–642 (2015).

6. Q. Lin, A. Armin, P. L. Burn, P. Meredith, Filterless narrowband visible photodetectors.

Nat. Photonics 9, 687–694 (2015).

7. Q. Chen, J. Wu, X. Ou, B. Huang, J. Almutlaq, A. A. Zhumekenov, X. Guan, S. Han, L. Liang, Z. Yi, J. Li, X. Xie, Y. Wang, Y. Li, D. Fan, D. B. L. Teh, A. H. All, O. F. Mohammed, O. M. Bakr, T. Wu, M. Bettinelli, H. Yang, W. Huang, X. Liu, All-inorganic perovskite nanocrystal scintillators. Nature 561, 88–93 (2018).

8. P. Odenthal, W. Talmadge, N. Gundlach, R. Wang, C. Zhang, D. Sun, Z.-G. Yu, Z. V. Vardeny, Y. S. Li, Spin-polarized exciton quantum beating in hybrid organic–inorganic perovskites.

Nat. Phys. 13, 894–899 (2017).

9. C. Zhang, D. Sun, C.-X. Sheng, Y. X. Zhai, K. Mielczarek, A. Zakhidov, Z. V. Vardeny, Magnetic field effects in hybrid perovskite devices. Nat. Phys. 11, 427–434 (2015). 10. D. Giovanni, H. Ma, J. Chua, M. Grätzel, R. Ramesh, S. Mhaisalkar, N. Mathews,

T. C. Sum, Highly spin-polarized carrier dynamics and ultralarge photoinduced magnetization in CH3NH3PbI3 perovskite thin films. Nano Lett. 15, 1553–1558

(2015).

11. M. Kepenekian, R. Robles, C. Katan, D. Sapori, L. Pedesseau, J. Even, Rashba

and Dresselhaus effects in hybrid organic–inorganic perovskites: From basics to devices.

ACS Nano 9, 11557–11567 (2015).

12. D. Niesner, M. Wilhelm, I. Levchuk, A. Osvet, S. Shrestha, M. Batentschuk, C. Brabec, T. Fauster, Giant Rashba splitting in CH3NH3PbBr3 organic-inorganic perovskite.

Phys. Rev. Lett. 117, 126401 (2016).

13. M. Kepenekian, J. Even, Rashba and Dresselhaus couplings in halide perovskites: Accomplishments and opportunities for spintronics and spin–orbitronics. J. Phys. Chem. Lett. 8, 3362–3370 (2017).

14. Y.-C. Hsiao, T. Wu, M. Li, B. Hu, Magneto-optical studies on spin-dependent charge recombination and dissociation in perovskite solar cells. Adv. Mater. 27, 2899–2906 (2015).

15. F. Pulizzi, Spintronics. Nat. Mater. 11, 367 (2012).

16. G. Long, C. Jiang, R. Sabatini, Z. Yang, M. Wei, L. N. Quan, Q. Liang, A. Rasmita, M. Askerka, G. Walters, X. Gong, J. Xing, X. Wen, R. Quintero-Bermudez, H. Yuan, G. Xing, X. R. Wang, D. Song, O. Voznyy, M. Zhang, S. Hoogland, W. Gao, Q. Xiong, E. H. Sargent, Spin control in reduced-dimensional chiral perovskites. Nat. Photonics 12, 528–533 (2018).

17. J. Wang, C. Zhang, H. Liu, R. McLaughlin, Y. Zhai, S. R. Vardeny, X. Liu, S. McGill, D. Semenov, H. Guo, R. Tsuchikawa, V. V. Deshpande, D. Sun, Z. V. Vardeny, Spin-optoelectronic devices based on hybrid organic-inorganic trihalide perovskites.

Nat. Commun. 10, 129 (2019).

18. B. Náfrádi, P. Szirmai, M. Spina, H. Lee, O. V. Yazyev, A. Arakcheeva, D. Chernyshov, M. Gibert, L. Forró, E. Horváth, Optically switched magnetism in photovoltaic perovskite CH3NH3(Mn:Pb)I3. Nat. Commun. 7, 13406 (2016).

19. B. Cai, X. Chen, M. Xie, S. Zhang, X. Liu, J. Yang, W. Zhou, S. Guo, H. Zeng, A class of Pb-free double perovskite halide semiconductors with intrinsic

ferromagnetism, large spin splitting and high Curie temperature. Mater. Horiz. 5, 961–968 (2018).

20. W. Ning, F. Wang, B. Wu, J. Lu, Z. Yan, X. Liu, Y. Tao, J.-M. Liu, W. Huang, M. Fahlman, L. Hultman, T. C. Sum, F. Gao, Long electron–hole diffusion length in high-quality lead-free double perovskite films. Adv. Mater. 30, e1706246 (2018).

21. W. Ning, X.-G. Zhao, J. Klarbring, S. Bai, F. Ji, F. Wang, S. I. Simak, Y. Tao, X.-M. Ren, L. Zhang, W. Huang, I. A. Abrikosov, F. Gao, Thermochromic lead-free halide double perovskites. Adv. Funct. Mater. 29, 1807375 (2019).

22. R. D. Shannon, Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. A 32, 751–767 (1976). 23. L. Vegard, Die Konstitution der Mischkristalle und die Raumfüllung der Atome. Z. Physik

5, 17–26 (1921).

24. F. Zheng, V. Pérez-Dieste, J. L. McChesney, Y.-Y. Luk, N. L. Abbott, F. J. Himpsel, Detection and switching of the oxidation state of Fe in a self-assembled monolayer. Surf. Sci. 587, L191–L196 (2005).

25. J. F. Moulder, W. F. Stickle, P. E. Sobol, K. D. Bombden, Handbook of X-Ray

Photoelectron Spectroscopy (Physical Electronics Division, Perkin-Elmer Corporation,

1992), pp. 80–81.

26. V. Petříček, M. Dušek, L. Palatinus, Crystallographic computing system JANA2006: General features. Z. Kristallogr. Cryst. Mater. 229, 345–352 (2014).

27. L. Schade, A. D. Wright, R. D. Johnson, M. Dollmann, B. Wenger, P. K. Nayak, D. Prabhakaran, L. M. Herz, R. Nicholas, H. J. Snaith, P. G. Radaelli, Structural and optical properties of Cs2AgBiBr6 double perovskite. ACS Energy Lett. 4, 299–305

(2019).

28. Y. Ishii, N. P. Wickramasinghe, S. Chimon, A new approach in 1D and 2D 13C

high-resolution solid-state NMR spectroscopy of paramagnetic organometallic complexes by very fast magic-angle spinning. J. Am. Chem. Soc. 125, 3438–3439 (2003).

29. R. J. Holmberg, T. Burns, S. M. Greer, L. Kobera, S. A. Stoian, I. Korobkov, S. Hill, D. L. Bryce, T. K. Woo, M. Murugesu, Intercalation of coordinatively unsaturated FeIII

ion within interpenetrated metal–organic framework MOF-5. Chem. 22, 7711–7715 (2016).

30. L. Kobera, S. Abbrent, L. Holcova, M. Urbanova, D. Kolousek, B. Dousova, J. Brus, Spying on Fe ions and their role in modified aluminosilicates during the sorption of anions using solid-state NMR spectroscopy. Micropor. Mesopor. Mat. 241, 115–122 (2017).

31. D. J. Kubicki, D. Prochowicz, A. Pinon, G. Stevanato, A. Hofstetter, S. M. Zakeeruddin, M. Grätzel, L. Emsley, Doping and phase segregation in Mn2+- and Co2+-doped lead

halide perovskites from 133Cs and 1H NMR relaxation enhancement. J. Mater. Chem. A 7,

2326–2333 (2019).

32. S. Kawaguchi, M. Takemoto, K. Osaka, E. Nishibori, C. Moriyoshi, Y. Kubota, Y. Kuroiwa, K. Sugimoto, High-throughput powder diffraction measurement system consisting of multiple MYTHEN detectors at beamline BL02B2 of SPring-8. Rev. Sci. Instrum. 88, 085111 (2017).

33. A. R. Haase, M. A. Kerber, D. Kessler, J. Kronenbitter, H. Krüger, O. Lutz, M. Müller, A. Nolle, Nuclear magnetic shielding and quadrupole coupling of 133Cs in cesium salt powders.

Z. Naturforsch. 32, 952–956 (1977).

34. E. L. Hahn, Spin Echoes. Phys. Rev. 80, 580–594 (1950).

35. H. Hamaed, M. W. Laschuk, V. V. Terskikh, R. W. Schurko, Application of solid-state 209Bi

NMR to the structural characterization of bismuth-containing materials. J. Am. Chem. Soc. 131, 8271–8279 (2009).

36. J. Brus, Heating of samples induced by fast magic-angle spinning. Solid State Nucl. Magn.

Reson. 16, 151–160 (2000).

Acknowledgments: We thank J. Klarbring, S. I. Simak, I. A. Abrikosov, and A. Wildes for the helpful discussions. Funding: This work was financially supported by Knut and Alice Wallenberg Foundation (Dnr KAW 2019.0082), the Swedish Energy Agency (2018-004357 and P43288-1), the Swedish Government Strategic Research Area in Materials Science on Functional Materials at Linköping University (faculty grant SFO-Mat-LiU no. 2009-00971), and the Grant Agency of the Czech Republic (grant GA19-05259S). We acknowledge the Diamond Light Source for access to beamline I19 (MT20805). We also thank L. Saunders and D. Allan for the technical support. L.W. acknowledges the China Scholarship Council (CSC) for the financial support. Work at Argonne (magnetic susceptibility measurements) was supported by the U.S. Department of Energy, Office of Science, Basic Energy Sciences, Materials Sciences and Engineering Division. Synchrotron radiation experiment was performed at the BL02B2 of SPring-8 with approval of the Japan Synchrotron Radiation Research Institute (JASRI) (proposal no. 2017B0074). We are grateful for the support of the European Community’s Seventh Framework Programme (FP7/2007–2013, grant agreement no. 312284) for the research at the Materials Science Beamline at the Elettra Synchrotron and the CERIC-ERIC Consortium for access to experimental facilities and financial support of the Czech Ministry of Education (LM2015057). G.A.C., M.C., and R.D.M. pay special thanks to N. Tsud and K. C. Prince at Elettra Synchrotron for assistance with the experiments. We also thank J. Bradley for assistance with measurements at the Elettra Synchrotron. M.C. and G.A.C. acknowledge the CERIC users’ grant for travel funding to visit the Elettra Synchrotron, while R.D.M. thanks the International Synchrotron Access Program of the Australian Synchrotron for travel funding to attend the experiment at the Elettra Synchrotron. F.G. is a Wallenberg Academy Fellow. Author contributions: W.N. and F.G. conceived the idea for the manuscript and designed the experiments. W.N. and F.J. developed the synthesis procedures and performed the basic chemical and physical characterization. J.B. performed the SQUID measurements and analyzed the results under the supervision of D.Y.C. and M.G.K. Y.P. and F.M. measured and analyzed the ESR results under the supervision of W.M.C. and I.A.B. L.K., S.A., and J.B. measured and

on January 4, 2021

http://advances.sciencemag.org/

(8)

analyzed the ssNMR results. S.S., S.K., H.I., and Y.K. helped with the specific heat and SPD measurements. L.W. performed the SC-XRD measurement and analysis under the supervision of L.S. M.C., R.D.M., and G.A.C. performed the NEXAFS measurements. W.N. wrote the paper with the contributions from the coauthors. F.G. supervised the project. All authors commented on the manuscript. Competing interests: The authors declare that they have no competing interests. Data and materials availability: All data needed to evaluate the conclusions in the paper are present in the paper and/or the Supplementary Materials. Additional data related to this paper may be requested from the authors.

Submitted 3 March 2020 Accepted 16 September 2020 Published 6 November 2020 10.1126/sciadv.abb5381

Citation: W. Ning, J. Bao, Y. Puttisong, F. Moro, L. Kobera, S. Shimono, L. Wang, F. Ji, M. Cuartero, S. Kawaguchi, S. Abbrent, H. Ishibashi, R. De Marco, I. A. Bouianova, G. A. Crespo, Y. Kubota, J. Brus, D. Y. Chung, L. Sun, W. M. Chen, M. G. Kanatzidis, F. Gao, Magnetizing lead-free halide double perovskites. Sci. Adv. 6, eabb5381 (2020).

on January 4, 2021

http://advances.sciencemag.org/

(9)

Magnetizing lead-free halide double perovskites

Yoshiki Kubota, Jiri Brus, Duck Young Chung, Licheng Sun, Weimin M. Chen, Mercouri G. Kanatzidis and Feng Gao Cuartero, Shogo Kawaguchi, Sabina Abbrent, Hiroki Ishibashi, Roland De Marco, Irina A. Bouianova, Gaston A. Crespo, Weihua Ning, Jinke Bao, Yuttapoom Puttisong, Fabrizo Moro, Libor Kobera, Seiya Shimono, Linqin Wang, Fuxiang Ji, Maria

DOI: 10.1126/sciadv.abb5381 (45), eabb5381.

6

Sci Adv

ARTICLE TOOLS http://advances.sciencemag.org/content/6/45/eabb5381

MATERIALS

SUPPLEMENTARY http://advances.sciencemag.org/content/suppl/2020/11/02/6.45.eabb5381.DC1

REFERENCES

http://advances.sciencemag.org/content/6/45/eabb5381#BIBL This article cites 35 articles, 0 of which you can access for free

PERMISSIONS http://www.sciencemag.org/help/reprints-and-permissions

Terms of Service Use of this article is subject to the

is a registered trademark of AAAS.

Science Advances

York Avenue NW, Washington, DC 20005. The title

(ISSN 2375-2548) is published by the American Association for the Advancement of Science, 1200 New

Science Advances

License 4.0 (CC BY-NC).

Science. No claim to original U.S. Government Works. Distributed under a Creative Commons Attribution NonCommercial Copyright © 2020 The Authors, some rights reserved; exclusive licensee American Association for the Advancement of

on January 4, 2021

http://advances.sciencemag.org/

References

Related documents

Detta är ett exempel på borgande för handlingsfrihet eftersom god förståelseav motståndaren syftar till ”att kunna möta oförutsedda händelseutvecklingar.” 114

Traditional gender patterns were observed when the form and communication of activities were well known for students and based on the logic of sport.. Other patterns were also

Ó¾ÔÕ

In the first case, the time taken to send a message across a link is taken to be 100 times faster than the average session time of nodes.. In the second case, it is 19

Keywords: Student writing, Academic Writing, University Writing, Writing Processes, Academic Literacies, Digital Literacies, Teacher Education, So- ciomateriality, Higher

In this study, we aim to investigate to what extent adolescents with social anxiety and/or compensation motives could develop supportive friendships online. We also aim to explore

Istället för att skapa relationer mellan människor, mellan vårdad och vårdare, där grunden är omtänksamhet och empati och där människor bemöter varandra med respekt och kan

Om vi, i enlighet med rollteoretikerna, byter ut begreppet kontroll mot rolltyd- lighet i Karaseks (1979) krav/kontroll-modell, ser vi att de inte längre befinner sig i det