efficient photoluminescence of isotropic rare-earth

28
doi.org/10.26434/chemrxiv.11494584.v1 Efficient Photoluminescence of Isotropic Rare-Earth Oxychloride Nanocrystals from a Solvothermal Route Guillaume Gouget, Morgane Pellerin, Lauriane Pautrot-d’Alençon, Thierry Le Mercier, Christopher B. Murray Submitted date: 02/01/2020 Posted date: 07/01/2020 Licence: CC BY-NC-ND 4.0 Citation information: Gouget, Guillaume; Pellerin, Morgane; Pautrot-d’Alençon, Lauriane; Mercier, Thierry Le; Murray, Christopher B. (2020): Efficient Photoluminescence of Isotropic Rare-Earth Oxychloride Nanocrystals from a Solvothermal Route. ChemRxiv. Preprint. https://doi.org/10.26434/chemrxiv.11494584.v1 Eu3+-doped sub-10 nm LaOCl nanocrystals with 43 % photoluminescence quantum yield were prepared by solvothermal synthesis from hydrated rare-earth chlorides. As-obtained nanocrystals are nearly spherical, monodisperse and stable as colloidal dispersions. These combined features should intensify the interest for nanocrystalline rare-earth oxyhalides and their optical properties. File list (3) download file view on ChemRxiv LaOCl_ChemRxiv.pdf (2.18 MiB) download file view on ChemRxiv GG_LaOCl_ChemRxiv.docx (5.25 MiB) download file view on ChemRxiv GG_LaOCl_SI_2020-01-02.docx (1.91 MiB)

Upload: others

Post on 26-Dec-2021

12 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Efficient Photoluminescence of Isotropic Rare-Earth

doi.org/10.26434/chemrxiv.11494584.v1

Efficient Photoluminescence of Isotropic Rare-Earth OxychlorideNanocrystals from a Solvothermal RouteGuillaume Gouget, Morgane Pellerin, Lauriane Pautrot-d’Alencon, Thierry Le Mercier, Christopher B. Murray

Submitted date: 02/01/2020 • Posted date: 07/01/2020Licence: CC BY-NC-ND 4.0Citation information: Gouget, Guillaume; Pellerin, Morgane; Pautrot-d’Alencon, Lauriane; Mercier, Thierry Le;Murray, Christopher B. (2020): Efficient Photoluminescence of Isotropic Rare-Earth Oxychloride Nanocrystalsfrom a Solvothermal Route. ChemRxiv. Preprint. https://doi.org/10.26434/chemrxiv.11494584.v1

Eu3+-doped sub-10 nm LaOCl nanocrystals with 43 % photoluminescence quantum yield were prepared bysolvothermal synthesis from hydrated rare-earth chlorides. As-obtained nanocrystals are nearly spherical,monodisperse and stable as colloidal dispersions. These combined features should intensify the interest fornanocrystalline rare-earth oxyhalides and their optical properties.

File list (3)

download fileview on ChemRxivLaOCl_ChemRxiv.pdf (2.18 MiB)

download fileview on ChemRxivGG_LaOCl_ChemRxiv.docx (5.25 MiB)

download fileview on ChemRxivGG_LaOCl_SI_2020-01-02.docx (1.91 MiB)

Page 2: Efficient Photoluminescence of Isotropic Rare-Earth

a. Department of Chemistry, University of Pennsylvania, Philadelphia, Pennsylvania 19104, United States.

b. Solvay, Research and Innovation Center Paris, F-93308, Aubervilliers, France.

c. Address here.

Electronic Supplementary Information (ESI) available: experimental section and additional data, Fig. S1-S9.

Efficient photoluminescence of isotropic rare-earth oxychloride nanocrystals from a solvothermal route

Guillaume Gouget,a Morgane Pellerin,

b Lauriane Pautrot‐D'Alençon,

b Thierry Le Mercier,

b and

Christopher B. Murray*a

Eu3+

-doped sub-10 nm LaOCl nanocrystals with 43 %

photoluminescence quantum yield were prepared by

solvothermal synthesis from hydrated rare-earth chlorides. As-

obtained nanocrystals are nearly spherical, monodisperse and

stable as colloidal dispersions. These combined features should

intensify the interest for nanocrystalline rare-earth oxyhalides and

their optical properties.

For several decades, rare-earth (RE) oxyhalides (REOX) have

been investigated for a variety of applications including bio-

imaging,1 small molecule sensing,

2 high-energy photon

detection,3,4

heterogeneous catalysis in the gas or liquid

phase,5–9

halide ion conduction,10

and light-emitting displays.11

REOXs are attractive for their luminescence properties: (i)

several REs can be readily combined in a single REOX structure

with various proportions, yielding optimized luminescence

features,1,4,12-16

and (ii) REOXs are low phonon energy hosts

(e.g. ca. 440 cm-1

for LaOCl),6,17

which prevents the excited RE-

emitting centers from luminescence quenching via multi-

phonon relaxation. PL excitation may occur in various energies

ranges, from infrared (IR) all the way to X-rays. LaOCl bandgap

is 4.17 eV according to density functional theory (DFT)

calculation,18

and it displays several absorption features in the

UV range.19,20

IR and visible light directly excite REs emitters,

whereas UV or X-Ray light excitation occur via the absorption

from LaOCl followed by energy transfer to the RE emitter.

LaOCl then sensitizes RE emission in the UV or X-Ray range.

The production of stable colloidal dispersions of REOX

nanocrystals (NCs) in the sub-20 nm regime is beneficial for

their processability (e.g. formation of thin close-packed films)

and for bio-compatibility.21

The search for routes to REOX NCs

has become more intense in the past ten years. They were

recently reviewed by Sarbajit Banerjee and co-workers.22

The

LaOCl NCs systematically display platelet morphologies, in

relation with their tetragonal structure (c axis orthogonal to

the most expressed facets). LaOCl structure (space group

P4/nmm, No. 129, Fig. 1) comprises layers of Cl anions

intercalated between slabs of LaO units (formally +1

charge·unit-1

) along the ab plane. Each La atom is coordinated

to 4 O and 5 Cl in a C4v symmetry environment. The separation

of cation sites along c limits cross-relaxation mechanisms

between emitting RE3+

centers.4,13,23

The lamellar structure is

also attractive for Cl ion conduction.10

Structural anisotropy is

thus beneficial for optical and transport properties. In 2009,

Du et al. reported ultra-thin Eu3+

-doped LaOCl nanoplates

(aspect ratio from 3.2 to 4.4). Hydrated RE trichloroacetates

single-source precursors were decomposed in the presence of

various amines.24

The relatively low PL quantum yield (PLQY)

of 5 % Eu3+

-doped LaOCl nanoplates (PLQY = 2.8 %) was

attributed to the predominant non-radiative pathways due to

high-frequency vibrations of long-chain amine ligands on the

surface of the nanoplates. Anisotropic morphologies can have

negative consequences on the photoluminescence (PL). More

isotropic (ideally cubic or spherical) morphologies would

provide a lower surface-to-volume ratio, i.e. a higher volume

for Eu3+

centers to be spatially separated from the interface,

and consequently higher PL efficiency.

Herein we report a synthetic route towards sub-10 nm

spherical LaOCl colloidal NCs, relying on the reaction of

hydrated chloride salts and oleylamine (OLA) at 220 °C.

Monodisperse samples were obtained, leading to self-

assembling of the dispersed crystals into close-packed films.

Fig. 1 Structure of LaOCl (P4/nmm, No. 129). The grey

polyhedron highlights La 9-fold coordinated environnement.

Page 3: Efficient Photoluminescence of Isotropic Rare-Earth

Using Eu3+

-doped (5 %) LaOCl NCs as a case study, PL efficiency

was improved, as detailed later. The correlations between

synthetic conditions, morphology and luminescence properties

of the NCs were discussed.

Details of the syntheses are reported in the Experimental

Section of Supporting Information. Briefly, hydrated RE

chlorides (starting either from LaCl3.7H2O only or a mix of

LaCl3.7H2O and EuCl3.6H2O (La:Eu molar ratio of 19:1)) were

dissolved in methanol (MeOH) prior to the addition of (OLA) in

air, at RT. The MeOH was evaporated and the mixture was

then degassed for 1 h at 120°C and 1 Torr. The reaction was

held at 220 °C during 1 h under N2, then oleic acid (OA) was

injected to the mixture and the reactant medium was cooled

to RT by blowing cool air over the vessel. Purification was

performed by two successive (i) additions of toluene and

MeOH followed by (ii) centrifugation to yield a precipitate and

(iii) removal of the supernatant. The final white pellet was

redispersed in chloroform, leading to a clear colorless solution.

X-ray diffraction (XRD) patterns in Fig. 2a were performed on

films obtained by dropcasting both dispersions on glass slides.

In both cases, LaOCl was unambiguously identified as a pure

phase (see also Supporting Information, Fig. S1). Significant

broadening of the peaks were attributed to small crystalline

domains. Using the Scherrer formula, apparent crystal sizes

from XRD diagrams were 6.6±0.3 and 7.1±0.6 nm for the Eu3+

-

doped and undoped samples, respectively.25

Energy dispersive

spectroscopy (EDS) (Fig. S2) indicated a molar ratio

RE:Cl = 1.00, in agreement with the expected value for REOCl.

Focusing on Eu3+

-doped LaOCl, the effective Eu insertion rate

(Eu:(La+Eu) molar ratio) was 4.8±0.2 %, according to elemental

analysis by inductive coupled plasma-optical emission

spectrometry (ICP-OES) measurements. The insertion rate of

Eu corresponds to the initial 19:1 (La:Eu) molar ratio.

Transmission electron microscopy (TEM, Fig. 2b and c) showed

ordered spherical NCs (see also Fig. S3). Selected area electron

diffraction (Fig. 2b, inset) confirmed that the NCs have the

oxychloride structure, as observed by XRD. Their average size

was 8.2±1.1 nm, according to TEM. This value was 24 % higher

than the apparent crystal size calculated from XRD, indicating

that the nanoparticles were monocrystalline. The deviation

was attributed to underestimation of apparent crystal size,

due to crystals strains contributing to XRD peak broadening.

The average aspect ratio was 1.1 according to TEM. The NCs

self-assembled upon dropcasting the dispersion on a TEM grid.

Monolayers (and bilayers, Fig. S3) with hexagonal packing

occured, as observed on the fast Fourier transform (FFT,

Fig. 2c, inset). The average center-to-center distance between

two neighboring NCs was 10.0 nm, so the shortest edge-to-

edge distance was 1.8 nm. This separation length is typically

observed for NCs capped with OLA or OA.26

Fourier transform

infra-red spectroscopy indicated that OLA binds to the surface

of the NCs after the purification steps (Fig. S4). OA was added

at the end of the heating step at 220 °C to avoid irreversible

precipitation in the early stages of the purification. It was

washed away in the final product, as confirmed by FTIR.

Hydrated lanthanum chloride was used for the first time as a

precursor to LaOCl NCs. The synthesis was performed at

220 °C. Alternative routes based on the solvothermal

decomposition of RE alkoxides or haloalkoxide intermediates

are reported between 300 and 330 °C.24,30

Other routes require

annealing steps above 500 °C.6,27–29

LaCl3.7H2O partially

dehydrates between 50 and 200 °C in argon atmosphere,31

and

it yields LaOCl and LaCl3 as competitive products.32

The

precursor contains La-Cl and La-O bonds already formed in the

initial precursor, as opposed to previously reported routes.22

We degassed the mixture of LaCl3.7H2O and OLA at 120 °C

prior to the reaction at 220 °C. At the end of the degassing

step, La probably maintained a Cl- and O-rich first coordination

sphere. A lower temperature of synthesis is then explained by

a reaction path involving successive dehydrations of the

intermediate, instead of the decomposition of an alkoxide.

Near-spherical LaOCl NCs (aspect ratio of 1.1) were

synthetized, while nanoplates were obtained in previous

reports.22

The anisotropic morphology was related to the

layered structure of LaOCl (tetragonal crystal cell) along the c

axis. Prefered growth along <110> directions was observed

when LaOCl ultrathin nanoplates (aspect ratio 4.0) were

synthetized from La(CCl3COO)3 with OLA.24

In contrast, our

study provided an example of isotropic morphology, which we

attributed to equal growth rates along the inequivalent

directions. We propose that the reactivity of intermediate

species dictates the relative growth rates, then resulting in the

difference in shapes between previous work24

and ours. The

nature of the intermediate species remains unclear, however

LaCl3.7H2O probably reacted into a Cl- and O-rich monomer,

which was responsible for lower temperatures of synthesis

and near-spherical morphologies. Future nvestigations on

intermediate species should be performed to better

understand and control shapes of LaOCl NCs. GdOCl and LaOBr

NCs were also synthetized by reacting gadolinium chloride and

lanthanum bromide (Fig. S5 and S6, respectively) in OLA.

Fig. 2 Structural characterization of LaOCl NCs. a) XRD patterns

of undoped and 5 % Eu3+

-doped samples. TEM pictures of b)

(inset: SAED) and c) (inset: FFT) 5 % Eu3+

-doped sample.

Page 4: Efficient Photoluminescence of Isotropic Rare-Earth

Fig. 3 Optical characterization of LaOCl NCs dispersed in

chloroform. a) mass exctinction coefficient of undoped and 5

% Eu3+

-doped samples, picture of a dispersion in inset. b) PL

(inset: dispersion irradiated at 300 nm) and c) PL excitation

spectra of 5 % Eu3+

-doped sample.

A controlled amount of Eu3+

doping was achieved by

introducing the desired ratio of RE precursors at the beginning

of the synthesis. The mass exctinction coefficient of the Eu3+

-

doped and undoped sample dispersions were compared on

Fig. 3a. Neither showed significant light absorption above 400

nm, yielding transparent solutions, as observed in inset. Both

doped and undoped samples displayed a peak at 286 nm,

before the mass exctinction coefficient dramatically increases

around 270 nm. These features were attributed to LaOCl, as

observed in previous work.18-20

The slope of absorption front

edge was sharper for the undoped sample than for the doped

one. An additional broad contribution was detected for the

Eu3+

-doped sample between 280 and 330 nm. This feature was

also observed in other Eu3+

-doped LaOCl materials.18,19

The

optical bandgap of undoped LaOCl sample determined by the

Tauc method (Fig. S7) was 4.11 eV.33

The value was close to

the one calculated for LaOCl structure from DFT (4.17 eV).18

The PL spectrum of Eu3+

-doped LaOCl spherical NCs excited at

λexc = 300 nm (Fig. 3b) showed narrow emission peaks in the

visible region, with a maximized intensity at λem = 618 nm. It

was responsible for the red color of the dispersion under UV

irradiation. Each peak was attributed to Eu3+

4f-4f inner

transitions (energy diagram in Fig. S8), and labeled accordingly.

Focusing on 5D0→

7F2 transition, two main peaks are detected

at 615 and 618 nm due to Stark splitting.34,35

In the case of

LaOCl nanoplates,24

inversion of relative intensity occured in

favour of the lower wavelength, which was attributed to a

lower symmetry of Eu3+

in a distorted environment compared

to bulk crystalline sample.36,37

In contrast, here the LaOCl NCs

display luminescence features similar to the bulk phase,

suggesting less distortion of Eu3+

environment from C4v

symmetry (Fig. 1). The PL excitation spectrum was performed

at λem = 618 nm (Fig. 3c). The main broad contribution

centered on 300 nm was related to charge-transfer band (CTB)

from O 2p to Eu 4f orbitals.29

The CTB was responsible for the

absorption feature between 280 and 330 nm in Eu3+

-doped

LaOCl (Fig. 3a).18

The narrower and weaker peaks above

350 nm were due to Eu3+

4f-4f excitations. The emission

spectrum at λexc = 395 nm (corresponding to Eu3+

7F0→

5L6

transition) in Fig. S9 showed a set of emission peaks similar to

the peaks observed at λexc = 300 nm (Fig. 3b).

The efficiency of radiative recombination pathways for Eu3+

in

LaOCl NCs dispersed in chloroform was probed by PL decay

and absolute PLQY measurements (Fig. 4). The PL decay at 618

nm was fit with a single exponential component (χ2 = 1.07).

The lifetime of excited states was 1.0 ms, which was 33 %

longer than for Eu3+

-doped LaOCl nanoplates (0.75 ms).24

The

lifetime of Eu3+

excited states increased, i.e. the rate of non-

radiative relaxation decreased. The absolute PLQY was

calculated from Equation (1), using an integrating sphere:

𝑃𝐿𝑄𝑌𝜆𝑒𝑥𝑐=

𝑁𝑒𝑚

𝑁𝑎𝑏𝑠 (1),

where Nabs is the number of photons absorbed by the sample,

Nem the number of photons emitted and λexc the wavelength of

excitation. These quantities were calculated as the intensity

difference between the Eu3+

-doped sample and the solvent as

reference, for the scattered light (Nabs) and the emitted light

(Nem). Intensities of absorption and emission were plotted as a

function of wavelength in Fig. 4b at λexc = 300 nm,

corresponding to the excitation in the O 2p→Eu 4f CTB. The

associated PLQY300 is 43 %. A similar experiment at

λexc = 395 nm (direct Eu3+

7F0→

5L6 excitation) leads to

PLQY395 = 14 %. The PLQY395 was 5 times higher than reported

perviously for LaOCl nanoplates.24

Higher PL efficiency is

observed for near-spherical NCs (aspect ratio of 1.1), as

evidenced by lifetime and PLQY measurements. PL quenching

is due to non-radiative recombination taking place at the

surface of the nanoparticles. The spherical morphology

minimizes the surface-to-volume ratio of the NCs. The increase

in lifetime and PLQY were then attributed to emissive centers

(Eu3+

) being, in average, further away from the surface.

Page 5: Efficient Photoluminescence of Isotropic Rare-Earth

Fig. 4 PL efficiency of Eu3+

emission in LaOCl NCs. a) PL decay

diagram fitted with a single exponential function. b) light

intensity absorbed (black) and emitted (red) by the sample

exposed at λexc = 300 nm in an integrating sphere.

In summary, monodisperse sub-10 nm LaOCl spherical NCs

have been obtained for the first time, starting from hydrated

RE chlorides and using lower temperatures than previously

reported. PL efficiency was improved by accessing isotropic

morphologies, as confirmed by longer PL lifetime and higher

PLQY. This work stresses the importance of exploring new

synthetic methods to diversify morphologies of LaOCl NCs,

which yields improved optical properties. In addition to their

processability, exemplified by the formation of close-packed

NCs films, isotropic LaOCl NCs doped with RE emitters should

be highly advantageous in various applications relying on their

optical properties. They can be further improved by different

strategies: (i) delineating conditions favoring isotropic

morphologies for REOX NCs (ii) improving the crystallinity of

the nanoparticles, and (iii) forming a protective crystalline shell

around the NCs. These strategies all highlight the need of more

experimental work on this new route, together with a better

understanding of the intermediate species involved during the

synthesis.

Conflicts of interest

There are no conflicts to declare.

Notes and references

1 T. Konishi, M. Shimizu, Y. Kameyama and K. Soga, J. Mater. Sci. Mater. Electron., 2007, 18, 183–186.

2 A. Marsal, G. Dezanneau, A. Cornet and J. R. Morante, Sensors Actuators, B Chem., 2003, 95, 266–270.

3 J. Rabatin, US patent, 375814, 1972.

4 G. R. Waetzig, G. A. Horrocks, J. W. Jude, G. V. Villalpando, L. Zuin and S. Banerjee, Inorg. Chem., 2018, 57, 5842–5849.

5 H. Yasuda, L.-N. He and T. Sakakura, J. Catal., 2002, 209, 547–550.

6 S. G. Podkolzin, E. E. Stangland, M. E. Jones, E. Peringer and J. A. Lercher, J. Am. Chem. Soc., 2007, 129, 2569–2576.

7 A. W. A. M. van der Heijden, A. J. M. Mens, R. Bogerd and B. M. Weckhuysen, Catal. Letters, 2008, 122, 238–246.

8 P. Afanasiev, M. Aouine, C. Deranlot and T. Epicier, Chem. Mater., 2010, 22, 5411–5419.

9 Q. Xie, Y. Wang, B. Pan, H. Wang, W. Su and X. Wang, Catal. Commun., 2012, 27, 21–25.

10 N. Imanaka, K. Okamoto and G. Adachi, Angew. Chemie Int. Ed., 2002, 41, 3890–3892.

11 G. Li and J. Lin, Chem. Soc. Rev., 2014, 43, 7099–7131. 12 D. Kim, J. R. Jeong, Y. Jang, J.-S. Bae, I. Chung, R. Liang, D.-K.

Seo, S.-J. Kim and J.-C. Park, Phys. Chem. Chem. Phys., 2019, 21, 1737–1749.

13 G. R. Waetzig, G. A. Horrocks, R. D. Davidson, J. W. Jude, G. V. Villalpando, L. Zuin and S. Banerjee, J. Phys. Chem. C, 2018, 122, 16412–16423.

14 G. Li, Z. Hou, C. Peng, W. Wang, Z. Cheng, C. Li, H. Lian and J. Lin, Adv. Funct. Mater., 2010, 20, 3446–3456.

15 J. Hölsä, Chem. Phys. Lett., 1984, 112, 246–248. 16 Z. Xia, J. Li, Y. Luo and L. Liao, J. Am. Ceram. Soc., 2012, 95,

3229–3234. 17 J. Hölsä, E. Kestilä, K. Koski and H. Rahiala, J. Alloys Compd.,

1995, 225, 193–197. 18 L. Lv, T. Wang, S. Li, Y. Su and X. Wang, CrystEngComm, 2016,

18, 907–916. 19 D. Kim, S. Park, S. Kim, S.-G. Kang and J.-C. Park, Inorg.

Chem., 2014, 53, 11966–11973. 20 A. A. Dakhel, J. Solid State Chem., 2016, 241, 219–224. 21 M. A. Boles, D. Ling, T. Hyeon and D. V. Talapin, Nat. Mater.,

2016, 15, 141–153. 22 M. Udayakantha, P. Schofield, G. R. Waetzig and S. Banerjee,

J. Solid State Chem., 2019, 270, 569–592. 23 G. R. Waetzig, G. A. Horrocks, J. W. Jude, L. Zuin and S.

Banerjee, Nanoscale, 2016, 8, 979–986. 24 Y. P. Du, Y. W. Zhang, L. D. Sun and C. H. Yan, J. Am. Chem.

Soc., 2009, 131, 3162–3163. 25 A. L. Patterson, Phys. Rev., 1939, 56, 978-982. 26 X. Ye, J. Chen, M. Engel, J. A. Millan, W. Li, L. Qi, G. Xing, J. E.

Collins, C. R. Kagan, J. Li, S. C. Glotzer and C. B. Murray, Nat. Chem., 2013, 5, 466–473.

27 E. Peringer, M. Salzinger, M. Hutt, A. A. Lemonidou and J. A. Lercher, Top. Catal., 2009, 52, 1220–1231.

28 G. Li, C. Li, C. Zhang, Z. Cheng, Z. Quan, C. Peng and J. Lin, J. Mater. Chem., 2009, 19, 8936.

29 Q. Kong, J. Wang, X. Dong, W. Yu and G. Liu, J. Mater. Sci. Mater. Electron., 2013, 24, 4745–4756.

30 S. W. Depner, K. R. Kort, C. Jaye, D. A. Fischer and S. Banerjee, J. Phys. Chem. C, 2009, 113, 14126–14134.

31 D. K. Sahoo, R. Mishra, H. Singh and N. Krishnamurthy, J. Alloys Compd., 2014, 588, 578–584.

32 J. E. Powell and H. R. Burkholder, J. Inorg. Nucl. Chem., 1960, 14, 65–70.

33 J. Tauc, R. Grigorovici and A. Vancu, Phys. status solidi, 1966, 15, 627–637.

34 K. Binnemans, Coord. Chem. Rev., 2015, 295, 1–45. 35 P. D. Dragic, M. Cavillon and J. Ballato, Appl. Phys. Rev.,

2018, 5, 041301. 36 H. Wang, M. Uehara, H. Nakamura, M. Miyazaki and H.

Maeda, Adv. Mater., 2005, 17, 2506–2509. 37 D. Tu, Y. Liu, H. Zhu, R. Li, L. Liu and X. Chen, Angew. Chemie

Int. Ed., 2013, 52, 1128–1133.

Page 6: Efficient Photoluminescence of Isotropic Rare-Earth

1

SUPPORTING INFORMATION

Efficient photoluminescence of isotropic rare-earth oxychloride nanocrystals

from a solvothermal route

Guillaume Gouget,a Morgane Pellerin,

b Lauriane Pautrot‐D'Alençon,

b Thierry Le Mercier,

b and Christopher B.

Murray*,a

aDepartment of Chemistry, University of Pennsylvania, Philadelphia, Pennsylvania 19104, United States

bSolvay, Research and Innovation Center Paris, F-93308, Aubervilliers, France

EXPERIMENTAL SECTION 2

Chemicals 2

Synthesis of LaOCl nanocrystals 2

Structural characterization 2

Optical characterization 3

ADDITIONAL DATA 3

Figure S1. Le Bail refinements of undoped and 5 % Eu3+-doped LaOCl nanocrystals. 3

Figure S2. SEM-EDS spectrum of a region of aggregated Eu3+-doped LaOCl NCs. 4

Figure S3. TEM of 5 % Eu3+-doped LaOCl. 4

Figure S4. FTIR spectrum of Eu3+-doped LaOCl nanocrystals. 4

Figure S5. Structural characterization of GdOCl nanocrystals. 5

Figure S6. Structural characterization of LaOBr nanocrystals. 5

Figure S7. Tauc plot of dispersed LaOCl NCs. 6

Figure S8. Partial energy diagram of Eu3+ (4f6). 6

Figure S9. Emission spectrum of 5 % Eu3+-doped LaOCl NCs excitated at λexc = 395 nm. 7

REFERENCES 7

Page 7: Efficient Photoluminescence of Isotropic Rare-Earth

2

EXPERIMENTAL SECTION

Chemicals. Heptahydrated lanthanum trichloride (GFS Chemicals, 99.9 %), hexahydrated europium

trichloride (GFS Chemicals, 99.9 %), oleylamine (OLA, Sigma-Aldrich, 70 %), oleic acid (OA, Aldrich,

90 %), methanol (MeOH), toluene and chloroform (Fisher, ACS grade) were stored in air prior to use.

Synthesis of LaOCl nanocrystals. Hydrated RE chlorides in various proportions (0.5 mmol total RE

molar content) were dissolved in MeOH (10 mL) prior to addition of OLA (20 mL) in air, at RT. MeOH

was evaporated and the mixture was then degassed 1 h at 120°C and 1 Torr. The temperature was

increased to 220 °C in 13 min under N2. After 1 h at 220 °C, OA (5 mL) was injected to the mixture,

the temperature suddenly dropped to 200 °C, and then reactant medium was further cooled down to

RT by blowing cool air on the vessel. 10 mL of toluene was added, then 10 mL of MeOH. The

precipitation of a white solid was observed. The precipitate was centrifuged at 8900 g during 3 min,

the supernatant was discarded and a white pellet was recovered. The pellet was redispersed in 10 mL

of toluene, and then the dispersion was reprecipitated with 15 mL of MeOH. After purification, the

pellet was redispersed in chloroform.

Structural characterization. Powder X-ray diffraction (XRD) diagrams were acquired on a Rigaku

Smartlab diffractometer with Cu Kα radiation (λ =1.5416 Å, 40 mA and 30 kV) and parallel beam

setting in the θ-2θ geometry. 2θ range was screened between 10 and 80 ° with 0.05° (2θ) steps and

during 35 s per step. Peak identification was performed using the cif files from the crystallographic

open database (COD), collection code 9009170 (LaOCl) and the inorganic crystal structure database

(ICSD), collection codes 77820 (GdOCl) and 84336 (LaOBr)). Apparent crystal size εhkl were calculated

using the Scherrer formula:

β is the integral width of the peak, λ is the X-ray source wavelength and θ is the incident beam angle.

β was calculated for each separate experimental peak (hkl) by fitting it with a pseudo-Voigt function,

using the WinPLOTR function (FullProfSuite).1 ε was then averaged from all values calculated on a

single diagram.

Energy dispersive spectroscopy (EDS) was performed on a JEOL 7500F high resolution scanning

electron microscope equipped with an EDS detector. Acceleration tension of the source was 10 kV.

RE/Cl atomic ratio was calculated from the average of EDS spectra measured on three distinct

100 x 50 µm2 areas of aggregated particles. Data were analyzed using the INCA software (Oxford).

The sample was dropcasted on a Si wafer, followed by drying, before a carbon layer was sputtered.

Inductive coupled plasma-optical emission spectrometry (ICP-OES) was performed on a Spectro

Genesis spectrometer with a concentric nebulizer. Calibrated solution of La and Eu were purchased

from Inorganic Ventures. Solid samples were dissolved in HNO3 (70 %mas.) at 60 °C.

Transmission electron microscopy (TEM) images of NCs were collected on a JEOL JEM-1400

microscope operating at 120 keV. Samples were prepared by depositing a drop of NCs in chloroform

(3.10-2 g.L-1) on a carbon-coated copper grid. Selected area electron diffraction (SAED) was performed

with a camera distance of 20 cm. Interreticular planes were deduced from distances calibrated with a

known sample of silicon particles. Size distribution data were extracted from at least 100 particles on

pictures taken on at least 3 distinct regions of a TEM grid.

Page 8: Efficient Photoluminescence of Isotropic Rare-Earth

3

Optical characterization. Absorption spectra of the dispersions were measured in a 10 mm quartz

cuvette, on a Cary 5000 UV−vis−NIR spectrophotometer (Agilent Technologies) from 800 to 200 nm

with a 1 nm step. The same cuvette filled with solvent was used as a reference. Photoluminescence

(PL) experiments were performed on a FLS1000 fluorimeter (Edinburgh Instruments) equipped with a

Xe lamp. PL spectrum was acquired from 310 to 800 nm with 1 nm steps and 0.1 s per step. Slit

bandwidths on both incident and emission sides were 1.1 nm. PL excitation spectrum was acquired

from 270 to 610 nm, 1 nm steps and 0.1 s per step. Slit bandwidths on both incident and emission

sides were 2.0 nm. PL decay measurements were performed with a pulsed laser source (MCS) at

378 nm, 125 Hz pulse frequency. For absolute PL quantum yield measurement, an integrating sphere

was adapted on the fluorimeter. Emission spectra were acquired in the same condition for the

reference (cuvette with solvent) and sample: λexc = 300 nm (or 395 nm), 10 nm slit bandwidth on the

incident side, 1.0 nm slit bandwidth on the emission side, 270 (or 370) to 800 nm, 1 nm steps and

0.5 s per step.

ADDITIONAL DATA

Figure S1. Le Bail refinements of a) undoped and b) 5 % Eu3+-doped LaOCl nanocrystals. The peak at 2θ = 44.5 ° is due to the aluminum plate. It is taken into account in the fit (Al, Fm-3m, ICSD .cif file code 64700) in the refinements. In both cases, refined cell parameters of LaOCl (P4/nmm) are similar to values reported (COD .cif file 9009170): a = 4.119 Å and c = 6.883 Å. Refinements were performed on JANA 2006.2

Page 9: Efficient Photoluminescence of Isotropic Rare-Earth

4

Figure S2. SEM-EDS spectrum of a region of aggregated Eu3+-doped LaOCl NCs. Si detection is attributed to a contamination from sample preparation or SEM chamber. It is not considered for quantification, nor O and C.

Figure S3. a) TEM picture of 5 % Eu3+-doped LaOCl nanocrystals forming close-packed monolayers and bilayers. b) SAED picture with diffraction rings indexed along with LaOCl structure. Lower left inset is a scheme reproducing (full lines) major and *(dashed) minor rings. c) Size distribution diagram of 5 % Eu3+-doped LaOCl nanocrystals (black) and gaussian function centered at 8.2 nm with a 1.1 nm standard deviation.

Figure S4. FTIR spectrum of Eu3+-doped LaOCl nanocrystals showing bond vibrations typical of OLA, as

opposed to OA.3,4

Page 10: Efficient Photoluminescence of Isotropic Rare-Earth

5

Figure S5. Structural characterization of GdOCl nanocrystals. The synthesis was adapted from LaOCl experiment, using hexahydrate gadolinium chloride as RE precursor, and without addition of OA at the end of the heating step. a) XRD diagram showed that GdOCl (P4/nmm, No. 129) is the only phase obtained. The various peak widths were due to strong shape anisotropy, in relation with crystalline structure. Especially, the narrowest peaks corresponded to (hk0) interreticular planes (blue), i.e. orthogonal to c axis. b-d) TEM pictures showed stacks of thin plates with rough edges, forming 200-500 nm spherical aggregates. e) Peak position and apparent crystal size (from Scherrer formula) of some interreticular planes (hkl). εhk0 = 16.6±1.1 nm and ε00l = 4.2±0.4 nm confirmed that GdOCl NCs were anisotropic with a smaller dimension in c direction, whereas facets orthogonal to c were the biggest. This is in agreement with the plate shape observed by TEM (b)).

Figure S6. Structural characterization of LaOBr nanocrystals. The synthesis was adapted from LaOCl experiment, using lanthanum bromide instead of the chloride, heating at 330 °C instead of 220 °C, and without addition of OA at the end of the heating step. a) XRD diagram showed that LaOBr (P4/nmm, No. 129, ICSD code 84336) was the only phase obtained. b-d) TEM pictures of nanoplates 15-20 nm thick and 30-100 nm long. Particles displayed an elongated hexagon-like shape when oriented along the edge. This shape was similar than LaOCl obtained by Depner et al..5 Rod-like arrays (single- and multi-linear) formed via face-to-face alignement of the nanoplates. e) Peak position and apparent crystal size (from Scherrer formula) of some interreticular planes (hkl). εhk0 = 20.6±0.1 nm ε00l = 15.1±0.5 nm confirmed that LaOBrl NCs are anisotropic with a smaller dimension in c direction, whereas the biggest facets were perpendicular to c axis. This was in agreement with the shape observed by TEM (b)). εhk0 was smaller than plates length observed by TEM. The size difference could be due to tha nanoplates being polycrystalline.

Page 11: Efficient Photoluminescence of Isotropic Rare-Earth

6

Figure S7. Tauc plot of dispersed LaOCl NCs.

For direct bandgap semiconductors such as LaOCl,6 optical bandgap measurement was based on the Tauc relationship (2) between the absorption coefficient α, the photon energy hν and the bandgap energy Eg:

where B was a constant for allowed transitions. For relatively low absorption A (A < 1), the absorption coefficient of a perfectly non-scattering dispersion followed the linear Beer-Lambert law:

with l the optical path length (1 cm).

Figure S8. Partial energy diagram of Eu3+ (4f6) taking into account interelectronic repulsion and spin-orbit coupling.7

Page 12: Efficient Photoluminescence of Isotropic Rare-Earth

7

Figure S9. PL spectrum of Eu3+-doped LaOCl NCs dispersed in chloroform, excitation at λexc = 395 nm. REFERENCES 1 T. Roisnel and J. Rodríquez-Carvajal, Mater. Sci. Forum, 2001, 378–381, 118–123. 2 V. Petrícek, M. Dušek and L. Palatinus, Zeitschrift fur Krist., 2014, 229, 345–352. 3 S. Mourdikoudis and L. M. Liz-Marzán, Chem. Mater., 2013, 25, 1465–1476. 4 I. O. Perez De Berti, M. V Cagnoli, G. Pecchi, J. L. Alessandrini, S. J. Stewart, J. F. Bengoa and S.

G. Marchetti, Nanotechnology, 2013, 24, 175601. 5 S. W. Depner, K. R. Kort, C. Jaye, D. A. Fischer and S. Banerjee, J. Phys. Chem. C, 2009, 113,

14126–14134. 6 L. Lv, T. Wang, S. Li, Y. Su and X. Wang, CrystEngComm, 2016, 18, 907–916. 7 K. Binnemans, Coord. Chem. Rev., 2015, 295, 1–45.

Page 14: Efficient Photoluminescence of Isotropic Rare-Earth

Efficient photoluminescence of isotropic rare-earth oxychloride nanocrystals from a solvothermal route

Guillaume Gouget,a Morgane Pellerin,b Lauriane Pautrot‐D'Alençon,b Thierry Le Mercier,b and Christopher B. Murray*a

a. Department of Chemistry, University of Pennsylvania, Philadelphia, Pennsylvania 19104, United States.

b. Solvay, Research and Innovation Center Paris, F-93308, Aubervilliers, France.c. Address here.Electronic Supplementary Information (ESI) available: experimental section andadditional data, Fig. S1‐S9.

Page 15: Efficient Photoluminescence of Isotropic Rare-Earth

Eu3+-doped sub-10 nm LaOCl nanocrystals with 43 %

photoluminescence quantum yield were prepared by

solvothermal synthesis from hydrated rare-earth chlorides. As-

obtained nanocrystals are nearly spherical, monodisperse and

stable as colloidal dispersions. These combined features

should intensify the interest for nanocrystalline rare-earth

oxyhalides and their optical properties.

For several decades, rare‐earth (RE) oxyhalides (REOX) have

been investigated for a variety of applications including bio‐

imaging,1 small molecule sensing,2 high‐energy photon

detection,3,4 heterogeneous catalysis in the gas or liquid

phase,5–9 halide ion conduction,10 and light‐emitting

displays.11 REOXs are attractive for their luminescence

properties: (i) several REs can be readily combined in a

single REOX structure with various proportions, yielding

optimized luminescence features,1,4,12‐16 and (ii) REOXs are

low phonon energy hosts (e.g. ca. 440 cm‐1 for LaOCl),6,17

which prevents the excited RE‐emitting centers from

luminescence quenching via multi‐phonon relaxation. PL

excitation may occur in various energies ranges, from

infrared (IR) all the way to X‐rays. LaOCl bandgap is 4.17 eV

according to density functional theory (DFT) calculation,18

and it displays several absorption features in the UV

range.19,20 IR and visible light directly excite REs emitters,

whereas UV or X‐Ray light excitation occur via the

absorption from LaOCl followed by energy transfer to the

RE emitter. LaOCl then sensitizes RE emission in the UV or

X‐Ray range.

The production of stable colloidal dispersions of REOX

nanocrystals (NCs) in the sub‐20 nm regime is beneficial for

their processability (e.g. formation of thin close‐packed

films) and for bio‐compatibility.21 The search for routes to

REOX NCs has become more intense in the past ten years.

They were recently reviewed by Sarbajit Banerjee and co‐

workers.22 The LaOCl NCs systematically display platelet

morphologies, in relation with their tetragonal structure (c

axis orthogonal to the most expressed facets). LaOCl

structure (space group P4/nmm, No. 129, Fig. 1) comprises

layers of Cl anions intercalated between slabs of LaO units

(formally +1 charge·unit‐1) along the ab plane. Each La atom

is coordinated to 4 O and 5 Cl in a C4v symmetry

environment. The separation of cation sites along c limits

cross‐relaxation mechanisms between emitting RE3+

centers.4,13,23 The lamellar structure is also attractive for Cl

ion conduction.10 Structural anisotropy is thus beneficial for

optical and transport properties. In 2009, Du et al. reported

ultra‐thin Eu3+‐doped LaOCl nanoplates (aspect ratio from

3.2 to 4.4). Hydrated RE trichloroacetates single‐source

precursors were decomposed in the presence of various

amines.24 The relatively low PL quantum yield (PLQY) of

5 % Eu3+‐doped LaOCl nanoplates (PLQY = 2.8 %) was

attributed to the predominant non‐radiative pathways due

to high‐frequency vibrations of long‐chain amine ligands on

the surface of the nanoplates. Anisotropic morphologies

can have negative consequences on the

photoluminescence (PL). More isotropic (ideally cubic or

spherical) morphologies would provide a lower surface‐to‐

volume ratio, i.e. a higher volume for Eu3+ centers to be

spatially separated from the interface, and consequently

higher PL efficiency.

Herein we report a synthetic route towards sub‐10 nm

spherical LaOCl colloidal NCs, relying on the reaction of

hydrated chloride salts and oleylamine (OLA) at 220 °C.

Monodisperse samples were obtained, leading to self‐

assembling of the dispersed crystals into close‐packed

films.

Fig. 1 Structure of LaOCl (P4/nmm, No. 129). The grey

polyhedron highlights La 9‐fold coordinated

environnement.

Using Eu3+‐doped (5 %) LaOCl NCs as a case study, PL

efficiency was improved, as detailed later. The correlations

between synthetic conditions, morphology and

luminescence properties of the NCs were discussed.

Details of the syntheses are reported in the Experimental

Section of Supporting Information. Briefly, hydrated RE

chlorides (starting either from LaCl3.7H2O only or a mix of

LaCl3.7H2O and EuCl3.6H2O (La:Eu molar ratio of 19:1)) were

dissolved in methanol (MeOH) prior to the addition of

(OLA) in air, at RT. The MeOH was evaporated and the

mixture was then degassed for 1 h at 120°C and 1 Torr. The

reaction was held at 220 °C during 1 h under N2, then oleic

acid (OA) was injected to the mixture and the reactant

medium was cooled to RT by blowing cool air over the

vessel. Purification was performed by two successive (i)

additions of toluene and MeOH followed by (ii)

centrifugation to yield a precipitate and (iii) removal of the

supernatant. The final white pellet was redispersed in

chloroform, leading to a clear colorless solution. X‐ray

diffraction (XRD) patterns in Fig. 2a were performed on

films obtained by dropcasting both dispersions on glass

slides. In both cases, LaOCl was unambiguously identified

as a pure phase (see also Supporting Information, Fig. S1).

Significant broadening of the peaks were attributed to

small crystalline domains. Using the Scherrer formula,

apparent crystal sizes from XRD diagrams were 6.6±0.3 and

7.1±0.6 nm for the Eu3+‐doped and undoped samples,

respectively.25 Energy dispersive spectroscopy (EDS) (Fig.

S2) indicated a molar ratio RE:Cl = 1.00, in agreement with

the expected value for REOCl. Focusing on Eu3+‐doped

LaOCl, the effective Eu insertion rate (Eu:(La+Eu) molar

Page 16: Efficient Photoluminescence of Isotropic Rare-Earth

ratio) was 4.8±0.2 %, according to elemental analysis by

inductive coupled plasma‐optical emission spectrometry

(ICP‐OES) measurements. The insertion rate of Eu

corresponds to the initial 19:1 (La:Eu) molar ratio.

Transmission electron microscopy (TEM, Fig. 2b and c)

showed ordered spherical NCs (see also Fig. S3). Selected

area electron diffraction (Fig. 2b, inset) confirmed that the

NCs have the oxychloride structure, as observed by XRD.

Their average size was 8.2±1.1 nm, according to TEM. This

value was 24 % higher than the apparent crystal size

calculated from XRD, indicating that the nanoparticles were

monocrystalline. The deviation was attributed to

underestimation of apparent crystal size, due to crystals

strains contributing to XRD peak broadening. The average

aspect ratio was 1.1 according to TEM. The NCs self‐

assembled upon dropcasting the dispersion on a TEM grid.

Monolayers (and bilayers, Fig. S3) with hexagonal packing

occured, as observed on the fast Fourier transform (FFT,

Fig. 2c, inset). The average center‐to‐center distance

between two neighboring NCs was 10.0 nm, so the shortest

edge‐to‐edge distance was 1.8 nm. This separation length is

typically observed for NCs capped with OLA or OA.26 Fourier

transform infra‐red spectroscopy indicated that OLA binds

to the surface of the NCs after the purification steps (Fig.

S4). OA was added at the end of the heating step at 220 °C

to avoid irreversible precipitation in the early stages of the

purification. It was washed away in the final product, as

confirmed by FTIR.

Hydrated lanthanum chloride was used for the first time as

a precursor to LaOCl NCs. The synthesis was performed at

220 °C. Alternative routes based on the solvothermal

decomposition of RE alkoxides or haloalkoxide

intermediates are reported between 300 and 330 °C.24,30

Other routes require annealing steps above 500 °C.6,27–29

LaCl3.7H2O partially dehydrates between 50 and 200 °C in

argon atmosphere,31 and it yields LaOCl and LaCl3 as

competitive products.32 The precursor contains La‐Cl and

La‐O bonds already formed in the initial precursor, as

opposed to previously reported routes.22 We degassed the

mixture of LaCl3.7H2O and OLA at 120 °C prior to the

reaction at 220 °C. At the end of the degassing step, La

probably maintained a Cl‐ and O‐rich first coordination

sphere. A lower temperature of synthesis is then explained

by a reaction path involving successive dehydrations of the

intermediate, instead of the decomposition of an alkoxide.

Near‐spherical LaOCl NCs (aspect ratio of 1.1) were

synthetized, while nanoplates were obtained in previous

reports.22 The anisotropic morphology was related to the

layered structure of LaOCl (tetragonal crystal cell) along the

c axis. Prefered growth along <110> directions was

observed when LaOCl ultrathin nanoplates (aspect ratio

4.0) were synthetized from La(CCl3COO)3 with OLA.24 In

contrast, our study provided an example of isotropic

morphology, which we attributed to equal growth rates

along the inequivalent directions. We propose that the

reactivity of intermediate species dictates the relative

growth rates, then resulting in the difference in shapes

between previous work24 and ours. The nature of the

intermediate species remains unclear, however LaCl3.7H2O

probably reacted into a Cl‐ and O‐rich monomer, which was

responsible for lower temperatures of synthesis and near‐

spherical morphologies. Future nvestigations on

intermediate species should be performed to better

understand and control shapes of LaOCl NCs. GdOCl and

LaOBr NCs were also synthetized by reacting gadolinium

chloride and lanthanum bromide (Fig. S5 and S6,

respectively) in OLA.

Fig. 2 Structural characterization of LaOCl NCs. a) XRD

patterns of undoped and 5 % Eu3+‐doped samples. TEM

pictures of b) (inset: SAED) and c) (inset: FFT) 5 % Eu3+‐

doped sample.

Fig. 3 Optical characterization of LaOCl NCs dispersed in

chloroform. a) mass exctinction coefficient of undoped and

5 % Eu3+‐doped samples, picture of a dispersion in inset. b)

PL (inset: dispersion irradiated at 300 nm) and c) PL

excitation spectra of 5 % Eu3+‐doped sample.

Page 17: Efficient Photoluminescence of Isotropic Rare-Earth

A controlled amount of Eu3+ doping was achieved by

introducing the desired ratio of RE precursors at the

beginning of the synthesis. The mass exctinction coefficient

of the Eu3+‐doped and undoped sample dispersions were

compared on Fig. 3a. Neither showed significant light

absorption above 400 nm, yielding transparent solutions, as

observed in inset. Both doped and undoped samples

displayed a peak at 286 nm, before the mass exctinction

coefficient dramatically increases around 270 nm. These

features were attributed to LaOCl, as observed in previous

work.18‐20 The slope of absorption front edge was sharper

for the undoped sample than for the doped one. An

additional broad contribution was detected for the Eu3+‐

doped sample between 280 and 330 nm. This feature was

also observed in other Eu3+‐doped LaOCl materials.18,19 The

optical bandgap of undoped LaOCl sample determined by

the Tauc method (Fig. S7) was 4.11 eV.33 The value was

close to the one calculated for LaOCl structure from DFT

(4.17 eV).18 The PL spectrum of Eu3+‐doped LaOCl spherical

NCs excited at λexc = 300 nm (Fig. 3b) showed narrow

emission peaks in the visible region, with a maximized

intensity at λem = 618 nm. It was responsible for the red

color of the dispersion under UV irradiation. Each peak was

attributed to Eu3+ 4f‐4f inner transitions (energy diagram in

Fig. S8), and labeled accordingly. Focusing on 5D0→7F2

transition, two main peaks are detected at 615 and 618 nm

due to Stark splitting.34,35 In the case of LaOCl nanoplates,24

inversion of relative intensity occured in favour of the lower

wavelength, which was attributed to a lower symmetry of

Eu3+ in a distorted environment compared to bulk

crystalline sample.36,37 In contrast, here the LaOCl NCs

display luminescence features similar to the bulk phase,

suggesting less distortion of Eu3+ environment from C4v

symmetry (Fig. 1). The PL excitation spectrum was

performed at λem = 618 nm (Fig. 3c). The main broad

contribution centered on 300 nm was related to charge‐

transfer band (CTB) from O 2p to Eu 4f orbitals.29 The CTB

was responsible for the absorption feature between 280

and 330 nm in Eu3+‐doped LaOCl (Fig. 3a).18 The narrower

and weaker peaks above 350 nm were due to Eu3+ 4f‐4f

excitations. The emission spectrum at λexc = 395 nm

(corresponding to Eu3+ 7F0→5L6 transition) in Fig. S9 showed

a set of emission peaks similar to the peaks observed at

λexc = 300 nm (Fig. 3b).

The efficiency of radiative recombination pathways for Eu3+

in LaOCl NCs dispersed in chloroform was probed by PL

decay and absolute PLQY measurements (Fig. 4). The PL

decay at 618 nm was fit with a single exponential

component (χ2 = 1.07). The lifetime of excited states was

1.0 ms, which was 33 % longer than for Eu3+‐doped LaOCl

nanoplates (0.75 ms).24 The lifetime of Eu3+ excited states

increased, i.e. the rate of non‐radiative relaxation

decreased. The absolute PLQY was calculated from

Equation (1), using an integrating sphere:

N|¿|

PLQY λexc=

N em

¿ (1),

where Nabs is the number of photons absorbed by the

sample, Nem the number of photons emitted and λexc the

Page 18: Efficient Photoluminescence of Isotropic Rare-Earth

wavelength of excitation. These quantities were calculated

as the intensity difference between the Eu3+‐doped sample

and the solvent as reference, for the scattered light (Nabs)

and the emitted light (Nem). Intensities of absorption and

emission were plotted as a function of wavelength in Fig.

4b at λexc = 300 nm, corresponding to the excitation in the

O 2p→Eu 4f CTB. The associated PLQY300 is 43 %. A similar

experiment at λexc = 395 nm (direct Eu3+ 7F0→5L6 excitation)

leads to PLQY395 = 14 %. The PLQY395 was 5 times higher

than reported perviously for LaOCl nanoplates.24 Higher PL

efficiency is observed for near‐spherical NCs (aspect ratio of

1.1), as evidenced by lifetime and PLQY measurements. PL

quenching is due to non‐radiative recombination taking

place at the surface of the nanoparticles. The spherical

morphology minimizes the surface‐to‐volume ratio of the

NCs. The increase in lifetime and PLQY were then attributed

to emissive centers (Eu3+) being, in average, further away

from the surface.

Fig. 4 PL efficiency of Eu3+ emission in LaOCl NCs. a) PL

decay diagram fitted with a single exponential function. b)

light intensity absorbed (black) and emitted (red) by the

sample exposed at λexc = 300 nm in an integrating sphere.

In summary, monodisperse sub‐10 nm LaOCl spherical NCs

have been obtained for the first time, starting from

hydrated RE chlorides and using lower temperatures than

previously reported. PL efficiency was improved by

accessing isotropic morphologies, as confirmed by longer

PL lifetime and higher PLQY. This work stresses the

importance of exploring new synthetic methods to diversify

morphologies of LaOCl NCs, which yields improved optical

properties. In addition to their processability, exemplified

by the formation of close‐packed NCs films, isotropic LaOCl

NCs doped with RE emitters should be highly advantageous

in various applications relying on their optical properties.

They can be further improved by different strategies: (i)

delineating conditions favoring isotropic morphologies for

REOX NCs (ii) improving the crystallinity of the

nanoparticles, and (iii) forming a protective crystalline shell

around the NCs. These strategies all highlight the need of

more experimental work on this new route, together with a

better understanding of the intermediate species involved

during the synthesis.

Conflicts of interest

There are no conflicts to declare.

Notes and references1 T. Konishi, M. Shimizu, Y. Kameyama and K. Soga, J. Mater.

Sci. Mater. Electron., 2007, 18, 183–186.

2 A. Marsal, G. Dezanneau, A. Cornet and J. R. Morante, Sensors Actuators, B Chem., 2003, 95, 266–270.

3 J. Rabatin, US patent, 375814, 1972.4 G. R. Waetzig, G. A. Horrocks, J. W. Jude, G. V. Villalpando,

L. Zuin and S. Banerjee, Inorg. Chem., 2018, 57, 5842–5849.

5 H. Yasuda, L.‐N. He and T. Sakakura, J. Catal., 2002, 209, 547–550.

6 S. G. Podkolzin, E. E. Stangland, M. E. Jones, E. Peringer and J. A. Lercher, J. Am. Chem. Soc., 2007, 129, 2569–2576.

7 A. W. A. M. van der Heijden, A. J. M. Mens, R. Bogerd andB. M. Weckhuysen, Catal. Letters, 2008, 122, 238–246.

8 P. Afanasiev, M. Aouine, C. Deranlot and T. Epicier, Chem. Mater., 2010, 22, 5411–5419.

9 Q. Xie, Y. Wang, B. Pan, H. Wang, W. Su and X. Wang, Catal. Commun., 2012, 27, 21–25.

10 N. Imanaka, K. Okamoto and G. Adachi, Angew. Chemie Int. Ed., 2002, 41, 3890–3892.

11 G. Li and J. Lin, Chem. Soc. Rev., 2014, 43, 7099–7131.12 D. Kim, J. R. Jeong, Y. Jang, J.‐S. Bae, I. Chung, R. Liang, D.‐

K. Seo, S.‐J. Kim and J.‐C. Park, Phys. Chem. Chem. Phys., 2019, 21, 1737–1749.

13 G. R. Waetzig, G. A. Horrocks, R. D. Davidson, J. W. Jude, G. V. Villalpando, L. Zuin and S. Banerjee, J. Phys. Chem. C, 2018, 122, 16412–16423.

14 G. Li, Z. Hou, C. Peng, W. Wang, Z. Cheng, C. Li, H. Lian and J. Lin, Adv. Funct. Mater., 2010, 20, 3446–3456.

15 J. Hölsä, Chem. Phys. Lett., 1984, 112, 246–248.16 Z. Xia, J. Li, Y. Luo and L. Liao, J. Am. Ceram. Soc., 2012,

95, 3229–3234.17 J. Hölsä, E. Kestilä, K. Koski and H. Rahiala, J. Alloys

Compd., 1995, 225, 193–197.18 L. Lv, T. Wang, S. Li, Y. Su and X. Wang, CrystEngComm,

2016, 18, 907–916.19 D. Kim, S. Park, S. Kim, S.‐G. Kang and J.‐C. Park, Inorg.

Chem., 2014, 53, 11966–11973.20 A. A. Dakhel, J. Solid State Chem., 2016, 241, 219–224.21 M. A. Boles, D. Ling, T. Hyeon and D. V. Talapin, Nat.

Mater., 2016, 15, 141–153.22 M. Udayakantha, P. Schofield, G. R. Waetzig and S.

Banerjee, J. Solid State Chem., 2019, 270, 569–592.23 G. R. Waetzig, G. A. Horrocks, J. W. Jude, L. Zuin and S.

Banerjee, Nanoscale, 2016, 8, 979–986.24 Y. P. Du, Y. W. Zhang, L. D. Sun and C. H. Yan, J. Am. Chem.

Soc., 2009, 131, 3162–3163.25 A. L. Patterson, Phys. Rev., 1939, 56, 978‐982.

Page 19: Efficient Photoluminescence of Isotropic Rare-Earth

26 X. Ye, J. Chen, M. Engel, J. A. Millan, W. Li, L. Qi, G. Xing, J. E. Collins, C. R. Kagan, J. Li, S. C. Glotzer and C. B. Murray, Nat. Chem., 2013, 5, 466–473.

27 E. Peringer, M. Salzinger, M. Hutt, A. A. Lemonidou and J. A. Lercher, Top. Catal., 2009, 52, 1220–1231.

28 G. Li, C. Li, C. Zhang, Z. Cheng, Z. Quan, C. Peng and J. Lin, J. Mater. Chem., 2009, 19, 8936.

29 Q. Kong, J. Wang, X. Dong, W. Yu and G. Liu, J. Mater. Sci. Mater. Electron., 2013, 24, 4745–4756.

30 S. W. Depner, K. R. Kort, C. Jaye, D. A. Fischer and S. Banerjee, J. Phys. Chem. C, 2009, 113, 14126–14134.

31 D. K. Sahoo, R. Mishra, H. Singh and N. Krishnamurthy, J. Alloys Compd., 2014, 588, 578–584.

32 J. E. Powell and H. R. Burkholder, J. Inorg. Nucl. Chem., 1960, 14, 65–70.

33 J. Tauc, R. Grigorovici and A. Vancu, Phys. status solidi, 1966, 15, 627–637.

34 K. Binnemans, Coord. Chem. Rev., 2015, 295, 1–45.35 P. D. Dragic, M. Cavillon and J. Ballato, Appl. Phys. Rev.,

2018, 5, 041301.36 H. Wang, M. Uehara, H. Nakamura, M. Miyazaki and H.

Maeda, Adv. Mater., 2005, 17, 2506–2509.37 D. Tu, Y. Liu, H. Zhu, R. Li, L. Liu and X. Chen, Angew.

Chemie Int. Ed., 2013, 52, 1128–1133.

Page 21: Efficient Photoluminescence of Isotropic Rare-Earth

SUPPORTING INFORMATION

Efficient photoluminescence of isotropic rare-earth oxychloride nanocrystals

from a solvothermal route

Guillaume Gouget,a Morgane Pellerin,b Lauriane Pautrot D'Alençon,‐ b Thierry Le Mercier,b and Christopher B.

Murray*,a

aDepartment of Chemistry, University of Pennsylvania, Philadelphia, Pennsylvania 19104, United StatesbSolvay, Research and Innovation Center Paris, F-93308, Aubervilliers, France

EXPERIMENTAL SECTION 2

Chemicals 2

Synthesis of LaOCl nanocrystals 2

Structural characterization 2

Optical characterization 3

ADDITIONAL DATA 3

Figure S1. Le Bail refinements of undoped and 5 % Eu3+-doped LaOCl nanocrystals. 3

Figure S2. SEM-EDS spectrum of a region of aggregated Eu3+-doped LaOCl NCs. 4

Figure S3. TEM of 5 % Eu3+-doped LaOCl. 4

Figure S4. FTIR spectrum of Eu3+-doped LaOCl nanocrystals. 4

Figure S5. Structural characterization of GdOCl nanocrystals. 5

Figure S6. Structural characterization of LaOBr nanocrystals. 5

Figure S7. Tauc plot of dispersed LaOCl NCs. 6

Figure S8. Partial energy diagram of Eu3+ (4f6). 6

Figure S9. Emission spectrum of 5 % Eu3+-doped LaOCl NCs excitated at λexc = 395 nm. 7

REFERENCES 7

1

Page 22: Efficient Photoluminescence of Isotropic Rare-Earth

EXPERIMENTAL SECTION

Chemicals. Heptahydrated lanthanum trichloride (GFS Chemicals, 99.9 %), hexahydrated europium

trichloride (GFS Chemicals, 99.9 %), oleylamine (OLA, Sigma-Aldrich, 70 %), oleic acid (OA, Aldrich,

90 %), methanol (MeOH), toluene and chloroform (Fisher, ACS grade) were stored in air prior to use.

Synthesis of LaOCl nanocrystals. Hydrated RE chlorides in various proportions (0.5 mmol total RE

molar content) were dissolved in MeOH (10 mL) prior to addition of OLA (20 mL) in air, at RT. MeOH

was evaporated and the mixture was then degassed 1 h at 120°C and 1 Torr. The temperature was

increased to 220 °C in 13 min under N2. After 1 h at 220 °C, OA (5 mL) was injected to the mixture, the

temperature suddenly dropped to 200 °C, and then reactant medium was further cooled down to RT

by blowing cool air on the vessel. 10 mL of toluene was added, then 10 mL of MeOH. The

precipitation of a white solid was observed. The precipitate was centrifuged at 8900 g during 3 min,

the supernatant was discarded and a white pellet was recovered. The pellet was redispersed in 10 mL

of toluene, and then the dispersion was reprecipitated with 15 mL of MeOH. After purification, the

pellet was redispersed in chloroform.

Structural characterization. Powder X-ray diffraction (XRD) diagrams were acquired on a Rigaku

Smartlab diffractometer with Cu Kα radiation (λ =1.5416 Å, 40 mA and 30 kV) and parallel beam

setting in the θ-2θ geometry. 2θ range was screened between 10 and 80 ° with 0.05° (2θ) steps and

during 35 s per step. Peak identification was performed using the cif files from the crystallographic

open database (COD), collection code 9009170 (LaOCl) and the inorganic crystal structure database

(ICSD), collection codes 77820 (GdOCl) and 84336 (LaOBr)). Apparent crystal size εhkl were calculated

using the Scherrer formula:

β is the integral width of the peak, λ is the X-ray source wavelength and θ is the incident beam angle.

β was calculated for each separate experimental peak (hkl) by fitting it with a pseudo-Voigt function,

using the WinPLOTR function (FullProfSuite).1 ε was then averaged from all values calculated on a

single diagram.

Energy dispersive spectroscopy (EDS) was performed on a JEOL 7500F high resolution scanning

electron microscope equipped with an EDS detector. Acceleration tension of the source was 10 kV.

RE/Cl atomic ratio was calculated from the average of EDS spectra measured on three distinct

100 x 50 µm2 areas of aggregated particles. Data were analyzed using the INCA software (Oxford). The

sample was dropcasted on a Si wafer, followed by drying, before a carbon layer was sputtered.

Inductive coupled plasma-optical emission spectrometry (ICP-OES) was performed on a Spectro

Genesis spectrometer with a concentric nebulizer. Calibrated solution of La and Eu were purchased

from Inorganic Ventures. Solid samples were dissolved in HNO3 (70 %mas.) at 60 °C.

Transmission electron microscopy (TEM) images of NCs were collected on a JEOL JEM-1400

microscope operating at 120 keV. Samples were prepared by depositing a drop of NCs in chloroform

(3.10-2 g.L-1) on a carbon-coated copper grid. Selected area electron diffraction (SAED) was performed

with a camera distance of 20 cm. Interreticular planes were deduced from distances calibrated with a

known sample of silicon particles. Size distribution data were extracted from at least 100 particles on

pictures taken on at least 3 distinct regions of a TEM grid.

2

Page 23: Efficient Photoluminescence of Isotropic Rare-Earth

Optical characterization. Absorption spectra of the dispersions were measured in a 10 mm quartz

cuvette, on a Cary 5000 UV−vis−NIR spectrophotometer (Agilent Technologies) from 800 to 200 nm

with a 1 nm step. The same cuvette filled with solvent was used as a reference. Photoluminescence

(PL) experiments were performed on a FLS1000 fluorimeter (Edinburgh Instruments) equipped with a

Xe lamp. PL spectrum was acquired from 310 to 800 nm with 1 nm steps and 0.1 s per step. Slit

bandwidths on both incident and emission sides were 1.1 nm. PL excitation spectrum was acquired

from 270 to 610 nm, 1 nm steps and 0.1 s per step. Slit bandwidths on both incident and emission

sides were 2.0 nm. PL decay measurements were performed with a pulsed laser source (MCS) at

378 nm, 125 Hz pulse frequency. For absolute PL quantum yield measurement, an integrating sphere

was adapted on the fluorimeter. Emission spectra were acquired in the same condition for the

reference (cuvette with solvent) and sample: λexc = 300 nm (or 395 nm), 10 nm slit bandwidth on the

incident side, 1.0 nm slit bandwidth on the emission side, 270 (or 370) to 800 nm, 1 nm steps and

0.5 s per step.

ADDITIONAL DATA

Figure S1. Le Bail refinements of a) undoped and b) 5 % Eu3+-doped LaOCl nanocrystals. The peak at2θ = 44.5 ° is due to the aluminum plate. It is taken into account in the fit (Al, Fm-3m, ICSD .cif filecode 64700) in the refinements. In both cases, refined cell parameters of LaOCl (P4/nmm) are similarto values reported (COD .cif file 9009170): a = 4.119 Å and c = 6.883 Å. Refinements were performedon JANA 2006.2

3

Page 24: Efficient Photoluminescence of Isotropic Rare-Earth

Figure S2. SEM-EDS spectrum of a region of aggregated Eu3+-doped LaOCl NCs. Si detection isattributed to a contamination from sample preparation or SEM chamber. It is not considered forquantification, nor O and C.

Figure S3. a) TEM picture of 5 % Eu3+-doped LaOCl nanocrystals forming close-packed monolayers andbilayers. b) SAED picture with diffraction rings indexed along with LaOCl structure. Lower left inset is ascheme reproducing (full lines) major and *(dashed) minor rings. c) Size distribution diagram of 5 %Eu3+-doped LaOCl nanocrystals (black) and gaussian function centered at 8.2 nm with a 1.1 nmstandard deviation.

4

Page 25: Efficient Photoluminescence of Isotropic Rare-Earth

Figure S4. FTIR spectrum of Eu3+-doped LaOCl nanocrystals showing bond vibrations typical of OLA, as

opposed to OA.3,4

Figure S5. Structural characterization of GdOCl nanocrystals. The synthesis was adapted from LaOClexperiment, using hexahydrate gadolinium chloride as RE precursor, and without addition of OA atthe end of the heating step. a) XRD diagram showed that GdOCl (P4/nmm, No. 129) is the only phaseobtained. The various peak widths were due to strong shape anisotropy, in relation with crystallinestructure. Especially, the narrowest peaks corresponded to (hk0) interreticular planes (blue), i.e.orthogonal to c axis. b-d) TEM pictures showed stacks of thin plates with rough edges, forming 200-500 nm spherical aggregates. e) Peak position and apparent crystal size (from Scherrer formula) ofsome interreticular planes (hkl). εhk0 = 16.6±1.1 nm and ε00l = 4.2±0.4 nm confirmed that GdOCl NCswere anisotropic with a smaller dimension in c direction, whereas facets orthogonal to c were thebiggest. This is in agreement with the plate shape observed by TEM (b)).

Figure S6. Structural characterization of LaOBr nanocrystals. The synthesis was adapted from LaOClexperiment, using lanthanum bromide instead of the chloride, heating at 330 °C instead of 220 °C,and without addition of OA at the end of the heating step. a) XRD diagram showed that LaOBr(P4/nmm, No. 129, ICSD code 84336) was the only phase obtained. b-d) TEM pictures of nanoplates15-20 nm thick and 30-100 nm long. Particles displayed an elongated hexagon-like shape whenoriented along the edge. This shape was similar than LaOCl obtained by Depner et al..5 Rod-like arrays(single- and multi-linear) formed via face-to-face alignement of the nanoplates. e) Peak position andapparent crystal size (from Scherrer formula) of some interreticular planes (hkl). εhk0 = 20.6±0.1 nmε00l = 15.1±0.5 nm confirmed that LaOBrl NCs are anisotropic with a smaller dimension in c direction,

5

Page 26: Efficient Photoluminescence of Isotropic Rare-Earth

whereas the biggest facets were perpendicular to c axis. This was in agreement with the shapeobserved by TEM (b)). εhk0 was smaller than plates length observed by TEM. The size difference couldbe due to tha nanoplates being polycrystalline.

Figure S7. Tauc plot of dispersed LaOCl NCs.

For direct bandgap semiconductors such as LaOCl,6 optical bandgap measurement was based on theTauc relationship (2) between the absorption coefficient α, the photon energy hν and the bandgapenergy Eg:

where B was a constant for allowed transitions. For relatively low absorption A (A < 1), the absorptioncoefficient of a perfectly non-scattering dispersion followed the linear Beer-Lambert law:

with l the optical path length (1 cm).

Figure S8. Partial energy diagram of Eu3+ (4f6) taking into account interelectronic repulsion and spin-orbit coupling.7

6

Page 27: Efficient Photoluminescence of Isotropic Rare-Earth

Figure S9. PL spectrum of Eu3+-doped LaOCl NCs dispersed in chloroform, excitation at λexc = 395 nm.

REFERENCES1 T. Roisnel and J. Rodríquez-Carvajal, Mater. Sci. Forum, 2001, 378–381, 118–123.2 V. Petrícek, M. Dušek and L. Palatinus, Zeitschrift fur Krist., 2014, 229, 345–352.3 S. Mourdikoudis and L. M. Liz-Marzán, Chem. Mater., 2013, 25, 1465–1476.4 I. O. Perez De Berti, M. V Cagnoli, G. Pecchi, J. L. Alessandrini, S. J. Stewart, J. F. Bengoa and S.

G. Marchetti, Nanotechnology, 2013, 24, 175601.5 S. W. Depner, K. R. Kort, C. Jaye, D. A. Fischer and S. Banerjee, J. Phys. Chem. C, 2009, 113,

14126–14134.6 L. Lv, T. Wang, S. Li, Y. Su and X. Wang, CrystEngComm, 2016, 18, 907–916.7 K. Binnemans, Coord. Chem. Rev., 2015, 295, 1–45.

7