Hydrophobic Organic Skin as a Protective Shield for Moisture

Feb 8, 2017 - A moisture-stable, red-emitting fluoride phosphor with an organic hydrophobic skin is reported. A simple strategy was employed to form a...
0 downloads 0 Views 1MB Size
Subscriber access provided by UB + Fachbibliothek Chemie | (FU-Bibliothekssystem)

Article

Hydrophobic Organic Skin as a Protective Shield for Moisture-Sensitive Phosphor-Based Optoelectronic Devices Paulraj Arunkumar, Yoon Hwa Kim, Ha Jun Kim, Sanjith Unithrattil, and Won Bin Im ACS Appl. Mater. Interfaces, Just Accepted Manuscript • DOI: 10.1021/acsami.6b14012 • Publication Date (Web): 08 Feb 2017 Downloaded from http://pubs.acs.org on February 9, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Applied Materials & Interfaces is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Hydrophobic Organic Skin as a Protective Shield for Moisture-Sensitive Phosphor-Based Optoelectronic Devices Paulraj Arunkumar, Yoon Hwa Kim, Ha Jun Kim, Sanjith Unithrattil, and Won Bin Im* School of Materials Science and Engineering and Optoelectronics Convergence Research Center, Chonnam National University, 77 Yongbong-ro, Buk-gu, Gwangju, 61186, Republic of Korea.

Corresponding Author *E-mail: [email protected]

Keywords: K2SiF6:Mn4+, moisture-resistant, oleic acid, solvothermal, and red phosphor.

ACS Paragon Plus Environment

1

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 37

ABSTRACT

A moisture-stable, red-emitting fluoride phosphor with an organic hydrophobic skin is reported. A simple strategy was employed to form a metal-free, organic, passivating skin using oleic acid (OA) as a hydrophobic encapsulant via solvothermal treatment. Unlike other phosphor coatings that suffer from initial efficiency loss, the OA-passivated K2SiF6:Mn4+ (KSF-OA) phosphor exhibited the unique property of stable emission efficiency. Control of thickness and a highly transparent passivating layer helped to retain the emission efficiency of the material after encapsulation. A moisture-stable KSF-OA phosphor could be synthesized because of the exceptionally hydrophobic nature of OA and the formation of hydrogen bonds (F····H) resulting from the strong interactions between the fluorine in KSF and hydrogen in OA. The KSF-OA phosphor exhibited excellent moisture stability and maintained 85 % of its emission intensity even after 450 h at high temperature (85°C) and humidity (85 %). As a proof-of-concept, this strategy was used for another moisture-sensitive SrSi2O2N2:Eu2+ phosphor which showed enhanced moisture stability, retaining 85 % of emission intensity after 500 h under the same conditions. White light-emitting devices were fabricated using surface-passivated KSF and Y3Al5O12:Ce3+ which exhibited excellent color rendering index of 86, under blue LED excitation.

ACS Paragon Plus Environment

2

Page 3 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

INTRODUCTION Phosphor-converted white light-emitting diodes (pc-WLEDs) have revolutionized artificial lighting since the development of efficient InGaN-based blue LEDs by the 2014 Nobel laureates.1 Commercial pc-WLEDs based on the combination of blue LEDs and yellow-emitting Y3Al5O12:Ce3+ (YAG:Ce3+) phosphors are efficient, but have limitations because of their poor color rendering indices (CRI < 80) due to the lack of a red spectral region. Hence, the search for red phosphors has become crucial for attaining high CRI, color saturation, and reproducibility for WLEDs. Red-emitting phosphors based on a Eu3+-activator, inorganic quantum dots (CsPbI3), and nitride phosphors (CaAlSiN3:Eu2+) are among the most promising red components for pcWLEDs.2-4 Nevertheless, poor absorption in the blue region and chemical instability of phosphor hindered the commercialization of former two candidates, while the commercialized nitride phosphor suffer from low CRI due to deep red emission, where the spectral sensitivity of human eye is poor.3, 5 Recently, a new class of Mn4+-doped A2MF6 (A = Li, Na, and K; M = Si and Ti) fluoride phosphors has attracted increasing attention as red components due to their excellent optical properties, including narrow-band emission within the spectral sensitivity of human eye, high luminous efficiency, and excellent color rendering indices (80 - 86).6-8 Despite their exceptional luminescent properties, sensitivity to atmospheric moisture has limited their applications. After long-term use under LED operating conditions of high temperature (85oC) and humidity (85 %), these fluoride phosphors turn brown because of the hydrolysis of MnF62- ions to hydrated MnO2, resulting in significant deterioration of their brightness and luminous efficiencies.7 Fluoride phosphors that are resistant to moisture-induced degradation are highly desirable, as they can prolong the lifetimes of LEDs. Typically, moisture sensitive phosphors are encapsulated with

ACS Paragon Plus Environment

3

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 37

hydrophobic materials to extend their lifetimes under LED operating conditions. Hydrophobic materials with high optical transparency in the visible region and controllable layer thicknesses are good at protecting moisture-sensitive phosphors from loss of quantum efficiency. Inorganic surface coatings such as oxides and metal phosphates, as well as encapsulation with glass and use of activator-free phosphor hosts significantly enhance the moisture stability of phosphors.7, 911

Nguyen et al.,10 reported a moisture-resistant K2SiF6:Mn4+ (KSF) phosphor with an alkyl

phosphate layer that used transition metals as cross linkers. Setlur et al.,7 reported an improvement in the moisture stability of KSF phosphor through surface encapsulation with a Mn4+-free phosphor host. Nevertheless, these inorganic coatings suffer from initial loss of luminous efficiency at room temperature. Therefore, organic coatings may be considered as an alternative approach that provides precise control over coating thickness, without an initial loss of luminous efficiency. Organic polymers are reported to exhibit improved moisture resistance, but are less stable at high operating temperatures.10,

12

A thermally stable organic molecule

(temperature ≥ 200oC) may be a promising candidate for hydrophobic coating of phosphors. Oleic acid, a fatty acid with a formula of CH3(CH2)7-CH=CH(CH2)7COOH, is a high boiling point organic molecules (360oC), which were used as surface-active agents for stabilization of quantum dots, nanocrystals, and cationic precursors. The polar (-COOH) group of oleic acid (OA) acts as a coordinating ligand, while the long alkyl tail (C18) serve as a hydrophobic moiety.13-15 OA is used to control particle growth, prevent nuclei agglomeration, and even bind and passivate inorganic CdSe QDs surfaces to maintain high quantum yield by minimizing surface bound electron traps.16-17 OA’s long alkyl chain, excellent optical transparency, and refractive index of 1.45 facilitate its use as a hydrophobic passivating layer for inorganic

ACS Paragon Plus Environment

4

Page 5 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

materials such as Fe2O3, upconversion NaYF4:Yb2+, Er3+ phosphors in biological applications.1721

In this study, we demonstrate a facile method for fabricating moisture-resistant, red-emitting KSF phosphors using OA as a hydrophobic encapsulant via a solvothermal process. A homogenous, hydrophobic passivating skin was formed on the KSF without any decrease in the initial luminous efficiency of the phosphor. The polar head (–COOH group) of the OA binds with fluorine from KSF through hydrogen bonding (F····H) in the core. OA’s hydrophobic tail forms a shell with a thin passivating layer that shields the phosphor from atmospheric moisture, and prevents the hydrolysis of MnF62-. The OA-passivated KSF phosphor exhibited excellent moisture resistance and may be a promising red component for color-stable pc-WLED lighting.

EXPERIMENTAL SECTION Sample preparation. Commercial KSF phosphor (denoted as KSF) without post treatment (coating) was purchased from Force4 Corporation, Gwangju, South Korea. The reference commercial KSF phosphor was purchased from Denka Company, Japan and used to evaluate the WLED device performance of KSF-OA phosphor, under high temperature (85oC) and humidity (85 %). The purchased Denka sample has a protective coating on the KSF phosphor which are sold for lighting applications. Oleic acid (Fluka, 4.3 ml, 13.5 mmol) was dissolved in 25 ml of anhydrous absolute ethanol (Daejung Chemicals, South Korea) and 1 g (4.5 mmol) of KSF powder was dispersed in the above solution via ultrasonication for 1 h. This mixture was loaded into a 100 ml hydrothermal reactor and heated at 140oC for 6 h. After the reaction, the mixture was allowed to cool and centrifuged at 7000 rpm for 10 min. The product was washed repeatedly with ethanol to remove excess OA and dried at 70oC for 4 h to produce OA-passivated KSF

ACS Paragon Plus Environment

5

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 37

(KSF-OA) powder. A control reaction was performed to understand the effect of solvothermal (ST) activity on the emission property of KSF powders, using ethanol as solvent medium. This reaction was carried without OA and the resulting powders were denoted as solvothermallytreated KSF (KSF-ST). Non-polar solvents namely hexane (Sigma-Aldrich) and octadecene (Aldrich) were also used as solvent medium to monitor the formation of OA passivation layer on the KSF phosphor. Material characterization. The phase identification of KSF, KSF-OA, and KSF-ST were carried out via powder X-ray diffraction and synchrotron diffraction. X-ray diffraction results was collected using a Philips X’Pert diffractometer with Cu Kα radiation, in the range of 10° ≤ 2θ ≤ 100°, with a step size of 0.02°, at room temperature. Synchrotron diffraction was collected using radiation source with wavelength of 1.5183 Å in the range of 10° ≤ 2θ ≤ 150°, with a step size of 0.02°, at Pohang Accelerator Laboratory (PAL), South Korea. The Rietveld refinement was performed using the General Structure Analysis System (GSAS) program, with the split Pseudo-Voigt profile function.22 The particle size and morphology were analyzed via field emission scanning electron microscopy (FE-SEM) and high-resolution transmission electron microscopy (HR-TEM). SEM images and elemental mapping were obtained using a Hitachi S4700, and TEM images were recorded using a FEI Tecnai F20 from the Korea Basic Science Institute (KBSI), Gwangju. Surface analyses were performed via X-ray photoelectron spectroscopy (XPS), using a VG Multilab 2000. Fourier transform infrared (FT-IR) spectra were recorded using a spectrometer (PerkinElmer Spectrum 400). The 1H magic angle spinning NMR spectra was measured using 400 MHz Avance III HD Bruker spectrometer (Korea Basic Science Institute; KBSI, Western Seoul center, South Korea) at Larmor frequency of 400.25 MHz. The samples were spun at 15 KHz, with a time interval of 10 s between successive scans and number

ACS Paragon Plus Environment

6

Page 7 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

of scans are fixed at 64. The

19

F NMR was measured using 400 MHz Avance III HD Bruker

spectrometer (Cooperative Center for Research Facility; CCRF, Gwangju, South Korea) at Larmor frequency of 376.55 MHz. Optical characterization. Photoluminescence spectra were measured using a Hitachi F-4500 spectrophotometer. Temperature-dependent fluorescent spectra were measured in the temperature range of 25 – 200oC using an integrated heater, temperature controller, and thermal sensor coupled to the Hitachi F-4500 fluorescent spectrometer. The internal quantum efficiency was measured via 450 nm excitation with a xenon laser (Hamamatsu C9920-02). The PL lifetimes of the phosphors were measured using a time-resolved photoluminescence system composed of a flash Xenon lamp and a Hamamatsu C10627 streak camera. The lifetimes were evaluated under pulsed laser irradiation at 450 nm. Diffuse reflectance measurements were carried out using a Hitachi U-4100-Vis/NIR spectrometer. WLED packaging and moisture tests. Commercial YAG:Ce3+ phosphors (Force4 Corporation, Gwangju, South Korea), KSF phosphors (KSF, KSF-OA, and KSF-ST) and blue LED chips (λmax = 450 nm) were used to fabricate WLEDs. Phosphors were mixed with silicone resin (Dow corning OE-6630 A and B) and coated over the surfaces of the LED chips. Moisture resistance tests were performed using LEDs fabricated with a KSF-OA phosphor, and they were subjected to high temperature (85oC) and humidity (85 %) conditions during measurement of the emission intensity. SrSi2O2N2:Eu2+ phosphor (Force4 Corporation, Gwangju, South Korea) was also used to demonstrate fabrication of moisture-resistant phosphors via encapsulation with a hydrophobic OA layer.

RESULTS AND DISCUSSION

ACS Paragon Plus Environment

7

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 37

Structural characterization of passivated KSF phosphor. The X-ray diffraction patterns of uncoated KSF (KSF), KSF-OA, and solvothermally-treated KSF (KSF-ST) phosphors were indexed to a cubic structure using the ICDD pattern 98-002-9407 (Supporting information, Figure S1). No impurities were observed in any of these samples. In addition, Rietveld refinement of X-ray diffraction for pristine KSF and KSF-OA samples was performed using Fm3m cubic structure, and is presented in Figure 1a and 1b. A lattice constant of 8.1519 Å is obtained for coated KSF-OA. As shown in Table 1, this is slightly larger than for the pristine KSF samples. The K+ ions in the KSF structure are 12-fold coordinated to the fluorine ions with a K-F distance of 2.9000 Å. Si4+ is surrounded by six fluorine ions at a distance of 1.6890 Å. These results are similar to the previously reported values.23-24 The Mn4+ activator occupies the Si4+ site in KSF, which is responsible for efficient red emission from the activator. SEM and TEM images of different KSF samples are shown in Figure 1c-1g. SEM images of KSF (Figure 1c and 1d) and KSF-OA (Figure 1e and 1f) show particle sizes of 10-20 µm with highly crystallized edges. The KSF particles are sensitive to the high-energy electron beam used for TEM measurement. This is indicated by their instability to high-energy electron beam, as reported in the literature.10 Hence, TEM images were measured using short electron beam exposure periods. The surface passivating organic layers of the KSF-OA samples appear to be about 10 - 20 nm thick. The sensitivity of organic layers to the electron beam hinders the precise determination of the thickness of surface passivating layer. However, TEM image of SrSi2O2N2:Eu2+-OA phosphor revealed the exact thickness of OA layer of ~37 nm (Supporting information, Figure S2). Elemental mapping of KSF-OA exhibits a homogeneous distribution of K, Si, and F elements on the phosphor surface with the stoichiometric atomic composition of

ACS Paragon Plus Environment

8

Page 9 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

2:1:6, respectively (Supporting information, Figure S3). The Mn4+ concentration in the KSF phosphor is ~0.25 %.

Optical studies of passivated KSF phosphor. Photoluminescence and excitation spectra of the pristine KSF, KSF-OA, and KSF-ST samples were measured and presented in Figure 2a. A narrow red emission in the 550 – 650 nm range was observed under λex = 450 nm due to spinforbidden 2Eg → 4A2g transitions of the Mn4+ activator. Multiple emission peaks at ~596, 612, 630, and 647 nm correspond to the different vibronic modes of the MnF62-octahedrons ν3 (t1u), ν6 (t2u), ν6 (t2u), and ν3 (t1u), respectively.6 The dominant emission peaks at 612, 630, and 647 nm are typical emission bands observed with Mn4+-activated phosphors.25 The excitation spectra under λem = 630 nm exhibit two broad bands at 360 and 460 nm, which are attributed to the spinallowed transitions of 4A2g→ 4T2g and 4A2g→ 4T1g, respectively. The emission and excitation peak positions of KSF-OA were the same as those of pristine KSF, revealing the high optical transparency of the thin OA layer in the visible region. Likewise, KSF-ST samples show the same emission and excitation spectral profiles as the pristine KSF sample. Furthermore, UVVisible spectra reveal no significant change in the absorption of KSF-OA and KSF-ST, compared to the pristine sample (Figure 2b), which is in agreement with the excitation spectra (Figure 2a). The optical transparency of the OA molecule was confirmed from UV-Vis spectra. The absorption band of OA is limited to the near-UV region (below 350 nm), thus, confirming its high optical transparency in the visible region range of 400 – 700 nm (Supporting information, Figure S4). OA does not alter the optical properties of the KSF phosphor, except by acting as a hydrophobic passivating layer.

ACS Paragon Plus Environment

9

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 37

The PL spectra reveal that the emission intensities of KSF-OA and KSF-ST are the same as that of pristine KSF under λex = 450 nm. Internal quantum efficiencies (IQEs) were measured to confirm that the emission intensity of encapsulated KSF-OA is unchanged. The KSF-OA phosphor exhibits an IQE value of 68.1, compared to 67.9 and 67.3 for pristine KSF and KSFST, respectively (Table 2). These results suggest that formation of an OA-based surface passivating layer and solvothermal treatment do not change or reduce the phosphors’ emission intensities. This unchanged luminous efficiencies of surface passivated phosphors is a unique property for the OA-based hydrophobic layer. It stands in contrast to the results reported with inorganic metal oxide coatings and alkyl phosphates with metal linkers, both of which show significant losses of luminous efficiency after coating.10 The luminous efficiency was sustained via the use of a thin organic passivating layer made from a hydrophobic material chosen for its optical transparency, good thickness control, and ease of processing. The OA encapsulation may also suppress the surface trap states which could slightly improve the IQE of KSF-OA phosphor. Generally, OA acting as a ligands effectively passivates the surface trap states of quantum dots and improves their QE by suppressing the trap-assisted non-radiative recombination.26 The PL decay characteristics of different KSF samples were monitored at 630 nm under λex = 450 nm, and are shown in Figure 2c. The decay time was determined using a single exponential decay mode, and the lifetime is estimated to be 9.1, 9.0, and 8.9 ms for KSF, KSF-OA, and KSFST, respectively. These results indicate that the OA-based hydrophobic layer doesn’t significantly alter the lifetime or PL properties of the KSF phosphor. This is an indirect verification of the stability of the luminescence efficiency after coating with a hydrophobic OA layer. Generally, a change in the luminescent lifetime is strongly correlated with a change in a phosphor’s luminescence intensity.10,

27

Some instances of increasing lifetimes have been

ACS Paragon Plus Environment

10

Page 11 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

reported, which result in an increase in the emission intensity due to the suppression of surface traps which cause non-radiative transitions.28 Therefore, these results further confirm that the OA surface passivation process does not reduce the emission intensity, unlike other systems that suffer from emission loss after surface coating. The temperature-dependent emission spectra of various KSF phosphors under λex = 450 nm in the temperature range of 25 – 200oC are shown in Figure 2d and 2e. The temperature dependence of the integrated area (550 – 700 nm) and emission peak intensities (630 nm) suggest a considerable improvement in the thermal stability of the KSF-OA phosphor. The relative emission intensity of the KSF-OA phosphor remains at 80 %, similar to pristine KSF at 200oC. The OA-passivated phosphor clearly exhibits reasonably good thermal stability than pristine KSF due to high boiling point of OA (360oC). Nevertheless, the thermal stability of KSF-OA phosphor is limited to 200oC, which is attributed to the decomposition reaction and thermal instability of organic OA layer at temperatures above 200oC (Supporting information, Figure S5). Surface analysis of the KSF-OA phosphor. FT-IR spectra were used to confirm the presence of OA (CH3(CH2)7-CH=CH(CH2)7COOH) on the surface of KSF particles, and are presented in Figure 3a. The FT-IR spectra of Pristine KSF, KSF-OA, and KSF-ST samples were measured in the solid form, while other OA samples, particularly pure OA, OA dissolved in ethanol (OA), and solvothermal-treated OA dissolved in ethanol (OA-ST) were measured in their liquid forms. All of the KSF samples exhibited typical bands at 714 and 483 cm-1, corresponding to the vibrations of Si-F bonds.29 The OA vibration modes in the KSF-OA sample were identified using the reference OA samples. The FT-IR spectra of pure OA, OA, and OA-ST were measured to monitor the influence of solvothermal treatment on OA, since the liquid changed from colorless

ACS Paragon Plus Environment

11

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 37

to yellow during solvothermal reaction (Supporting information, Figure S6). However, there are no significant differences between the FT-IR spectra of different OA samples, thus, indicating that the chemical nature of the OA is retained even after solvothermal treatment. The only spectral difference noted is the presence of a wider -OH vibration band from carboxyl groups at 3500 cm-1 in the OA and OA-ST samples and also due to the association of –OH groups of ethanol and OA.30 This strongly refutes the possibility of side or decomposition reactions during solvothermal process. Moreover, the absence of sharp band at 3590 cm-1 in OA and OA-ST sample, corresponding to free -OH of ethanol suggests possible formation of intermolecular hydrogen bonding with OA through –OH of ethanol.31 The FT-IR spectrum of the KSF-OA sample shows prominent bands at 2927, 2857, and 1710 cm-1, corresponding to anti-symmetric and symmetric stretching of the C-H bonds of long alkyl chains, and the C=O bond from the carboxyl group in OA, respectively confirming the presence of OA on the phosphor surface. This is similar to the spectrum of the OA-ST sample.32 This indicates that the surfaces of the KSF particles are encapsulated by the OA ligands. The interaction between OA and the KSF phosphor may be presumed to occur through two possible mechanisms. Two of the mechanisms involve chemisorption: metal-oleate and hydrogen bond formation. Metal-oleate complex could be formed through the potassium in KSF and –-COOgroup in OA resulting in the -COOK bond.31, 33 Hydrogen bonds may be formed between F-, with its strong partial negative charge and the proton from the -COOH group in OA. Hydrogen bonding is a relatively stronger interaction between atoms or molecules. The nature of interaction between the KSF and OA can be identified from the vibration band of FT-IR spectra. The characteristic vibration bands of metal-oleates are in the range of 1650 – 1510 cm-1 (precisely at 1610 and 1520 cm-1) for the bidentate coordination.31 In the KSF-OA sample, the absence of

ACS Paragon Plus Environment

12

Page 13 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

peaks in the 1650 – 1510 cm-1 and 1410 cm-1 corresponding to the -COO- stretching, completely precludes the metal-oleate formation (–COOK). In addition, the major band at 1710 cm-1, can be assigned to the carbonyl group (C=O) of OA on the surface of KSF particles.31 Moreover, KSF phosphor having a highly electronegative fluorine as anion on its crystal structure, fluorine would be the most favorable site for interaction with reactive ligands. These suggests that the hydrogen bond formation between the fluorine in KSF and hydrogen in OA might be the most favorable interaction. Therefore, the interaction between OA encapsulant and KSF phosphor may be via strong hydrogen bond formation. To further examine the chemical nature and surface composition of the KSF-OA phosphor, the XPS spectra of the C1s, K2p, Si2p, and F1s core levels were measured. The C1s spectrum is shown in Figure 3b. XPS is a surface-sensitive technique, and can primarily probe the outermost 5-10 nm of a sample.34 The strong C1s peak from the KSF-OA sample reveals the presence of a carbon-containing hydrophobic OA shell at 284.6 eV, which corresponds to aliphatic carbon chains (C-C).33 The absence of a C1s peak from the carboxylate group (-COOH) may be ascribed to complete encapsulation of the group in the core layer, alongside KSF. Only the shell, which is composed of long, aliphatic carbon chains from OA, is present on the surface. Moreover, a shift in the F1s peak to a lower binding energy (0.9 eV) in the KSF-OA sample suggests that fluorine from the KSF phosphor is linked to the electropositive hydrogen in the OA carboxyl group (Supporting information, Figure S7). Thus, fluorine-terminated KSF may be linked to the electropositive hydrogen, through hydrogen bonding. These results further confirm the absence of metal-oleate, for which a shift in K2p peak to a lower binding energy is expected. Therefore, it could be confirmed that OA forms a hydrophobic skin over the KSF phosphor through the

ACS Paragon Plus Environment

13

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 37

formation of hydrogen bond in the core. The long alkyl chain from OA forms a hydrophobic sheath on the phosphor surface, resulting in the moisture-resistant KSF phosphor. The formation of hydrogen bond between KSF and OA encapsulant in KSF-OA was further substantiated through 1H MAS and 19F NMR spectroscopy (Supporting information, Figure S8). A feature of 1H MAS NMR resonance of KSF-OA, shifting downfield (7.7 ppm) compared to pristine KSF (7.5 ppm) exposes the deshielding effect of electronegative fluorine of KSF coordinated to the hydrogen nuclei of pendant –COOH group of OA, suggesting the formation of hydrogen bond (-OH····F). Similarly, the 19F NMR resonance of KSF-OA are shifted upfield to 130.15 ppm compared to pristine KSF (-130.25 ppm), describing the shielding effect of hydrogen of pendant -COOH of OA coordinated to the fluorine nuclei of KSF. These results further confirms the formation of intermolecular hydrogen bonding between the fluorine of KSF and hydrogen of (-COOH) of OA layer. Li et al., has reported a similar hydrogen bond formation between halide ion in the hybrid perovksite material and hydrogen in the phosphonic acid crosslinker, for solar cell applications.35 An improved stability of perovskite solar cells was attained due to the strong hydrogen bonding and effective passivation of perovskite surface, rendering immunity to moisture which proved to be beneficial for the long-term stability of solar cells. This further corroborates the possibility of hydrogen bond formation in the current work, between hydrogen of carboxylic acid in OA and fluorine in KSF phosphor rendering excellent moisture stability to the KSF-OA-based optoelectronic devices. The surface elemental composition was also calculated by integrating the XPS peaks to estimate the content of OA encapsulant on the phosphor surface, and the result is given in Table 3. The atomic compositions of the core elements namely K, Si, and F are in stoichiometry, within the KSF samples. In addition, a significant proportion of carbon (~14 – 16 %) and oxygen (~3

ACS Paragon Plus Environment

14

Page 15 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

%) were observed even in the pristine KSF and KSF-ST, which were attributed to the absorbed CO2from the atmosphere, as reported earlier.36-37 A sharp decrease in the concentration of core elements were apparent in the KSF-OA, attributed to the presence of OA sheath on the phosphor surface. The atomic% of core elements in KSF-OA reduced upto 30 %, 47 %, and 40 %, for K, Si, and F, respectively. The total sum of carbon and oxygen content in the KSF-OA also increased to 50 % from 20 % of the pristine KSF. These results strongly evidences the presence of sheath of OA encapsulant (~30 %) on the phosphor surface. The robustness and water tolerance of KSF-OA phosphor was investigated under harsh water condition at room temperature and presented in the Figure 4a and 4b. An extremely harsh condition was employed in which phosphor was dispersed in water at a phosphor concentration of 2.50 × 105 ppm, where phosphor comes in direct contact with water as shown in Figure 4a. The pristine KSF particles form a brown solution in water within 5 min. In contrast, the KSF-OA sample remains unchanged after 30 min, suggesting that OA-encapsulated KSF has enhanced moisture stability. After 4 h, the orange color of the pristine KSF becomes completely brown due to hydrolysis of MnF62- to MnO2, while only 25 % of the KSF-OA sample becomes brown. PL after 4h in deionized water was measured to further monitor the water stability of KSF-OA sample in comparison with pristine KSF (Supporting information, Figure S9). The pristine KSF exhibited only 4 % of the emission intensity with respect to KSF-OA sample after 4 h in deionized water. This shows that the KSF-OA sample is highly water-tolerant, due to the presence of a hydrophobic, passivating OA sheath over the KSF phosphor. Therefore, OA surface passivation could be a promising approach for obtaining hydrophobic coating of phosphors.

ACS Paragon Plus Environment

15

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 37

Water tolerance test for pristine KSF and KSF-OA samples were conducted by measuring PL in water after 15 days, using a phosphor concentration of 6.25 × 103 ppm as shown in Figure 4b. After 15 days, the red emission of the KSF-OA sample still remained. The retention of orange color of the encapsulated phosphor reveals its excellent moisture resistance. In contrast, no emission is observed for pristine KSF due to the complete hydrolysis of Mn4+ to hydrated MnO2, which precipitates as a brown powder. These results suggest that the KSF-OA phosphor with a hydrophobic, passivating OA skin exhibits good stability in water. Images of KSF and KSF-OA samples under visible and blue light excitation (λex = 450 nm) are shown in Figure 4c. The orange KSF samples exhibit a strong red emission under blue light excitation. The ionic conductivity were measured for the pristine KSF and KSF-OA in deionized water to monitor the change in conductivity with respect to time (Supporting information, Table S1). A large increase in the ionic conductivity (28 % increase, 40 mS/m) was detected for uncoated KSF sample, compared to the negligible increase in the ionic conductivity value (~1 % increase, 2 mS/m) for KSF-OA. These results strongly suggest the effective passivation of KSF phosphor by the hydrophobic OA layer from the surrounding moisture. A schematic illustration of the nature of the interaction between the core KSF phosphor and the OA passivating shell is presented in Figure 4d. Oleic acid is an excellent organic hydrophobic

material

with

a

high

boiling

point

and

the

formula

CH3(CH2)7-

CH=CH(CH2)7COOH, where COOH is the polar head group. The remaining long alkyl chain behaves as a strongly hydrophobic tail. The polar KSF molecule, which contains fluorineterminated moieties, interacts with the polar head group (-COOH) of OA through the formation of F····H via hydrogen bonding. This has been confirmed via FT-IR and XPS (Figure 4a and 4b). The core undergoes polar interactions between fluorine and hydrogen, forming F····H bonds.

ACS Paragon Plus Environment

16

Page 17 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

The shell, which is composed of long alkyl chains, provides strong resistance to atmospheric moisture. In addition, the bending configuration of the alkyl chain of the cis-form of OA ensures complete encapsulation of KSF particles with a hydrophobic, passivating skin. These results have implications for strong interactions between the KSF phosphor and the organic passivating OA skin that enhances the moisture stability of the phosphors. The OA encapsulation of KSF phosphor were also carried out by simple mixing of OA and KSF without solvent, and in the presence of ethanol as solvent. The resulting KSF powder exhibit moisture instability owing to improper adherence of OA on the phosphor surface. It is to be noted that a sticky powder was obtained due to the presence of large amount of OA on the phosphor surface, making it unsuitable for commercial applications. Hence, extensive ethanol washing were employed to remove excess OA, for obtaining a free flowing KSF powder. In addition, OA encapsulation was carried out under refluxing condition, also exhibited poor moisture stability caused by non-uniform encapsulation. An homogeneous OA encapsulation layer was achieved only through solvothermal method in ethanol medium. This suggests that solvothermal treatment is necessary for the formation of homogeneous, thin hydrophobic OA layer which could improve the moisture stability of KSF phosphor. Moreover, polar solvent especially ethanol was identified as an ideal solvent medium for initiating an interaction between the polar group of OA (–COOH) and KSF (F- anion). The use of non-polar solvents such as hexane and octadecene in solvothermal condition, suffered serious loss in the initial emission intensity due to thick formation of OA layer (not shown). Therefore solvothermal processing using ethanol medium is mandatory for the fabrication of moisture-resistant KSF phosphor with a thin hydrophobic, passivating OA layer.

ACS Paragon Plus Environment

17

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 37

WLED fabrication. The electroluminescence (EL) spectra of WLEDs made using redemitting KSF-OA and yellow-emitting YAG:Ce3+ under blue LED light (λmax = 450 nm) are presented in Figure 5a. The WLEDs cover the entire visible spectrum, including a white emission which lies within the sensitivity of the human eye (