Download as pdf or txt
Download as pdf or txt
You are on page 1of 184

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/248475071

Recent Progress in Quantum Cutting Phosphors

Article in Progress in Materials Science · July 2010


DOI: 10.1016/j.pmatsci.2009.10.001

CITATIONS READS

294 583

2 authors, including:

Xiaoyong Huang
Taiyuan University of Technology
32 PUBLICATIONS 1,296 CITATIONS

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

phosphor View project

Novel rare-earth-free luminescent materials for lightings and displays View project

All content following this page was uploaded by Xiaoyong Huang on 29 September 2015.

The user has requested enhancement of the downloaded file.


Accepted Manuscript

Recent progress in quantum cutting phosphors

Q.Y. Zhang, X.Y. Huang

PII: S0079-6425(09)00093-0
DOI: 10.1016/j.pmatsci.2009.10.001
Reference: JPMS 242

To appear in: Progress in Materials Science

Received Date: 25 October 2008


Revised Date: 12 August 2009
Accepted Date: 22 September 2009

Please cite this article as: Zhang, Q.Y., Huang, X.Y., Recent progress in quantum cutting phosphors, Progress in
Materials Science (2009), doi: 10.1016/j.pmatsci.2009.10.001

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

Recent progress in quantum cutting phosphors

Q. Y. Zhanga) and X. Y. Huang

Key Laboratory of Specially Functional Materials of Ministry of Education and Institute of

Optical Communication Materials, South China University of Technology, Guangzhou

510641, People’ Republic of China

Abstract

Luminescent materials with the quantum efficiency (QE) higher than unity could

be playing a significant role in the progress of lighting industry and certain electronic display

systems. The recent demonstration of an efficient visible quantum cutting (QC) in vacuum

ultraviolet (VUV)-excited LiGdF4:Eu phosphors (Wegh RT, et al. Science 283(1999)663) has

provided an exciting and interesting trends in the development of several potentially

important luminescent materials and devices. The possibility of the higher QE depends on the

principle of QC in phosphors which could generate two or more low-energy photons for every

incident high-energy photon that is being absorbed by phosphors. Investigation on QC

systems has started on single ions doped-fluorides capable of a cascade emission from ions

such as Pr3+, Tm3+, Er3+ and Gd3+. The focus has now been shifted to the combination of two

ions, where the energy of the donor ion could be transferred stepwise to two acceptor ions via

a downconversion. A well-known example is the Gd3+-Eu3+ dual ions. QC via

downconversion has now been widely witnessed in many rare-earths (RE)-based phosphors,

the interesting and appreciable QE in the visible spectral region has earlier been reported from

a)
Author to whom correspondence should be addressed; Tel:+86-20-87113681; Fax:
+86-20-87114204; Electronic mail: qyzhang@scut.edu.cn (Q. Y. Zhang)

1
ACCEPTED MANUSCRIPT

LiGdF4:Eu (190%) and BaF2:Gd,Eu (194%) phosphors. QC materials could also be used in

solar cells, if conversion of one UV-visible photon into two near-infrared (NIR) photons is

realized, and energy loss due to thermalization of electron-hole pairs is minimized. The

present article reviews on the recent progress made on: (a) materials and developments in the

fields of UV-visible QC phosphors and the mechanism involved, including QC in single RE

ion activated fluorides- and oxides-based phosphors, energy transfer and downconversion, QC

in dual/ternary ions activated phosphors; and (b) NIR QC in RE3+-Yb3+ (RE=Tb, Tm, and Pr)

dual ions doped phosphors via cooperative energy transfer. Appropriate discussions have been

made on materials, materials synthesis and characterization, the structural and luminescence

properties of various QC luminescent materials via different synthesis techniques. In addition,

applications, challenge and future advances of the visible- and NIR- QC phosphors have also

been dealt with.

Keywords: Phosphors; Luminescent materials; Rare-earths; Fluorides; Oxides; Solid-state

reaction; Hydrothermal synthesis; Combustion synthesis; Wet-chemical synthesis; Glasses;

Glass-ceramics; Photon cascade emission; Quantum cutting; Quantum splitting;

Downconversion; Upconversion; Decay lifetime; Energy transfer; Cooperative energy

transfer.

2
ACCEPTED MANUSCRIPT

Contents

1. Introduction 7

1.1. Scope and arrangement of this review 7

1.2. Overview 8

1.3. Challenge and objectives 15

2. Fundamental aspects of quantum cutting phosphors 16

2.1. Basic concepts 16

2.1.1. Basics of energy transfer between rare-earth ions 16

2.1.2. Quantum cutting 19

2.2. Experimental description: materials synthesis and fabrication 20

2.2.1. Solid-state reaction 21

2.2.2. Hydrothermal synthesis 22

2.2.3. Solution-based chemical synthesis 23

2.2.4. Luminescent glass and glass-ceramics by melting-quenching method 25

2.3. Materials characterization and properties 26

2.3.1. Structural properties and morphology 26

2.3.2. Thermal analysis 29

2.3.3. Spectral analysis: Raman and IR 29

2.3.4. Photoluminescence 30

3. Single ion activated phosphors 32

3.1. Quantum cutting in Pr3+ ion activated phosphors 32

3.1.1. Principle 32

3.1.2. Pr3+ -doped quantum cutting materials: fluorides and oxides 41

3.1.3. Visible quantum efficiency 44

3.2. Tm3+ ion activated phosphors 47

3.3. Er3+ ion activated phosphors 48

3
ACCEPTED MANUSCRIPT

3.4. Gd3+ ion activated phosphors 49

4. QC in dual/ternary ions activated phosphors 51

4.1. Energy transfer and downconversion 51

4.2. Gd3+-involved phosphors 51

4.2.1. Gd3+–Eu3+ dual ions activated phosphors 52

4.2.1.1. Materials, photoluminescence and quantum efficiency 52

4.2.1.2. Sensitization 57

4.2.2. Gd3+–Tb3+ dual ions activated phosphors 60

4.2.3. Gd3+–Er3+–Tb3+ ternary ions activated phosphors 61

4.3. Pr3+- involved phosphors 62

4.3.1. Pr3+–Eu3+ dual ions activated phosphors 63

4.3.2. Pr3+–Er3+ dual ions activated phosphors 64

4.3.3. Pr3+–Mn2+ dual ions activated phosphors 65

4.3.4. Pr3+–Cr3+ dual ions activated phosphors 68

4.4. Downconversion in Tb3+-activated phosphors 71

5. Near-infrared quantum cutting phosphors 74

5.1. Basic concepts 74

5.2. Materials, photoluminescence and quantum efficiency 77

6. Perspectives and future advances 86

7. Concluding remarks 88

8. Acknowledgements 90

9. References 91

Table captions and tables 133

Figure captions and figures 138

4
ACCEPTED MANUSCRIPT

Nomenclature
α Constant
β0 Scan aperture of the diffractometer
βel Constant
ћ Planck constant
ћωmax The maximum phonon energy
θ Diffraction angle
ηET Energy transfer efficiency
Ω2, Ω4, Ω6 Judd-Ofelt intensity parameters
τ The intrinsic lifetime of the luminescent ions
The gamma function
Γ( y )
2 Judd-Ofelt reduced matrix elements square

ω Frequency
λ Wavelength
λemi Emission wavelength
λexc Excitation wavelength
AFM Atom force microscope
A.R. Analytical reagent
CA The acceptor concentration
CR Cross relaxation
CRET Cross relaxation energy transfer
CRT Cathode ray tube
CT Charge transfer
CTB Charge transfer band
CTS Charge transfer state
DC Downconversion
DESY Deutsches Elektronen-Synchrotron
DSC Differential scanning calorimetry
DTA Differential thermal analysis
Eg Band-gap energy
EDS Energy-dispersive spectrometer
ET Energy transfer
ETE Energy transfer efficiency
ΔE Energy difference between different energy levels
e-h Electron-hole
eV Electronvolt
FT-IR Fourier-transform infrared spectroscopy
IR Infrared
J-O Judd-Ofelt
MFFL Mercury-free fluorescent lamp
NIR Near-infrared

5
ACCEPTED MANUSCRIPT

PCR Probability of cross relaxation


PDT Probability of direct energy transfer
PET+ Probability of the desired energy transfer
PET- Probability of the undesired energy transfer
PCE Photon cascade emission
PDP Plasma display panel
PL Photoluminescence
PLE Photoluminescence excitation
PV Photovoltaic
QC Quantum cutting
QE Quantum efficiency
QS Quantum splitting
R Radius
R1, R2, R3 Rare-earth ions
R(5D0/5D1,2,3) The ratio of the 5D0 and the 5D1,2,3 emission intensities
R(5D4 /rest) The ratio of emission intensity of 5D4 to that attributed to 5D3 of
Tb3+ and 6P7/2 of Gd3+
The intensity ratios of Er3+ 4S3/2 emission to all remaining emission
R ( 4 S3/ 2 / rest ) Er 3+
upon Er3+ ƒ→ d excitation
The intensity ratios of Er3+ 4S3/2 emission to all remaining emission
R ( 4 S3/ 2 / rest )Gd 3+
upon Gd3+ 6IJ excitation
RE Rare earth
RT Room temperature
rCharT Charge transfer rate
rCoopT Cooperative transfer rate
rD Energy diffusion rate between identical ions
rex Pump rate of Tb3+ 5D4 level
rNR Non-radiative rate
rR Spontaneous emission rate
SEM Scanning electron microscopy
STE Self-trapped exciton
TEM Transmission electron microscopy
TGA Thermogravimetric analysis
UC Upconversion
UV Ultraviolet
VUV Vacuum ultraviolet
WNR Nonradiative transition rate
XRD X-ray diffraction

6
ACCEPTED MANUSCRIPT

1. Introduction

1.1. Scope and arrangement of this review

The contribution will start with Section 1 that presents a brief overview on the

developments of quantum cutting (QC) phosphors: materials and properties. The visible- and

near-infrared (NIR)-QC phenomena in phosphors are highlighted. Understanding the nature

and factors dominating the general trends and limitations of QC in phosphors is of

fundamental importance for advancing technological applications. Sections 2-5 will describe

the basic concepts and mechanism of QC in phosphors, experimental observation and

investigation including materials synthesis and characterizations, the structural and

luminescence properties, energy transfer (ET) and downconversion (DC), applications,

question and challenge of the visible- and NIR-QC phosphors are comparatively analyzed.

Emphasis will be given to:

(i) Basic concept of QC, materials synthesis and characterization.

(ii) QC of Pr3+ individual ion activated fluorides- and oxides-based phosphors:

materials, materials synthesis and characterization, quantum efficiency (QE)

and mechanism.

(iii) ET and DC.

(iv) Dual/ternary rare-earth (RE) ions activated QC phosphors, i.e. Gd3+–Eu3+,

Gd3+–Tb3+, Pr3+–Mn2+ dual ions activated phosphors: materials synthesis and

characterizations, luminescence properties and QE.

and,

7
ACCEPTED MANUSCRIPT

(v) NIR-QC phosphors: basic concepts, materials, characterization, and

applications.

In Sections 6 and 7, perspectives and a summary of the main conclusions will be given in

responding to the challenges addressed in Section 1. The report will end with

recommendations for future directions in extending the developed knowledge and the

associated approaches.

1.2. Overview

Over the past several years, phosphors have been considered as key and

technologically important components as the prerequisites to the functionality and success of

many lighting and display systems [1,2]. At present, RE-based phosphors with efficiencies

close to the theoretical maximum (100%) are employed in different fluorescent tubes, X-ray

imaging and color televisions [3,4]. Such applications depend on the luminescent properties

of RE ions, eg. sharp lines, high efficiency and high lumen equivalent. However, a good

phosphor for electronic or UV excitation is not necessarily a good choice for excitation in

vacuum ultraviolet (VUV) (E> 50 000 cm–1, λ < 200 nm) [5]. High-performance plasma

display panels (PDP) and mercury-free fluorescent lamps (MFFL), where the Xe/Ne

discharge is used to generate VUV photons from 147 nm to 190 nm, rely critically on

phosphors that should have great luminescence efficiency in the VUV region. Still there are

efficiency constraints with regard to QE in relation to those phosphors used in fluorescent

lamps or cathode-ray tubes (CRT) due to too much energy is lost when converting one VUV

photon into one visible photon (λ= 400 to 700 nm) [6]. For instance, the screen efficiency of

8
ACCEPTED MANUSCRIPT

current PDP phosphors is only half that of CRT used in today’s television sets [7-9]. It is well

known that QE is the most important factor for phosphors. Therefore, new VUV phosphors

with visible QE higher than 100% are required in applications such as PDPs and MFFLs to

remedy this deficiency.

Since the energy of a VUV photon is more than twice that of a visible photon, it is

theoretically possible to achieve two visible photon emissions for every incident VUV photon

with a QE of up to 200%. This two-photon luminescence phenomenon is called as QC,

quantum splitting (QS) or photon cascade emission (PCE), which was predicted by Dexter as

early as 1957 [10]. The efficiency gain in QC materials is based on the principle that a QC

phosphor is able to emit two visible photons for every VUV photon absorbed. Extensive

research on VUV energy levels and spectroscopy of RE ions has recently been reported

[11-47]. Especially, research on the 4ƒn energy levels has extended the existing Dieke

diagrams into the VUV region [48-50], and high resolution excitation VUV spectra have also

been measured and a model has been developed explaining the 4ƒn-15d energy level structure

[51-54]. Furthermore, synchrotron radiation as an intense tunable source of UV/VUV regions

plays a crucial role in the VUV spectroscopy studying [55]. All these studies and

experimental facilities provide prerequisite conditions for investigating the QC effects.

Research on QC systems started on single RE ions which are capable of a cascade

emission. Two conditions must be fulfilled for efficient visible QC on single ions [34]: (i) the

energy gap between adjacent levels must be large enough to prevent multiphonon relaxation;

and (ii) the branching ratio of visible emission must be high. The energy level scheme of Pr3+

favors QC by cascade emission is shown in Fig. 1. Incident VUV photons are absorbed by

9
ACCEPTED MANUSCRIPT

Pr3+ from its ground state (3H4) into the 4ƒ5d configuration, when the lowest 4ƒ5d levels are

above the Pr3+ 1S0 state, the excitation decays to the 1S0 state situated at 47 000 cm–1. Then 1S0

is de-populated through two-step 4ƒ2→ 4ƒ2 transitions, involving transitions 1S0→ 1I6, 3PJ

(~400 nm) followed by 3P0→ 3FJ, 3HJ (480–700 nm). Consequently, a system with internal QE

greater than 100% could therefore be achieved. In 1970s the QC effect was first observed in

Pr3+-doped YF3 with a QE of about 140% under the excitation of 185 nm [56-58]. QC was

also observed in Pr3+-doped oxides such as SrAl12O19 [59], LaMgB5O10 [60], LaB3O6 [61]

under VUV excitation. It is interesting to note that QC can also be occurring upon excitation

with X-ray [62-64].

In Pr3+ ion, QC occurs which depends critically on the relative energy position

of the 4ƒ5d states with respect to the 1S0 state. In order to observe QC, the 1S0 level must be

below the 4ƒ5d states. According to Dorenbos [65,66], it is possible to calculate the lowest

energy position of 4ƒ5d states for RE ions in many host matrices. Based on Dorenbos's works,

van der Kolk and coworkers [67,68] have made an attempt to predict in which compounds

QC can be expected. Hosts with weak crystal field, low phonon energies, large band gap,

large cation–anion distance, and large coordination number for the substitution site are

considered important to exhibit QC phenomenon [63]. Up to now, QC has been reported in a

variety of Pr3+-doped fluorides and several oxides. And it has been concluded that host lattice,

temperature, and Pr3+ concentration greatly affect the luminescence property of Pr3+-doped

materials. According to Judd-Ofelt (J-O) calculations [69,70], it is possible based on the

theory to achieve a visible QE of 199% from Pr3+ [71]. However, Pr3+ singly doped materials

as such could not be employed in lamp and display phosphors since the first photon emission

10
ACCEPTED MANUSCRIPT

corresponding to the 1S0→ 1I6 transition (~405 nm) is located in the near UV region, which is

not sensitive to the human eyes. Moreover, the unwanted process of self-trapped exciton

(STE)-mediated ET induced by the host absorption might strongly reduce the visible QE of

Pr3+-doped QC phosphors [67].

QC phenomenon and its possibilities in other RE ions such as Tm3+ [72], Gd3+

[34,73-76], and Er3+ [77,78] have also been investigated. However, up until now, it has been

verified from literature no visible QE higher than 100% could be obtained in phosphors based

on a single RE ion, due to the competing invisible emissions in the NIR or UV region.

In order to achieve an efficient visible QC, a special emphasis has been made

on the combinations of two or three RE ions since visible QC could be possible through ET,

which enables the excited ion transfer partial excitation energy to the two acceptor ions, and

each ion emits a visible photon. This ET-assisted process is generally known as DC [79],

which is opposite to the upconversion (UC) phenomenon [80]. DC has now been regarded as

a major requirement to realize visible QC. Recently, an efficient visible QC in VUV-excited

LiGdF4:Eu3+ phosphors with internal QE as great as 190% has been demonstrated [79,81]. In

the LiGdF4:Eu3+ phosphors, two red photons (~590 nm) are emitted via different Eu3+ ions

corresponding to 7F1→ 5D0 transitions for every VUV photon absorbed by Gd3+ ion. Through

ET from Gd3+ to Eu3+, the UV emission due to Gd3+ 6PJ→ 8S7/2 (~313 nm) is transferred

efficiently to a red photon (~590 nm) due to the transition 7F1→ 5D0 of Eu3+. An efficient

visible QC has also been obtained in Gd3+–Tb3+–Er3+ system with a maximum QE of 130%

[82]. Gd3+ so far, has been found in a number of DC systems [83-93], where the Gd3+ ion is an

element of the host lattice and the ET generally occurs via Gd3+ ion. In addition, to obtain an

11
ACCEPTED MANUSCRIPT

efficient visible QC, attempts to sensitize the absorption of Gd3+ by using Er3+ [82], Tm3+ [94],

Pb2+ [95], Pr3+ [96], and Nd3+ [97,98] have been employed. However, no promising phosphor

for practical applications has so far been found. On the other hand, Eu3+ [99,100], Er3+ [101],

Tm3+ [102], Mn2+ [103-110], and Cr3+ [111-114] have been investigated as an appropriate

codopant for Pr3+ ion which might convert the 405 nm photon due to the 1S0→ 1I6 transition of

Pr3+ to proper visible photon through ET.

To fulfill the requirements of QC, the host materials should have a band gap

larger than 3.0 eV, otherwise the material will absorb visible light; the excitation energy must

be higher than 6 eV (λ<200 nm) so that the high energy levels (E> 50000 cm–1) can be located

within the band gap; and the energy of phonons should be as low as possible in order to

reduce the probabilities of multi-phonon relaxations between spaced energy levels of RE ions

[92,115,116]. Fluorides are having wide band gap and low phonon energy and hence are

potential hosts for QC as they could ensure an excellent performance of about 200% QE. In

fact, QC phenomenon was first observed in Pr3+-doped fluorides [56-58]. However, fluoride

phosphors are generally considered to be unstable and not suitable for lighting or display

applications [117,118]. In contrast, oxides have advantages over fluorides, such as high VUV

absorption efficiency, high stability, easy fabrication and low cost, which are more suitable

for practical applications [107]. As the penetration depth of VUV photons is extremely

shallow [119], a reduction in the grain size and an improvement in morphology for phosphors

would significantly improve the luminescence efficiency under VUV excitation [120]. On the

other hand, nanophosphors have been extensively investigated due to their potential

applications in high-performance displays [121]. Accordingly, QC nanophosphors could be

12
ACCEPTED MANUSCRIPT

identified as more promising materials for PDPs and MFFLs. Solid-state reaction,

hydrothermal synthesis, chemical coprecipitation, wet-chemical sol-gel process and

combustion synthesis have conventionally been employed to synthesize QC phosphors. An

efficient QE could be achieved if the material synthesis procedure could be optimized [79].

All the aforementioned DC systems are based on resonant ET [122]. This is the

first-order DC, which requires a good overlap between emission spectrum of the donor and

the excitation spectrum of the acceptor [10,123]. Splitting of the energy could be possible by

means of the population of an intermediate energy level of the donor. If there is no such

overlap, a second-order DC could be found to be more dominant relaxation process in

competing with a spontaneous emission [122]. In this process a donor simultaneously excites

two acceptors. This cooperative sensitization process was predicted by Dexter [10] as early as

1957, however, only recently the occurrence of cooperative sensitization (second-order DC)

was reported by Basiev et al. [124]. Subsequently, a cooperative DC has been observed from

Tb3+–Yb3+ dual ions in KYb(WO4)2 [125,126] and YPO4 [122]. The second-order DC is

regarded as a promising approach to produce QC phosphors [120].

Although, QC phosphors via DC have been investigated for decades for their

use in the lighting and display industry, only recently their applications in solar cells have

been realized with enhanced conversion efficiency [122,127-136]. The most widely used solar

cells are based on crystalline silicon (Si). The thermalization of electron-hole pairs, generated

by the absorption of high-energy photons, is one of the major energy loss mechanisms in a

conventional solar cell. Thermalization losses can be reduced by using DC whereby, for

example, a photon with twice the energy of the band gap becomes transformed into two

13
ACCEPTED MANUSCRIPT

photons with an exactly the same band gap energy. Theoretically, for a Si-based solar cell in

conjunction with an ideal DC layer, conversion efficiency up to 38.6% could be achieved

from the sunlight, which is a significant improvement over the limiting efficiency of 30.9%

for conventional solar cells under the same assumption of only radiative recombination [127].

As DC can be a linear process, it will be easier to gain a performance advantage using the

solar spectrum as the illumination source [128]. If conversion of one UV/visible photon into

two NIR photons is realized, QC phosphors could greatly benefit the development of solar

cells as their emission energy is just above the band gap of Si (Eg=1.12 eV, λ=1100 nm).

The Tb3+–Yb3+ dual ion combination is one of the promising systems to realize

NIR QC, since the Tb3+ 5D4→ 7F6 transition (~485 nm) is located at approximately twofolds

the energy of the Yb3+ 2F7/2→ 2F5/2 transition (~980 nm) and Yb3+ has no other energy level up

to the UV region. Moreover, the 2F7/2→ 2F5/2 emission (~980 nm) is just above the band-gap

energy of Si. Vergeer et al. [122] reported cooperative QC for Tb3+–Yb3+ couple in YPO4 with

a maximum QE of 188% and showed that those Tb3+–Yb3+ cooperative QC phosphors may

increase the efficiency of Si solar cells by downconverting the green-to-UV part of the solar

spectrum to ~1000 nm photons, with almost two times the number of photons. Subsequently,

Zhang et al. [131-135] observed an efficient NIR QC for Tb3+–Yb3+ couple in GdAl3(BO3)4,

GdBO3, Y3Al5O12, and Zn2SiO4 phosphors. Upon excitation of Tb3+ with a blue-visible photon

at 485 nm, two NIR photons could be emitted by Yb3+ through an efficient cooperative ET

from one Tb3+ to two Yb3+ with an optimal QE close to 200%, before reaching the

concentration quenching threshold. Furthermore, efficient NIR QC for RE3+–Yb3+ (RE= Pr3+,

Tm3+, Tb3+) couple in GdAl3(BO3)4 and Y3Al5O12 has also been demonstrated [133,134].

14
ACCEPTED MANUSCRIPT

These RE3+–Yb3+ (RE= Pr3+, Tm3+, Tb3+) cooperative QC systems belong to second-order DC.

More recently, Ye et al. [137] reported NIR QC in Tb3+–Yb3+ codoped transparent

glass-ceramics containing CaF2 nanocrystals with an optimum QE of 155%, and suggested

that this glass-ceramics could be a promising DC layer material to enhance efficiency of Si

solar cell.

1.3. Challenge and objectives

An overwhelming contribution has successfully been made to the development

of QC phosphors from the advent of technology such as micro- and nano-solid materials

synthesizing, functioning and characterizing, development of VUV science and technology,

QC and DC mechanism investigating as well as new luminescent materials development.

However, a deep insight into the factors dominating the general trends of applications of QC

phosphors and the mechanism behind is still in its infancy.

The main objective of this contribution is to present the very recently made

progress and advances on: (a) materials and developments in the fields of UV-visible QC and

the mechanism involved, including QC in single RE ions and in ion pairs activated phosphors:

QC materials, materials synthesis and characterization, ET and DC; and (b) NIR QC via

cooperative DC in RE3+–Yb3+ (RE=Tb, Tm, and Pr) dual ions based phosphors. Appropriate

discussions have been made on new materials, materials synthesis and characterization, the

structural and luminescence properties of various QC luminescent materials via different

synthesis techniques. In addition, applications, challenge and future advances of the visible-

and NIR QC phosphors have also been dealt with.

15
ACCEPTED MANUSCRIPT

2. Fundamental aspects of quantum cutting phosphors

2.1. Basic concepts

2.1.1. Basics of energy transfer between rare-earth ions

ET between RE ions finds wide application in sensitizing solid state lasers,

infrared quantum counters as well as infrared to visible convertors. ET between RE ions also

plays important role in QC processes. In this chapter, the basic processes of ET and their

applications to QC are presented. Aside from ET by movement of charge transports, there

remain four basic mechanisms involved in ET processes between RE ions: (a) resonant

radiative transfer through emission of sensitizer (S) and reabsorption by activator (A); (b)

non-radiative transfer associated with resonance between absorber (sensitizer) and emitter

(activator); (c) multiphonon assisted ET; and (d) cross-relaxation (CR) between two identical

ions. A schematic diagram to illuminate the different ET processes between two ions is

presented in Fig. 2 [80]. The efficiency of radiative transfer (Fig. 2a) depends on how

efficiently the activator fluorescence is excited by the sensitizer emission. It requires a

significant spectral overlap of the emission region of sensitizer and the absorption region of

activator and an appreciable intensity of the absorption of activator. If the radiative ET takes

place predominantly, then the decay time of sensitizer fluorescence does not vary with the

activator concentration. In contrast to the non-radiative ET (Fig. 2b), the radiative ET

accompanies the significant decrease in decay time of sensitizer fluorescence with the

activator concerntration. Becasue of the requirement for a considerable absorption capability

of activator, the radiative ET can usually be neglected relative to non-radiative ET in most

inorganic systems. Only in a few cases, the requirements for radiative ET are satified [118].

16
ACCEPTED MANUSCRIPT

ET may occur if the energy difference between the ground and excited states of donor (or

sensitizer) is equal to that of acceptor (or activator) and there exists a suitable interaction

between systems [118]. Non-resonant ET can also take place by the assist of phonon unless

the difference between the ground and excited states of donor and acceptor is large. The ET

rate between a donor and an acceptor is derived by Dexter [118,138] as follows:

2π 2
WDA =
h
a ' b H DA ab' ∫g D ( E ) g A ( E )dE (1)

where D and A are a donor and an acceptor, respectively. The factors gD(E) and gA(E)

represent the normalized shape of the donor emission and acceptor absorption spectra,

respectively. The matrix elements in Eq. (1) can be expressed as a function of the distance

between donor and acceptor, so that the ET probability depends upon the distance between

donor and acceptor. The distance varies with the interaction type. For exchange interaction it

is exponential, while it is of the type R-n for the multipolar interactions [118,138].

If two RE ions are with different excited states, as shown in Fig.2c, the probability

for ET should drop to zero, where the overlap integral ∫g D ( E ) g A ( E )dE vanishes [80].

However, it is experimentally found that ET can take place without phonon-broadened

electronic overlap provided that overall energy conservation is maintained by production or

annihilation of phonons with energies approaching kΘ d , where Θ d is the Debye

temperature of the host matrix [139]. Then for small energy mismatch ( ≈ 100 cm-1), ET

assisted by one or two phonons can take place [140]. However, in ET between RE ions,

energy mismatches as high as several thousand reciprocal centimeters are encountered. This is

much higher than the Debye cut off frequency found in normally encountered hosts, so that

multiphonon phenomena have to be considered. This was done by Miyakawa and Dexter

17
ACCEPTED MANUSCRIPT

[141,142]. In their theoretical analysis of multiphonon processes, Miyakawa and Dexter

derived a comparative relaxation analogue of the multiphonon gap dependence. According to

their theory, the probability of phonon-assisted transfer (PAT) is expressed by,

WPAT (ΔE ) = WPAT (0)e − βΔE (2)

where ΔE is the energy gap between the electronic levels of donor and acceptor ions and β is a

parameter determined by the strength of electron-lattice coupling as well as by the nature of

the phonon involved. The above equation has the same form as that for the energy gap

dependence of the multiphonon relaxation (MPR) rate, which is also given by the

Miyakawa-Dexter theory as:

WMPR (ΔE ) = WMPR (0)e − βΔE (3)

It is further indicated that the parameter α is given by:

α=
1
[ln{N / g (n + 1)} − 1] (4)

1
where α and β are connected with each other as: β = α − γ , and γ = ln(1 + g S / g A ) ,

where g is the electron-lattice coupling constant, suffixes S and A are sensitizer and activator

ions respectively, n is the number of phonons excited at the temperature of the system, ћω is

the phonon energy which contributes dominantly to these multiphonon processes and N is the

number of phonons emitted in the processes, namely, N = ΔE / ћω. Nonresonant

phonon-assisted ET between various trivalent RE ions in yttrium oxide crystals were

thoroughly studied by Yamada et al. [143]. In their experiments the energy gap between the

sensitizer and activator system varied in a wide range of energies up to 4000 cm-1. The

probability of phonon-assisted ET was observed to obtain the exponential dependence on

energy gap predicted by the Miyakawa-Dexter theory [142]. It was revealed that the phonons

18
ACCEPTED MANUSCRIPT

of about 400 cm-1 which produce the highest intensity in the vibronic side bands of yttrium

oxide contribute dominantly to the phonon-assisted process [142].

CR terminology usually refers to all types of DC ET between identical ions. In

such a case the same kind of ion is both a sensitizer and an activator. As shown in Fig.2d, CR

may give rise to the diffusion process already considered between sensitizers when the levels

involved are identical or to self-quenching when they are different. In the first case there is no

loss of energy, whereas in the second there is a loss or a change in the energy of the emitted

photons [80].

2.1.2. Quantum cutting

The energy of a VUV photon is twice more than that of a visible photon.

Theoretically, it is possible to generate two visible photons for a single VUV photon absorbed;

this two-photon luminescence phenomenon is called as QC, QS or PCE. The efficiency gain

in QC materials is based on the principle that a QC phosphor can result in two visible photons

for each absorbed VUV photon. QC has been demonstrated based on different mechanisms

[71,81,129]. Fig. 3 illustrates the energy level diagrams for two (hypothetical) types of RE

ions (I and II) showing the concept of DC [81]. Efficient visible QC via two-photon emission

from a high energy level for a single RE ion is theoretically possible as shown in Fig. 3(a).

However, competing emissions in the IR and UV (the thin lines of Fig. 3(a)) can also occur

and prevent efficient visible QC on a single RE ion. Figs. 3(b)-3(d) presents generalized

energy level diagrams for three DC mechanisms involving ET between two different RE ions

I and II. Type I is an ion for which emission from a high-lying level occurs. Type II is an

19
ACCEPTED MANUSCRIPT

activator ion to which ET takes place. Fig. 3(b) indicates two photon emission from ion pairs

by CR from ions I to II (denoted by ①), and ET from ions I to II (denoted by ②) with

emission from ion II. Figs. 3(c) and 3(d) show a CR mechanism followed by the emission of

photons from both ions I and II. In all three cases, if the two-step ET process is efficient, a

theoretical visible QE of 200% can be achieved because the previous IR and UV losses

existing in a single ion can be avoided.

2.2. Experimental description: materials synthesis and fabrication

It is well known that materials’ performances are closely related to the ways that

are processed [144]. Synthesis method of phosphors has played a significant role in

determining the microstructural, luminescence properties and QE of phosphors. Phosphors

were conventionally synthesized via a solid-state reaction process, using powders raw

materials as the starting materials. Due to their relatively rough grains, these phosphors

require relatively high sintering temperature to obtain phosphors with designed compositions

and desired performances. For non-oxides phosphors, they are very likely to undergo changes

in their properties during the sintering at high temperatures due to the high volatility and

unstable chemical properties. To reduce the sintering temperature, it is necessary to use

powders of phosphors compounds in very small grain sizes and narrow size distribution. For

this purpose, submicron or even nanosized products have been synthesized by various

wet-chemistry methods over the past several years, including hydrothermal synthesis,

chemical coprecipitation, sol-gel process, combustion synthesis, etc. Although significant

progresses have been achieved, there are problems. For example, sol-gel process uses metal

20
ACCEPTED MANUSCRIPT

alkoxides as the starting materials, which are very expensive and extremely sensitive to the

environmental conditions such as moisture, light and heat. Moisture sensitivity makes it

necessary to conduct the experiment in dry boxes or clean rooms. Co-precipitation processes

involve repeated washing in order to eliminate the anions coming from the precursor salts

used, making the process complicated and very time consuming. Furthermore, it is difficult to

produce large batches by using most of the wet-chemical solution processing routes [144].

Therefore, exploring alternative methods for the preparation of phosphors is still of

technological as well as scientific significances.

2.2.1. Solid-state reaction

Solid-solid reaction, which occurs between powders in the solid state, has been

used partly to synthesize inorganic materials because of the facile operation. Up to now, this

method has been the most widely and intensively used approach for the preparation of QC

phosphors including fluorides and oxides, because it is comparatively simple and very

suitable for mass production. Generally, solid-solid reaction technique to obtain

fluoride-based QC phosphors consists of mixing, heating above their melting points, and

cooling the fluorides of each element incorporated in the synthesized fluoride under an inert

gas atmosphere or vacuum [34,89,104,115,116]. Fluorination by SF6, HF or F2 is also

performed in order to extract oxygen contamination. Because of refractory nature of alumina

and RE oxides, conventional solid-solid reaction synthesis of oxide-based QC phosphors

requires temperature higher than 1000 °C. The mechanism of solid state reactions is diffusion

control reaction and hence, repeated grinding and repeated heating are required

21
ACCEPTED MANUSCRIPT

[59,74,109,114]. Furtheremore, the controlled atmosphere is necessary to master the valence

of the activator and the stoichiometry of the host lattice.

There are several limitations in synthesis of QC phosphors by a conventional

solid-state method. Firstly, this process often results in poor homogeneity and requires high

calcinations temperature. Secondly, the grain size of phosphor powders prepared by this

method is in several tens of micrometers. Phosphors of small particles must be obtained by

grinding the larger phosphor particles. Furthermore, preparation of single phase compound is

difficult by the conventional solid-state method. This process easily introduces additional

impurities and defects which would greatly reduce luminescence efficiency [145].

2.2.2. Hydrothermal synthesis

Hydrothermal method is a promising synthetic route, which can be better

controlled from the molecular precursor to the reaction parameters, such as the reaction time

and temperature, to give highly pure and homogeneous materials. The technique allows low

reaction temperature and controllable size, phase and morphology of the products [146-151].

Fluorides are promising host materials to realize the QC phenomenon for its low phonon

energy and wide band-gap; the interesting and appreciable QE in the visible spectral region

has earlier been reported from LiGdF4:Eu (190%) [79] and BaF2:Gd,Eu (194%) phosphors

[86]. The most attractive advantage of hydrothermal techniques in synthesis of fluorides lies

in that the oxygen impurities can be prevented from entering the compound lattices [152-155].

Recently, efforts have been done to develop fluoride-based QC phosphors through

hydrothermal synthesis [83-85,156-158]. These experimental results have shown that the

22
ACCEPTED MANUSCRIPT

critical factors for the pure products are dependent on the pH value, the initial composition

and the reaction temperature [85], while the controlled atmosphere is not necessarily required

[83].

2.2.3. Solution-based chemical synthesis

Solution-based chemical syntheses, namely wet-chemical methods, are defined

as techniques which do not comprise of the normal mixing, calcinations and grinding

operations [159]. In fact, the solution-based chemical synthesis such as hydrothermal, sol-gel,

co-precipitation and combustion have been received considerable attention since they offer

the possibilities for controlling homogeneity, purity of phase, size distribution, surface area,

and microstructural uniformity of the phosphors. Hydrothermal and sol-gel methods have

been successfully employed to synthesize fluoride-based QC phosphors.

Sol-gel synthesis: the beginning of the sol–gel methods can be dated back to as

early as 1846 when Ebelenen discovered the formation of SiO2 gel by hydrolyzing Si(OEt)4.

Only in the 1930s did this method begin its further development. The basic processing steps

of this method can be summarized as follows: Precursor → hydrolysis → reactive monomer

→ condensation → sol gelation → gel → further treatment. Based on this synthetic route,

variations of precursors, solvents, ligands, different addition sequences of compounds and

further treatment as well as other changes in sol–gel methods have been reported and widely

applied in the preparation of, e.g., glasses, ceramics, inorganic fillers and coatings [160-164].

An intense level of research activity has provided a special focus on the development of

luminescent materials via sol-gel method in the past years [165-189]. Nevertheless, so far

23
ACCEPTED MANUSCRIPT

most inorganic compounds based on the sol-gel process consist almost only of oxide networks.

Fluorides are promising host materials to show QC effect, however, some problems might

arise from the synthesis procedure. Therefore, there is a real need to prepare fluorides by

means of convenient sol-gel method. More recently, fluoride materials prepared through

sol-gel method have been reported in the literature [190-194], and efficient QC has been

demonstrated in the sol–gel derived LiGdF4:Eu3+ powder phosphors [194].

Chemical coprecipitation: coprecipitation has also been used to fabricate

fluoride-based [195] and oxide-based QC phosphors [136,196]. Co-precipitation synthesis

may be regarded as an attractive method for the preparation of complex fluorides, because it

does not require expensive high purity RE fluorides as starting materials, and results in

smaller grains and more homogeneous distribution of dopant ion. However, co-precipitation

processes involve repeated washing in order to eliminate the anions coming from the

precursor salts used, making the process complicated and very time consuming. Moreover, to

obtain fluorides, the post-fabrication thermal treatment in an inert or reactive atmosphere is

indispensable for removing of water molecules adsorbed at a surface of the powder and OH−

groups incorporated into lattice. And calcination is required to get crystalline phosphors.

Liquid phase method: pure NaGdF4:Eu3+ nanocrystals in the cubic and the

hexagonal crystal phase were synthesized in the high-boiling coordinating solvent

N-(2-hydroxyethyl)-ethylenediamine, by only adjusting the molar ratio between metal and

fluoride ions in the synthesis, recently [197]. When this molar ratio is close to stoichiometric,

the hexagonal phase is formed, whereas the cubic phase is obtained in the presence of excess

metal ions. The optical properties of NaGdF4:Eu3+ nanocrystals are different for the two

24
ACCEPTED MANUSCRIPT

crystal phases. The results indicate an increased number of oxygen impurities close to Eu3+

ions, if excess metal ions are used in the synthesis.

Combustion synthesis: solution combustion synthesis is another wet-chemical

method, which has been proved to be an excellent technique for preparing several grams

micro/nanocrystalline phosphors due to its short processing time, low processing temperature,

low cost and high yield as well as good ability to achieve high purity in making single or

multiphase complex oxide powders at the as-synthesized state [198-202]. This technique is

based on exothermic redox reactions that undergo self-sustaining combustion. Mixtures of

metal nitrates (oxidizers) and a fuel (reducer; e.g., urea, citric acid or glycine) undergo

spontaneous combustion under heating, and the chemical energy from the exothermic reaction

heats the precursor mixture to high temperatures. Such a high temperature leads to formation

and crystallization of phosphor materials. In comparison with other methods, the products

obtained by the combustion synthesis method are generally more homogeneous, have less

impurity, and have higher surface areas than powders prepared by conventional methods

[203]. Till now, this technique has been employed to produce a variety of materials such as

oxides, borates, silicates, and aluminates [204-224].

2.2.4. Luminescent glass and glass-ceramics by melting-quenching method

RE-doped glasses have been extensively investigated due to their potential

applications in solid state lasers, fiber amplifiers, and color displays. Among them,

oxyfluoride glass-ceramics are more appropriate for practical applications due to their low

phonon energies compared to oxide glasses and also for their possessing excellent chemical

25
ACCEPTED MANUSCRIPT

durability and mechanical strength compared to fluoride glasses [137]. The oxyfluoride

glasses were prepared by employing the melting-quenching technique using powders of the

raw materials as the starting materials. The melts were then poured onto a cold brass plate and

annealed at around glass transition temperature, Tg, for several hours to remove thermal

strains. For the preparation of glass-ceramics, the as-made glass was then underwent thermal

treatment at a given temperature and time according to the thermal analysis measurements.

2.3. Materials characterization and properties

The synthesized QC luminescent materials were characterized by X-ray

diffraction (XRD), thermal analysis, scanning electron microscopy (SEM), transmission

electron microscopy (TEM), Raman spectrum, and IR spectroscopy. The luminescent

properties including excitation spectra, emission spectra and fluorescence decay curves in the

VUV-visible-NIR wavelength ranges were performed with spectrophotometers.

2.3.1. Structural properties and morphology

XRD, SEM, and TEM are the most important characterization tools used in solid

state chemistry and materials science. XRD has been extensively employed to characterize the

phase composition and structure of QC phosphors synthesized. SEM is especially useful for

convenient observation of grain and grain boundary structures. And TEM has been widely

used to characterize the morphology, defects, phase structure of QC phosphors, especially for

those in nanometer scales.

XRD: X-rays are electromagnetic radiations having the wavelength around 0.1

26
ACCEPTED MANUSCRIPT

nm. The wavelength of X-rays is about the same as the interatomic distances in crystals.

When high energy electrons strike an anode (e.g., copper, iron or molybdenum) in a sealed

vacuum, X-rays are generated. X-rays discovered in 1895 provides a fundamental condition

to probe crystalline structure at the atomic level. XRD has been used widely for identifying

the crystal phases and understanding the structures of crystalline solids. Each crystalline solid

has its unique characteristic X-ray powder diffraction pattern which may be used as a

fingerprint for its identification. Once the material has been identified, XRD patterns may be

used to determine its structure. For instance, we can calculate the size and shape of a unit cell

from the positions of the XRD peaks and we can determine the positions of the atoms in the

unit cell from the intensities of the diffraction peaks. Furthermore, it is possible to determine

an average crystallite size from the broadening of the diffraction peaks using the Scherrer’s

0.9λ
formula [225]: D = , where D denotes the average size of the crystallites, β
cos θ β 2 − β 02

is the observed full width at half maximum of a diffraction line located at θ, β0 represents the

scan aperture of the diffractometer, and λ stands for the X-ray radiation wavelength. It should

be noted that the estimate of the crystallite size is volume averaged in the direction

perpendicular to the plane of diffraction and therefore provides no information about the

crystallite size distribution. In general, XRD measurements are carried out using a θ–2θ

diffractometer with CuKa (λ= 0.15406 nm) radiation. Besides of identifying the crystal phases,

crystalline structures and estimating the mean size of the crystals, XRD can also be used to

identify the occurrence of lattice shrinkage in RE-doped QC phosphors, which resulted from

the replacement of host ions with smaller or biger RE ions. For example, one can observe that

the XRD diffraction peaks move towards the higher angle side in comparison with the

27
ACCEPTED MANUSCRIPT

standard XRD pattern of BaF2 crystals, due to the replacement of Ba2+ with smaller ions Tm3+

or Ho3+ in BaF2: Tm3+, Ho3+ QC phosphors.

SEM: SEM usually uses an electron beam spot of about 1 µm in diameter, which

is scanned repeatedly over the surface of samples. Slight variations in surface topography

produce marked variations in the strength of the beam of secondary electrons, which are

ejected from the surface of samples by the force of collision with primary electrons from the

electron beam. The magnification of SEM is less than that of TEM but much better than that

of an optical microscope. Usually, SEM practically attached with an electron

energy-dispersive spectrometer (EDS) could be carried out to obtain detailed information

about the morphology of surface and interfaces, structure, composition and the homogeneity

of the synthesized QC phosphors [90,120,132,136,144].

TEM: the design principle of TEM is similar to that of a conventional optical

microscope. The only difference is that TEM uses a beam of electrons focused by

electromagnets, while optical microscopes use a beam of light focused by glass lenses. The

wavelike nature of electron is the basis of TEM. The monochromatic wavelength of electron

beam is about five orders of magnitude smaller than the wavelength of visible light (400–700

nm) used in optical microscopes. As a result, TEM can resolve much smaller structural details

than optical microscopes. The resolution of TEM is about 1 nm or less, as compared to about

0.25 µm, the best resolution achieved by optical microscopes. In comparison, the resolution

of the SEM is about an order of magnitude poorer. Therefore, TEM is indispensable for the

structural imaging of nanometer-sized features, and is used to obtain structural information

from specimens thin enough to transmit electrons. Lattice imaging and diffraction yields

28
ACCEPTED MANUSCRIPT

crystallographic and orientation effect information on structural features, defects, and phases

of various QC materials [131,196,197,226]. The analytical capability of EDS equipped with

TEM allows elemental identification through measurement of characteristic X-ray energies.

2.3.2. Thermal analysis

The most widely used thermal analysises include differential thermal analysis

(DTA), differential scanning calorimetry (DSC), and thermogravimetry analysis (TGA). DTA

and DSC monitor the difference in heat flow to or from a sample and to or from a reference as

a function of temperature or time, while the sample is subjected to a controlled temperature

program. TGA is a technique by means of the mass of the sample is monitored as a function

of temperature or time. Thermal analysises have been employed to characterize phase

formation, decomposition, thermal stability, and sintering behavior of QC luminescent

materials. For QC glass-ceramics, DTA together with XRD measurements were employed to

analyze the temperature difference between the nanocrystals and the bulk amorphous matrix,

to provide the possibility to control the crystallization and growth of the nanocrystals.

2.3.3. Spectral analysis: Raman and IR

Raman spectrum [120,131,227] and IR spectroscopy [135,226,227] have been

used to confirm the phase formation of synthesized QC phosphors and identify the presence

of impurities and defects. Raman spectroscopy is a spectroscopic technique used to study

vibration, rotational, and other low-frequency modes in a material system. It relies on

inelastic scattering, or Raman scattering of monochromatic light, usually from a laser source

29
ACCEPTED MANUSCRIPT

in the visible, NIR, or near UV range. Phonons or other excitations in the material tested are

absorbed or emitted by the laser light, resulting in the energy of the laser photons being

shifted up or down. The shift in energy gives information about the phonon modes in the

system. IR spectroscopy yields similar, but complementary information.

2.3.4. Photoluminescence

In order to investigate the QC process and ET mechanism, the high-resolution

excitation and emission spectra in the VUV region have usually been measured with a

synchrotron radiation as excitation source. Synchrotron radiation can offer an intense and

tunable VUV-UV excitation light. Furthermore, the temperature of the sample could be varied

between liquid helium and room temperature (RT), so the low-temperature excitation and

emission spectra also can be recorded using synchrotron radiation. Zimmerer [55] recently

provided a brief review on discussing the history, highlights and future of luminescence

spectroscopy with synchrotron radiation, and emphasized the important role of synchrotron

radiation in the investigation of QC effect. Ever since the pioneering experiment was carried

out by Wegh et al. [79] using synchrotron radiation and the equipment of the HIGITI

experimental station of the Synchrotronstrahlungslabor HASYLAB at DESY (Deutsches

Elektronen-Synchrotron) in Hamburg (Germany), large numbers of spectroscopic

investigations on QC effect have been performed at DESY [34,64,67,68,74,75,82,94,95].

Details of the experimental setup have been described elsewhere [228,229]. In addition,

several measurements of VUV excitation and emission spectra have been carried out at the

Spectroscopy Station of Beijing Synchrotron Laboratory (China) [83-86], and National

30
ACCEPTED MANUSCRIPT

Synchrotron Radiation Laboratory in University of Science and Technology of China

[108-110,112-114,156]. Low-resolution emission and excitation spectra at RT in the

VUV-visible wavelength range have been carried out on a conventional spectrophotometer

with a continuous Xe lamp [87-89,104,230,231] or a Deuterium lamp [90,94-96,115,145]. For

the NIR QC effect, the luminescent spectra and time-resolved measurements in the

UV-visible-NIR region are generally obtained from the Triax320 spectrometer (Jobin–Yvon

corp.) equipped with a Xe-lamp as the excitation source [131-136].

31
ACCEPTED MANUSCRIPT

3. Single ion activated phosphors

3.1. Quantum cutting in Pr3+ ion activated phosphors

The luminescence of the Pr3+ has been investigated thoroughly for various

applications, such as fiber optical communications [232-235], field emission display devices

[117,236-238], laser applications [239-243], scintillator applications [244-246], and tunable

UV lasers [247-249]. The 4ƒ energy level scheme of Pr3+ ion proposed by Dieke [250] in

1968 has also been used for the conversion of one UV photon into two visible photons

[56-58]. A special focus has been made on the PCE or QC phenomenon from the 1S0 state of

Pr3+. The PCE process of Pr3+ can be observed only when the lowest energy 4ƒ5d states are

situated above the 1S0 state. Hosts with weak crystal fields, low phonon energies, large band

gap energies, large cation–anion distances, and large coordination numbers for the

substitution could be desired [63]. Up to now, PCE from Pr3+ has been found in a number of

fluoride hosts and several oxides, which will be discussed in detail in the following parts.

However, an efficient quantum cutter based on Pr3+ single ion is still not possible.

3.1.1. Principle

In principle, two types of emission transitions from Pr3+, namely

inter-configurational 4ƒ5d→ 4ƒ2 or intra-configurational 4ƒ2→ 4ƒ2, could take place with a

VUV excitation, as shown in Fig. 4. The emission transition of Pr3+ depends on the host

lattice and the energetic location of the 4ƒ5d states relative to the 1S0 state, which is located at

about 47 000 cm–1. If the lowest 4ƒ5d state lies below the 1S0 state, the high energetic

excitation will stimulate the broadband emission from the lowest 4ƒ5d state, which is

32
ACCEPTED MANUSCRIPT

parity-allowed inter-configurational 4ƒ5d→ 4ƒ2 transition, as illustrated in Fig. 4(b). Thus no

QC is observed. The 4ƒ5d→ 4ƒ2 emission is usually situated at the UV region, which may be

interesting for application in fast scintillator materials. In most hosts so far has been

investigated, the 1S0 lies above the 4ƒ5d states, for example in LiYF4 [156], CaF2 [57,58,251],

BaF2 [251], K2YF5 [252], Cs2NaYCl6 [253], Y3Al5O12 [254,255], YAlO3 [255], and CaAl4O7

[256]. Fig. 5 shows the excitation and emission spectra of Pr3+ in YPO4 and YBO3 at 293 K

[257]. Both the emission spectra are found to be dominated by the broad-band 4ƒ5d→ 4ƒ2

transitions of Pr3+, implying that the 1S0 state lies above the lowest 4ƒ5d state. The narrow

band peaking at 489 nm in YPO4:Pr3+ is assigned to the 3P0→ 3H4 emission which may

originate from the lowest 4ƒ5d state.

In contrast, when the lowest 4ƒ5d state is located above the 1S0 state, the 1S0 will

be populated under excitation into 4ƒ5d levels. After the 1S0 de-populated to the ground state,

QC takes place due to a two-step intra-configurational 4ƒ2→ 4ƒ2 transitions: 1S0→ (1I6, 3PJ)

(~400 nm) followed by 3PJ→ (3FJ, 3HJ) (480–700 nm) as shown in Fig. 4(a). The most intense

second step can either be 3P0→ 3H4 (~480 nm) or 1D2→ 3H4 (~600 nm) depending upon the

radiative decay possibilities from 3P0 to 1D2. Generally, the process could be determined based

on the maximum phonon energy of the host-material. Table 1 summarizes the maximum

phonon energies of different type’s inorganic host-materials. For a QC material with small

maximum phonon energy, the nonradiative decay from 3P0 to 1D2 through multiphonon

relaxation has been found to be inefficient. The rate of multiphonon relaxation can be

calculated using the modified energy gap law of van Dijk and Schuurmans: [258,259]:

W NR = β el exp(−α (ΔE − 2hω max )) (5)

33
ACCEPTED MANUSCRIPT

where WNR is the nonradiative transition rate, the βel and α are constants for a given host

lattice, ΔE is the energy difference between the energy levels considered, and ћωmax is the

maximum phonon energy.

Fluorides exhibit phonon frequencies as low as ~500 cm–1. The nonradiative

transition rate of the 3P0→ 1D2 transition of Pr3+ in fluorides was estimated about 53 s–1 [260],

which was much smaller than the radiative transition rate of the 3P0 level ~2×104 s–1 [68]. As a

result, QC observed in fluorides involves typically two-step transitions: 1S0→ 1I6 (~400 nm)

and 3P0→ 3H4 (~480 nm). Such QC has been identified in a variety of fluorides such as YF3

[56-58], BaMgF4 [261], KMgF3 [262], NaMgF3 [263], SrAlF5 [64], LiSrAlF6 [264], and

NaYF4 [156]. Fig. 6 shows the excitation and emission spectra of Pr3+-doped YF3 and KMgF3

[262]. The presence of the emission from 1S0 indicated that the 1S0 level in KMgF3:Pr3+ and

YF3:Pr3+ is lower than the states of the 4ƒ5d configuration. The QC with a two-step emission

from the transitions of 1S0→ 1I6 and 3P0→ 3H4 of Pr3+ occurs in both YF3 and KMgF3.

However, luminescence quenching has been observed at higher Pr3+ concentration

[68,112,114,261-264]. For YF3:Pr3+ [265], decrease of the 3P0→ 3H4 emission with increasing

Pr3+ concentration has clearly been observed as shown in Fig. 7. It is noted that the blue

emission from the 3P0→ 3H4 transition was found to be almost absent when the doping

concentration increases to 10% Pr3+, which might be attributed to the efficient nonradiative

ET processes: (3P0, 3H4)→ (3F2, 1D2) and (3P0, 3H4)→ (1G4, 1G4) [265].

QC has also been observed in many oxides, including aluminates (SrAl12O19

[59], CaAl12O19 [113,114], CaMgAl14O23 [257], and Sr0.7La0.3Al11.7Mg0.3O19 [266]), borates

(LaMgB5O10 [60], LaB3O6 [61], and SrB4O7 [107,266,267], and SrB6O10 [108]) and sulfates

34
ACCEPTED MANUSCRIPT

(BaSO4 [67] and SrSO4 [67]). LaMgB5O10:Pr3+ [60] and LaB3O6:Pr3+ [61] showing efficient

emissions for the first step of the cascade process (1S0→ 1I6 transition, 400 nm), but no

emission from the 3P0 level was observed. A similar phenomenon was found in SrB4O7:Pr3+

[266], as shown in Fig. 8. The absence of emission from the 3P0 state could be attributed to an

efficient 3P0→ 1D2 multiphonon relaxation resulting from the high maximum phonon energies

(1400 cm−1) in these borates compared to the energy difference between the 3P0 and 1D2 levels

(3400 cm−1) [59]. According to the modified energy gap law of van Dijk and Schuurmans (Eq.

(5)), a large nonradiative decay rate of about 1.9 × 105 s–1 for 3P0→ 1D2 was obtained from

LaB3O6 [260]. As for BaSO4 and SrSO4, which possess relatively high maximum phonon

energies ~ 1100 cm−1 [268], efficient 1D2→ 3H4 emission was also observed [67]. However, in

some aluminates, such as Sr0.7La0.3Al11.7Mg0.3O19 [266], SrAl12O19 [59,113] and CaAl12O19

[113,114], transitions due to 1S0→ lI6 (~400 nm) and 3P0→ 3H4 (~490 nm) constitute the

cascade emission process, as shown in Fig. 9. This is because the 3P0→ 1D2 multiphonon

relaxation is inefficient due to their low phonon energies (~ 650 cm−1) with respect to the

3
P0→ 1D2 energy gap [59]. The phonon energies of oxides in general are greatly higher than

that of fluorides. Consequently, most of the oxides did not show any emission originating

from the 3P0 state except some aluminates. However, the effect of concentration quenching of

the 3P0 state at high Pr3+ concentration was also observed in SrAl12O19 [112,269], and

CaAl12O19 [114]. It was concluded that the emission from the 3P0 level in Pr3+-doped materials

depends on both the host and Pr3+ concentration. Obviously, luminescence quenching of the

3
P0 state leads to a strong reduction of QE.

It should be noted that the emission from the 1D2 level is easily quenched at the

35
ACCEPTED MANUSCRIPT

high Pr3+ concentration. Concentration quenching of the 1D2 emission has been found in

fluorides, for example in BaSiF6:Pr3+ [68]. Luminescence quenching of the 1D2 state of Pr3+

has also been found in oxide hosts. For instance, a red emission due to the 1D2→ 3H4

transition (~600 nm) has been observed in SrB4O7:0.1% Pr3+ [100], however, the red 1D2→

3
H4 emission was almost absent in BaSO4:1% Pr3+ [67]. Such an emission was absent for

SrB4O7:1% Pr3+ reported in Ref. [266]. In the case of BaSO4:1% Pr3+ [67], the luminescence

quenching of the 1D2 state has been ascribed to: (i) multiphonon relaxation, (ii) CR between

Pr3+ ions, and (iii) energy migration to quenching centres [270-272]. The last two processes

depend on the Pr3+ concentration. According to van der Kolk et al. [68,266], the 1D2→ 1G4

multiphonon relaxation was less efficient because of the large energy gap between the energy

levels (about 6400 cm–1 for BaSiF6 and 6500 cm–1 for SrB4O7). The concentration quenching

effect by CR process is therefore more likely to explain the absence of the 1D2→ 3H4 emission.

At higher Pr3+ concentrations, CR became more probable because Pr3+ ions enter the host

preferentially as pairs. It is well known that CR paths exist between two Pr3+ ions, (1D2,

3
H4)→ (3F4, 1G4), which can result in the 1D2 luminescence completely quenching even at a

1% Pr3+ concentration [68,273,264]. The radiative transition probability from the 1D2 state

was decreased due to CR. Consequently, the intensity of the 1D2→ 3H4 emission becomes

weak or even absent in most of the oxide hosts.

It was found that under excitation in the 4ƒ5d bands, both inter-configurational

4ƒ5d→ 4ƒ2 and intra-configurational 4ƒ2→ 4ƒ2 transitions simultaneously occurred in

Pr3+-doped Sr0.7La0.3Al11.7Mg0.3O19 [266], BaSO4 [67] and SrSO4 [67]. Such observation has

been reported for Pr3+-doped fluorides [56-58,252,275], which was explained by the presence

36
ACCEPTED MANUSCRIPT

of two or more different Pr3+ sites in theses hosts. In Pr3+-doped Sr0.7La0.3Al11.7Mg0.3O19, the

simultaneous observation of 1S0 line emission and the 4ƒ5d→ 4ƒ2 broad band emission

originates from two different Pr3+ sites [266]. However, there was only one type of Pr3+ site

exists in BaSO4 and SrSO4 [67]. For BaSO4:Pr3+, the intensity of the 4ƒ5d emission relative to

the 4ƒ2 emission was found to be temperature-dependent, which increased with the increase of

the temperature [276], as shown in Fig. 10. Meanwhile, the decay time of the 1S0→ 1I6

transition decreased from 190 to 56 ns with an increase of the temperature. This typical

behavior was explained by thermal population of the 4ƒ5d state from the 1S0 state [276]. At a

higher temperature, some of electrons will cross the energy barrier between the 1S0 and the

lowest 4ƒ5d state. Energy barrier of about 323 cm–1 was determined by means of the

measurements of both the intensity and the decay-times. The value was about ten times lower

than that of the energy difference between the 1S0 and the lowest 4ƒ5d state (3400 cm–1),

which was attributed to the Stokes shift of ~4700 cm–1 [277]. Such a small energy barrier then

results in with a short decay time (190 ns) for the 1S0→ 1I6 transition at 10 K due to an

admixture of 4ƒ5d wave functions in the 4ƒ2 1S0 state. Thermal population was also found in

BaSO4:Eu2+ and SrSO4:Eu2+ where the 5d state is reached by a thermal excitation from the

6
P7/2 state, resulting in both 4ƒ7 6P7/2→ 8S7/2 line and 4ƒ65d→ 4ƒ7 broad band emission from

the same Eu2+ center [278]. Generally, the process of thermal population is presented for all

Pr3+-doped materials which show the QC processes. However, thermal population can only be

observed at RT for those materials because the 1S0 level was relatively close to the 4ƒ5d states.

More recently, the pressure-dependent luminescence of the mixing 4ƒ2 1S0 and 4ƒ5d states of

Pr3+ in BaSO4 has been investigated, and a reduced admixture with an increase of pressure has

37
ACCEPTED MANUSCRIPT

been reported [279]. The mixing of the 4ƒ2 1S0 state with the 4ƒ5d states was also identified in

Pr3+ doped SrAl12O19 [280,281], however, no 4ƒ5d→ 4ƒ2 emission was observed.

An interesting phenomenon has been found in some Pr3+-doped materials: the

luminescence from the 3P0 level was excitated more efficiently than the 1S0 level under host

(band-to-band) excitation at low temperatures. Taking SrAlF5:Pr3+ [64] as a typical example,

only weak emission due to the 3P0→ 3H4 transition of Pr3+ has been observed, whereas the 1S0

emission was absent under the host excitation (λexc = 111 nm) at T = 10 K, as shown in Fig.

11. To explain this, a mechanism of recombination and transfer of the energy from excitonic

states to Pr3+ is proposed [64,282]. A cross-band-gap excitation generates a number of free

electrons in the conduction band and holes in the valence band of the crystal. The free holes

will be converted to self-trapped holes instantly, which can move at RT in the host. Some of

them can reach activator centers and participate in their excitation processes. Others can

capture free electrons from conduction band, creating STEs. The STEs can decay radiatively

or transfer their energies to the activators. However, after relaxation, most STEs have enough

energy to populate the 3P0 state, but not enough for feeding the 1S0 level. As a result, the

luminescence from 3P0 is excitated more efficiently than 1S0 under band-to-band excitation.

This mechanism of ET from host lattice to Pr3+ is referred to as STE-mediated ET, as

illustrated in Fig. 12. Similar mechanism has been found in Ce3+-doped scintillator materials

upon excitation with ionizing radiations [283-285]. In fact, the STE-mediated ET exists in

almost all Pr3+-doped materials including fluorides [62,263,286,287], aluminates

[113,256,266,282], borates [266], and sulfates [67]. Upon host lattice excitation, the STE

-mediated ET will result in a preferential ET from the host to lower-lying Pr3+ 4ƒ2 levels

38
ACCEPTED MANUSCRIPT

instead of 4ƒ5d states. Since QC does not occur, the STE-mediated ET is a highly unfavorable

process which can prevent the efficient selective Pr3+ excitation and then decrease the visible

QE of QC process. As a consequence, these QC phosphors are unsuitable for application in

PDPs and MFFLs.

ET from host to Pr3+ can give rise to visible QC process. QC has been observed

in PbWO4: Pr3+ [288] crystal, which consisted of the 1S0→ 1D2 (~325 nm) and 1D2→ 3H4

(~604 nm) transitions as shown in Fig. 13. The efficient ET between the Pr3+ ion and the host

crystal converted the UV emission from the 1S0→ 1D2 transition into the visible emissions

from the 3PJ (J = 0, 1) and 1D2 multiplets. Fig. 14 shows a schematic diagram of the ET

processes. Upon 204 nm excitation, the Pr3+ ions are excited to the 4ƒ5d configuration then

populate the 1S0 multiplet through a fast nonradiative relaxation. Efficient ET from Pr3+ ions

to the tungstate anions WO42– takes place (ET1 in Fig. 14), resulting in a weak emission from

the 1S0→ 1D2 transition and populated lower-lying 1D2 level of Pr3+. Subsequently, the excited

tungstate anions [WO42– ]* transfer the energy back to the ground state of Pr3+ ions through

another ET process (ET2 in Fig. 14), and the Pr3+ ions populate the 3PJ (J = 0, 1, 2) and 1D2

states. Finally, intense emissions from the 3PJ (J = 0, 1) and 1D2 states of Pr3+ ions would

occur in the visible range of 490–650 nm. Thus visible QC involving two–photon emission in

the range of 490–650 nm could be realized for PbWO4:Pr3+ crystal under an UV or a VUV

excitation.

Research on Pr3+-doped materials under X-ray excitation has been recently

performed in a variety of host materials [63,64,289], and QC has been observed in SrB4O7,

SrAl12O19, SrA1F5, and LiSrA1F6 under X-ray excitation [289]. Srivastava and Duclos [62]

39
ACCEPTED MANUSCRIPT

reported identical emission spectra of the YF3:Pr3+ both under VUV and X-ray excitation at

RT and stated that the STE accounts for the observation of the QC effect under X-ray

excitation. Fig. 15 shows the emission spectra of SrAlF5:Pr3+ under X-ray excitation at T=100

K and 350 K [64]. Obviously, QC takes place at 350 K, whereas emission from the 1S0 level

is absent at 100 K. But emissions originated from the lower-lying Pr3+ levels, such as the 3P0

and 1D2 levels, have clearly been observed at the both temperatures. The broadband emission

around 450 nm obtained at 100 K could be assigned to emission of the STE. ET from the STE

to the Pr3+ ions resulted in emissions from the 3P0 and 1D2 levels at T=100 K. This ET was

more efficient at a higher temperature. The QC process in SrAlF5:Pr3+ exhibited also a

temperature-dependent behavior. This is attributed to the electron in the conduction band,

which is first trapped before it transfers its energy to Pr4+. This ET mechanism was also found

in other Pr3+-doped materials upon X-ray excitation [63,289]. It is shown that there is a strong

correlation between exciton emission and the intensity of the 3P0→ 3H4 emission. The exciton

emission is usually suppressed by the high Pr3+ concentration and temperature, in this case the

3
P0 level is fed by the 1S0→ 1I6 transition.

In summary, we conclude that the observation of QC in Pr3+-doped materials

requires: (1) the lowest 4ƒ5d state above the 1S0 state; (2) emissions from the 3P0 and 1D2

levels usually encounter concentration quenching via CR process; (3) the simultaneous

observation of the 4ƒ2→ 4ƒ2 line emission and the 4ƒ5d→ 4ƒ2 broadband emission originate

from different Pr3+ sites or thermal population process; (4) host absorption induced

STE-mediated ET can reduce the visible QE of QC phosphors; (5) cascade emission depends

on host lattice, temperature and Pr3+ concentration as well as excitation source, and (6) QC

40
ACCEPTED MANUSCRIPT

can also occur under an X-ray excitation.

3.1.2. Pr3+ -doped quantum cutting materials: fluorides and oxides

The efficiency of the PCE strongly depends on the ion-host interactions. In order

to observe PCE or QC instead of the 4ƒ5d→ 4ƒ2 emission, the 4ƒ5d states of Pr3+ must have a

higher energy than the 1S0 state. However, it is difficult to calculate the position of 4ƒ5d

levels because the properties of the mixed 4ƒ5d configuration depend on the various electron

interactions including contributions of crystal field splitting, as well as the effect of covalency,

chemical binding, ligand polarization, and anisotropy of the luminescence centre [63].

To obtain high energy 4ƒ5d states and for the occurrence of QC effect from Pr3+,

two conditions need to be met: (i) a high centroid energy, namely a small centroid shift, and

(ii) a small crystal field splitting [63,67,68]. The first condition is characteristic of compounds

with a high electronegativity of anions [63]. F– ion is preferential in this case because of its

high electronegativity. Furthermore, centroid shift tends to increase in a manner from fluoride

compounds to the sulfate, carbonate, phosphate, borate, silicate, and aluminate compounds

respectively [67]. Thus, fluorides are most likely to exhibit QC effect. In fact, the first

observation of the Pr3+ 1S0 luminescence takes place from Pr3+-doped LaF3 [254].

The magnitude of the crystal field splitting depends strongly on the shape and

size of the anion coordination polyhedron around the Pr3+ ion [67]. Recently, Dorenbos have

discussed on the relationship between the size and shape of the coordination polyhedron and

the energy of the 5d crystal-field states of trivalent RE ions [290,291]. It was found that

cuboctahedral, fivecapped, and tricapped trigonal prism coordination always appear to

41
ACCEPTED MANUSCRIPT

produce small crystal field splitting whereas cubal and octahedral type coordination yield 2 to

3 times larger crystal field splitting. The former three polyhedral coordinations therefore are

favorable for observing QC phenomena. A higher coordination number would result in a

more elevated 4ƒ5d configuration and a reduced splitting of the energy levels. Clearly, hosts

with large coordination number are desirable for the occurrence of QC. For instance, the

aluminate SrAl12O19 with high coordination number (12) supports QC [59]. A large radius of

the substituted lattice site leads to a small crystal field splitting and therefore favors a cascade

emission. In fact, QC has been observed for Pr3+-doped BaSO4 and SrSO4 with Pr3+ on large

metal ion sites like Ba2+ and Sr2+ [67]. In addition, it was found that a small radius and high

charge of a second cation presented in the host supports QC. For example, QC has been

observed in LaZrF7:Pr3+ where the size of Zr4+ is small [286]. This observation was explained

by the nephelauxetic effect [292]. The second cation interacts strongly with the nearest

neighbor Pr3+ ion and then gives rise to a small centroid shift which leads to QC.

Hosts with low photon energy always generate cascade emissions. As is

considered, fluorides are promising hosts. Aluminates with low phonon frequencies are also

of interest. Both steps of cascade emission have been observed in SrAl12O19 because of its low

phonon energy [59]. In contrast, borates are not convenient materials for QC, because they

have high phonon energy, which leads to the absence of emission from the 3P0 level due to the

efficient 3P0→ 1D2 multiphonon relaxation [60,61]. In a word, hosts with weak crystal field,

low phonon energy, and large coordination number for the substitution site are the potential

materials with QC effect of Pr3+.

In addition, possible quantum cutters can be found based on the comparison of

42
ACCEPTED MANUSCRIPT

the spectroscopic properties of Pr3+ and Ce3+ [63,67,68,292]. It is well established that there is

a constant energy difference between the 4ƒ→ 5d transition energy of Ce3+ and that of the

4ƒ2→ 4ƒ5d transitions of Pr3+, which is independent on the type of host lattice [65,66]. The

energetic position E(Pr3+, 4ƒ5d) of the lowest 4ƒ5d level of Pr3+ can be calculated from the

lowest energy E(Ce3+, 5d) of the 5d level of Ce3+ as follows [65,66]:

E(Pr3+,4ƒ5d) = E(Ce3+,5d) + 12 240 cm−1 ± 750 cm−1 (6)

By shifting the energy scale of the Ce3+ spectrum relative to the Pr3+ scale using this equation,

the similarities between the two spectra can be revealed and the corresponding crystal-field

components of the Pr3+ 4ƒ5d states can be identified.

Recently, van der Kolk et al. [68] have made a reliable prediction whether PCE

of Pr3+ can be expected in the same host lattice based on the excitation energy of 4ƒn→

4ƒn-15d of Ce3+ and 4ƒ7→ 4ƒ65d of Eu2+. In Fig. 16, the lowest 4ƒn→ 4ƒn-15d transition energy

ΔE(Ln, A) of Ce3+, Pr3+ and Eu2+ for a number of different host lattices A, is plotted as

function of the lowest energy 4ƒn→ 4ƒn-15d transition of Ce3+, ΔE(Ce3+, A). The horizontal

lines show the energy of the 1S0 state of Pr3+ (upper) and the 6P7/2 state of Eu2+ (lower). Note

that the intersection of these lines with the corresponding 4ƒn–15d states is accidentally at the

same energy (about 35 000 cm–1), which means that all Eu2+-doped hosts showing 6P7/2 line

emission also show 1S0 emission when doped with Pr3+. Such hosts are therefore expected to

be potential quantum cutters. Similarly, all Ce3+-doped hosts on the right side of the vertical

line (at about 35 000 cm–1) should be considered as potential quantum cutters when doped

with Pr3+. The Ce3+-doped and Eu2+-doped host lattices selected in this way are listed in Table

2 and 3, respectively. All these hosts are expected to show 1S0 emission when doped with Pr3+.

43
ACCEPTED MANUSCRIPT

Indeed, 1S0 emission had been observed in BaSiF6 [68] which were selected on the basis of

the observed Eu2+ line emission.

The doping concentration of Pr3+ is usually less than 1%, so that there will not

be any luminescence quenching of the 3P0 and 1D2 levels. Both the 3P0 and 1D2 levels could be

de-excited by nonradiative CR processes such as (3P0, 3H4)→ (1D2, 3H6), (3P0, 3H4)→ (3F3,

1
D2), (3P0, 3H4)→ (1G4, 1G4), (1D2, 3H4)→ (3F4, 1G4), resulting in a strong reduction in QE.

Therefore, low Pr3+ concentrations are required in order to keep the 3P0 or 1D2 emission

efficient. In principle, a low concentration of Pr3+ would not be a serious drawback for the

Xe-excited Pr3+-doped phosphors owing to its strong absorption in the 4ƒ5d levels [293]. The

optimal concentration is between 0.1% and 1% in order to obtain a high absorption efficiency

into the 4ƒ5d bands and simultaneously maintain a substantially high QE of the second step in

the cascade emission (3P0→ 3H4 or 1D2→ 3H4) [265].

Materials synthesized at the nanoscale have been known to possess certain novel

physical properties due to quantum confinement [294]. Loureiro et al. [196] recently reported

the observation of QC effect in nanocrystalline SrAl12O19:1%Pr,Mg phosphor and compared it

with the microcrystalline sample. The results show that luminescence properties in

microcrystalline samples at low temperatures and at RT are maintained at the nanoscale and

that no property degradation due to impurity or surface-induced loss mechanisms commonly

reported for other nanocrystalline systems was found [295-303].

3.1.3. Visible quantum efficiency

QE of a luminescent material is defined as the ratio of the number of photons

44
ACCEPTED MANUSCRIPT

emitted to the number of photons absorbed. QE has long been used as a criterion for the

selection of luminescent materials for applications in fluorescent lighting and PDPs [304-309].

However, QE measurements itself is a complicated work since there are many sources for

errors those should be considered [310-312]. Recently, Rohwer and Martin [313] developed a

measurement system and mathematical procedure for determining the absolute QE of

luminescent materials, but it seems to contain technical deficiencies and literature

inconsistencies and omissions [314]. Moreover, this technique is only applied to fluorescent

laser dyes and conventional phosphors. The major problem in the QE measurement

techniques is the lack of an acceptable calibration standard [313].

Here, for Pr3+-doped QC phosphors, an important factor with respect to

applications is the visible QE which is defined as the number of emitted photons in the visible

spectral range divided by the number of excitation photons [292]. The visible QE in principle

can be determined experimentally, but due to experimental difficulties, especially in the VUV

range, this is not an easy task and hence it has not been a common procedure. Furthermore,

especially in the first stages of the development of a new compound, the measurements might

be disturbed by impurities and contamination of the compounds. QE of QC phosphors are

usually be determined by using synchrotron radiation since the excitation wavelength can be

freely tuned over a wide energy range [315]. Kuck et al. [265] determined experimentally the

visible QE for YF3:Pr3+ by measurements of the total photon flux. The measurements was

performed with synchrotron radiation, and sodium salicylate (C7H5NaO3) was chosen as a

reference. The QE of sodium salicylate was assumed to be 50% in the VUV region [316]. The

absolute visible QE for YF3:Pr3+ as great as 128% was obtained, however, with a large

45
ACCEPTED MANUSCRIPT

uncertainty mainly caused by the sodium salicylate sample [265]. Moine et al. [307] recently

showed that the luminescence efficiency of sodium salicylate is not constant in the VUV

spectral region. Other sources of error might be alignment of the samples, the age and the

storage condition of the samples, the status of sample degradation, the condition of surface,

and so on [265].

J-O intensity theory has now been an essential method for predicting visible QE

of Pr3+-doped materials [59,78,156,261,262,317], since Pappalardo [72] successfully

employed J-O theory to calculate QE of LaF3:Pr3+. If one takes into account the reduced

2
matrix elements square U λ and oscillator strengths for all transitions involved in the QC

process, it follows that the ratios of J-O parameters Ω2/Ω4 and Ω4/Ω6 determine the visible QE.

Smaller are these ratios, higher is the branching ratio of the 1S0→ 1I6 transition and thus a

visible QE. The Ω2/Ω4 and Ω4/Ω6 can be obtained by fitting the experimentally observed

branching ratios from the emission spectra of Pr3+-doped materials. However, the validity of

the J-O calculations for the Pr3+ ion is questionable [78,292,318]. The J-O theory explains the

intensities of the forbidden 4ƒ2→ 4ƒ2 transitions by taking into account the admixture of

configurations of opposite parity 4ƒ5d configurations into the 4ƒ2 configuration. It is assumed

that the proximity of the 5d levels has a strong influence on the transition probabilities and

energy-level wavefunctions. Thus the approximations used in the J-O theory might be found

no longer valid. Because the 1S0 is energetically very close to the 5d levels, it will have a

strong influence on these 5d levels. As a result, a discrepancy between theory and experiment

will be brought about. For example, the visible QE for YF3:Pr3+ was calculated to be about

157% by using J-O theory [72], which is larger than that of about 140% obtained

46
ACCEPTED MANUSCRIPT

experimentally by Piper et al. [56]. This is because that the calculation method does not

consider the losses of energy due to impurities and other nonradiative processes that can

lower the actual QE in a real luminescent material [78].

The visible QE calculated from J-O theory and the branching ratio of the 1S0→

1
I6 emission, strongly depends on the host. A small radius of the second cation, a large

substituted lattice site, and a high coordination number may give rise to a high visible QE

[292]. J-O calculations show that for Pr3+ it is possible in theory to achieve a visible QE of

199% [71]. However, the observation of PCE in a Pr3+-activated material does not necessarily

imply a QE of greater than unity. For example, the visible QE of SrAl12O19:Pr3+ [59] was

expected to be lower than unity, which was attributed to the dominated UV emission (1S0→

1
G4, 1S0→ 3F4) from 1S0 state. Moreover, the 1S0→ 1I6 transition is the most important step in

the PCE process. This transition has a branching ratio of about 80% in YF3:Pr3+ [56].

Nevertheless, the emission corresponding to this transition (~405 nm) cannot be directly used

for applications due to the inadequate color-rendering index [3]. Furthermore, it has been

found that as a result of the excitation of the hosts, the host absorption induced STE-mediated

ET could significantly reduce the visible QE of Pr3+-doped QC phosphors [67]. Thus, an

efficient QC could not be possible from singly Pr3+ doped materials.

3.2. Tm3+ ion activated phosphors

The Tm3+ ion is a possible candidate for QC due to the favorable energy

distribution of its excited electronic levels. Fig. 17 presents the energy level scheme of Tm3+

[319]. Because Tm3+ and Pr3+ lie in the same RE series, the general energy level distribution in

47
ACCEPTED MANUSCRIPT

Tm3+ therefore can be roughly viewed as that of Pr3+, expanded in energy by a factor of two,

due to the doubled strength of the spin-oribit interaction changing from Pr3+ to Tm3+ [72].

However, different from Pr3+, the 1S0 level of Tm3+ which is located in the far UV at 75000

cm–1 is not available for QC because it is severely interfered by the lower 5d levels. As a

consequence, the energy level scheme of Tm3+ offers the possibility to return from the 3P2

excited state via the 1G4 state to the ground state, giving emission of two visible photons [72].

Such a QC effect has been observed in LaF3:Tm3+ with visible QE less than 50% [72]. The

visible QE is greatly lowered by IR emissions due to the multiphoton relaxation 3F2,3→ 3F4.

Tanner et al. [320] recently reported the QC effect in Cs2LiTmCl6 under the 476.5 nm argon

ion laser excitation which populates the 1G4 levels of Tm3+. Visible emission from this level

was not observed due to CR quenching effect and only NIR emission in the range of

12534–10991 cm–1 (798–910 nm) was observed. The higher-energy emission was assigned to

the 3F4→ 3H6 transition of Tm3+ while the lower-energy emission was attributed to the 4F3/2→

4
I9/2 transition of the trace Nd3+ impurities. Thus this is a DC system via CRET between Tm3+

and Nd3+.

3.3. Er3+ ion activated phosphors

The Er3+ ion has also got an energy level structure which could be found suitable

for QC in the visible region, as shown in Fig. 18. QC has been observed in Er3+-doped LiYF4

and LaF3, however the most intense emissions were located at the UV region of the spectrum

[77]. Based on the J–O theory, the maximum visible QE of Er3+-doped LaF3 was found to be

112% [78], which is not sufficient for a Xe-discharge lamp phosphor, due to a significant part

48
ACCEPTED MANUSCRIPT

of the emission intensity is located in the UV region. Nevertheless, if a codopant can

efficiently convert this UV emission into visible light, efficient QC phosphors codoped with

Er3+ and RE3+ might be possible in a potential way.

3.4. Gd3+ ion activated phosphors

Wegh et al. [34] in 1997 made a systematic spectroscopic study of the 4ƒ7 energy

levels of Gd3+ in LiYF4 at the VUV spectral region (50 000–70 000 cm–1). Fig. 19 depicts the

schematic energy level diagram of Gd3+ showing the possibilities for QC. The large energy

gap of 8000 cm–1 between the 6G7/2 level and the next lower 6DJ levels leads to low

multiphonon-relaxation rate and hence to radiative decay. QC was indeed observed in

LiYF4:5% Gd3+ upon the 8S7/2→ 6GJ excitation when a red emission due to the 6GJ→ 6PJ

transition (~600 nm) is followed by a UV emission due to the 6PJ→ 8S7/2 transition (~313 nm)

[34]. However, the visible QE is expected to be low because the second photon corresponding

to the 6PJ→ 8S7/2 transition (~313 nm) is located in UV region, and a Gd3+ ion has several

possibilities to decay from the 6GJ levels, of which the 6GJ→ 6PJ transition is the only

emission in the visible region.

QC of Gd3+ ion has also been observed in GdBaB9O16 [73], ScPO4 [74,75], and

Na(Y,Gd)FPO4 [76]. The typical cascade emissions from the transitions of 6GJ→ 6PJ (~600

nm) and 6PJ→ 8S7/2 (~313 nm) are clearly observed in ScPO4:1% Gd3+ as shown in Fig. 20.

The QC mechanism is schematically presented in Fig. 21. After the creation of the e-h pair by

the absorption of the VUV photon in an above band gap transition, a STE is formed. When

Gd3+ is introduced, an ET from the STE to the excited states of Gd3+ takes place, as shown in

49
ACCEPTED MANUSCRIPT

Fig. 21. The replacement of the STE emission with the 6GJ emission of Gd3+ strongly suggests

a STE→ Gd3+ ET process involving the 6GJ state. This ET process is known as host

sensitization of Gd3+ [321-324]. It has been demonstrated that the ET from the STE to Gd3+ is

thermally activated probably due to exciton mobility. When at low temperatures, ET from the

STE is dominated by transfer to killer centers, whereas at RT the transfer to Gd3+ effectively

dominates the ET. The experimental results are well described with a model that assumes two

STE states split by 280 cm−1, a lower triplet and an upper singlet, whose radiative rates differ

by about 2 orders of magnitude. The comparison with the model yields a thermal activation

energy of 970 cm−1. To use STE sensitization of the 6GJ state of Gd3+ for QC, it will be

necessary to identify materials for which STE emission occurs at even shorter wavelengths

than in the case of ScPO4 so that a larger fraction of the ET occurs to the 6GJ state or even

higher-lying states of Gd3+. However, the efficiency of ET to the 6GJ state is only about 30%,

resulting in a low QE of about 92%, which is far beyond the desired 200%. Therefore,

materials based on the emission of Gd3+ alone might not be found suitable as visible quantum

cutter. If the UV emission due to the 6PJ→ 8S7/2 transition (~313 nm) can be transferred

efficiently to another ion with visible emission, then Gd3+-doped materials could be

potentially more efficient QC phosphors.

50
ACCEPTED MANUSCRIPT

4. QC in RE dual/ternary ions activated phosphors

4.1. ET and DC

The possibility of achieving two-photon emission via ET between ions was

predicted by Dexter [10] in the 1950s. This process is the opposite of the Addition de Photons

par Transfert d’Energie (APTE) UC mechanism discovered by Auzel [80] in 1966, and it is

called as DC [79]. DC is a new route to realize visible QC effect. Fig. 3(b-d) illustrates

generalized energy level diagrams for three DC mechanisms involving ET between two

different RE ions (I and II). Type I is an ion for which emission from a high-lying level

occurs. Type II is an activator ion to which ET takes place. Fig. 3(b) indicates two photon

emission from ion pairs by CR from ions I to II (denoted by ①) and ET from ions I to II

(denoted by ②) with emission from ion II. Fig. 32(c,d) shows a CR mechanism followed the

emission of photons from both ions I and II. In all three cases, if the two-step ET process is

efficient, a theoretical visible QE of 200% can be achieved because the previous IR and UV

losses existing in a single ion can be avoided. DC including the first-order DC and

second-order DC are found in Gd3+ and Pr3 as well as Tb3+-activated phosphors and are

discussed below.

4.2. Gd3+-involved phosphors

Phosphors based on a single Gd3+ ion can not be used as a visible quantum

cutter as discussed in the aforementioned Section 3.4. However, if the UV emission due to the

6
PJ→ 8S7/2 transition (~313 nm) can be transferred efficiently to another ion along with

emitting visible light in Gd3+-activated materials, they could be applied in MFLLs and PDPs.

51
ACCEPTED MANUSCRIPT

RE ions, such as Eu3+ and Tb3+, have been considered useful in visible QC phosphors due to

their efficient visible emissions. In recent times, a great deal of investigations have been

undertaken in evaluating possibilities in obtaining better QC upon using some other RE3+ ions

(RE= Eu, Tb, and Er) as the co-dopants by modifying the emission properties of Gd3+-based

QC phosphors. The design of QC phosphors requires appropriate combinations of R13+–R23+

or R13+–R23+–R33+ (R1, R2, and R3 are RE ions) as activator ion-pairs or couples and relevant

VUV spectral data. Visible QC through DC process has successfully been observed in

Gd3+–Eu3+ [79], Gd3+–Tb3+ [91], and Gd3+–Er3+–Tb3+ [82] co-doped fluoride-based systems.

In particular, the systems of nMF–mGdF3 (M=Li [79,82,97,98,193], Na [83,84], K

[84,87,88,91], Rb [85], and Cs [89]) and xAF2–yGdF3 (A being a group-II element) [86,92,95]

have currently been of great interest as effective host-matrices to realize QC. Investigations

on realizing visible QC in Gd3+–Tm3+ [325], Gd3+–Er3+ [325], Dy3+–Tb3+ [115,116],

Gd3+–Er3+–Dy3+ [158], and Gd3+–Tm3+–Dy3+ [326] systems and the ET mechanisms involved

have also been made over the past several years, however, the results are still not successful.

4.2.1. Gd3+–Eu3+ dual ions activated phosphors

4.2.1.1. Materials, PL and QE

Wegh et al. [79,81] in 1999 reported visible QC in LiGdF4:Eu3+ through DC

with an interesting and appreciable QE as great as 190%. The powder sample of LiGdF4:Eu3+

was prepared by firing stoichiometric mixtures of LiF, GdF3, and EuF3 at 550 °C in a nitrogen

atmosphere. Fig. 22 shows a schematic energy-level diagram of the Gd3+–Eu3+ system,

illustrating the possibility of visible QC through two-step ET from Gd3+ to Eu3+ upon

52
ACCEPTED MANUSCRIPT

excitation in the 6GJ levels of Gd3+. It is noticed that the energy of the 6GJ→ 6PJ (~590 nm)

transition of Gd3+ matches well with the 7FJ→ 5D0 (~590 nm) excitation energy of Eu3+. Upon

excitation in the 6GJ levels of Gd3+, firstly, a part of excitation energy is transferred by CR

between Gd3+ in the 6GJ state and Eu3+ in the 7FJ state, resulting in Gd3+ in the 6PJ state and

Eu3+ in the 5D0 state. Subsequently, Eu3+ emits a visible photon due to the 5D0→ 7FJ transition.

Secondly, the remaining excitation energy of Gd3+ ion in the 6PJ state is transferred to a high

excited state of another Eu3+ ion through migration via the Gd3+ sublattice. After a fast

relaxation from the high excited state to the 5DJ states, a second visible photon takes place due

to the 5DJ (J = 0, 1, 2, 3)→ 7FJ transition with a normal branching ratio which can be

determined from an emission spectrum upon excitation in the 6IJ levels of Gd3+ (seen in Fig.

23). As a consequence, there will be a significant increase in the relative luminescence

intensity of the 5D0 emission from the 6GJ levels, if the expected CRET process takes place. It

is reported that this two-step ET process is very efficient even at RT [81]. Fig. 24 shows the

emission spectra of LiGdF4:0.5% Eu3+ upon 8S7/2→ 6GJ (202 nm) excitation on Gd3+ at 7 K (a),

50 K (b) and 300 K (c). From the low-temperature (7 K) emission spectrum some extra lines

are observed at 592 nm, 612 nm, 635 nm and 636 nm, which are very weak at 50 K and are

absent at the RT. These are assigned to the Gd3+ 6GJ→ 6PJ emissions. At low temperatures,

although two-step ET does not occur, visible QC takes place: upon excitation in the 6GJ levels

of Gd3+, visible Gd3+ 6G7/2→ 6PJ emission is followed by Eu3+ 5DJ→ 7FJ emission after ET.

However, this visible QC is not very efficient, because the Gd3+ 6G7/2→ 6PJ emission intensity

is much smaller than the total Eu3+ emission intensity. The CR step is temperature activated

and becomes efficient at elevated temperatures. Consequently, CR probability such as Gd3+

53
ACCEPTED MANUSCRIPT

(6G7/2) + Eu3+ (7F1)→ Gd3+ (6P7/2) + Eu3+ (5D0) becomes high and can compete very well with

radiative decay from the 6G7/2 state (which is parity forbidden) or direct ET of all Gd3+

excitation energy to a high energy level of Eu3+.

Efficiency of CRET between Gd3+ in the 6GJ states and Eu3+ in the 7FJ states can

be calculated by evaluating the 5D0 and 5D1,2,3 integrated emission intensities using the

following equation [81]:

PCR R( 5 D0 / 5D1, 2,3 ) 6 G − R( 5 D0 / 5D1, 2,3 ) 6 I


= J J
(7)
PCR + PDT R ( 5 D0 / 5D1, 2,3 ) 6 I + 1
J

where PCR is the probability for CR, and PDT is the probability for the direct ET from Gd3+ to

Eu3+, R(5D0/5D1,2,3) is the ratio of the 5D0 and the 5D1,2,3 emission intensities, and the subscript

(6GJ or 6IJ) indicates the excitation level for which the ratio is obtained. This formula is

applicable for all the Gd3+–Eu3+ systems. If an assumption is made that there are no

nonradiative losses due to energy migration to quenching centers (defects and impurities), the

QE will be given by PCR/(PCR+PDT) +1. For LiGdF4:0.5% Eu3+, the efficiency of the CRET

step was determined about 90% using the intensity ratios obtained from the emission spectra,

resulting in a visible QE as great as 190%.

Feldmann et al. [145] had determined the external QE of LiGdF4: Eu3+ phosphors

for the first time, by making use of two independent methods. On the one hand, the external

QE was determined by quantitative comparison of the emission spectra of LiGdF4:Eu3+ and

Y2O3:Eu3+. On the other hand, the external QE of LiGdF4:Eu3+ was determined based on

reflection spectra. Both methods yielded an external QE of 32% upon excitation into the 6GJ

states of Gd3+ (202 nm), although the internal QE is as high as 195%. This is due to the fact

that a substantial amount of the UV photons (202 nm) are not absorbed by Gd3+ but by the

54
ACCEPTED MANUSCRIPT

host lattice itself. The VUV absorption for Gd3+ ions is very weak because the transition

between 4ƒn levels involved in the QC process are parity and spin forbidden, and finally the

luminescence efficiency is less than one. The host lattice absorption resulting in nonradiative

relaxation is probably caused by lattice defects or impurities which might be significantly

reduced by optimizing the synthesis of the material.

Fluorides with wide band gap and low phonon energy are the ideal host lattices

for QC [92,115,116]. Visible QC via DC has so far been identified in many Eu3+-doped

fluorides as summarized in Table 4, since the fact that the 4ƒn energy levels are located at

roughly the same positions in all the fluorides. It is found that QE varies with different host

due to the ET process depends strongly on host lattice. The obtained QE of BaF2 is as great as

194% [86], which is significantly higher than that of KLiGdF5 (140%) [88]. The

sample-synthesis method is another important factor to affect QE. Nonradiative losses in

defects and impurities could deduce QE. Wegh et al. [79] suggested that improvement in

method of synthesis would increase the QE. The complex fluorides were synthesized

conventionally by a solid-state reaction method at high temperatures, however, this approach

might be causing some impurities and defects that could significantly affect luminescence. In

addition, complicated setup under an atmosphere of F2 or HF mixed with an inert carrier gas

to avoid possible contamination from oxygen is necessary for solid-state reaction [90]. For

instance, in GdF3:Eu3+ [81] single crystal grown by Bridgman method, broadband emission

(400–475 nm) assigned to the 4ƒ65d→ 4ƒ7 transition of Eu2+ ions is presented in the emission

spectra due to the reducing atmosphere during the preparation of the sample by using a carbon

crucible, which consequently makes calculation of the efficiency of the two-step ET process

55
ACCEPTED MANUSCRIPT

more complicated than that for LiGdF4:Eu3+. Hydrothermal synthesis is regarded as a

convenient method to prepare fluoride phosphors in high purity [152-157]. Recently, Hua et

al. [90] reported a visible QE of 170% in GdF3:Eu3+ nanocrystals synthesized by a

microemulsion-mediated hydrothermal process. Lepoutre et al. [192-194] showed that sol-gel

synthesis route can also sufficiently improve the efficiencies of fluoride materials. Recently,

Hachani et al. [227] investigated the opportunity of QC effect for Gd3+–Eu3+ dual ions in

phosphates. Unfortunately, no QC process has been observed in these materials because the

position of the charge transfer band (CTB) of Eu3+ overlaps the 6GJ levels of Gd3+, as shown

in Fig. 25. Instead of CR process between Gd3+ and Eu3+, an efficient ET from the CTB of

Eu3+ to Gd3+ 6PJ levels has been observed.

In principle, ET from Gd3+ to Eu3+ in QC is possible at any Eu3+ concentration.

However, the ET process clearly exhibits a strong Eu3+ concentration-dependent behavior

[87,88,93,327]. Takeuchi et al. [327] interpreted this behavior in terms of the fixed distance

between the donor (Gd3+) and the acceptor (Eu3+) in the crystal. The

concentration-dependence ET process can be explained as follows according to Kodama and

Oishi [88]: at a low Eu3+ concentrations, the CR probability between Eu3+ ions is low,

resulting in Eu3+ emission from all excited 5DJ levels of the Eu3+ ion fed by the first step in the

QC process and emission from the 5D0 level only in the second step. Meanwhile, energy

migration over the Gd3+ sublattice in concentrated Gd compounds occurs, providing efficient

ET at low Eu3+ concentrations through the efficient trapping behavior of Eu3+ in energy

migration over the Gd3+ sublattice, whereas, at a higher Eu3+ concentrations, CR takes place

between Eu3+ ions through direct ET from Gd3+ to Eu3+ level at 50 000 cm−1 or direct

56
ACCEPTED MANUSCRIPT

absorption of Eu3+ instead of Gd3+, which prevent CR. Therefore, an optimal Eu3+

concentration must greatly improve the QE for Gd3+–Eu3+ QC system.

Time-resolved luminescence spectra have recently been recorded in order to

analyze the dynamics of QC of Gd3+–Eu3+ dual ions although a great deal of work has been

focused on QC of Gd3+–Eu3+ due to CRET [230]. It was found that upon Gd3+ 6GJ excitation,

a relatively fast increase in the time-resolved signal of the Eu3+ 5D0 emission is expected by

comparing the 5D1 emission intensity, due to the 5D0 level is populated directly by the CR step.

However, the Eu3+: 5DJ emissions show a strong instant signal upon excitation into the 6IJ

levels of Gd3+ in NaGdF4:Eu3+. Subsequently, Vergeer et al. [328] investigated the dynamics

of the Gd3+ 6PJ, Eu3+ 5D1 and the Eu3+ 5D0 emissions for LiGdF4:Eu3+ upon Gd3+ 6PJ and 6GJ

excitation. Analysis of the RT luminescence decay curves shows that the experimental results

can be explained by rate equations based on the DC model. A good agreement has been

obtained between the various decay curves and fitted curves based on the magnitudes of the

parameters such as decay rates and ET rates for the various levels involved.

4.2.1.2. Sensitization

An ideal VUV phosphors should posses both high QE and intense absorption

[1,7,329]. The calculated QE of Gd3+–Eu3+ dual ions based on the DC mechanism is quite

efficient, which can reach a maximum of 190% [79]. However, since only 4ƒ levels are

involved in the process, the VUV absorption is very low due to the transitions between ƒ-ƒ

levels are parity and spin forbidden, and finally the luminescence efficiency is significantly

less than 100% [145]. Thus, a successful QC phosphor based on Gd3+–Eu3+ pair will require

57
ACCEPTED MANUSCRIPT

sensitization of the high energy 6G7/2 state of Gd3+. A so-called sensitizer for which a strong

absorption in the VUV region (147-190 nm), a good overlap of the spectrum in the VUV

region with Gd3+, and be able to transfer the energy efficiently to the Gd3+ ions is necessary

[94,95]. Investigations in this direction have currently been undertaken.

Wegh et al. [82] reported Er3+ as a sensitizer to improve the absorption of Gd3+,

but the QE is only about 110% due to a large amount of nonradiative relaxation losses, so this

material is not a useful QC phosphor. Vergeer et al. [100] reported recently that Pr3+ acts as a

sensitizer, wherein the luminescence of Pr3+ was quenched by Eu3+. Feofilov et al. [96]

reported sensitization of Gd3+ with Pr3+ in GdF3:Pr,Eu. However, the maximum QE was only

20%, falling far below the desired goal of 200%, due to that the ET from Pr3+ to Gd3+ ions

occurs predominantly to the 6IJ state of Gd3+, but not to the 6GJ state. Hirai et al. [45] found

that Pr3+ ions in NaGdF4:Pr3+,Eu3+ can transfer the absorbed VUV excitation energy to Eu3+

ions through three-step ET process: (i) from Pr3+ to Gd3+ ions; (ii) from Gd3+ to Gd3+ ions; and

(iii) from Gd3+ to Eu3+ ions. Solarz et al. [330-333] also reported Pr3+ as a sensitizer of Gd3+

excitation and showed that the concentration of the Pr3+ should be controlled carefully

because Pr3+ ions can also act as luminescence quenchers. Peijzel et al. [94] successfully used

Tm3+ as a sensitizer in LiGdF4:Eu3+,Tm3+ but simultaneously found that the CR step from

Gd3+ to Eu3+ was quenched by more efficient CR process from Gd3+ to Tm3+. A similar

competing CR from Gd3+ to Nd3+ was observed by Jia et al. [97] and Zhou et al. [98] in

LiGdF4:Nd3+. Recently, van der Kolk et al. [334] also used the strong 4ƒ→ 5d absorption

transitions of Nd3+ and Tm3+ to sensitize the 6G7/2 state of Gd3+ in the phosphors NaGdF4:Nd3+

and NaGdF4:Tm3+. In both systems, QC effect was observed which resulted in the emission of

58
ACCEPTED MANUSCRIPT

two IR photons for each absorbed VUV photon, but therefore they are not useful for visible

quantum cutters. Hirai et al. [46,335] reported that Tb3+ ions as sensitizers in

NaGdF4:Tb3+,Eu3+ absorbed the VUV light (170–220 nm) due to the dipole-allowed 4ƒ8→

4ƒ75d transition and efficient ET from Tb3+ to Eu3+ through Gd3+ was indeed observed,

however, whether or not the QC process via the ET from Tb3+ to Eu3+ occurs was unclear.

Babin et al. [95] suggested Pb2+ could serve as a sensitizer for the Gd3+–Eu3+, but it seems that

sensitization of the 6GJ state will be difficult.

More recently, Lee et al. [336] reported K2GdF5:Eu3+ QC phosphor by using Pr3+

as a sensitizer. Results show that Pr3+ ion offers intrinsically effective absorption in the UV

and VUV spectral ranges, and K2GdF5:Eu3+,Pr3+ is red-emitting QC phosphors of which the

theoretical QE was improved from 107% to 138% on codoping Pr3+ as a sensitizer through

mechanisms of CR and direct ET. Fig 26 presents the PL spectra for K2GdF5:Eu3+,Pr3+

phosphors upon excitation at 274, 210, and 172 nm, respectively. It was observed that, upon

excitation on the 4ƒ5d state of Pr3+ at 210 and 172 nm, the relative intensity of emissions due

to Eu3+ 5D1→ 7FJ and 5D0→ 7FJ multiplet transition become significantly more intense in

K2GdF5:Eu3+,Pr3+ phosphor than for K2GdF5:Eu3+, which clearly indicates that Pr3+ might act

as an effective sensitizer for the phosphor system containing Gd3+–Eu3+–Pr3+. Pr3+

concentration dependent QE was calculated, and the theoretical QE was found to be

increasing from 107% for K2GdF5:5%Eu3+ to an optimal value 138% for

K2GdF5:5%Eu3+,1%Pr3+ before Pr3+ reaching concentration quenching threshold under

excitation at 210 nm. Fig. 27 provides the energy level diagrams of Gd3+, Eu3+ and Pr3+ ions,

showing the possible mechanisms for QC and ET in the K2GdF5:Eu3+,Pr3+ phosphors. Under

59
ACCEPTED MANUSCRIPT

excitation at 274 nm, the Gd3+ ion is excited into the 6IJ state, following nonradiative

relaxation to the 6PJ state subsequently, direct ET might occur from 6PJ to neighboring Eu3+

ions, or a radiative relaxation of Gd3+ from 6PJ to the 8S7/2 ground state occurs, thus no QC is

expected. In contrast, under excitation at 210 nm, Pr3+ pumping dominates when Pr3+ is

excited to the 4ƒ5d state, as the energy released from transitions 1S0→ 3P1 + 1I6 matches and

feeds the excitation transition 7FJ→ 5D3; ET from Pr3+ to neighboring Eu3+ ions might proceed

by CR (step ①), which will result in emission of multiplet transitions 5D0,1,2,3→ 7FJ (J=0–6).

The remaining energy in 3P1 or 3P0 of Pr3+ is subsequently transferred (step ②) directly to

nearby Eu3+ ions, from which emissions assigned to transitions 5D0,1→ 7FJ (J=0–6) occur.

Such ET mechanism involved in K2GdF5:Eu3+,Pr3+ phosphors gives rise to the greatest visible

QE of about 138% for K2GdF5:5%Eu3+,1%Pr3+ [336].

4.2.2. Gd3+–Tb3+ dual ions activated phosphors

Besides visible QC in Gd3+–Eu3+ dual ions, visible QC has also been observed in

Tb3+:K2GdF5 [91] and Tb3+:BaGdF5 [92]. However, in the Gd3+–Tb3+ dual ions doped

phosphors QC can only be observed upon excitation of Tb3+ with VUV photons. Fig. 28

shows emission spectra of K2GdF5:5%Tb3+ with the VUV-UV excitation. It is noted that upon

excitation of Tb3+ at 212 nm or 172 nm, two green photons due to Tb3+ 5D4→ 7FJ are emitted

through a two-step process, i.e. CR and direct ET from one Tb3+ to a neighboring Tb3+ or Gd3+.

However, no such QC effect was observed for UV excitation at 274 nm. The VUV-UV

excitation spectra of K2GdF5:5%Tb3+ monitored by both the 5D4→ 7F5 emission at 542 nm

and the 5D3→ 7F6 emission at 415 nm of Tb3+ were shown in Fig. 29. The strong broad

60
ACCEPTED MANUSCRIPT

excitation bands peaked at 212 and 172 nm in the 140–230 nm region are assigned to the

spin-allowed transitions of 4ƒ8 (7F6) to the low-spin 4ƒ75d states of Tb3+. And the weak

emission peaked at 274 nm was assigned to the transitions from the ground state 8S7/2 to the 6IJ

state of Gd3+. The relative mechanisms are explained in Fig. 30 to rationalize the QC effect

from Gd3+–Tb3+, which indicate that there is no QC occurrence upon excitation with λexc =274

nm (Fig. 30a), while upon excitation with 212 nm or 172 nm of Tb3+, visible QC takes place

via a two-step ET process (Fig. 30b and 30c). The efficiency of CR between Tb3+ and a

neighboring Tb3+ for Gd3+–Tb3+ pair can be calculated using the following equation [91,92]:

PCR R( 5 D4 / rest ) Tb 3+ − R( 5 D4 / rest ) Gd 3+


= (8)
PCR + PDT R ( 5 D4 / rest ) Tb3+ + 1

Here, PCR represents the probability for CR and PDT is the probability for direct ET. R(5D4

/rest) is the ratio of photoluminescence intensity of 5D4 to that attributed to 5D3 of Tb3+ and

6
P7/2 of Gd3+; the subscript indicates excitation from Tb3+ or Gd3+. Eq. (8) was modified from

the aforementioned Eq. (7). For K2GdF5:11%Tb3+, the calculated visible QE was found to be

189% and 187% for VUV excitations at 212 and 172 nm, respectively. In the case of

BaGdF5:15%Tb3+, the calculated QE was found to be 168% and 180% upon excitation with

215 and 187 nm, respectively. The high QEs of K2GdF5:Tb3+ and BaGdF5:Tb3+ phosphors

make them very promising for application in PDPs and MFFLs [91].

4.2.3. Gd3+–Er3+–Tb3+ ternary ions activated phosphors

Wegh et al. [82] in 2000 reported visible QC of Gd3+–Tb3+–Er3+ system in

LiGdF4 lattice. The simplified energy level diagram of Gd3+–Tb3+–Er3+ system has been

shown in Fig. 31. In this system QC takes place on Er3+ upon excitation in the 4ƒ105d levels

61
ACCEPTED MANUSCRIPT

which are situated in the VUV. After nonradiative relaxation to the lowest 4ƒ105d level,

energy can be transferred by CR, exciting Gd3+ into the 6PJ, 6IJ or 6DJ states and leaving Er3+ in

a level just above the 4S3/2 level. Subsequently, a green photon is emitted efficiently due to the

4
S3/2→ 4I15/2 transition of Er3+. The energy transferred to the Gd3+ will be reaching to the

Gd3+-sublattice and then to the Tb3+ ion. After ET from Gd3+ to Tb3+ the characteristic blue or

green Tb3+ emission due to 5DJ→ 7FJ transition emits. Fig. 32 shows the emission spectra of

LiGdF4:Er3+, Tb3+ (1.5%, 0.3%). The efficiency of this QC process can be estimated by

measuring the increase in the 4S3/2 emission intensity under VUV excitation (when QC can

occur) compared to the emission spectrum under UV excitation (no QC and no ‘extra’ 4S3/2

emission). The QC efficiency can be calculated using the following expression [82]:

PET + R( 4 S 3 / 2 / rest ) Er 3+ − R( 4 S 3 / 2 / rest ) Gd 3+


= (9)
PET + + PET − R( 4 S 3 / 2 / rest ) Gd 3+ + 1

Here, PET+ and PET- are the probabilities of the desired and the undesired ET possibilities from

the 4ƒ105d states of Er3+ to Gd3+, respectively. R ( 4 S3/ 2 / rest ) Er 3+ and R ( 4 S3/ 2 / rest )Gd 3+

are the intensity ratios of Er3+ 4S3/2 emission to all remaining emission upon Er3+ ƒ→ d

excitation and upon Gd3+ 6IJ excitation, respectively. Eq. (9) is similar to the aforementioned

Eq. (7) for calculating the QE of Gd3+–Eu3+ ion pair. If the nonradiative losses due to energy

migration at the defects and impurities are ignored, the QE then can be given by PET+ /( PET+ +

PET-) +1. With help of the spectra of Fig. 32 and Eq. (9), the efficiency for the CR steps of

LiGdF4:Er3+, Tb3+ (1.5%, 0.3%) was calculated to be 30%, resulting in a maximum QE of

130%.

4.3. Pr3+-involved phosphors

62
ACCEPTED MANUSCRIPT

Ever since the first demonstration of QC phenomenon in the deep blue region

for lanthanide ions Pr3+ doped fluorides in the early 1970s [56-58], intense level of research

activity has provided a special focus on RE doped VUV QC phosphors. So far, QC has been

widely witnessed in a variety of RE-doped phosphors. However, none of Pr-doped QC

phosphors was proved to be useful for practical application because of the violet emission

from the 1S0→ 1I6 transition resulted in a low color rendering index [7,117]. A promising

approach has been employed to deal with this problem with an appropriate co-dopant, which

can convert the energy of the first step violet photon into a proper visible photon through ET.

The requirements for such a transfer partner ion in principle are [138,262]: (i) it should have a

strong transition around 400 nm, i.e. in resonance to the 1S0→ 1I6 transition of the Pr3+; (ii) it

should emit predominantly in the visible spectral range with high sensitivity of the human

eyes, and (iii) the whole energy level scheme of the co-dopant ion should not have any energy

levels interfering with the Pr3+ energy level scheme, whereas, the cascade emission could be

affected or even prevented. Intensive investigation has recently been performed in many

systems [100-114] in order to find suitable co-dopants and excellent hosts to improve the

color rendering index of Pr3+ emissions.

4.3.1. Pr3+–Eu3+ dual ions activated phosphors

For the Pr–Eu dual ions, efficient ET through CR process of Pr3+: 1S0→ 1I6 and

Eu3+ : 7F0,1→ 5D3, 5L6 is expected since spectral overlap between the Pr3+ emission and the

Eu3+ absorption lines exist at 400 nm as shown in Fig. 33. However, no red/orange emission

from Eu3+ was observed upon excitation in the 4ƒ5d bands of Pr3+ [99], although there is a

63
ACCEPTED MANUSCRIPT

favorable spectral overlap.

To determine whether the absence of Eu3+ emission is due to the low ET rates,

the critical distance between Pr and Eu ions in YF3: Pr3+, Eu3+ was calculated and was found

to be 0.40 nm [100], which is similar to the distance between nearest-neighbors in the YF3

host lattice (0.359 nm). It indicates that for nearest neighbor pairs of Pr–Eu the transfer rate is

comparable to the radiative decay rate from the 1S0 level. Therefore ET and subsequent Eu3+

emission should occur. To investigate the observed absence of Eu3+ emission upon excitation

in the 1S0 level of Pr3+, luminescence spectra was measured for YF3: 1%Pr3+, x%Eu3+ (x = 0, 5

and 10) as shown in Fig. 34. It is noted that the emission intensity of Pr3+ has significantly

been decreasing due to the presence of Eu3+ ions. The intensity of the Pr3+:1S0 emission has

reduced to its 10% with 5% Eu3+ doping by comparison with YF3: Pr3+ sample. Experimental

variations in the measured luminescence intensity however cannot explain the observed

strong reduction in intensity. Meanwhile, with incorporation of Eu3+, the decay lifetime

decreases very fast. The changes in the time-resolved signal can be explained by the

introduction of nonradiative decay when incorporating Eu3+ into the host. Both the emission

spectra and the time-resolved luminescence measurements reveal that Eu3+ ions act as

quenching centers for the Pr3+:1S0 luminescence. As a result, the interaction between Eu3+ and

Pr3+ does not lead to efficient Eu3+ emission. This quenching process is attributed to

nonradiative relaxation via a Eu2+–Pr4+ metal-to-metal charge-transfer state (CTS).

4.3.2. Pr3+–Er3+ dual ions activated phosphors

Er3+ is considered as a good codopant because of its strong green emission and

64
ACCEPTED MANUSCRIPT

some of its absorption bands overlap with the Pr3+ 1S0 emissions. Temperature- and

time-dependent ET processes in Pr3+ and Er3+ codoped CaAl12O19 crystal from 12 to 290 K

has been investigated [101]. Fig. 35 shows the excitation spectrum of the 4S3/2→ 4I15/2

emission of Er3+ (dashed line) and emission spectrum (solid line) of CaAl12O19:Pr3+, Er3+

under the excitation of 205 nm. It was found that the most of the emission lines come from

the transitions of Pr3+ except in the 510–560 nm range, where the emissions from Pr3+ and

Er3+ were overlapped. Clearly, emissions centered at 253 nm, 273 nm, 400 nm, 486 nm and at

the 520–550 nm regions, assigned to the transitions of 1S0→ 3F4, 1S0→ 1G4,1S0→ 1I6, 3P0→

3
H4, and 3P1/3P0→ 3H5 of Pr3+, overlapped with the Er3+ excitations 4I15/2→ 4D7/2, 4I15/2→ 2H11/2,

4
I15/2→ 2H9/2, 4I15/2→ 4F7/2, and 4I15/2→ 2H11/2/4S3/2, respectively. ET from Pr3+ to Er3+ ions has

been observed with the ET efficiency of only about 25% [101], indicating the ET can partially

convert the UV emission (1S0→ 1I6) of Pr3+ into the characteristic green emission of Er3+.

However, since both Pr3+ and Er3+ ions replaced Ca2+ ions into the lattice, the large

donor–acceptor separation results in a small interaction, leading to a small ET rate.

4.3.3. Pr3+–Mn2+ dual ions activated phosphors

Mn2+ ion has been recognized as a suitable co-doped ion for Pr3+ because it has

many absorption transitions within the 3d5 configurations, which likely to coincide with the

Pr3+ emission and also because it exhibits efficient visible emission from green to red. Fig. 36

shows the energy level scheme of Pr3+ and Mn2+. It is clear that energy of the transitions of

1
S0→ 1I6, 3PJ (~405 nm) and 1S0→ 1D2 (~340 nm) of Pr3+ overlap very well with that of the

transitions of 6A1g→ 4Eg, 4A1g (~410 nm) and 6A1g→ 4T2g (~350 nm) of Mn2+, respectively,

65
ACCEPTED MANUSCRIPT

providing the conditions for the ET from Pr3+ to Mn2+. If this ET indeed occurs upon the

excitation of Pr3+:4ƒ5d band, then the Mn2+ related red emission due to the 4T1g→ 6A1g

transition will be arising along with the second photon from Pr3+ in the visible region, leading

to visible QE larger than unity.

A USA patent stated that QC can be achieved in Pr3+–Mn2+ co-doped SrAl12O19

[103], however, since there is not spectral overlaps in SrAl12O19:Pr3+, Mn2+ system, efficient

ET may not takes place [337]. van der Kolk et al. [104] has investigated the ET processes of

Pr3+–Mn2+ dual ions in SrAlF5, CaAlF5, and NaMgF3. No efficient ET from Pr3+ to Mn2+ could

be obtained in SrAlF5 due to the weak Mn2+ emission [104], although there is a favorable

spectra overlap between the 1S0→ 1I6 Pr3+ emission and the 6A1→ 4Eg, 4A1g Mn2+ absorption,

Similar results have been obtained in Pr–Mn co-doped KMgF3 [105], SrY2F8 [106], YF3 [106],

CaF2 [106], LiBaF3 [106], and SrB6O10 [108]. The absence of efficient ET from Pr3+ to Mn2+

might be due to the selection rules, or the total spin of the system change, or there is no empty

excited state with suitable symmetry near the 1S0 state, which becomes the highest excited

state within the f–manifold [71]. These compounds therefore seem not to be a suitable host

material with respect to the realization of the Pr3+–Mn2+-based VUV phosphor.

It should be mentioned here that visible QC due to CRET for Pr3+–Mn2+ dual

ions was recently successfully identified in SrB4O7 [107,108] and LaMgB5O10 [109,110]. Fig.

37 shows emission and excitation spectra of SrB4O7:Pr3+, SrB4O7:Mn2+, and SrB4O7:Pr3+,

Mn2+. The emission from both the transitions 1S0→ 1I6, 3PJ (~400 nm) and 1D2→ 3H4, 5 (~600

nm, 675 nm) are clearly observed in the emission spectrum of SrB4O7:Pr3+, indicating the

cascade emission occurs. The emission from 3P0 level was not observed due to the large

66
ACCEPTED MANUSCRIPT

energy phonons available in the borate host (1400 cm−1) which can bridge the energy gap

between 3P0 and 1D2 levels by multiphonon relaxation [61]. The expected spectral overlaps

between the emission spectra of the SrB4O7:Pr3+ sample and the excitation spectra of the

SrB4O7:Mn2+ sample have clearly been observed in the region of 330–430 nm, which

encourages the ET from Pr3+ to Mn2+. Such an ET can be determined by comparing Figs. 37(b)

with 37(d). It is noted that SrB4O7:Mn2+ shows almost no emission from Mn2+. While in

SrB4O7:Pr3+, Mn2+, besides the line emissions of Pr3+, a broadband emission at 615 nm was

observed which was assigned to the emission of Mn2+. This implies that in SrB4O7:Pr3+, Mn2+,

the emission of Mn2+ is not from the direct absorption by Mn2+ itself, but from the ET

between Pr3+ and Mn2+. As a result, the unpractical UV photon from Pr3+ 1S0 can be converted

into the red Mn2+ emission. An energy level scheme of Pr3+ and Mn2+ showing the transitions

involved in the QC via a two-step ET from Pr3+ to Mn2+ and the emission processes is

presented in Fig. 38. The comparison of Figs. 37(a) and 37(b) also provides a simple method

for the estimation of the ET efficiency. Calculation shows an ET efficiency of 43% can be

obtained, resulting in a visible QE of 143%. However, before using SrB4O7:Pr3+, Mn2+ as

VUV QC phosphors in the practical applications, the nonradiative energy losses from Pr3+ in

the subsequent processes following the first step photon emission from 1S0 level or the ET

from Pr3+ to Mn2+ should be minimized. The current results indicate that the nonradiative

processes are more remarkable in SrB4O7:Pr3+, Mn2+ than in SrB4O7:Pr3+, as suggested by the

lower total intensity of the emission spectra in Figs. 37(b) relative to 37(a).

ET process for Pr3+–Mn2+ couple in LaMgB5O10 [109,110] was confirmed from

the emission and excitation spectra of the co-doped sample as well as from comparison of the

67
ACCEPTED MANUSCRIPT

decay curves of 1S0→ 1I6, 3PJ transition of Pr3+ between Pr3+ singly doped and Pr3+–Mn2+

co-doped samples. The ET efficiency from Pr3+ to Mn2+ shows a strong concentration or

distance dependence on Pr3+ and Mn2+. For LaMgB5O10:5%Pr3+,5%Mn2+ sample, the ET

efficiency was estimated only about 30%, which means that only a fraction of energy

transferring from Pr3+ to Mn2+ leads to Mn2+ emission. This is probably due to the emission of

Mn2+ is a spin forbidden transition. By analyzing the ET process, it was found that the ET

process in LaMgB5O10:Pr3+,Mn2+ was likely of resonant ET and the re-absorption process can

be excluded. The critical distances of ET based on the electric dipole–dipole interaction and

electric dipole–quadrupole interaction were calculated to be 4.78 and 9.46 Å in

LaMgB5O10:Pr3+,Mn2+, respectively, which are smaller than the mean distance of Pr3+ and

Mn2+ (17 Å) from high concentration doped materials. The near neighboring Pr3+–Mn2+

clusters formed in the LaMgB5O10 host is responsible for the ET process. High doping

concentration is expected to enhance the ET efficiency, whereas it also can lead to

concentration quenching in LaMgB5O10. In addition, in LaMgB5O10:Pr3+, the second step

emission of Pr3+ is very weak relative to the first step emission due to the quenching of 1D2 at

high concentration of Pr3+. As a result, LaMgB5O10 could not be found as a suitable host to

realize the Pr3+–Mn2+-based quantum cutter.

4.3.4. Pr3+–Cr3+ dual ions activated phosphors

Nie and co-workers [111-114] recently reported CRET process of Pr3+–Cr3+ dual

ions doped SrAl12O19 and CaAl12O19. The Cr3+ ion was chosen as a suitable co-activator ion

because it has abundant absorption transitions to match the Pr3+ 1S0 emissions. Unlike

68
ACCEPTED MANUSCRIPT

Pr3+–Er3+ pair in CaAl12O19 [101], Pr3+ and Cr3+ ions replace Ca2+ (Sr2+) and Al3+ in SrAl12O19

and CaAl12O19, respectively. The average Ca–Al distance (~4.64 Å) is shorter than the Ca–Ca

distance (~10.95 Å) [338,339], which may prompt efficient ET from Pr3+ to Cr3+.

Part of energy level diagram for the Pr3+–Cr3+ pair in SrAl12O19 (or CaAl12O19)

system, showing the possibility of QC via a two-step ET, is presented in Fig. 39. Upon

excitation to the 4ƒ5d states of Pr3+, in the first step, part of the excitation energy is

transferred to a Cr3+ by CR, leading to Cr3+ in 4T1(F) or 4T1(P) excited states and the Pr3+ in 1I6

or 1G4 levels. In the second step, the populations quickly relax from 1I6 down to 3P0, where the

remaining excitation is transferred to another Cr3+. Consequently, the two-step ET results in

two Cr3+ ions in excitation states, and thereby finally converting a VUV photon absorbed by a

Pr3+ into two red photons due to 2E→ 4A2 transitions of two Cr3+ ions. It is to be noted that the

CRET in the first step can depopulate the Pr3+ 1S0 state, feed the high-energy states and

simultaneously populate the 1I6 and 1G4 of Pr3+. As a result, an increase of the relative

emission intensity of 3P0 or 1G4 is expected in Pr3+ and Cr3+ doubly doped sample compared

with the Pr3+ singly doped sample. Thus the presence of above-stated CRET can be confirmed

by comparing the emission spectra with an excitation to Pr3+ 4ƒ5d states in the samples with

and without Cr3+ codoping.

Fig. 40(a) shows the VUV excitation and emission spectrum of CaAl12O19:1%

Pr3+. The emissions originating from 1S0→ 1I6 and 3P0→ 3H4 were simultaneously observed,

which demonstrates the occurrence of QC process. The UV and VUV excitation and emission

spectra of CaAl12O19:1% Cr3+ is shown in Fig. 40(b). The broad band dominated at 184 nm is

ascribed to the O2–→ Cr3+ related CTB [340]. The emission of CaAl12O19:1% Cr3+ consists of

69
ACCEPTED MANUSCRIPT

a 2E→ 4A2 zero-phonon line peaking at 686 nm with some vibronic sidebands [341,342]. The

comparison of the emission spectrum of CaAl12O19:1% Pr3+ and the excitation spectrum of

CaAl12O19:1% Cr3+ indicates that the two-step Pr3+ cascade emissions 1S0→ 1I6 at 400 nm and

3
P0→ 3H4 at 485 nm overlap Cr3+ excitations 4A2(4F) → 4T1(4F) and 4A2(4F) → 4T1(4P) states,

respectively. ET from Pr3+ 1S0 and 3P0 states to Cr3+ is possible under this condition.

Fig. 41 shows the emission spectra of Pr3+ in CaAl12O19:1% Pr3+, x% Cr3+ (x= 0,

1, 2, 3, 5) with excitation of 205 and 465 nm [114]. Upon excitation at 465 nm the 3P0

emission decreases with increasing Cr3+ doping concentration, which suggests the occurrence

of ET from the 3P0 to Cr3+. On the contrary, the expected increase in the relative emission

intensity of 3P0 to 1S0 is observed clearly as Cr3+ doping concentration increases upon

excitation of 205 nm, though the ET from 3P0 to Cr3+ can simultaneously depopulate the 3P0

level resulting in the decreasing relative intensity. This observation supports the existence of

ET from Pr3+ 1S0 to Cr3+. In addition, with 465 nm excitation, the decay curves of Pr3+ 3P0→

3
H4 emission in CaAl12O19:Pr3+, Cr3+ are found to be faster and nonexpontial, which also

indicates the occurrence of ET from the 3P0 to Cr3+. Similarly, the CRET for Pr3+–Cr3+ pair in

SrAl12O19 was also confirmed. Therefore, upon VUV excitation of the 4ƒ5d states of Pr3, DC

will take place via CRET in (Ca, Sr)Al12O19:Pr3+, Cr3+. Calculation results show that a

theoretical visible QE as great as 143% and 147% can be obtained for

CaAl12O19:1%Pr3+,5%Cr3+, and for SrAl12O19:2%Pr3+,5%Cr3+, respectively. In both hosts, the

visible QE is dependent on the dopant concentrations of Pr3+ and Cr3+ and their concentration

ratios. However, in a practical QC phosphor based on the Pr3+–Cr3+ pair, the Pr3+ ion and not

the Cr3+ ion should absorb the VUV emission from the Xe discharge. As shown in Fig. 40, it

70
ACCEPTED MANUSCRIPT

becomes evident that Cr3+ ions will be involved in the absorption apart from the Pr3+ ions.

Furthermore, for (Ca, Sr)Al12O19:Pr3+,Cr3+, most of the absorptions of the Pr3+ 4ƒ5d states are

located at wavelengths longer than 180 nm. Thus, the Xe discharge mostly excites the host

absorption band and the CTB of Cr3+ instead of the intense Pr3+ 4ƒ5d states. Consequently, in

the VUV region the host absorption band and the Cr3+-related CTB have an unfavorable

overlap with the Pr3+ 4ƒ5d states, preventing efficient selective Pr3+ excitation under radiation

of Xe discharge. Such direct Cr3+ excitation leads to a one-photon emission process, which

will reduce the actual visible QE of Pr3+–Cr3+ pair based QC phosphors used in practical

applications.

4.4. Downconversion in Tb3+-activated phosphors

Wide-band gap materials doped with RE ions have become the focus of

intensive research owing to their potential prospect as new VUV phosphors for lighting and

displays [102,343,344]. Similar to fluorides, CaSO4 has a wide band gap (>10 eV) and a low

crystal field, and hence provides an excellent opportunity to develop QC materials suitable for

applications with VUV radiation as an excitation source. Excited SO42– complexes have been

reported to create a STE-like state [67]. The unique property of CaSO4 arises from the

localized ET through CTS mediated anion excitons that are created within SO42– complexes.

However, such STE-mediated ET in various hosts, such as BaSO4 [67], CaAl12O19 [113],

Sr0.7La0.3Al11.7Mg0.3O19 [266], has so far been regarded as a highly unwanted process for

two-photon cascade emission since this process will result in the nonoccurrence of the QC

effect and reduce the possibilities of QC phenomenon (see Section 3.1.1). Direct absorption

71
ACCEPTED MANUSCRIPT

of VUV radiation by RE activator ions without any interference from the host lattice has been

considered as the efficient mode for two-photon emission, especially for the first-order DC

process [79,145]. Nevertheless, the attempt to obtain a two-photon emission by the interaction

of an anion exciton, directly created by an exciting photon, with a close pair of RE3+ ions in a

host doped with one type of RE3+, is considered as a promising method to realize the

second-order DC process. Recently, Lakshmanan et al. [120] reported visible QC effect via

the second-order DC in green-emitting CaSO4:Tb,Na phosphors by exploiting the ET from

anion excitons SO42– to Tb3+ activator ions. In CaSO4:Tb3+,Na+ phosphors, Tb3+ causes

fluorescence and Na+ serves as a charge compensator.

In CaSO4:Tb3+,Na+, the first-order ET from relaxed excitons to Tb3+ ions

involving a large Stokes shift is considered to be unlikely because the host absorption occurs

at a much higher energy (~10 eV) than the 4ƒ→ 5d energy absorption in Tb3+, as shown in

Fig. 42. A direct CTS-mediated VUV absorption is far off resonance in CaSO4:Tb3+. The

mobility of anion excitons which are self-trapped is restricted, unlike that of host-relaxed free

excitons. This makes the Stokes-shifted luminescence under the 147 nm excitation, a highly

unlikely and inefficient process in CaSO4 host.

Lakshmanan et al. [120] then proposed that the STE states created by excited

SO42– complexes with a maximum emission energy of 8.44 eV (147 nm) can trigger a

second-order DC to neighboring Tb3+ ions, specifically to 5H5 level of Tb3+ which occurs at

4.22 eV, exactly half that of the STE resonance energy. This is the upper limit of energy

absorption by the two Tb3+ acceptor ions as depicted in Fig. 43, satisfying the required

resonance condition for the second-order DC process. From the emission spectrum of

72
ACCEPTED MANUSCRIPT

CaSO4:Tb,Na in Fig. 42, the lower limit can be determined, which corresponds to 5D3 level of

Tb3+ ions at 3.26 eV. As the STEs start hopping through the CaSO4 crystal, their energy

relaxation will commence. But the ET by the second-order DC process, with the host anions

(SO42–) serving as donors while Tb3+ ions serving as acceptors, will continue to be operative

as there are several close lying 4ƒ levels of Tb3+ (as many as 9 levels between 5H5 and 5D3,

some of which are wide bands, see Fig. 43) between 3.26 and 4.22 eV, capable of absorbing

the relaxed energy emission from the STEs. Hence the resonance condition can be satisfied

over the entire STE-emission energy range (6.52–8.44 eV).

The value of QE for CaSO4:4%Tb3+,12%Na+ phosphor was found to be 117

±8% by comparing its integrated spectral power outputs (320–720 nm area) with that of the

standard YBO3:Tb3+ phosphor. A reduction in the grain size and an improvement in its

morphology would increase its luminescence efficiency further from the present value of

117% under VUV excitation. Thus, the CaSO4:Tb,Na phosphor is a promising QC phosphor

for application in PDPs. However, factors such as the spectral overlap of host-anion excitons

and charge compensator, the luminescence quenching by flux materials, and the concentration

quenching effects operating at energy states higher than 5D4 or 5D3 levels in Tb3+ ions due to

the CR process may prevent the CaSO4:Tb,Na phosphor to demonstrate a theoretical

maximum QE of 200 %.

73
ACCEPTED MANUSCRIPT

5. Near-infrared quantum cutting phosphors

5.1. Basic concepts

While QC phosphors via DC have been investigated for decades in the lighting

and display industry, DC phosphors only recently have been reported to possess of potential

application in photovoltaic (PV) cells [122,127–135]. The thermalization of electron-hole

pairs with energy greater than the band gap of Si (Eg=1.12 eV, λ=1100 nm), generated by the

absorption of high-energy photons, is one of the major energy loss mechanisms in a

conventional Si-based solar cell, which need to be overcome to significantly enhance device

efficiencies [127]. These energy losses can be efficiently reduced by spectrum modification

via DC. Each high-energy photon impinging on a DC layer, which overlies the solar cell,

results in the emission of more than one low-energy photon. To achieve two-photon emission

via DC, each incident high-energy photon must possess energy of at least twice the Si band

gap, or a wavelength of shorter than 550 nm. In this way, it’s possible to overcome the

classical efficiency limit of Si solar cells with doubling the incident photons.

Fig. 44 plots air-mass global spectrum (AM1.5) showing the fraction that is

currently absorbed by a thick Si device and the additional regions of the spectrum that can

contribute towards UC and DC [127-129,345-349]. Clearly, if the maximum fraction

available for DC can be efficiently utilized by Si, the conversion efficiency of Si solar cell

will be greatly improved. There are two important and key benefits in successfully applying

DC to PV technologies. First, DC components are passive, optical devices with carrier

collection still performed via the single PV junction. Therefore, Si solar cells with DC layers

have a distinct advantage over tandem solar cell, where the photocurrents generated in upper

74
ACCEPTED MANUSCRIPT

and lower cells must match throughout the day in order to avoid significant mismatch losses.

Second, the application of DC layers to PV does not require modification of the existing solar

cell since the layers are passive and purely optical in operation. Schematic diagram of the DC

system of a solar cell in combination with a down-converter is shown in Fig. 45. High-energy

photons with energy ћω>2Eg are absorbed by the converter and efficiently down-converted

into two lower energy photons with ћω>Eg, which can both be absorbed by the solar cell.

Using detailed balance calculations, for Si solar cell with this DC system, a maximum

conversion efficiency of 38.6% can be achieved [127].

RE-doped DC phosphors have been reported to achieve QE as great as 190%

[79,81], however, these require excitation with VUV photons (λ≤200 nm), which are not

present in the solar spectrum. Until now, it has not been identified any promising first-order

DC systems that would be applicable to Si solar cell. In contrast, the second-order DC

mechanism [350,351] seems promising. Absorbed by suitable DC phosphors, a photon with

twice the energy of the band gap (λ<550 nm) can be down-converted into two photons with

exactly the energy of the band gap (Eg=1.12 eV, λ=1100 nm), resulting in a great

improvement in conversion efficiency of Si solar cell. Dexter [10] has discussed the

possibility of cooperative sensitization of two acceptors by a single donor when the excited

state of the donor is at twice the energy of the acceptors. Basiev et al. [124] observed the

occurrence of second-order DC in the compound La1−xCexF3: 0.3%Nd3+, where the Nd3+ 4F3/2

level exciting two Ce3+ ions to their 2F7/2 level. Unfortunately, the Ce3+ 2F7/2→ 2F5/2 emission,

in the IR (5000 nm), is not suitable for the Si solar cell. Zhang et al. [131-133] in 2007

presented the concept of NIR QC involving the emission of two NIR photons for an absorbed

75
ACCEPTED MANUSCRIPT

visible photon. The NIR QC system is based on cooperative sensitization ET mechanism

[122,125].

Despite the Tb3+–Yb3+ pair has been studied extensively in many UC systems

[352-357], it is still a more promising system to study second-order ET processes involving

the combination of one Tb3+ and two Yb3+ ions: the Tb3+ 5D4→ 7F6 transition is located at

approximately twice the energy of the Yb3+ 2F5/2→ 2F7/2 transition and Yb3+ has no other

levels up to the UV region (Fig. 46). Upon excitation with blue lights (Tb3+: 7F6→ 5D4

transition), GdAl3(BO3)4:Tb,Yb might display a cooperative ET from Tb3+ to two Yb3+ ions:

5
D4→ 2F5/2 + 2F5/2 [131-135]. The narrow Yb3+ 2F5/2→ 2F7/2 emission peak at ~970 nm, is

found to be just above the band edge of crystalline Si. Part of the energy-level diagram of

RE3+ (RE= Tb, Pr, and Tm) and Yb3+ in GdAl3(BO3)4 are shown in Fig. 47 to illuminate the

possible ET between RE3+ (RE= Tb, Pr, and Tm) and Yb3+. Another promising system to

study the cooperative sensitisation process is the combination of one Pr3+ and two Yb3+ ions.

The Pr3+: 3P0 → 3H4 transition is located at approximately twice the energy of the Yb3+: 2F5/2→

2
F7/2 transition and Yb3+ has no other levels up to the UV region. The cooperative sensitisation

process may become the dominant relaxation process in this system [10,122,133]. The

concept of NIR QC through a cooperative ET from the GdAl3(BO3)4:RE3+,Yb3+ (RE= Tb, Pr,

and Tm) system has also been illustrated in Fig. 47. In the GdAl3(BO3)4:Pr,Yb system, the

proposed NIR QC occurs upon excitation at the 3P0 level of a Pr3+ ion. The energy of the

excited Pr3+ ions is transferred to two different Yb3+ ions through DC mechanism: 3P0→ 2F5/2

+ 2F5/2. Similarly, an excitation with visible lights (Tm3+:3H6→ 1G4 transition) in the

GdAl3(BO3)4:Tm,Yb system might exhibit a cooperative ET from Tm3+ to two Yb3+ ions:

76
ACCEPTED MANUSCRIPT

1
G4→ 2F5/2 + 2F5/2, which is opposite to the UC phenomenon of the RE3+-Yb3+ (RE= Tb, Pr,

and Tm) system [358-369].

5.2. Materials, photoluminescence and quantum efficiency

Up to now, NIR QC phenomena have been identified in many systems including

powder phosphors [122,131-134,136,370-373], crystals [125,126], trinuclear complex [374],

thin-film-phosphors [135], glasses [375-379], and glass ceramics [380-383].

Strek et al. [125,126] investigated cooperative DC in Tb3+-doped KYb(WO4)2.

The crystal was excited with a 308 nm laser, which was chosen to pump only the Tb3+ ions.

From the emission spectra shown in Fig. 48, it is clear that not only the visible emission

(16,000–21,000 cm–1) from the Tb3+ ions (5D4→ 7FJ) but also an NIR emission from the Yb3+

ions (2F5/2→ 2F7/2) have been observed. Furthermore, the intensity of the Yb3+ emission is

significantly greater at RT, indicating that phonon-assisted ET is required for Tb3+→ Yb3+.

However, the energy level diagram for Tb3+–Yb3+ pair (see Fig. 49) shows that such CR is

highly non-resonant with an energy mismatch of ΔE ≈ 4200 cm–1 for the 5D4→ 7F0 and 2F5/2→

2
F7/2 transitions. As the maximum phonon energy in the KYb(WO4)2 crystal is only about

1070 cm–1, the CR process for Tb3+→ Yb3+ therefore requires at least four maximum energy

phonons to balance the energy gap ΔE. Thus, Strek et al. [126] measured the power

dependencies of Tb3+ and Yb3+ emissions, and found that the Yb3+ emission intensity increases

with power more slowly than the Tb3+ intensities. This is assumedly due to the fact that only a

part of the excited Tb3+ ions transfer their energy to Yb3+ ions. Hence, apart from the

phonon-assisted process, the DC process may also take place, in which the energy of the

77
ACCEPTED MANUSCRIPT

excited Tb3+ ion is distributed between two different Yb3+ ions, 5D4→ 2F5/2 + 2F5/2, as

presented schematically in Fig. 49. Cooperative DC process was also observed in Tb3+–Yb3+

co-doped trinuclear complex when directly excited in the Tb3+ 5D4 level at 488 nm [374]. The

observed 980 nm luminescence was assigned to the Yb3+ ion, however, further more

investigations are considered necessary for having a better understanding of the ET process.

Vergeer et al. [122] in 2005 reported cooperative QC in solid-state reaction

synthesized YPO4:Tb3+,Yb3+ phosphor. In order to verify the ET process from Tb3+ to Yb3+,

an emission spectrum upon excitation in the Tb3+ 5D4 level has been carried out. Fig. 50

shows the emission spectrum of YPO4:Tb3+,Yb3+ at the RT. Upon Tb3+ 7F6→ 5D4 excitation

(489 nm), an emission band due to Yb3+ 2F5/2→ 2F7/2 transition is observed in the wavelength

region of 900–1100 nm, showing an ET from Tb3+ to Yb3+ occurs. Excitation spectra of Tb3+

and Yb3+ emission in Yb0.25Y0.74Tb0.01PO4 were also recorded to give convincing evidence for

the presence of Tb3+→ Yb3+ ET as shown in Fig. 51. The observation of the Tb3+ 7F6→ 5D4

lines in the excitation spectrum of Yb3+ shows that ET from Tb3+ to Yb3+ is present.

To elucidate ET mechanism of couple Tb3+–Yb3+, the time-resolved

luminescence measurements are compared with theories for phonon-assisted ET [141] and

second-order DC through a cooperative [122] or an accretive ET mechanism [350,351] by

exact calculations and simulations using Monte Carlo methods. Auzel [351] has proposed two

mechanisms for second-order ET: a cooperative and an accretive mechanism as shown in Fig.

46. After excitation of the donor (Tb3+: 7F6→5D4), both mechanisms require polarization

induced via a virtual state of opposite parity. For the cooperative pathway, the virtual state is

located on the donor (Tb3+) ion and this ion acts twice as a donor of energy. For the accretive

78
ACCEPTED MANUSCRIPT

pathway, the virtual state is located on one of the acceptor (Yb3+) ions. The latter accepts

energy from the donor ion, after which part of the energy is transferred to the second acceptor

ion [122]. For both mechanisms, Andrews et al. [350] have calculated the matrix element M

needed as input in Fermi’s golden rule to determine the ET rate:


γ ETr = M ρ
2
(10)
h
For the cooperative process, the cooperative ET rate is given by

1
γ coET = C coET ∑∑ 6 6
(11)
a c>a r r
aTb Tbc

which is independent of angular. The constant CcoET is fitted by means of Monte Carlo

simulations and the analytical solutions describe the time dependence of the donor emission

intensity. The ET rate of the accretive ET process can be treated analogously. However, since

there are two accretive pathways, where the energies are accrued at b and c, respectively, the

matrix element M for the accretive pathway is the sum of two contributions. Upon evaluation

of Eq. (10) one may note that the terms of quantum interference vanish in the orientational

averaging [122]. This leads to:

1 1
γ acET = C acET ∑∑ ( 6
+ 6 6)
6
(12)
a c >a r r rTbc rbc
Tbb bc

where b and c run over all Yb3+ positions in the lattice. Again, CacET is fitted by means of

Monte Carlo simulations. The third ET mechanism, phonon-assisted ET, is an ET process in

which the energy difference between the transition on the donor (Tb3+) and the transition on

the acceptor (Yb3+) is dissipated by multiphonon emission. For an dipole-dipole ET

mechanism, orientational averaging over the transition dipole moments yields 2/3 , so that the

transfer rate can be written as [122]:

79
ACCEPTED MANUSCRIPT

1
γ paET = C paET ∑ 6
(13)
a rTba

where the sum runs over all Yb3+ ions in the lattice. The parameter CpaET is fitted by means of

Monte Carlo simulations.

Fig. 52 shows luminescence decay curves of the Tb3+ 5D4→ 7F5 emission (544

nm) for various Yb3+ concentrations with simulation results. To obtain the computed curves,

the radiative decay rate and the ET rate in YbPO4 (as determined from single exponential fits

to the decay curves of YPO4 and YbPO4 in Fig. 52) were used as input parameters. Therefore,

at the Yb3+ concentration of 99%, the three simulated curves all match well with the

experimental curve. However, the cooperative model matches the experiments for all Yb3+

concentrations while the accretive and phonon-assisted ET model deviate substantially. At all

the intermediate Yb3+ concentrations, the decay profiles of the phonon-assisted model and the

accretive model fall off faster than the cooperative model. An excellent agreement has been

achieved between the experimentally measured luminescence decay curves. And, it is found

that the calculated curves for cooperative ET via dipole–dipole interaction thus provide strong

evidence that this mechanism could be found operational in the YPO4:Tb3+,Yb3+ system. The

ET efficiency and the QE are determined from the luminescence decay curves. A maximum

ET efficiency (ETE) of 88% achieves in YbPO4, resulting in a peak QE of 188%. Table 5

summarizes NIR QC materials, luminescence properties and QE. These Tb–Yb cooperative

QC phosphors might have prospects for increasing the energy efficiency of crystalline silicon

solar cells by down-converting the green-to-UV part of the solar spectrum to ~1000 nm

photons with almost complete doubling of the number of photons.

Zhang et al. [131-134] reported the NIR QC in Tb3+–Yb3+ codoped GdBO3,

80
ACCEPTED MANUSCRIPT

GdAl3(BO3)4, and Y3Al5O12 phosphors through a combustion synthesized route, which is

considered as a good method to synthesis inorganic oxides with improved properties for their

photonic applications. Luminescence measurements and decay curves of Tb3+ and Yb3+

emissions in (Gd0.99-xYbxTb0.01)BO3 nanophosphors were recorded to give convincing

evidence for the presence of Tb3+-Yb3+ ET. Fig. 53 shows the photoluminescence excitation

and emission spectra of (Gd0.99-xYbxTb0.01)BO3 nanophosphors. In the excitation spectrum [Fig.

53, left], an intense band observed at 486 nm, assigned to the 7F6→5D4 transition of Tb3+, has

been measured by monitoring both the 5D4→ 7F5 transition of Tb3+ at 543 nm and 2F5/2 → 2F7/2

transition of Yb3+ at 970 nm. The observation of the Tb3+ 7F6→5D4 lines in the excitation

spectrum of Yb3+ shows that ET from Tb3+ to Yb3+ is present. Excitation at 486 nm lights

gave rise to emissions both from Tb3+ and Yb3+ ions as shown in the right side of Fig. 53. The

peaks at 543 nm, 585 nm, 622 nm, 659 nm, 665 nm and 683 nm in the region of 500–700 nm

have been assigned to the electronic transitions of 5D4→ 7FJ (J = 5, 4, 3, 2, 1 and 0) of Tb3+

ions, respectively. The emissions from Tb3+ ions were reduced in intensity with incorporation

Yb3+ into the nanophosphors. Meanwhile, it is noteworthy to mention that a clear emission

centred at 970 nm along with a shoulder at 1015 nm was observed, which corresponds to the

2
F5/2→ 2F7/2 transition of Yb3+ ions.

The dependence of the green- and NIR-emission intensities on Yb3+

doping-concentration of the (Gd0.99-xYbxTb0.01)BO3 nanophosphors upon excitation at 486 nm

is also shown in Fig. 54. It is noticed that the emisison intensity of the NIR-emission at 970

nm increases rapidly with an increase of Yb3+ content which ranges from to 1 to 10 mol.%

while a green-emission at 543 nm decreases monotonically. However, at higher Yb3+

81
ACCEPTED MANUSCRIPT

concentrations, CR takes place between Yb3+ ions, which effectively reduces luminescence

yield. A decrease of the NIR-emission at 970 nm has been clearly observed when Yb3+

concentration is set to be more than 10 mol.%, due to concentration quenching. Since the Yb3+

luminescence is less affected by CR process than other RE ions for a same array, larger

amount of Yb3+ as great as 10–20 mol.% could be introduced before reaching concentration

quenching threshold, which increases the visible-NIR QE of Tb3+ quite significantly.

Fig. 55 shows the luminescence decay curves of Tb3+ 5D4→ 7F5 emission

intensity for GdBO3:Tb3+,Yb3+ doped with 1 mol.% Tb3+ and 0–75 mol.% Yb3+ upon

excitation at 486 nm. The decay curves of a singly doped (Gd0.99Tb0.01)BO3 nanophosphors fit

well to a single exponential function with a decay time of 4.25 ms. However, the decay

lifetimes decrease rapidly from 4.25 to 1.17 ms with increasing Yb3+ doping-concentration in

the range of 0–75 mol.%. The faster decline in the lifetime as a function of Yb3+ concentration

can be explained by the introduction of extra decay pathways due to the Yb3+-doping: ET

from Tb3+ 5D4 to Yb3+ enhances the Tb3+ 5D4 decay rate. These results indicate that the

emission of two NIR photons per absorbed visible photon is possible with the Tb3+–Yb3+ dual

ions via a cooperative ET. The Tb3+ 5D4→ 7F6 transition is located at approximately twice the

energy of the Yb3+ 2F7/2→ 2F5/2 transition, and the 2F5/2→ 2F7/2 emission is situated around

1000 nm, just above the band edge of crystalline Si. An energy-level diagram to explain the

concept of NIR QC through a cooperative ET in the GdBO3 nanophosphors codoped with

Tb3+ and Yb3+ is shown in the inset of Fig. 54. In this system, QC occurs upon excitation at

the 5D4 level of a Tb3+ ion. The energy of the excited Tb3+ ions is transferred to two different

Yb3+ ions. The downconversion process could be expressed as: 5D4→ 2F5/2 + 2F5/2.

82
ACCEPTED MANUSCRIPT

The Tb3+ decay curves have been fitted to the generalization of the

Yokota–Tanimoto model [80,356]:

⎡ t 4π S −3 / S − 2

( S ) 3 / S ⎛ 1 + a1 X + a 2 X ⎞
2
3
I (t ) = I (0) exp ⎢− − C A Γ(1 − ) × (C DA t ) ⎜ ⎜ ⎟⎟ ⎥ (14)
⎢⎣ τ 3 S ⎝ 1 +b1 X ⎠ ⎥⎦
−2 / S t 1− 2 / S
with X = D[C DA
(S )
] , where τ is the intrinsic lifetime of the luminescent ions, S=6, 8,

10,...depending on the interaction character (dipole–dipole, dipole–quadrupole,

quadrupole–quadrupole,…), CA is the acceptor concentration, Γ( y ) is the gamma function,

(S )
C DA is the donor-acceptor energy-transfer parameter, and D is the diffusion parameter which

characterizes the transfer between donors. The values for the ai and bi coefficients in Eq. (14)

are introduced as reported earlier in Ref. [356]. In this generalized model, energy migration

between donors is considered with whatever multipole interaction prevails upon the donors

and acceptors. The obtained best fitting parameter S is 6, indicating that ET originates from

the dipole-dipole interaction. From the luminescence decay curves [shown in Fig. 55], the ET

efficiency and QE can be determined. The ET efficiency, ηx%Yb, is defined as the ratio of Tb3+

ions that depopulate by ET to Yb3+ ions over the total number of Tb3+ ions excited. The total

QE, η, could be defined as the ratio of the number of photons emitted to the number of

photons that are absorbed, assuming that all excited Yb3+ ions decay radiatively. This

assumption leads to an upper limit of QE. By dividing the integrated intensity of the decay

curves of the (Gd0.99-xYbxTb0.01)BO3 nanophosphor samples to the integrated intensity of the

(Gd0.99Tb0.01)BO3 curve, the ET efficiency could be obtained as a function of the Yb3+

concentration:

η x %Yb = 1 −
∫I x %Yb dt
(15)
∫I 0%Yb dt

83
ACCEPTED MANUSCRIPT

where I denotes intensity and x%Yb stands for the Yb3+ concentration. The relation between

the ET efficiency and the total QE is defined as [122]:

η = η Tb (1 − η x %Yb ) + 2η x %Yb (16)

where QE for the Tb3+ ions, ηTb, is set to 1.

Fig. 56 shows an ET efficiency of GdBO3:Tb3+,Yb3+ nanophosphors as a

function of Yb3+ doping-concentration. The ET efficiencies, QE and decay lifetimes of the

nanophosphors are also summarized in the inset of Fig. 55. It is found that the ET efficiency

of the nanophosphors shows an increase up to a maximum value of 98% in

(Gd0.24Yb0.75Tb0.01)BO3 nanophosphors with an increase in the Yb3+ doping-concentration in

the samples according to the aforementioned equations. For (Gd0.24Yb0.75Tb0.01)BO3

nanophosphors a QE has been found to be about 198%, which is noticed to be quite

satisfactory, indicating that depopulation of the Tb3+ 5D4 level proceeds 98 out of 100 times

by excitation of two Yb3+ ions to the 2F5/2 level. However, it should be mentioned here that

concentration quenching occurs at Yb3+ doping-concentration beyond 10 mol.% as illustrated

in Fig. 54. The exact maximum QE should be 182% at the Yb3+ doping-concentration of 10

mol.%. Although the absolute QE has not been determined and the nonradiative losses still

remain unknown, the very week Tb3+ emissions of the 5D4→ 7FJ transitions upon excitation of

486 nm compared to the total emission intensity suggests that ET from Tb3+ to Yb3+ is

extremely efficient.

NIR QC has also been observed in Zn2SiO4:Tb3+,Yb3+ transparent thin films

[135], Y2O3:Tb3+,Yb3+ [136], and Pr3+–Yb3+ or Tm3+–Yb3+ codoped GdAl3(BO3)4 and

Y3Al5O12 phosphors (Fig. 57) [133,134]. More recently, NIR QC has been achieved in

84
ACCEPTED MANUSCRIPT

Tb3+–Yb3+ codoped transparent glass ceramics with nanocrystals [380-383]. By taking into

account the excellent chemical and mechanical properties as well as high transparency to

visible light, these glass ceramics could be considered as more promising DC layer candidates

to enhance efficiency of Si solar cell.

85
ACCEPTED MANUSCRIPT

6. Perspectives and future advances

Major advances have been made by many research groups around the world

recently in demonstrating QC and NIR QC, fabricating new luminescent materials,

developing high efficiency QC phosphors and the mechanism involved. There are still many

areas that need additional work, including:

(1) Development of new QC materials, to realize efficient visible- or NIR- QC.

This will require a better understanding of ET and DC of the visible- and NIR- QC

mechanism, and also a better understanding of materials properties and new materials

fabrication process.

(2) Realization of higher VUV absorption and effectively ET in QC phosphors

with good characteristics. This will require suitable sensitizer ions.

(3) Improved QE in dual/ternary ions activated QC phosphors. This will need a

better understanding of mechanism of DC, ET, CR and nonradiative relaxation in ion pairs

such as Pr3+–Mn2+, Gd3+–Eu3+, Gd3+–Tb3+, etc. It is also necessary to understand that nature of

the role of sensitizer ions, new QC materials and its characteristics: synthesis, characterization

and optical properties.

(4) Efficient NIR QC: mechanism, materials and applications. A desirable goal is

to achieve an efficient NIR QC thin films or materials, which could highly enhance the

efficiency of state-of-the-art solar cells.

(5) Clear and direct demonstrations of ET and DC between ions. It is necessary

to develop new methods for the measurements of QC luminescent properties including QE.

It is likely that many of these areas could be possible in attaining more

86
ACCEPTED MANUSCRIPT

significant progress in near future concerning the QC phosphors for their use in VUV-excited

PDPs, MFFLs, and even in the development of efficient solar cells converter becomes more

widely known.

87
ACCEPTED MANUSCRIPT

7. Concluding remarks

In summary, we have noticed that a recent demonstration of efficient visible QC

in VUV-excited LiGdF4:Eu phosphors have been considered to be more important and

exciting scenario towards the development of superior luminescent materials and devices.

Ever since the first report has appeared on QC phenomenon in the deep blue region of Pr3+

-doped YF3 and NaYF4 in the early 1970s, an in-depth research activity has been commenced

with a special focus on its potential applications in efficient VUV-excited plasma displays and

mercury-free fluorescent tubes. The efficient gain in QC materials is based on the principle

that a QC phosphor can emit two visible photons for every VUV photon absorbed. The

excitation energy is divided over the two photons, leading to the necessary red-shift of the

absorbed radiation without losing energy efficiency. Investigation on QC systems has started

on single ions capable of a cascade emission from ions such as Pr3+, Tm3+, Gd3+ and Er3+. The

focus has now been shifted to the combination of two ions, where the energy of the donor ion

could be transferred in stepwise to two acceptor ions via a DC process. It has widely been

known now that, QC via DC could be found in many RE-based phosphors, especially in

fluoride-based phosphors. Due to their wide band gap and low photon energy, fluoride

materials containing trivalent RE ions could provide more possibilities in the development of

newer optical materials that are suitable for variety of applications with VUV radiation as an

excitation source. However, an in-depth investigation on oxide-based QC phosphors is

necessary for their potential applications in various high-performance display and devices.

On the other hand, QC materials could also be applied in solar cells. For this

application, the concept of NIR QC involving the emission of two NIR photons for an

88
ACCEPTED MANUSCRIPT

absorbed visible photon has been proposed. Through NIR QC process, energy loss due to

thermalization of electron-hole pairs could be reduced significantly. The development of solar

cells could greatly benefit from QC phosphors with the energy of its emission located just

above the band gap of silicon. RE3+–Yb3+ (RE= Pr3+, Tm3+, Tb3+) dual-ions-based QC

phosphors have potential prospect in realizing high efficiency Si solar cells by converting the

green-to-UV part of the solar spectrum to ~1000 nm photons with almost double the number

of photons.

A great number of newer concepts and novel materials are still in the research

stage. Some of them may lead to much higher efficiency and lower cost in the coming

decades. A long way is still ahead in achieving the goal from efficient QC phosphors.

However, the prospects of reaching this goal have been found to be satisfactory and more

interesting in the back drop of emergence of different and promising materials and concepts.

89
ACCEPTED MANUSCRIPT

8. Acknowledgements

Helpful discussions with Professors S. Buddhudu and C. Q. Sun, and editorial

advice given by Professor T. Mohri are all gratefully acknowledged. Thanks are also due to

the anonymous referee for his/her thorough reviewing of the manuscript and his/her

constructive suggestions to revise the manuscript. QYZ acknowledges funding support from

NFSC (Grant Nos. U0934001 and 50872036). Permission for reprinting diagrams from

Elsevier Science, John Wiley, Springer, IOP, APS, and AIP is also acknowledged.

90
ACCEPTED MANUSCRIPT

9. References

[1] Feldmann C, Jüstel T, Ronda CR, Schmidt PJ. Inorganic luminescent materials: 100 years

of research and applications. Adv Funct Mater 2003; 13: 511-516.

[2] Blasse G, Grabmaier BC. Luminescent materials. Berlin: Springer; 1994.

[3] Ronda CR. Phosphors for lamps and displays: an applicational view. J Alloys Compd

1995; 225: 534-538.

[4] Henderson B, Imbusch GF. Optical spectroscopy in inorganic solids. Oxford: Clarendon;

1998.

[5] Moine B, Bizarri G. Why the quest of new rare earth doped phosphors deserves to go on?

Opt Mater 2006; 28: 58-63.

[6] Oskam KD, Wegh RT, Donker H, van Loef EVD, Meijerink A. Downconversion: a new

route to visible quantum cutting. J Alloys Compd 2000; 300–301: 421-425.

[7] Ronda CR, Jüstel T, Nikol H. Rare earth phosphors: fundamentals and applications. J

Alloys Compd 1998; 275–277: 669-676.

[8] Sommerer TJ. Model of a weakly ionized, low-pressure xenon dc positive column

discharge plasma. J Phys D: Appl. Phys. 1996; 29: 769-778.

[9] Jüstel T, Nikol H. Optimization of luminescent materials for plasma display panels. Adv

Mater 2000; 12: 527-530.

[10] Dexter DL. Possibility of luminescent quantum yields greater than unity. Phys Rev 1957;

108: 630-633.

[11] Elias LR, Heaps WS, Yen WM. Excitation of UV fluorescence in LaF3 doped with

trivalent cerium and praseodymium. Phys Rev B 1973; 8: 4989-4995.

91
ACCEPTED MANUSCRIPT

[12] Heaps WS, Elias LR, Yen WM. Vacuum-ultraviolet absorption bands in trivalent

lanthanides LaF3. Phys Rev B 1976; 13: 94-104.

[13] Heaps WS, Elias LR, Yen WM. In, Proc. Fourth Inter. Conf. Vacuum Ultraviolet Phys.;

Koch, E. E., Ed.; Hamburg, Germany [DESY, Hamburg], 1974.

[14] Carnall WT, Fields PR, Sarup R. Optical absorption spectra of Er3+:LaF3 and ErCl3·6H2O.

J Chem Phys 1972; 57: 43-51.

[15] Yang KH, DeLuca JA. VUV fluorescence of Nd3+-, Er3+, and Tm3+-doped trifluorides and

tunable coherent sources from 1650 to 2600 Å. Appl Phys Lett 1976; 29: 499-501.

[16] Yang KH, DeLuca JA. Vacuum-ultraviolet excitation studies of 5d14fn-1 to 4fn and 4fn to

4fn transitions of Nd3+-, Er3+-, and Tm3+-doped trifluorides. Phys Rev B. 1978; 17:

4246-4255.

[17] Devyatkova LI, Ivanova ON, Mikhailin VV, Rudnev SN, Chernov SP. High-energy states

of Er3+ and Ho3+ ions in fluoride-crystals. OPTIKA I SPEKTROSKOPIYA 1987; 62:

464-466.

[18] Devyatkova KM, Ivanova ON, Seiranyan KB, Tamazyan SA, Chernov SP. Vacuum

ultraviolet properties of a new fluoride matrix. Doklady Akademii Nauk SSSR 1990; 310:

72-74.

[19] Sarantopoulou E, Cefalas AC, Dubinskii MA, Nicolaides CA, Abdulsabirov RYu,

Korableva SL, Naumov AK, Semashko VV. VUV and UV fluorescence and absorption

studies of Nd3+ and Ho3+ ions in LiYF4 single-crystals. Opt Commun 1994; 107:

104-110.

[20] Sarantopoulou E, Cefalas AC, Dubinskii MA, Kollia Z, Nicolaides CA, Abdulsabirov

92
ACCEPTED MANUSCRIPT

RY, Korableva SL, Naumov AK, Semashko VV. VUV and UV fluorescence and

absorption studies of Tb3+ and Tm3+ trivalent ions in LiYF4 single crystal hosts. J Mod

Opt 1994; 41: 767-775.

[21] Sarantopoulou E, Kollia Z, Cefalas AC, Dubinskii MA, Nicolaides CA, Abdulsabirov

RY, Korableva SL, Naumov AK, Semashko VV. Vacuum ultraviolet and ultraviolet

fluorescence and absorption studies of Er3+-doped LiLuF4 single crystals. Appl Phys Lett

1994; 65: 813-815.

[22] Sarantopoulou E, Kollia Z, Cefalas AC. 4ƒ95d→4ƒ10 spin-allowed and spin-forbidden

transitions of Ho3+ ions in LiYF4 single crystals in the vacuum ultraviolet. Opt Commun

1999; 169: 263-274.

[23] Sarantopoulou E, Kollia Z, Cefalas AC, Semashko VV, Abdulsabirov RY, Naumov AK,

Korableva SL. On the VUV and UV 4ƒ7(8S)5d→4ƒ8 interconfigurational transitions of

Tb3+ ions in LiLuF4 single crystal hosts. Opt Commun 1998; 156: 101-111.

[24] Waynant RW, Klein PH. Vacuum ultraviolet laser emission from Nd3+:LaF3. Appl Phys

Lett 1985; 46: 14-16.

[25] Solarz P, Dominiak-Dzik G, Lisiecki R, Ryba-Romanowski W. Conversion of VUV to

UV and visible in K5Li2LnF10 containing rare-earth from cerium group (Ln = La3+, Ce3+,

Pr3+, Nd3+). Radiat Meas 2004; 38: 603-606.

[26] Becker J, Gesland JY, Kirikova NY, Krupa JC, Makhov VN, Runne M, Queffelec M,

Uvarova TV, Zimmerer G. VUV emission of Er3+ and Tm3+ in fluoride crystals. J Lumin

1998; 78: 91-96.

[27] Khaidukov NM, Kirm M, Lam SK, Lo D, Makhov VN, Zimmerer G. VUV spectroscopy

93
ACCEPTED MANUSCRIPT

of KYF4 crystals doped with Nd3+, Er3+ and Tm3+. Opt Commun 2000; 184: 183-193.

[28] Makhov VN, Khaidukov NM, Kirikova NY, Kirm M, Krupa JC, Ouvarova TV,

Zimmerer G. VUV emission of rare-earth ions doped into fluoride crystals. J Lumin 2000;

87-89: 1005-1007.

[29] Kirm M, Stryganyuk G, Vielhauer S, Zimmerer G, Makhov VN, Malkin BZ, Solovyev

OV, Abdulsabirov RY, Korableva SL. Vacuum-ultraviolet 5d-4f luminescence of Gd3+

and Lu3+ ions in fluoride matrices. Phys Rev B 2007; 75: 075111.

[30] Becker J, Gesland JY, Kirikova NY, Krupa JC, Makhov VN, Runne M, Queffelec M,

Uvarova TV, Zimmerer G. Fast VUV emission of rare earth ions (Nd3+, Er3+, Tm3+) in

wide bandgap crystals. J Alloys Compd 1998; 275-277: 205-208.

[31] Wegh RT, Meijerink A. Spin-allowed and spin-forbidden 4fn↔4fn-15d transitions for

heavy lanthanides in fluoride hosts. Phys Rev B 1999; 60: 10820-10830.

[32] Wegh RT, van Klinken W, Meijerink A. High-energy 2G(2)9/2 emission and 4f25d→4f3

multiphonon relaxation for Nd3+ in orthoborates and orthophosphates. Phys Rev B 2001;

64: 045115.

[33] Wegh RT, Donker H, Meijerink A. Spin-allowed and spin-forbidden fd emission from

Er3+ and LiYF4. Phys Rev B 1998; 57: R2025-2028.

[34] Wegh RT, Donker H, Meijerink A, Lamminmaki RJ, Hölsä J. Vacuum-ultraviolet

spectroscopy and quantum cutting for Gd3+ in LiYF4. Phys Rev B 1997; 56:

13841-13848.

[35] Reid MF, van Pieterson L, Wegh RT, Meijerink A. Spectroscopy and calculations for

4fnÆ 4fn-15d1 transitions of lanthanide ions in LiYF4. Phys Rev B 2000; 62:

94
ACCEPTED MANUSCRIPT

14744-14749.

[36] van Pieterson L, Reid MF, Meijerink A. Reappearance of fine structure as a probe of

lifetime broadening mechanisms in the 4fN→ 4fN-15d excitation spectra of Tb3+, Er3+, and

Tm3+ in CaF2 and LiYF4. Phys Rev Lett 2002; 88: 067405.

[37] Peijzel PS, Vermeulen P, Schrama WJM, Meijerink A, Reid MF, Burdick GW.

High-resolution measurements of the vacuum ultraviolet energy levels of trivalent

gadolinium by excited state excitation. Phys Rev B 2005; 71: 125126.

[38] Peijzel PS, Schrama WJM, Reid MF, Meijerink A. Probing vacuum ultraviolet energy

levels of trivalent gadolinium by two-photon spectroscopy. J Lumin 2003; 102–103:

211-215.

[39] Krupa JC, Queffelec M. UV and VUV optical excitations in wide band gap materials

doped with rare earth ions: 4ƒ–5d transitions. J Alloys Compd 1997; 250: 287-292.

[40] Belsky AN, Krupa JC. Luminescence excitation mechanisms of rare earth doped

phosphors in the VUV range. Displays 1999; 19: 185-196.

[41] Kwon IE, Yu BY, Bae H, Hwang YJ, Kwon TW, Kim CH, Pyun CH, Kim SJ.

Luminescence properties of borate phosphors in the UV/VUV region. J Lumin. 2000;

87-89: 1039-1041.

[42] Rao RP, Devine DJ. RE-activated lanthanide phosphate phosphors for PDP applications.

J Lumin 2000; 87-89: 1260-1263.

[43] Drozdowski W, Wojtowicz AJ. Radiative recombination in BaF2:Pr. J Alloys Compd

2000; 300-301: 261-266.

[44] Hirai T, Yoshida H, Sakuragi S, Hashimoto S, Ohno N. Transfer of excitation energy

95
ACCEPTED MANUSCRIPT

from Pr3+ to Gd3+ in YF3:Pr3+,Gd3+. Jpn J Appl Phys 2007; 46: 660-663.

[45] Hirai T, Ohno N, Hashimoto S, Sakuragi S. Abstracts of the 203rd Meeting of the

Electrochemical Society, 2003 (No. 2461).

[46] Hirai T, Ohno N, Hashimoto S, Sakuragi S. Luminescence of NaGdF4:Tb3+,Eu3+ under

vacuum ultraviolet excitation. J Alloys Compd 2006; 408–412: 894-897.

[47] Tanner PA, Mak CSK, Faucher MD, Kwok WM, Phillips DL, Mikhailik V. 4ƒ-5d

transitions of Pr3+ in elpasolite lattices. Phys Rev B 2003; 67: 115102.

[48] Peijzel PS, Meijerink A, Wegh RT, Reid MF, Burdick GWJ. A complete 4fn energy level

diagram for all trivalent lanthanide ions. J Solid State Chem 2005; 178: 448-453.

[49] Peijzel PS, Vergeer P, Meijerink A, Reid MF, Boatner LA, Burdick GW. 4fnÆ 4fn-15d

emission of Ce3+, Pr3+, Nd3+, Er3+, and Tm3+ in LiYF4 and YPO4. Phys Rev B 2005; 71:

045116.

[50] Wegh RT, Meijerink A, Lamminmaki RJ, Hölsä J. Extending Dieke's diagram. J Lumin

2000; 87-89: 1002-1004.

[51] van Pieterson L, Reid MF, Wegh RT, Soverna S, Meijerink A. 4fnÆ 4fn-15d transitions of

the light lanthanides: Experiment and theory. Phys Rev B 2002; 65: 045113.

[52] van Pieterson L, Reid MF, Burdick GW, Meijerink A. 4fnÆ 4fn-15d transitions of the

heavy lanthanides: Experiment and theory. Phys Rev B 2002; 65: 045114.

[53] van Pieterson L, Wegh RT, Meijerink A, Reid MF. Emission spectra and trends for

4fn-15d↔4fn transitions of lanthanide ions: Experiment and theory. J Chem Phys 2001;

115: 9382-9392.

[54] van Pieterson L, Reid MF, Wegh RT, Meijerink A. 4fnÆ 4fn-15d transitions of the

96
ACCEPTED MANUSCRIPT

trivalent lanthanides: experiment and theory. J Lumin 2001; 94–95: 79-83.

[55] Zimmerer G. Luminescence spectroscopy with synchrotron radiation: history, highlights,

future. J Lumin 2006; 1: 119–120.

[56] Piper WW, Ham FS. Cascade fluorescent decay in Pr3+-doped fluorides: Achievement of

a quantum yield greater than unity for emission of visible light. J Lumin 1974; 8:

344-348.

[57] Sommerdijk JL, Bril A, de Jager AW. Two photon luminescence with ultraviolet

excitation of trivalent praseodymium. J Lumin 1974; 8: 341-343.

[58] Sommerdijk JL, Bril A, de Jager AW. Luminescence of Pr3+-activated fluorides. J Lumin

1974; 9: 288-296.

[59] Srivastava AM, Beers WW. Luminescence of Pr3+ in SrAl12O19:observation of two

photon luminescence in oxide lattice. J Lumin 1997; 71: 285-290.

[60] Srivastava AM, Doughty DA, Beers WW. Photon cascade luminescence of Pr3+ in

LaMgB5O10. J Electrochem Soc 1996; 143: 4113-4116.

[61] Srivastava AM, Doughty DA, Beers WW. On the vacuum-ultraviolet excited

luminescence of Pr3+ in LaB3O6. J Electrochem Soc 1997; 144: L190-192.

[62] Srivastava AM, Duclos SJ. On the luminescence of YF3-Pr3+ under vacuum ultraviolet

and X-ray excitation. Chem Phys Lett 1997; 275: 453-456.

[63] Rodnyi PA, Mikhrin SB, Dorenbos P, van der Kolk E, van Eijk CWE, Vink AP, Avanesov

AG. The observation of photon cascade emission in Pr3+-doped compounds under X-ray

excitation. Opt Commun 2002; 204: 237-245.

[64] Vink AP, Dorenbos P, de Haas JTM, Donker H, Rodnyi PA, Gavanesov A, van Eijk CWE.

97
ACCEPTED MANUSCRIPT

Photon cascade emission in SrAlF5:Pr3+. J Phys: Condens Matter 2002; 14: 8889-8899.

[65] Dorenbos P. The 4fnÆ 4fn-15d transitions of the trivalent lanthanides in halogenides and

chalcogenides. J Lumin 2000; 91: 91-106.

[66] Dorenbos P. The 5d level positions of the trivalent lanthanides in inorganic compounds. J

Lumin 2000; 91: 155-176.

[67] van der Kolk E, Dorenbos P, Vink AP, Perego RC, van Eijk CWE, Lakshmanan AR.

Vacuum ultraviolet excitation and emission properties of Pr3+ and Ce3+ in MSO4 [M=Ba,

Sr, and Ca] and predicting quantum splitting by Pr3+ in oxides and fluorides. Phys Rev B

2001; 64: 195129.

[68] van der Kolk E, Dorenbos P, van Eijk CWE, Vink AP, Fouassier C, Guillen F. VUV

excitation and 1S0 emission of Pr3+ in BaSiF6 and selecting host lattices with 1S0 Pr3+

emission. J Lumin 2002; 97: 212-223.

[69] Judd BR. Optical absorption intensities of rare-earth ions. Phys Rev 1962; 127: 750-61.

[70] Ofelt GS. Intensities of crystal spectra and decay of Er3+ fluorescence in LaF3. J Chem

Phys 1962; 37: 511-519.

[71] Ronda C. Luminescent materials with quantum efficiency larger than 1, status and

prospects. J Lumin 2002; 100: 301-305.

[72] Pappalardo R. Calculated quantum yield for photon cascade emission [PCE] for Pr3+ and

Tm3+ in fluoride hosts. J Lumin 1976; 14: 159-193.

[73] Yang Z, Lin JH, Su MZ, Tao Y, Wang W. Photon cascade luminescence of Gd3+ in

GdBaB9O16. J Alloys Compd 2000; 308: 94-97.

[74] Feofilov SP, Zhou Y, Seo HJ, Jeong JY, Keszler DA, Meltzer RS. Host sensitization of

98
ACCEPTED MANUSCRIPT

Gd3+ ions in yttrium and scandium borates and phosphates: Application to quantum

cutting. Phys Rev B 2006; 74: 085101.

[75] Zhou Y, Feofilov SP, Seo HJ, Jeong JY, Keszler DA, Meltzer RS. Energy transfer to Gd3+

from the self-trapped exciton in ScPO4:Gd3+: Dynamics and application to quantum

cutting. Phys Rev B 2008; 77: 075129.

[76] Tian ZF, Liang HB, Han B, Su Q, Tao Y, Zhang GB, Fu YB. Photon Cascade Emission

of Gd3+ in Na(Y,Gd)FPO4. J Phys Chem C 2008; 112: 12524-12529.

[77] Wegh RT, van Loef, EVD, Burdick GW, Meijerink A. Luminescence spectroscopy of

high-energy 4f[11] levels of Er3+ in fluorides. Mol Phys 2003; 101: 1047-1056.

[78] Peijzel PS, Meijerink A. Visible photon cascade emission from the high energy levels of

Er3+. Chem Phys Lett 2005; 401: 241-245.

[79] Wegh RT, Donker H, Oskam KD. Meijerink A. Visible quantum cutting in LiGdF4: Eu3+

through downconversion. Science 1999; 283: 663-666.

[80] Auzel F. Up-conversion in RE-doped Solids. In, Spectroscopic Properties of Rare Earths

in Optical Materials. Eds. G. Liu and B. Jacquier. Springer, 2005. p. 266-319.

[81] Wegh RT, Donker H, Oskam KD, Meijerink A. Visible quantum cutting in Eu3+-doped

gadolinium fuorides via downconversion. J Lumin 1999; 82: 93-104.

[82] Wegh RT, van Loef EVD, Meijerink A. Visible quantum cutting via downconversion in

LiGdF4:Er3+, Tb3+ upon Er3+ 4f11->4f105d excitation. J Lumin 2000; 90: 111-122.

[83] You FT, Wang Y, Lin JH, Tao Y. Hydrothermal synthesis and luminescence properties of

NaGdF4:Eu. J Alloys Compd 2002; 343: 151-155.

[84] You FT, Huang SH, Liu SM, Tao Y. VUV excited luminescence of MGdF4: Eu3+ [M = Na,

99
K, NH4]. J Lumin 2004; 110: 95-99.

[85] You FT, Huang SH, Liu SM, Tao Y. Synthesis, structure and VUV luminescent properties

of rubidium rare-earth fluorides. J Solid State Chem 2004; 177: 2777-2782.

[86] Liu B, Chen Y, Shi C, Tang H, Tao Y. Visible quantum cutting in BaF2:Gd,Eu via

downconversion. J Lumin 2003; 101: 155-159.

[87] Kodama N, Watanabe Y. Visible quantum cutting through downconversion in Eu3+-doped

KGd3F10 and KGd2F7 crystals. Appl Phys Lett 2004; 84: 4141-4143.

[88] Kodama N, Oishi S. Visible quantum cutting through downconversion in KLiGdF5:Eu3+

crystals. J Appl Phys 2005; 98: 103515.

[89] Karbowiaka M, Mecha A, Romanowski WR. Optical properties of Eu3+:CsGd2F7

downconversion phosphor. J Lumin 2005; 114: 65-70.

[90] Hua RN, Niu JH, Chen BJ, Li MT, Yu TZ, Li WL. Visible quantum cutting in GdF3:Eu3+

nanocrystals via downconversion. Nanotechnology 2006; 17: 1642-1645.

[91] Lee TJ, Luo LY, Diau EWG, Chen TM, Cheng BM, Tung CY. Visible quantum cutting

through K2GdF5:Tb3+ phosphors. Appl Phys Lett 2006; 89: 131121.

[92] Tzeng HY, Cheng BM, Chen TM. Visible quantum cutting in green–emitting

BaGdF5:Tb3+ phosphors via downconversion. J Lumin 2007; 122-123: 917-920.

[93] Wegh RT, Donker H, van Loef EVD, Oskam KD, Meijerink A. Quantum cutting through

downconversion in rare-earth compounds. J Lumin 2000; 87-89: 1017-1019.

[94] Peijzel PS, Schrama WJM, Meijerink A. Thulium as a sensitizer for the Gd3+/Eu3+

quantum cutting couple. Mol Phys 2004; 102: 1285-1290.

[95] Babin V, Oskam KD, Vergeer P, Meijerink A. The role of Pb2+ as a sensitizer for

100
Gd3+–Eu3+ downconversion couple in fluorides. Radiat Meas 2004; 38: 767-770.

[96] Feofilov SP, Zhou Y, Jeong JY, Keszler DA, Meltzer RS. Sensitization of Gd3+ and the

dynamics of quantum splitting in GdF3:Pr,Eu. J Lumin 2007; 122-123: 503-505.

[97] Jia W, Zhou Y, Feofilov SP, Meltzer RS, Jeong JY, Keszler, D. Quantum splitting and its

dynamics in GdLiF4:Nd. Phys Rev B 2005; 72: 075114.

[98] Zhou Y, Feofilov SP, Jeong JY, Keszler DA, Meltzer RS. Quantum cutting in GdxY1-xLiF4:

Nd-dynamics and mechanisms. J Lumin 2006; 119-120: 264-270.

[99] Zachau M, Zwaschka F, Kummer, F. in: C. R. Ronda, T. Welker [Eds.], Proc.Sixth Intern.

Conf. Lumin. Mater., Electrochem Soc 1998, p.314.

[100] Vergeer P, Babin V, Meijerink A. Quenching of Pr3+ 1S0 emission by Eu3+ and Yb3+. J

Lumin 2005; 114: 267-274.

[101] Wang XJ, Huang S, Lu L, Yen WM, Srivastava AM, Setlur AA. Energy transfer in Pr3+

and Er3+ codoped CaAl12O19 crystal. Opt Commun 2001; 195: 405-410.

[102] Jia W, Zhou Y, Keszler DA, Jeong JY, Jang KW, Meltzer RS. Relaxation of the 4ƒn-15d1

electronic states of rare earth ions in YPO4 and YBO3. Phys Stat Sol C 2005; 2: 48-52.

[103] Park W, Summers CJ, Do Y, Park D, Yang H. US Patent, 6669867 B2, Dec.30, 2003.

[104] van der Kolk E, Dorenbos P, van Eijk CWE, Vink AP, Weil M, Chaminade JP.

Luminescence excitation study of the higher energy states of Pr3+ and Mn2+ in SrAlF5,

CaAlF5, and NaMgF3. J Appl Phys 2004; 95: 7867-7872.

[105] Kück S, Sokólska I. Spectroscopic investigation of Mn2+, Pr3+ codoped KMgF3 under

vacuum-ultraviolet excitation. J Phys: Condens Matter 2006; 18: 5447-5457.

[106] Meijerink A, Wegh RT, Vergeer P, Vlugt T. Photon management with lanthanides. Opt

101
Mater 2006; 28: 575-581.

[107] Chen YH, Shi CS, Yan WZ, Qi ZM, Fu YB. Energy transfer between Pr3+ and Mn2+ in

SrB4O7:Pr,Mn. Appl Phys Lett 2006; 88: 061906.

[108] Chen YH, Yan WZ, Shi CS. Energy transfer in Pr3+ and Mn2+ co-doped SrB6O10 and

SrB4O7. J Lumin 2007; 122-123: 21-24.

[109] Fu YB, Zhang GB, Qi ZM, Wu WQ, Shi CS. Energy transfer in LaMgB5O10:Pr3+,Mn2+

under VUV excitation. J Lumin 2007; 124: 370-374.

[110] Fu YB, Zhang GB, Wu WQ, Qi ZM, Chen YH, Wang DW, Shi CS. The energy transfer

processes in LaMgB5O10:Pr3+, Mn2+. J Solid State Chem 2007; 68: 1779-1784.

[111] Nie ZG, Zhang JH, Zhang X, Luo YS, Lu SZ, Wang XJ. Energy transfer in Pr3+- and

Cr3+-codoped SrAl12O19 system. J Lumin 2006; 119-120: 332-336.

[112] Nie ZG, Zhang JH, Zhang X, Ren XG. Evidence for visible quantum cutting via energy

transfer in SrAl12O19 : Pr, Cr. Opt Lett 2007; 32: 991-993.

[113] Nie ZG, Zhang JH, Zhang X, Ren XG, Di WH, Zhang GB, Zhang DH, Wang XJ.

Spectroscopic investigation of CaAl12O19:M3+ upon UV/vacuum–UV excitation: a

comparison with SrAl12O19:M3+ [M = Pr, Cr]. J Phys: Condens Matter 2007; 19: 076204.

[114] Nie ZG, Zhang JH, Zhang X, Lu SZ, Ren XG, Zhang GB, Wang XJ. Photon cascade

luminescence in CaAl12O19:Pr, Cr. J Solid State Chem 2007; 180: 2933-2941.

[115] Moine B, Beauzamy L, Gredin P, Wallez G, Labeguerie J. Research of green emitting

rare-earth doped materials as potential quantum-cutter. Opt Mater 2008; 30: 1083-1087.

[116] Beauzamy L, Moine B, Gredin P. Energy transfers between dysprosium and terbium in

YF3. J Lumin 2007; 127: 568-574.

102
[117] Jüstel T, Nikol H, Ronda C. New developments in the field of luminescent materials for

lighting and displays. Angew Chem Int Ed 1998; 37: 3084-3103.

[118] Nalwa HS, Rohwer LS. Handbook of luminescence, display materials, and devices.

Volume 2: inorganic display materials. American Scientific Publishers, 2003.

[119] Moine B, Bizarri G. Rare-earth doped phosphors: oldies or goldies? Mat Sci Eng B

2003; 105: 2-7.

[120] Lakshmanan AR, Kim SB, Jang HM, Kum BG, Kang BK, Heo S, Seo D. A

Quantum-Splitting Phosphor Exploiting the Energy Transfer from Anion Excitons to

Tb3+ in CaSO4:Tb,Na. Adv Funct Mater 2007; 17: 212-218.

[121] Chander H. Development of nanophosphors-A review. Mater Sci Eng R 2005; 49:

113-155.

[122] Vergeer P, Vlugt TJH, Kox MHF, Den Hertog MI, van der Eerden JPJM, Meijerink A.

Quantum cutting by cooperative energy transfer in YbxY1−xPO4:Tb3+. Phys Rev B 2005;

71: 014119.

[123] Forster T. "Zwischenmolekulare energiewanderung und fluoreszenz," Ann Phys 1948;2:

55-75. An English translation of Förster’s original work is provided by R. S. Knox,

“Intermolecular energy migration and fluorescence,” in Biologial physics, E. Mielczarek,

R. S. Knox, and E. Greenbaum, eds., 148-160, New York; American Institute of Physics,

1993.

[124] Basiev TT, Doroshenko ME, Osiko VV. Cooperative nonradiative cross-relaxation in

crystals of La[1-x]CexF3 solid solutions. JETP Lett 2000; 71: 8-11.

[125] Strek W, Deren P, Bednarkiewicz A. Cooperative processes in KYb[WO4]2 crystal

103
doped with Eu3+ and Tb3+ ions. J Lumin 2000; 87–89: 999-1001.

[126] Strek W, Bednarkiewicz A, Deren PJ. Power dependence of luminescence of

Tb3+-doped KYb[WO4]2 crystal. J Lumin 2001; 92: 229-235.

[127] Trupke T, Green MA, Würfel P. Improving solar cell efficiencies by down-conversion

of high-energy photons. J Appl Phys 2002; 92: 1668-1674.

[128] Richards BS, Shalav A. The role of polymers in the luminescence conversion of

sunlight for enhanced solar cell performance. Synth Metal 2005; 154: 61-64

[129] Richards BS. Enhancing the performance of silicon solar cells via the application of

passive luminescence conversion layers. Solar Energy Mater Sol Cells 2006; 90:

2329-37.

[130] Richards BS. Luminescent layers for enhanced silicon solar cell performance:

Down-conversion. Sol Energy Mater Sol Cells 2006; 90: 1189-1207.

[131] Zhang QY, Yang CH, Pan YX. Cooperative quantum cutting in one-dimensional

YbxGd1−xAl3[BO3]4:Tb3+ nanorods. Appl Phys Lett 2007; 90: 021107.

[132] Zhang QY, Yang CH, Jiang ZH, Ji XH. Concentration-dependent near-infrared in

GdBO3:Tb3+,Yb3+ nanophosphors. Appl Phys Lett 2007; 90: 061914.

[133] Zhang QY, Yang GF, Jiang ZH. Cooperative downconversion in GdAl3[BO3]4:

RE3+,Yb3+ [RE=Pr, Tb, and Tm]. Appl Phys Lett 2007; 91: 051903.

[134] Zhang QY, Liang XF. Cooperative quantum cutting in Y3Al5O12:RE,Yb [RE=Tb, Tm,

and Pr] nanophosphors. J Soc Inform Display 2008; 16: 755-758.

[135] Huang XY, Zhang QY. Efficient near-infrared downconversion in Zn2SiO4:Tb3+,Yb3+

thin-films, J Appl Phys 2009; 105: 053521.

104
[136] Yuan JL, Zeng XY, Zhao JT, Zhang ZJ, Chen HH, Yang XX. Energy transfer

mechanisms in Tb3+, Yb3+ codoped Y2O3 downconversion phosphor. J Phys D: Appl.

Phys. 2008; 41: 105406.

[137] Ye S, Zhu B, Chen JX, Luo J, Qiu JR. Infrared quantum cutting in Tb3+,Yb3+ codoped

transparent glass ceramics containing CaF2 nanocrystals. Appl Phys Lett 2008; 92:

141112.

[138] Dexter DL. A theory of sensitized luminescence in solids. J Chem Phys 1953; 21:

836-850.

[139] Axe JD, Weller PF. Fluorescence and Energy Transfer in Y2O3:Eu3+. J Chem Phys 1964;

40: 3066-3069.

[140] Orbach R. In, Optical properties of ions in solids, ed. B. DiBartolo, Plenum Press, New

York, 1975, p. 445.

[141] Miyakawa T, Dexter DL. Phonon sidebands, multiphonon relaxation of excited states,

and phonon-assisted energy transfer between ions in solids. Phys Rev B 1970; 1:

2961-2969.

[142] Reisfeld R. Excited states and energy transfer from donor cations to rare earths in the

condensed phase. Structure and Bonding, 1976; 30: 65-97.

[143] Yamada NS, Shionoya, Kushida T. J. Phys. Soc. Jpn. 1972; 32:1577.

[144] Kong LB, Zhang TS, Ma J, Boey F. Progress in synthesis of ferroelectric ceramic

materials via high-energy mechanochemical technique. Prog Mater Sci 2008; 53:

207-322.

[145] Feldmann C, Jüstel T, Ronda CR, Wiechert DU. Quantum efficiency of

105
down-conversion phosphor LiGdF4: Eu. J Lumin 2001; 92: 245-254.

[146] Liang LF, Wu H, Hua HL, Wu MM, Su Q. Enhanced blue and green upconversion in

hydrothermally synthesized hexagonal NaY1−xYbxF4:Ln3+ (Ln3+ = Er3+ or Tm3+). J

Alloys Compd 2004; 368: 94-100.

[147] Reis KP, Ramanan A, Whittingham MS. Low-temperature hydrothermal reduction of

ammonium paratungstate. J Solid State Chem 1991; 91: 394-396.

[148] Zhao H, Feng S. Hydrothermal synthesis and oxygen ionic conductivity of codoped

nanocrystalline Ce1-xMxBi0.4O2.6-x, M = Ca, Sr, and Ba. Chem Mater 1999; 11: 958-964.

[149] Li G, Feng S, Li L, Li X, Jin W. Mild hydrothermal syntheses and thermal behaviors of

hydrogarnets Sr3M2(OH)12 (M = Cr, Fe, and Al). Chem Mater 1997; 9: 2894-2901.

[150] Li G, Li L, Feng S, Wang M, Zhang L, Yao X. An effective synthetic route for a novel

electrolyte: nanocrystalline solid solutions of (CeO2)1-x(BiO1.5)x. Adv Mater 1999; 11:

146-149.

[151] Byrappa K, Adschiri T. Hydrothermal technology for nanotechnology. Prog Cryst

Growth Ch 2007; 53: 117-166

[152] Zhao C, Feng S, Xu R, Shi C, Ni J. Hydrothermal synthesis of the complex fluorides

LiBaF3 and KMgF3 with perovskite structures under mild conditions. Chem Commun

1996; 14: 1641-1642.

[153] Xun X, Feng S, Wang J, Xu R. Hydrothermal synthesis of complex fluorides NaHoF4

and NaEuF4 with fluorite structures under mild conditions. Chem Mater 1997; 9:

2966-2968.

[154] Xun X, Feng S, Xu R. Hydrothermal synthesis of complex fluorides LiHoF4 and LiErF4

106
with scheelite structures under mild conditions. Mater Res Bull 1998; 33: 369-375.

[155] Liang LF, Xu HF, Su Q, Konishi H, Jiang YB, Wu MM, Wang YF, Xia DY.

Hydrothermal synthesis of prismatic NaHoF4 microtubes and NaSmF4 nanotubes. Inorg

Chem 2004; 43: 1594-1596.

[156] Wang DW, Huang SH, You FT, Qi SQ, Fu YB, Zhang GB, Xu JH, Huang Y. Vacuum

ultraviolet spectroscopic properties of Pr3+ in MYF4 [M =Li, Na, and K] and LiLuF4. J

Lumin 2007; 122–123: 450-452.

[157] Li CX, Yang J, Yang PP, Lin J. Hydrothermal synthesis of lanthanide fluorides LnF3 (Ln

= La to Lu) nano-/microcrystals with multiform structures and morphologies. Chem

Mater 2008; 20: 4317-4326.

[158] Belsky AN, Khaidukov NM, Krupa JC, Makhov VN, Philippov A. Luminescence of

CsGd2F7: Er3 +, Dy3 + under VUV excitation. J Lumin 2001; 94–95: 45-49.

[159] Rao CNR. Chemical Approaches to the Synthesis of Inorganic Materials. New Delhi:

Wiley; 1994.

[160] Fu LJ, Liu H, Li C, Wu YP, Rahm E, Holze R, Wu HQ. Electrode materials for lithium

secondary batteries prepared by sol–gel methods. Prog Mater Sci 2005; 50: 881-928

[161] Brincker CJ, Scherer GW. Sol-Gel Science, Academic Press, New York, 1990.

[162] Santos AMM, Vasconcelos WL. Properties of porous silica glasses prepared via sol-gel

process. J Non-Cryst Solids, 2000; 273: 145-149.

[163] Lin J, Sänger DU, Mennig M, Bärner K, Sol-gel synthesis and characterization of

Zn2SiO4:Mn phosphor films. Mater Sci Eng B 1999; 64: 73-78.

[164] Rao RP. Growth and characterization of Y2O3:Eu3+ phosphors films by sol-gel process.

107
Solid State Commun 1996; 99: 439-443.

[165] Yang P, Song CF, Lü MK, Yin X, Zhou GJ, Xu D, Yuan DR. The luminescence of PbS

nanoparticles embedded in sol-gel silica glass. Chem Phys Lett 2001; 345: 429-434.

[166] Zhang QY, Pita K, Kam CH. Sol-gel derived zinc silicate phosphor films for full color

display applications. J Phys Chem Solids 2003; 64: 333-338.

[167] Que WX, Zhang QY, Chan YC, Kam CH. Sol-gel derived hard optical coatings via

organic/inorganic composites. Composite Sci Technol 2003; 63: 347-351.

[168] Yang PP, Quan ZW, Li CX, Yang J, Wang HA, Liu XM, Lin J. Fabrication and

luminescent properties of the core-shell structured YNbO4:Eu3+/Tb3+@SiO2 spherical

particles. 2008; 181: 1943-1949.

[169] Selomulya R, Ski S, Pita K, Kam CH, Zhang QY, Buddhudu S. Luminescence properties

of Zn2SiO4:Mn2+ thin-films by a sol-gel process. Mater Sci Eng B 2003; 100: 136-141.

[170] Zhang QY, Pita K, Ye W, Que WX, Kam CH, Effects of composition and structure on

spectral properties of Eu3+-doped yttrium silicate transparent nanocrystaline films by

metallorganic decomposition method. Chem Phys Lett 2002; 356: 161-167.

[171] Zhang QY, Pita K, Ye W, Que WX. Influence of annealing atmosphere and temperature

on photoluminescence of Tb3+ or Eu3+-activated zinc silicate thin film phosphors via

sol-gel method. Chem Phys Lett 2002; 351: 163-170.

[172] Zhang QY, Pita K, Buddhudu S, Kam CH. Luminescent properties of rare earth ions

doped yttrium silicate thin films phosphors for a full color display. J Phys D: Appl Phys

2002; 35: 3085-3090.

108
[173] Zhang QY, Shen J, Wu GM, Wang J, Chen LY. ZrO2 thin films and ZrO2/SiO2 optical

reflection filters deposited by sol-gel method. Mater Lett 2000; 45: 311-314.

[174] Lu HF, Yan B. Molecular design and fluorescent whitening emission from novel

lanthanide activated organic-inorganic covalently hybrid micro-particles. J Fluoresc

2008; 18: 763-769.

[175] Xiao F, Xue YN, Zhang QY. Warm white light from Y4MgSi3O13:Bi3+,Eu3+

nanophosphor for white light-emitting diodes. Spectrochim Acta A: Mol Biomol

Spectrosc 2009; doi:10.1016/j.saa.2009.06.050.

[176] Wang FF, Yan B. Co-luminescence effect of heterometallic terbium-gadolinium hybrid

molecular materials constructed by covalent grafting. J Fluoresc 2007; 17: 418-426.

[177] Yan B, Xu S, Lu HF. Photophysical properties of luminescent quaternary lanthanide

molecular hybrid systems with chemical bonds from the cooperative design and

assembly of structure and function. J Fluoresc 2007; 17: 155-161.

[178] Lin YJ, Wu PH, Tsai CL, Liu CJ, Lin ZR, Chang HC, Lee CT. Effects of Mg

incorporation on the optical properties of ZnO prepared by the sol-gel method. J Appl

Phys 2008; 103: 113709.

[179] Liu XM, Zhu L, Wang LL, Yu CC, Lin J. Synthesis and luminescent properties of

Lu3Ga5O12:RE3+ (RE=Eu, Tb, and Pr) nanocrystalline phosphors via sol-gel process. J

Electrochem Soc 2008; 155: P21-P27.

[180] Li XJ, Yu QM, Yang JH, Zeng ZZ, Jing XP. Particle growth of Y3Al5O12:Tb3+ phosphor

in sol-gel preparation. J Electrochem Soc 2007; 154: H726-H729.

[181] Lee YK, Oh JR, Do YR. Enhanced extraction efficiency of Y2O3:Eu3+ thin-film

109
phosphors coated with hexagonally close-packed polystyrene nanosphere

monolayers. Appl Phys Lett 2007; 91: 041907.

[182] Lu CH, Hsu WT, Cheng BM. Luminescence characteristics of sol-gel derived

Y3Al5O12:Eu3+ phosphors excited with vacuum ultraviolet. J Appl Phys 2006; 100:

063535.

[183] Wang D, Li YX, Xiong YH, Yin QR. SrAl2O4:(Eu2+, Dy3+) phosphor thin films derived

from the sol-gel process. J Electrochem Soc 2005; 152: H12-H14.

[184] Wei ZG, Sun LD, Liao CS, Jiang XC, Yan CH, Tao Y, Hou XY, Ju X. Size dependence

of luminescent properties for hexagonal YBO3:Eu nanocrystals in the vacuum

ultraviolet region. J Appl Phys 2003; 93: 9783-9788.

[185] Lu CH, Jagannathan R. Cerium-ion-doped yttrium aluminum garnet nanophosphors

prepared through sol-gel pyrolysis for luminescent lighting. Appl Phys Lett 2002; 80:

3608-3610.

[186] Ravichandran D, Roy R, White WB. Low-temperature synthesis and particle-size

control in yttrium-based phosphors. J Soc Inform Display 1997; 5: 107.

[187] Wang YH, He L. Synthesis of Ca4GdO(BO3)3:Eu3+, Al3+ by sol-gel processing and its

luminescence properties. J Electrochem Soc 2006; 153: H78-H81.

[188] Biswas A, Maciel GS, Kapoor R, Friend CS, Prasad PN. Er3+-doped multicomponent

sol-gel-processed silica glass for optical signal amplification at 1.5 µm. Appl Phys Lett

2003; 82: 2389-2391.

[189] Potdevin A, Chadeyron G, Boyer D, Mahiou R. Optical properties upon vacuum

ultraviolet excitation of sol-gel based Y3Al5O12:Tb3+, Ce3+ powders, J Appl Phys 2007;

110
102: 073536.

[190] Boyer D, Mahiou R. Powders and coatings of LiYF4:Eu3+ obtained via an original way

based on the sol-gel process, Chem Mater 2004; 16: 2518-2521.

[191] Boyer D, Kharbache H, Mahiou R. Comparison between the optical properties of

sol-gel process and solid state reaction derived LiYF4:Eu3+ powders. Opt Mater 2006;

28: 53-57.

[192] Lepoutre S, Boyer D, Potdevin A, Dubois M, Briois V, Mahiou R. Structural

investigations of sol-gel-derived LiYF4 and LiGdF4 powders, J Solid State Chem 2007;

180: 3049-3057.

[193] Lepoutre S, Boyer D, Mahiou R. Structural and optical characterizations of sol-gel

based fluorides materials: LiGdF4:Eu3+ and LiYF4:Er3+. Opt Mater 2006; 28: 592-596.

[194] Lepoutre S, Boyer D, Mahiou R. Quantum cutting abilities of sol-gel derived

LiGdF4:Eu3+ powders. J Lumin 2008; 128: 635-641.

[195] Mech A, Karbowiak M, Kępiński L, Bednarkiewicz A, W. Stręk. Structural and

luminescent properties of nano-sized NaGdF4:Eu3+ synthesised by wet-chemistry route.

J Alloys Compd 2004; 380: 315-320.

[196] Loureiro SM, Setlur A, Heward W, Taylor ST, Comanzo H, Manoharan M, Srivastava A,

Schmidt P, Happek U. First observation of quantum splitting behavior in

nanocrystalline SrAl12O19:Pr, Mg phosphor. Chem Mater 2005; 17: 3108-3113.

[197] Ptacek P, Schäfer H, Kömpe K, Haase M. Crystal phase control of luminescing

NaGdF4:Eu3+ nanocrystals. Adv Funct Mater 2007; 17: 3843-3848.

[198] Kingsley J, Patil KC. A novel combustion process for the synthesis of fine particle

111
α-alumina and related oxide materials. Mater Lett 1988; 6: 427-432.

[199] Patil KC, Aruna ST, Ekambaram S. Combustion synthesis. Curr Opin Solid State Mater

Sci 1997; 2: 158-165.

[200] Ekambaram S, Maaza M. Synthesis of lamp phosphors: facile combustion approach. J

Alloys Compd 2005; 393: 81-92.

[201] Rakov N, Ramos FE, Hirata G, Xiao MF. Strong photoluminescence and

cathodoluminescence due to ƒ–ƒ transitions in Eu3+ doped Al2O3 powders prepared by

direct combustion synthesis and thin films deposited by laser ablation. Appl Phys Lett

2003; 83: 272-274.

[202] Rakov N, Maciel GS, Lozano WB, Araújo CB. Investigation of Eu3+ luminescence

intensification in Al2O3 powders codoped with Tb3+ and prepared by low-temperature

direct combustion synthesis. Appl Phys Lett 2006; 88: 081908.

[203] Kingsley JJ, Pederson LR, Energetic materials in ceramics synthesis. Mater Res Soc

Symp Proc 1993; 296: 361-366.

[204] Yang CH, Pan YX, Zhang QY, Jiang ZH. Cooperative energy transfer and frequency

upconversion in Yb3+–Tb3+ and Nd3+–Yb3+–Tb3+ Codoped GdAl3(BO3)4 Phosphors. J

Fluoresc 2007; 17: 500-504.

[205] Yang CH, Pan YX, Yang GF, Zhang QY. Synthesis and spectroscopic properties of

GdAl3(BO3)4 poly-crystals codoped with Yb3+ and Eu3+. J Fluoresc 2009; 19: 105–09.

[206] Pan YX, Zhang QY. Comparative investigation on nanocrystal structure and

luminescence properties of gadolinium molybdates codoped with Er3+/Yb3+. J Fluoresc

2007; 17: 444-448.

112
[207] Pan YX, Zhang QY. White upconverted luminescence of rare earth ions codoped

Gd2(MoO4)3 nanocrystals. Mater Sci Eng B 2007; 138: 90-95.

[208] Zhang QY, Yang CH, Pan YX. Enhanced white light emission from GdAl3(BO3)4:Dy3+,

Ce3+ nanorods. Nanotechnology 2007; 18: 145602.

[209] Liang XF, Huang XY, Zhang QY. Gd2(MoO4)3:Er3+ nanophosphors for an enhancement

of silicon solar-cell near-infrared response. J Fluoresc 2009; 19: 285-289.

[210] Yang CH, Pan YX, Zhang QY. Enhanced white light emission from Dy3+/Ce3+ codoped

GdAl3(BO3)4 phosphors by combustion synthesis. Mater Sci Eng B 2007; 137:

195-199.

[211] Patil KC, Aruna ST, Mimani T. Combustion synthesis: an update. Curr Opin Solid State

Mater Sci 2002; 6: 507-512.

[212] Bhatkar VB, Omanwar SK, Moharil SV. Combustion synthesis of silicate phosphors.

Opt Mater 2007; 29: 1066-1070.

[213] Kinsley JJ, Suresh K, Patil KC. Combustion synthesis of fine particle rare earth

orthoaluminates and yttrium aluminum garnet. J Solid State Chem 1990; 88: 435-442.

[214] Rakov N, Maciel GS. Enhancement of luminescence efficiency of ƒ–ƒ transitions from

Tb3+ due to energy transfer from Ce3+ in Al2O3 crystalline ceramic powders prepared by

low temperature direct combustion synthesis. Chem Phys Lett 2005; 400: 553-557.

[215] Hirata GA, Ramos F, Garcia R, Bosze EJ, McKittrick J, Ponce FA. A new combustion

synthesis method for GaN:Eu3+ and Ga2O3:Eu3+ luminescent powders. Phys Stat Sol (a)

2001; 188: 179-182.

[216] Tanner PA, Pan ZF, Rakov N, Maciel GS. Luminescence of Eu3+ in α-Al2O3 powders. J

113
Alloys Compd 2006; 424: 347-349.

[217] Gomes J, Pires AM, SerraOA. Nanocrystalline RE2O3:Tm3+ (RE: Gd3+, Y3+) blue

phosphors synthesized via the combustion method. J Fluoresc 2006; 16: 411-421.

[218] Jacobsohn LG, Bennett BL, Muenchausen RE, Tornga SC, Thompson JD, Ugurlu O,

Cooke DW, Sharma ALL. Multifunction Gd2O3:Eu nanocrystals produced by solution

combustion synthesis: Structural, luminescent, and magnetic characterization. J Appl

Phys 2008; 103: 104303.

[219] Sun LD, Yao J, Liu CH, Liao CS, Yan CH. Rare earth activated nanosized oxide

phosphors: synthesis and optical properties, J Lumin 2000; 87–89: 447-450.

[220] Luo XX, Cao WH. Ethanol-assistant solution combustion method to prepare

La2O2S:Yb,Pr nanometer phosphor. J Alloys Compd 2008; 460: 529-534.

[221] Zhao CL, Chen DH, Yuan YH, Wu M. Synthesis of Sr4Al14O25:Eu2+,Dy3+ phosphor

nanometer powders by combustion processes and its optical properties. Mater Sci Eng

B 2006; 133: 200-204.

[222] McKittrick J, Shea LE, Bacalski CF, Bosze EJ. The influence of processing parameters

on luminescent oxides produced by combustion synthesis. Displays 1999; 19: 169-172.

[223] Cheng BM, Yu LX, Duan CK, Wang HS, Tanner PA. Vacuum ultraviolet and visible

spectra of ZnO:Eu3+ prepared by combustion synthesis. J Phys: Condens Matter 2008;

20: 345231.

[224] Mukherjee S, Sudarsan V, Vatsa RK, Godbole SV, Kadam RM, Bhatta UM, Tyagi AK.

Effect of structure, particle size and relative concentration of Eu3+ and Tb3+ ions on the

luminescence properties of Eu3+ co-doped Y2O3:Tb nanoparticles. Nanotechnology

114
2008; 19: 325704.

[225] Klug P, Alexander LE. X-ray Diffraction Procedure. New York: Wiley; 1954.

[226] Karbowiak M, Mech A, Bednarkiewicz A, Stręk W, Kępiński L. Comparison of

different NaGdF4:Eu3+ synthesis routes and their influence on its structural and

luminescent properties. J Phys Chem Solids 2005; 66: 1008-1019.

[227] Hachani S, Moine B, El-akrmi A, Férid M. Luminescent properties of some ortho- and

pentaphosphates doped with Gd3+–Eu3+: potential phosphors for vacuum ultraviolet

excitation. Opt Mater 2009; 31: 678-684.

[228] Zimmerer G. Status report on luminescence investigations with synchrotron radiation at

HASYLAB. Nucl Instrum Methods Phys Res A 1991; 308: 178-186.

[229] Hahn U, Schwentner N, Zimmerer G. A system for time and energy resolved VUV

luminescence spectroscopy using synchrotron radiation for excitation. Nucl Instrum

Methods 1978; 152: 261-264.

[230] Kondo H, Hirai T, Hashimoto S. Dynamical behavior of quantum cutting in alkali

gadolinium fluoride phosphors. J Lumin 2004; 108: 59-63.

[231] Kondo H, Hirai T, Hashimoto S. Energy migration and relaxation through Gd3+

sublattice in NaGdF4. J Lumin 2003; 102–103: 727-732.

[232] Digonnet MJF, Sadowski RW, Shaw HJ, Pantell RH. Resonantly enhanced nonlinearity

in doped fibers for low-power all-optical switching: A review. Opt Fiber Technol 1997;

3: 44-64.

[233] Wang J, Hector JR, Brady D, Hewak D, Brocklesby B, Kluth M, Moore R, Payne DN.

Halide-modified Ga–La sulfide glasses with improved fiber-drawing and optical

115
properties for Pr3+-doped fiber amplifiers at 1.3 µm. Appl Phys Lett 1997; 71:

1753-1755.

[234] Ballman AA, Porto SPS, Yariv A. Calcium niobate Ca(NbO3)2–a new laser host crystal.

J Appl Phys 1963; 34: 3155-3156.

[235] Krasteva V, Machewirth D, Sigel GH. Pr3+-doped Ge–S–I glasses as candidate materials

for 1.3 μm optical fiber amplifiers. J Non-Cryst Solids 1997; 213–214: 304-310.

[236] Shionoya S, Yen WM. Phosphor Handbook. New York: CRC press; 1998.

[237] Duverger C, Ferrari M, Mazzoleni C, Montagna M, Pucker G, Turrell S. Optical

spectroscopy of Pr3+ ions in sol–gel derived GeO2–SiO2 planar waveguides. J

Non-Cryst Solids 1999; 245: 129-134.

[238] Birkhahn R, Garter M, Steckl AJ. Red light emission by photonluminescence and

electroluminescence from Pr-doped GaN on Si substrates.1999; 74: 2161-163.

[239] Okamoto S, Kobayashi H, Yamamoto HJ. Enhancement of characteristic red emission

from SrTiO3:Pr3+ by Al addition. Appl Phys 1999; 86: 5594-5597.

[240] Kaminskii AA. Laser Crystals, Springer Series in Optical Sciences; Springer: Berlin,

1981; vol. 14.

[241] Abe K, Takebe H, Morinaga K. Preparation and properties of Ge-Ga-S glasses for laser

hosts. J Non-Cryst Solids 1997; 212: 143-150.

[242] Balda R, Fernández J, Ocáriz IS, Voda M, García AJ. Laser spectroscopy of Pr3+ ions in

LiKY1-xPrxF5 single crystals. Phys Rev B 1999; 59: 9972-9980.

[243] Koch ME, Kueny AW, Case WE. Photon avalanche upconversion laser at 644 nm Appl

Phys Lett 1990; 56: 1083-1085.

116
[244] Dorenbos P, Marsman M, van Eijk CWE, Korzhik MV, Minkov BI. Scintillation

properties of Y2SiO5:Pr crystals. Radiat Effects Defects Solids 1995; 135: 325-328.

[245] van Eijk CWE, Dorenbos P, Visser R. Nd3+ and Pr3+ Doped Inorganic Scintillators.

IEEE Trans Nucl Sci 1994; 41: 738-741.

[246] Dujardin C, Pedrini C, Gâcon JC, Petrosyan A G, Belsky AN, Vasiľev AN.

Luminescence properties and scintillation mechanisms of cerium- and

praseodymium-doped lutetium orthoaluminate. J Phys: Condens Matter 1997; 9:

5229-5243.

[247] Laroche M, Braud A, Girard S, Doualan JL, Moncorge R, Thuau M, Merkle LD.

Spectroscopic investigations of the 4ƒ5d energy levels of Pr3+ in fluoride crystals by

excited-state absorption and two-step excitation measurements. J Opt Soc Am B 1999;

16: 2269-2277.

[248] Nicolas S, Laroche M, Girard S, Moncorge R, Guyot Y, Joubert MF, Descroix E,

Petrosyan, AG. 4f[2] to 4f5d excited state absorption in Pr3+:YAlO3. J Phys: Condens

Matter 1999; 11: 7937-7946.

[249] Laroche M, Bettinelli M, Girard S, Moncorge R. ƒ–d Luminescence of Pr3+ and Ce3+ in

the chloro-elpasolite Cs2NaYCl6. Chem Phys Lett 1999; 311: 167-172.

[250] Dieke GH. Spectra and Energy Levels of Rare Earth Ions in Crystals; Interscience: New

York, 1968.

[251] Pedrini C, Bouttet D, Dujardin C, Moine B, Bill H. Spectroscopic studies of the f-d

ultraviolet transitions of Pr3+ in alkaline-earth fluorides. Chem Phys Lett 1994; 220:

433-436.

117
[252] Makhov VN, Khaidukov NM, Lo D, Kirm M, Zimmerer G. Spectroscopic properties of

Pr3+ luminescence in complex fluoride crystals. J Lumin 2003; 102–103: 638-643.

[253] Laroche M, Nettinelli M, Girard S, Momcorge R. ƒ-d luminescence of Pr3+ and Ce3+ in

the chloro-elpasolite Cs2NaYCl6. Chem Phys Lett 1999; 311: 167-172.

[254] Levey CG, Glynn TJ, Yen WM. Excitation of the 1S0 state of Pr3+ in crystal hosts. J

Lumin 1984; 31–32: 245-247.

[255] Gumanskaya EG, Korzhik MV, Smirnova SA, Pawlenko VB, Fedorov AA.

Interconfiguration luminescence of Pr3+ ions in Y3Al5O12 and YAlO3 single-crystals.

OPTIKA SPEKTROSKOPIYA 1992; 72: 155-159.

[256] Heerdt MLHT, van der Kolk E, Yen WM, Srivastava AM. Vacuum ultraviolet

spectroscopy of Pr3+ in CaAl4O7, LaMgAl11O19 and SrLaAlO4. J Lumin 2002; 100:

107-113.

[257] You FT, Huang SH, Meng CX, Wang DW, Xu JH, Huang Y, Zhang GB. 4ƒ5d

configuration and photon cascade emission of Pr3+ in solids. J Lumin 2007; 122–123:

58-61.

[258] van Dijk JMF, Schuurmans MFH. On radiative and non-radiative decay-rates and a

modified exponential energy-gap law for 4f-4f transitions in rare-earth ions. J Chem

Phys 1983; 78: 5317-5323.

[259] Schuurmans MFH, van Dijk JMF. On radiative and non-radiative decay times in the

weak coupling limit. Physica B&C1984; 23: 131-155.

[260] Vink AP, Dorenbos P, van Eijk CWE. Observation of the photon cascade emission

process under 4f15d1 and host excitation in several Pr3+-doped materials. J Solid State

118
Chem 2003; 171: 308-312.

[261] Kück S, Sokólska I. Observation of photon cascade emission in Pr3+-doped LuF3 and

BaMgF4. Chem Phys Lett 2002; 364: 273-278.

[262] Sokólska I, Kück S. Observation of photon cascade emission in Pr3+-doped pervoskite

KMgF3. Chem Phys 2001; 270: 355-362.

[263] Masson NJML, Vink AP, Dorenbos P, Bos AJJ, Van Eijk CWE, Chaminade JP. Ce3+ and

Pr3+ 5d-energy levels in the [pseudo] perovskites KMgF3 and NaMgF3. J Lumin 2003;

101: 175-183.

[264] Kück S, Sokólska I. High energetic transitions in Pr3+-doped polycrystalline LiCaAlF6

and LiSrAlF6. J Electrochem Soc 2002; 149: J27-J30.

[265] Kück S, Sokólska I, Henke M, Osiac E. Quantum efficiency of 1S0 and 3P0,1 levels of

Pr3+ doped YF3. Chem Phys 2005; 310: 139-144.

[266] van der Kolk E, Dorenbos P, van Eijk CWE. Vacuum ultraviolet excitation of 1S0 and

3
P0 emission of Pr3+ in Sr0.7La0.3Al11.7Mg0.3O19 and SrB4O7. J Phys: Condens Matter

2001; 13: 5471-5486.

[267] Liang HB, Wang J, Ye X, Ti ZF, Lin HH, Su Q, Tao Y, Xu JH, Huang Y, Zhang GB, Fu

YB. The VUV-vis luminescent properties of Ln3+ [Ln = Ce, Pr, Tb] in

Sr0.96Na0.02Ln0.02B4O7. J Alloys Compd 2006; 425: 307-313.

[268] Brixner LH, Crawford MK, Blasse G. Optical luminescence of electronic and vibronic

transitions in Gd2-xYx[SO4]3·8H2O. J Solid State Chem 1990; 85: 1-7.

[269] Huang SH, Wang XJ, Chen BJ, Jia D, Yen WM. Photon cascade emission and quantum

efficiency of the 3P0 level in Pr3+-doped SrAl12O19 system. J Lumin 2003; 102–103:

119
344-348.

[270] Donega CDM. Vibronic spectroscopy of Pr3+ in solids. Thesis: Utrecht University, The

Netherlands, 1994.

[271] Balda R, Fernández J, Saéz de Ocáriz I, Voda, M, García AJ, Khaidukov N. Laser

spectroscopy of Pr3+ ions in LiKY1-xPrxF5 single crystals. Phys Rev B 1999; 59:

9972-9980.

[272] Malinowski M, Woliński W, Wolski R, Strçek W. Excited-state kinetics and

energy-transfer in Pr3+-doped YAG. J Lumin 1991; 48–49: 235-238.

[273] Dornauf H, Heber J. Concentration-dependent fluorescence-quenching in La1−xPrxP5O14.

J Lumin 1980; 22: 1-16.

[274] Lorenzo A, Bausá LE, Solé JG. Optical spectroscopy of Pr3+ ions in LiNbO3. Phys Rev

B 1995; 51: 16643-16650.

[275] Oskam KD, Houtepen AJ, Meijerink A. Site selective 4ƒ5d spectroscopy of CaF2:Pr3+. J

Lumin 2002; 97: 107-114.

[276] Vink AP, Dorenbos P, van Eijk CWE. Thermal population of the 4f15d1 state in

BaSO4:Pr3+. Phys Rev B 2002; 66: 075118.

[277] Vink AP, van der Kolk E, Dorenbos P, van Eijk CWE. Luminescent properties of Ce3+

in MSO4 [M: Ca, Sr and Ba] and the effect of Na+ co-doping. Opt Comm 2002; 210:

277-284.

[278] Yamashita N, Yamamoto I, Ninagawa K, Wada T, Yamashita Y, Nakao Y. Investigation

of TLD phosphors by optical-excitation-luminescence of Eu2+ centers in MgSO4,

CaSO4, SrSO4 and BaSO4. Jpn. J Appl Phys 1985; 24: 1174-1180.

120
[279] Wang JW, Turos-Matysiak R, Grinberg M, Yen WM, Meltzer RS. Mixing of the ƒ2[1S0]

and 4ƒ5d states of Pr3+ in BaSO4 under high pressure. J Lumin 2006; 119–120:

473-477.

[280] Huang SH, Wang XJ, Meltzer RS, Srivastava AM, Setlur AA, Yen WM. The mixing of

the 4ƒ2 1S0 state with the 4ƒ5d states in Pr3+ doped SrAl12O19. J Lumin 2001; 94–95:

119-122.

[281] Huang SH, Lu L, Jia W, Wang XJ, Yen WM, Srivastava AM, Setlur AA. The spectra

property of the 1S0 state in SrAl12O19:Pr. Chem Phys Lett 2001; 348: 11-16.

[282] Rodnyi PA, Dorenbos P, Stryganyuk GB, Voloshinovskii AS, Potapov AS, van Eijk

CWE. Emission of Pr3+ in SrAl12O19 under vacuum ultraviolet synchrotron excitation J

Phys: Condens Matter 2003; 15: 719-729.

[283] Guillot-Noёl O, de Haas JTM, Dorenbos P, van Eijk CWE, Krämer K, Güdel HU.

Optical and scintillation properties of cerium-doped LaCl3, LuBr3 and LuCl3. J Lumin

1999; 85: 21-35.

[284] Visser R, Dorenbos P, van Eijk CWE, Meijerink A, Blasse G, den Hartog HW.

Energy-transfer process involving diffent luminescence centers in BaF2:Ce. J Phys:

Condens Matter 1993; 5: 1659-1680.

[285] van't Spijker JC, Dorenbos P, van Eijk CWE, Krämer K, Güdel HU. Scintillation and

luminescence properties of Ce3+ doped K2LaCl5. J Lumin 1999; 85: 1-10.

[286] van der Kolk E, Dorenbos P, van Eijk CWE. Vacuum ultraviolet excitation and quantum

splitting of Pr3+ in LaZrF7 and α-LaZr3F15. Opt Comm 2001; 197: 317-326.

[287] Wojtowicz AJ, Szupryczynskii P, Glodo J, Drozdowski W, Wisniewski D.

121
Radioluminescence and recombination processes in BaF2:Ce. J Phys: Condens Matter

2000; 12: 4097-4124.

[288] Xiong FB, Lin Y F, Chen YJ, Luo ZD, Ma E, Huang YD. Visible quantum cutting in

Pr3+:PbWO4 crystal through energy transfer between the host and Pr3+ ions. Chem Phys

Lett 2006; 429: 410-414.

[289] Rodnyi PA, van Eijk CWE, Mishin AN, Mikhrin SB, Avanesov AG, Potapov AS.

Interconfiguration and intraconfiguration transitions of Pr3+ in some oxides and

fluorides. Proc of SPIE 2002; 4766: 165-170.

[290] Dorenbos P. 5d-level energies of Ce3+ and the crystalline environment. I. Fluoride

compounds. Phys Rev B 2000; 62: 15640-15649.

[291] Dorenbos P. 5d-level energies of Ce3+ and the crystalline environment. II. Chloride,

bromide, and iodide compounds. Phys Rev B 2000; 62: 15650-15659.

[292] Kück S, Sokólska I, Henke M, Scheffler T, Osiac E. Emission and excitation

characteristics and internal quantum efficiencies of vacuum-ultraviolet excited

Pr3+-doped fluoride compounds. Phys Rev B 2005; 71: 165112.

[293] Lupei V, Lupei A. Energy transfer effects in the VUV-to-VIS quantum cutting in

Praseodymium-activated phosphors. Proc SPIE 2004; 5581: 238-244.

[294] Alivisatos AP. Perspectives on the physical chemistry of semiconductor nanocrystals. J

Phys Chem 1996; 100: 13226-13239.

[295] Riwotzki K, Haase M. Wet-chemical synthesis of doped colloidal nanoparticles:

YVO4:Ln (Ln = Eu, Sm, Dy). J Phys Chem B 1999; 102: 10129-10135.

[296] Riwotzki K, Meyssamy H, Kornowski A, Haase M. Liquid-phase synthesis of doped

122
nanoparticles: colloids of luminescing LaPO4:Eu and CePO4:Tb particles with a narrow

particle size distribution. J Phys Chem B 2000; 104: 2824-2828.

[297] Riwotzki K, Haase M. Colloidal YVO4:Eu and YP0.95V0.05O4:Eu nanoparticles:

luminescence and energy transfer processes. J Phys Chem B 2001; 105: 12709-12713.

[298] Krebs JK, Feofilov SP, Kaplyanskii AA, Zakharchenya RI, Happek U. Non-radiative

relaxation of Yb3+ in highly porous γ-Al2O3. J Lumin 1999; 83–84: 209-213.

[299] Trojan-Piegza J, Zych E, Hreniak D, Stręk W. Comparison of spectroscopic properties

of nanoparticulate Lu2O3:Eu synthesized using different techniques. J Alloys Compd

2004; 380:123-129.

[300] Wei ZG, Sun LD, Jiang XC, Liao CS, Yan CH, Tao Y, Zhang J, Hu TD, Xie YN.

Correlation between size-dependent luminescent properties and local structure around

Eu3+ ions in YBO3:Eu nanocrystals: An XAFS Study. Chem Mater 2003; 15:

3011-3017.

[301] Schmechel R, Kennedy M, Seggern HV, Winkler H, Kolbe M, Fischer RA, Li XM,

Benker A, Winterer M, Hahn H. Luminescence properties of nanocrystalline Y2O3:Eu3+

in different host materials. J Appl Phys 2001; 89: 1679-1686.

[302] Kömpe K, Borchert H, Storz J, Lobo A, Adam S, Möller T, Haase M. Green-emitting

CePO4:Tb/LaPO4 core-shell nanoparticles with 70 % photoluminescence quantum yield.

Angew Chem Int Ed 2003; 42: 5513-5516.

[303] Feldmann C. Polyol-mediated synthesis of nanoscale functional materials. Adv Funct

Mater 2003; 13: 101-107.

[304] Srivastava AM. Encyclopedia of Physical Science and technology, 3rd ed, Academic

123
Press: 2002; Vol. 11, p. 855. Srivastava AM. Handbook of Luminescence, Display

Materials, and Devices; Nalwa HS., Rowher LS., Eds, American Scientific Publishers:

2003; Vol 3, p 79.

[305] Alexander G, Ramakrishnan P, Mukherjee TK, Nambi KSV. Measurement of quantum

and cathodoluminescent efficiencies of inorganic powder phosphors using silicon

photodiode detector. Ind. J Pure Appl Phys 1993; 31: 531-538.

[306] Digonnet MJF. Rare-Earth-Doped Fiber Lasers and Amplifiers. New York: Marcel

Dekker; 2001.

[307] Moine B, Bizarri G, Varrel B, Rivoire JY. VUV-extended measurements of quantum

efficiency of sodium salicylate and of some NBS standard phosphors. Opt Mater 2007;

29: 1148-1152.

[308] Do YR, Bae JW. Application of photoluminescence phosphors to a phosphor-liquid

crystal display. J Appl Phys 2000; 88: 4660-4665.

[309] Murase N, Li CL. Consistent determination of photoluminescence quantum efficiency

for phosphors in the form of solution, plate, thin film, and powder. J Lumin 2008; 128:

1896-1903.

[310] Wang XJ, Huang SH, Lu LZ, Yen WM, Srivastava AM, Setlur AA. Measurement of

quantum efficiency in Pr3+-doped CaAl4O7 and SrAl4O7 crystals. Appl Phys Lett 2001;

79: 2160-2162.

[311] Lima SM, Andrade AA, Lebullenger R, Hernandes AC, Catunda T, Baesso ML.

Multiwavelength thermal lens determination of fluorescence quantum efficiency of

solids: Application to Nd3+-doped fluoride glass. Appl Phys Lett 2001; 78: 3220-3222.

124
[312] Baesso ML, Bento AC, Andrade AA, Sampaio JA, Nunes LAO, Catunda T, Gama S.

Absolute thermal lens method to determine fluorescence quantum efficiency and

oncentration quenching of solids. Phys Rev B 1998; 57: 10545-10549.

[313] Rohwer LS, Martin JE. Measuring the absolute quantum efficiency of luminescent

materials. J Lumin 2005; 115: 77-90.

[314] Abrams BL, Wilcoxon JP. Comment on “Measuring the absolute quantum efficiency of

luminescent materials by L. Rohwer, J. Martin, J Lumin 2005;115:77-90”. J Lumin

2009; 129: 329-330.

[315] Jüstel T, Krupa J, Wiechert DU. VUV spectroscopy of luminescent materials for plasma

display panels and Xe discharge lamps J Lumin 2001; 93: 179-189.

[316] Nygaard KJ. The variation of the quantum efficiency of sodium salicylate with

thickness of material. Br J Appl Phys 1964; 15: 597-599.

[317] Kück S, Sokólska I, Henke M, Döring M, Scheffler T. Photon cascade emission in

Pr3+-doped fluorides. J Lumin 2003; 102–103: 176-181.

[318] Bowlby BE, di Bartolo B. Applications of the Judd-Ofelt theory to the praseodymium

ion in laser solids. J Lumin 2002; 100: 131-139.

[319] Hölsä J, Lastusaari M, Maryško M, Tukia M. A few remarks on the simulation and use

of crystal field energy level schemes of the rare earth ions. J Solid State Chem 2005;

178: 435-440.

[320] Tanner PA, Mak CSK, Kwok WM, Phillips DL, Joubert MF. Luminescence from the 3P2

State of Tm3+. J Phys Chem B 2002; 106: 3606-3611.

[321] Lin J, Su Q. Luminescence and energy migration in the oxyapatite Ca2Gd8(SiO4)6O2

125
doped with several rare earth and mercury-like ions. J Alloys Compd 1994; 210:

159-263.

[322] Kiliaan HS, Kotte JFAK, Blasse G. Energy-transfer in the luminescent system Na(Y,

Gd)F4–Ce, Tb. J Electrochem Soc 1987; 134: 2359-2364.

[323] Blasse G, Kiliaan HS, De Vries A J. A study of the energy transfer processes in

sensitized gadolinium phosphors. J Less-Common Met 1986; 126: 139-146.

[324] Blasse G. The physics of new luminescent materials. Mater Chem Phys 1987; 16:

201–236.

[325] Khaidukov NM, Lam SK, Lo D, Makhov VN, Suetin NV. Luminescence spectroscopy

from the vacuum ultra-violet to the visible for Er3+ and Tm3+ in complex fluoride

crystals. Opt Mater 2002;19:365-376.

[326] Liu CX Liu JY, Lu SZ, Chen BJ, Zhang JH. Energy migration and transfer of

Tm3+-Gd3+-Dy3+ system in NaGdF4 under VUV and UV excitations. J Lumin 2007;

122–123: 970-972.

[327] Takeuchi N, Ishida S, Matsumura A, Ishikawa Y. Time-resolved study of luminescence

in LiGd1-xF4: Eux. J Phys Chem B 2004; 108: 12397-12403.

[328] Vergeer P, van den Pol E, Meijerink A. Time and temperature dependence of the

emissions from the quantum-cutting phosphor LiGdF4:Eu3+. J Lumin 2006; 121:

456-464.

[329] Jung KY, Lee DY, Kang YC. Improved thermal resistance of spherical

BaMgAl10O17:Eu blue phosphor prepared by spray pyrolysis. J Lumin 2005; 115:

91-96.

126
[330] Solarz P. Pr3+ as a sensitiser of red Eu3+ luminescence in K5Li2GdF10:Pr3+,Eu3+ upon

VUV–UV excitation. Opt Mater 2008; 31: 114-116.

[331] Solarz P, Ryba-Romanowski W. Energy transfer processes in K5Li2GdF10:Eu,Pr. Radiat

Meas 2007; 42: 759-762.


[332] Ryba-Romanowski W, Solarz P, Gusowski M, Dominiak-Dzik G. Luminescence and

excitation energy transfer in newfluoride crystals containing rare earth ions. Radiat

Meas 2007; 42: 798-802.

[333] Gusowski MA, Dzik GD, Solarz P, Lisiecki R, Ryba-Romanowski W. Luminescence

and energy transfer in K3GdF6:Pr3+. J Alloys Compd 2007; 438: 72-76.

[334] van der Kolk E, Dorenbos P, Krämer K, Biner D, Güdel HU. High-resolution

luminescence spectroscopy study of down-conversion routes in NaGdF4:Nd3+ and

NaGdF4:Tm3+ using synchrotron radiation. Phys Rev B 2008; 77: 125110.

[335] Hirai T, Hashimoto S, Sakuragi S, Ohno N. 4ƒ–5d absorption of gadolinium ions in

sodium gadolinium tetrafluorides. Chem Phys Lett 2007; 446: 138-141.

[336] Lee TJ, Luo LY, Cheng BM, Diau WG, Chen TM. Investigation of Pr3+ as a sensitizer in

quantum-cutting fluoride phosphors. Appl Phys Lett 2008; 92: 081106.

[337] Liu B, Shi CS, Qi ZM, Tao Y. Luminescence spectra of SrAl12O19:Pr3+, Mn2+ under

VUV-UV excitation. Chin Phys Lett 2005; 22: 2677-2679.

[338] Utsunomiya A, Tanaka K, Morikawa H, Marumo F. Structure refinement of Ca0.6Al2O3.

J Solid State Chem 1988; 75: 197-200.

[339] Kimura K, Ohgaki M, Tanaka K, Morikawa H, Marumo F. Study of the bipyramidal site

in magnetoplumbite-like compounds, SrAl12O19, SrFe12O19, SrGa12O19. J Solid State

Chem 1990; 87: 186-194.

127
[340] Tippins HH. Charge-transfer spectra of transition-metal ions in corundum. Phys Rev B

1970; 1: 126-131.

[341] Henderson B, Imbush GF. Optical Spectroscopy of Inorganic Solids. Oxford: Oxford

Science Publications; 1989.

[342] Sugano SJ, Tanabe Y, Kamikura H. Multiplets of Transitions Metals Ions in Crystals.

New York: Academic;1970.

[343] Lushchik A, Lushchik C, Kotlov A, Kudryavtseva I, Maaroos A, Nagirnyi V,

Vasil’chenko E. Spectral transformers of VUV radiation on the basis of wide-gap

oxides. Radiat Meas 2004; 38: 747-752.

[344] Lakshmanan AR. Photoluminescence and thermostimulated luminescence processes in

rare-earth-doped CaSO4 phosphors. Prog Mater Sci 1999; 44: 1-187

[345] Green MA. Third Generation Photovoltaics, Advanced Solar Energy Conversion. Berlin

(Germany): Springer; 2003; Luque A, Hegedus S. Handbook of Photovoltaic Science

and Engineering. Chichester (UK): Wiley; 2003.

[346] van Sark WGJHM. Enhancement of solar cell performance by employing planar

spectral converters. Appl Phys Lett 2005; 87: 151117.

[347] Maruyama T, Kitamura R. Transformations of wavelength of the light incident upon

solar cells. Sol Energy Mater Sol Cells 2001; 69: 207-216.

[348] Timmerman D, Izeddin I, Stallinga P, Yassievich IN, Gregorkiewicz T.

Space-separated quantum cutting with silicon nanocrystals for photovoltaic

applications. Nature Photon 2008; 2: 105-109.

[349] Strümpel C, McCann M, Beaucarne G, Arkhipov V, Slaoui A, Švrček V, del Cañizo C,

128
Tobias I. Modifying the solar spectrum to enhance silicon solar cell efficiency—An

overview of available materials. Sol Energy Mater Sol Cells 2007; 91: 238-249.

[350] Andrews DL, Jenkins RD. A quantum electrodynamical theory of three-center energy

transfer for upconversion and downconversion in rare earth doped materials. J Chem

Phys 2001; 114: 1089-1100.

[351] Auzel F. Upconversion procession in coupled ion systems. J Lumin 1990; 45: 341-345.

[352] Auzel F. Upconversion and anti-stokes processes with f and d ions in solids. Chem Rev

2004; 104: 139-173.

[353] Maciel GS, Biswas A, Kapoor R, Prasad PN. Blue cooperative upconversion in

Yb3+-doped multicomponent sol-gel processed silica glass for three-dimensional

display. Appl Phys Lett 2000; 76: 1978-1980.

[354] Martins E, de Araújo CB, Delben JR, Gomes ASL, da Costa BJ, Messaddeq Y.

Cooperative frequency up conversion in Yb3+-Tb3+ codoped fluoroindate glass. Opt

Commun 1998; 158: 61-64.

[355] Ostermayer FW, Jr Van Uitert LG. Cooperative energy transfer from Yb3+ to Tb3+ in YF3.

Phys Rev B 1970; 1: 4208-4212.

[356] Martin IR, Yanes AC, Mendez-Ramos J, Torres ME, Rodriguez VD. Cooperative

energy transfer in Yb3+–Tb3+ codoped silica sol-gel glasses. J Appl Phys 2001; 89:

2520-2524.

[357] Salley GM, Valiente R, Güdel HU. Cooperative Yb3+-Tb3+ dimer excitations and

upconversion in Cs3Tb2Br9:Yb3+. Phys Rev B 2003; 67: 134111.

[358] Zhang QY, Yang ZM, Yang GF, Jiang ZH. Enhanced blue-green-red up-conversion and

129
1.3-mm emission of Pr3+/Yb3+-codoped oxyhalide tellurite glasses with PbCl2 doping. J

Phys Chem Solids 2005; 66: 1281-1286.

[359] Zhang QY, Li T, Jiang ZH, Ji XH, Buddhudu S. 980 nm laser-diode-excited intense blue

upconversion in Tm3+/Yb3+-codoped gallate-bismuth-lead glasses. Appl Phys Lett 2005;

87: 171911.

[360] Zhang QY, Li T, Shi DM, Yang GF, Yang ZM, Jiang ZH, Buddhudu S. Effects of PbF2

doping on structure and spectroscopic properties of Ga2O3–GeO2–Bi2O3–PbO glasses

doped with rare earths. J Appl Phys 2006; 99: 033510.

[361] Zhao C, Zhang QY, Yang GF, Jiang ZH. Laser-diode-excited blue upconversion in

Tm3+/Yb3+-codoped TeO2-Ga2O3-R2O (R=Li,Na,K) glasses. J Fluoresc 2008; 18: 87-91.

[362] Oliveira AS, Gouveia EA, de Araujo MT, Gouveia-Neto AS, de Araújo CB, Messaddeq

Y. Twentyfold blue upconversion emission enhancement through thermal effects in

Pr3+/Yb3+-codoped fluoroindate glasses excited at 1.064 µm. J Appl Phys 2000; 87:

4274-4278.

[363] Osiac E, Heumann E, Huber G, Kück S. Orange and red upconversion laser pumped by

an avalanche mechanism in Pr3+,Yb3+:BaY2F8. Appl Phys Lett 2003; 82: 3832-3834.

[364] Suyver JF, Grimm J, van Veen MK, Biner D, Krämer KW, Güdel HU. Upconversion

spectroscopy and properties of NaYF4 doped with Er3+, Tm3+ and/or Yb3+. J Lumin

2006; 117: 1-12.

[365] Matsuura D. Red, green, and blue upconversion luminescence of trivalent-rare-earth

ion-doped Y2O3 nanocrystals. Appl Phys Lett 2002; 81: 4526-4528.

[366] Patra A, Saha S, Alencar MARC, Rakov N, Maciel GS. Blue upconversion emission of

130
Tm3+–Yb3+ in ZrO2 nanocrystals: role of Yb3+ ions. Chem Phys Lett 2005; 407:

477-481.

[367] Pandozzi F, Vetrone F, Boyer JC, Naccache R, Capobianco JA, Speghini A, Bettinelli

M. A spectroscopic analysis of blue and ultraviolet upconverted emissions from

Gd3Ga5O12:Tm3+,Yb3+ nanocrystals. J Phys Chem B 2005; 109: 17400-17405.

[368] Shi DM, Zhang QY. Enhanced 1.47 mu m emission and lowered upconversion of

Tm3+-doped gallate-germanium-bismuth-lead glass by codoping rare earths. J Appl

Phys 2008; 104: 123517.

[369] Zhang F, Wan Y, Shi YF, Tu B, Zhao DY. Ordered mesostructured rare-earth fluoride

nanowire arrays with upconversion fluorescence. Chem Mater 2008; 20: 3778-3784.

[370] Wang YH, Xie LC, Zhang HJ. Cooperative near-infrared quantum cutting in Tb3+, Yb3+

codoped polyborates La0.99−xYbxBaB9O16:Tb0.01. J Appl Phys 2009; 105: 023528.

[371] Xie LC, Wang YH, Zhang HJ. Near-infrared quantum cutting in YPO4:Yb3+, Tm3+ via

cooperative energy transfer. Appl Phys Lett 2009; 94: 061905.

[372] van der Ende BM, Aarts L, Meijerink A. Near-infrared quantum cutting for

photovoltaics, Adv Mater 2009; 21: 3073.

[373] X. P. Chen, X. Y. Huang, Zhang QY. Concentration-dependent near-infrared quantum

cutting in NaYF4:Pr3+,Yb3+ phosphor. J Appl Phys 2009; 106: 063518.

[374] Faulkner S, Pope SJA. Lanthanide-sensitized lanthanide luminescence: terbium-

sensitized ytterbium luminescence in a trinuclear complex. J Am Chem Soc 2003; 125:

10526-10527.

[375] Lakshminarayana G, Yang HC, Ye S, Liu Y, Qiu JR. Co-operative downconversion

luminescence in Tm3+/Yb3+: SiO2–Al2O3–LiF–GdF3 glasses. J Phys D: Appl Phys 2008; 41:

131
175111.

[376] Lakshminarayana G, Yang HC, Ye S, Liu Y, Qiu JR. Cooperative downconversion

luminescence in Pr3+/Yb3+: SiO2–Al2O3–BaF2–GdF3 glasses. J Mater Res 2008; 23:

3090-3095.

[377] Lakshminarayana G, Qiu JR. Near-infrared quantum cutting in RE3+/Yb3+ (RE = Pr, Tb,

and Tm): GeO2–B2O3–ZnO–LaF3 glasses via downconversion. J Alloys Compd 2009;

481: 582-589.

[378] Liu X, Ye S, Qiao Y, Dong G, Zhu B, Chen D, Lakshminarayana G, Qiu J. Cooperative

downconversion and near-infrared luminescence of Tb3+–Yb3+ codoped lanthanum

borogermanate glasses. Appl Phys B 2009; 96: 51-55.

[379] Chen D, Wang Y, Yu Y, Huang P, Weng F. Quantum cutting downconversion by

cooperative energy transfer from Ce3+ to Yb3+ in borate glasses. J Appl Phys 2008; 104:

116105.

[380] Chen JX, Ye S, Wang X, Qiu JR. Cooperative quantum cutting of nano-crystalline

BaF2:Tb3+,Yb3+ in oxyfluoride glass ceramics. Chin Phys Lett 2008; 25: 2078-2080.

[381] Ye S, Zhu B, Luo J, Chen JX, Lakshminarayana G, Qiu JR. Enhanced cooperative

quantum cutting in Tm3+-Yb3+ codoped glass ceramics containing LaF3 nanocrystals.

Opt Exp 2008; 16: 8989-8994.

[382] Chen DQ, Wang YS, Yu YL, Huang P, Weng FY. Near-infrared quantum cutting in

transparent nanostructured glass ceramics. Opt Lett 2008; 33: 1884-1886.

[383] Chen DQ, Yu YL, Wang YS, Huang P, Weng FY. Cooperative energy transfer

up-conversion and quantum cutting down-conversion in Yb3+:TbF3 nanocrystals

embedded glass ceramics. J Phys Chem C 2009; 113: 6406-6410.

132
Table and Table captions

Table 1. Maximum phonon energies ћωmax (cm-1) of several inorganic materials.

Materials ћωmax (cm-1)

Borate 1400

Phosphate 1100

Silicate 1000-1100

Germanate 800-975

Tellurite 600-850

Fluoride 500-600

Chalcogenide 200-300

Bromide 175-190

Iodide 160

133
Table 2. Predicted lowest energy 4ƒ2→ 4ƒ5d transition of Pr3+ in host lattices that are

expected to give 1S0 emission. (Adapted from Ref. [68].)

Host lattice Predicted 4f2→4f5d Host lattice Predicted 4f2→4f5d


Energy (cm-1) Energy (cm-1)
Free ion 61 580 La2(SO4)3 49 277

KMgF3 54 975 K2NaScF6 49 140

BaSiF6 53 192 KGdF4 49 140

LaF3 52 829 YMg5O10 49 140

NaYF4 52 726 GdB3O6 49 054

LiBaF3 52 563 LaMgB5O10 49 012

Aqueous-[Ce(OH2)9]3+ 51 766 CaMgAl11.33O19 49 005

YF3 51 664 KMgF3(C4υ) 49 005

NaGdF4 51 456 LaPO4 48 802

BaThF6 51 303 LiCaAlF6 48 236

La(C2H5SO4)3·9H2O 51 303 LiLaP4O12 48 211

LuF3 51 006 Sr2B5O9Cl 47 954

PbThF6 50 850 LaCl3 47 827

LiMgAlF6 50 702 SrB4O7 47 500

BaMgF4 50 554 La9.33□0.67(SiO4)6O2:4f 47 328

SrAl12O19 50 530 KYF4 47 083

CaAl12O19 49 976 BaLu2F8 46 962

CsY2F7 49 976 BaF2 46 940

SrAlF5 49 693 Gd9.33□0.67(SiO4)6O2:4f 46 842

LiSrAlF6 49 415 SrB6O10 46 842

LaB3O6 49 294 LaP3O9 46 772

LaMgAl11O19 49 277 K2NaYF6 46 729

134
Table 3. Observed lowest energy 4ƒ7→ 4ƒ65d transition of Eu2+ in host lattices that are

expected to give 1S0 emission. (Adapted from Ref. [68].)

Host lattice Observed 4ƒ7→ 4ƒ65d energy of Eu2+


BaSiF6 33 000
SrSiF6 32 787
KMgF3 32 500
KMgF3 (С4υ) 30 769
α-BaCaAlF4 30 121
BaMg(SO4)2 30 121
β-SrBeF4 30 030
γ-SrBeF4 30 030
LiBaF3 30 000
α-BaAlF5 29 940
RbMgF3 29412
BaFCl 29 325
SrAlF5 28 990
BaBe2(BO3)2 28 986
SrFCl 28 986
BaBeF4 28 818
BaMgF4 28 090
SrB4O7 28 090
CaFCl 28 011
CaBeF4 27 701
SrBe6O10 27 550
CaB2O(Si2O7) 27 030
BaF2 26 530
Ca2B5SiO9(OH)5 26 385
RbCl 26 316
SrF2 25 970
6
SrBe2Si2O7 P7/2→ 8S7/2 emission

135
Table 4. Gd3+–Eu3+ dual ions activated phosphors: materials, synthetic methods and

properties.

ion pair host lattices synthetic methods Eu3+ concentration (%) QE Ref.
LiGdF4 solid-state reaction 0.5 190% [79]
LiGdF4 sol-gel method 5.0 175% [194]
NaGdF4 hydrothermal method 0.5 160% [83]
BaF2 solid-state reaction 1.0 194% [86]
KGd3F10 solid-state reaction 2.0 165% [87]
Gd3+–Eu3+
RbGd3F10 hydrothermal method 0.5 150% [85]
KGdF4 hydrothermal method 1.0 175% [84]
KLiGdF5 solid-state reaction 2.0 140% [88]
CsGd2F7 solid-state reaction 0.5 150% [89]
GdF3 hydrothermal method 0.5 170% [90]

136
Table 5. NIR QC phosphors: materials, luminescence properties and QE.

Couple Host Preparation Excitation Maximum Ref.


method wavelength QE (%)
(nm)
YPO4 solid-state 489 188 [122]
KYb(WO4)2 crystal growth 308 - [125,126]
GdAl3(BO3)4 combustion 485 179.4 [133]
GdBO3 combustion 486 198 [132]
Y3Al5O12 combustion 487 191 [134]
Y2O3 coprecipitation 304 137 [136]
LaBaB9O16 solid-state 481 152.5% [370]
Zn2SiO4 sol-gel 485 162.2 [135]
GeO2–B2O3–ZnO–LaF3 glass melting 483 166.3 [377]
glasses
Tb3+-Yb3+ lanthanum glass melting 484 146 [378]
borogermanate glasses
glass-ceramics containing glass melting 485 145 [380]
BaF2 nanocrystals
glass-ceramics containing glass melting 484 155 [137]
CaF2 nanocrystals
glass-ceramics containing glass melting 485 141 [383]
TbF3 nanocrystals
GdAl3(BO3)4 combustion 475 180.8 [133]
Y3Al5O12 combustion 473 165 [134]
YPO4 hydrothermal 474 172.8% [371]
SiO2–Al2O3–LiF–GdF3 glass melting 467 187 [375]
Tm3+-Yb3 glasses
GeO2–B2O3–ZnO–LaF3 glass melting 467 191.4 [377]
glasses
glass-ceramics containing glass melting 468 162 [381]
LaF3 nanocrystals
GdAl3(BO3)4 combustion 489 195.4 [133]
Y3Al5O12 combustion 488 173 [134]
SrF2 solid-state 441 140 [372]
Pr3+-Yb3+ NaYF4 hydrothermal 420-500 133.6 [373]
GeO2–B2O3–ZnO–LaF3 glass melting 482 183.6 [377]
glasses
glass-ceramics containing glass melting 482 194.3 [382]
YF3 nanocrystals
Ce3+-Yb3+ B2O3–BaO–CaO–La2O3 glass melting 330 174 [379]
glasses

137
Figure captions

Fig. 1. Schematic energy-level diagram of Pr3+ ion, showing the concept of QC or PCE with

VUV excitation.

Fig. 2. A schematic diagram to illuminate the different ET processes between two ions: (a)

resonant radative transfer through emission of sensitizer (S) and reabsorption by activator (A);

(b) non-radative transfer associated with resonance between absorber (sensitizer) and emitter

(activator); (c) multiphonon assisted ET; and (d) CR between two identical ions [80].

Reproduced from Spectroscopic Properties of Rare Earths in Optical Materials, 2005,

Copyright © 2005, Springer.

Fig. 3. Energy level diagrams for two (hypothetical) types of RE ions (I and II) showing the

concept of DC. Type I is an ion for which emission from a high energy level can occur. Type

II is an ion to which ET takes place. (a) QC on a single ion I by the sequential emission of

two visible photons. (b) The possibility of QC by a two-step ET. In the first step (indicated by

①), a part of the excitation energy is transferred from ion I to ion II by CR. Ion II returns to

the ground state by emitting one photon of visible light. Ion I is still in an excited stated and

can transfer the remaining energy to a second ion of type II (indicated by ②), which also

emits a photon in the visible spectral region, giving a QE of 200%. (c and d) The remaining

two possibilities involve only one ET step from ion I to ion II. This is sufficient to obtain

visible QC if one of two visible photons can be emitted by ion I [81]. Reproduced from J.

Lumin., 1999, 82, 93. Copyright © 1999, Elsevier Science B.V.

Fig. 4. Schematic representation of the excited states of Pr3+. The most probable radiative

transitions are indicated by the vertical arrows for the situation where the lowest 4ƒ5d state

has a higher energy (a) or a lower energy (b) than the 1S0 state [67]. Reproduced from Phys.

Rev. B, 2001, 64, 195129. Copyright © 2001, The American Physical Society.

138
Fig. 5. Excitation and emission spectra of YPO4:Pr3+ and YBO3:Pr3+ at 293 K. (a) λemi=263

nm, (b) λexc=188 nm, (c) λemi=276 nm, (d) λexc=185 nm [257]. Reproduced from J. Lumin.,

2007, 122&123, 58. Copyright © 2006, Elsevier Science B.V.

Fig. 6. Excitation and emission spectra of Pr3+-doped YF3 and KMgF3 measured at RT.

Excitation spectra (corrected) were taken at a detection wavelength of about 400 nm

(corresponding to the 1S0→ 1I6 transition). Emission spectra (uncorrected) were recorded for

excitation in 4ƒ5d states (λexc =180 nm), resolution 3.2 nm [262]. Reproduced from Chem.

Phys., 2001, 270, 355. Copyright © 2001, Elsevier Science B.V.

Fig. 7. Corrected emission spectra of the Pr3+ doped YF3 samples normalized to the emission

peak at 400 nm of the 1S0→ (1I6, 3P0,1,2) transition [265]. Reproduced from Chem. Phys., 2005,

310, 139. Copyright © 2004, Elsevier Science B.V.

Fig. 8. The time-integrated emission spectrum of SrB4O7: 1%Pr3+ recorded at liquid-He

temperature at 189 nm pulsed excitation [266]. Reproduced from J. Phys.: Condens. Matter.,

2001, 13, 5471. Copyright © 2001, IOP Publishing Ltd.

Fig. 9. Emission spectra of CaAl12O19:1% Pr3+ (a) and SrAl12O19:1% Pr3+ (b) upon 4f5d

excitation at 198 nm [113]. Reproduced from J. Phys.: Condens. Matter., 2007, 19, 076204.

Copyright © 2007, IOP Publishing Ltd.

Fig. 10. Emission spectra (λexc= 188 nm) of BaSO4:Pr3+ at T=10 K (dashed line) and 292 K

(solid line) [276]. Reproduced from Phys. Rev. B, 2002, 66, 075118. Copyright © 2002, The

American Physical Society.

139
Fig. 11. VUV-excited emission spectra of SrAlF5:Pr3+ measured at T = 10 K. The solid

spectrum (λexc = 190 nm) shows the PCE process, whereas the dotted spectrum (λexc = 111 nm)

shows the typical STE band around 360 nm [64]. Reproduced from J. Phys.: Condens.

Matter., 2002, 14, 8889. Copyright © 2002, IOP Publishing Ltd.

Fig. 12. A schematic representation of direct electron–hole pair recombination involving the

4ƒ5d states leading to Pr3+ 1S0 excitation (process I) and Pr3+ 3P0 excitation via an intermediate

STE state (process II) [266]. Reproduced from J. Phys.: Condens. Matter., 2001, 13, 5471.

Copyright © 2001, IOP Publishing Ltd.

Fig. 13. Polarized emission spectra under excitation at 204 nm (a) and excitation spectrum of the

325 nm emission (b) of Pr3+:PbWO4 crystal at room temperature [288]. Reproduced from Chem.

Phys. Lett., 2006, 429, 410. Copyright © 2006, Elsevier Science B.V.

Fig. 14. Partial energy levels diagram of Pr3+ ions and tungstate anions indicating QC and

energy transfers between Pr3+ ions and tungstate anions [288]. Reproduced from Chem. Phys.

Lett., 2006, 429, 410. Copyright © 2006, Elsevier Science B.V.

Fig. 15. X-ray-excited emission spectra of SrA1F5:Pr3+ measured at T = 100 K (dotted curve)

and T = 350 K (solid curve) [64]. Reproduced from J. Phys.: Condens. Matter., 2002, 14,

8889. Copyright © 2002, IOP Publishing Ltd.

Fig. 16. The lowest energy 4ƒn→ 4ƒn-15d transition E, of Ce3+, Pr3+ and Eu2+ plotted against

that of Ce3+ in the same host A plotted in such a way that the Ce3+ data points are on a straight

line [68]. Reproduced from J. Lumin., 2002, 97, 212. Copyright © 2002, Elsevier Science

B.V.

140
Fig. 17. Possible QC schemes for the Pr3+, Eu3+, Gd3+,Tb3+ and Tm3+ ions [319]. Reproduced

from J. Solid State Chem., 2005, 178, 435. Copyright © 2004, Elsevier Inc.

Fig. 18. Schematic energy level diagram for LaF3:Er3+ up to 65 000 cm–1. Energy levels from

which visible emission may be observed are indicated with a filled semi-circle [78].

Reproduced from Chem. Phys. Lett., 2005, 401, 241. Copyright © 2004, Elsevier B.V.

Fig. 19. Energy levels of Gd3+. The arrows and associated labels indicate the wavelengths of

the relevant transitions [74]. Reproduced from Phys. Rev. B, 2006, 74, 085101. Copyright ©

2006, The American Physical Society.

Fig. 20. Emission spectra excited at 160 nm of undoped and 1% Gd-doped ScPO4 at T=300 K.

The Gd3+ transitions and the STE are identified [75]. Reproduced from Phys. Rev. B, 2008, 77,

075129. Copyright © 2008, The American Physical Society.

Fig. 21. Schematic showing the creation of the initial e-h pair, the relaxation to form the STE,

and the subsequent energy flow to the excited states of Gd3+. The relevant energy levels of

Gd3+, along with the Gd3+ transitions observed are identified [75]. Reproduced from Phys. Rev.

B, 2008, 77, 075129. Copyright © 2008, The American Physical Society.

Fig. 22. Energy level diagram of the Gd3+–Eu3+ system, showing the possibility of visible QC

by two-step ET from Gd3+ to Eu3+ upon excitation in the 6GJ levels of Gd3+. The two ET steps

are indicated by ① (CR step) and ② (transfer of remaining energy on Gd3+ to Eu3+ ) [81].

Reproduced from J. Lumin., 1999, 82, 93. Copyright © 1999, Elsevier Science B.V.

Fig. 23. Emission spectra of LiGdF4:Eu3+ (0.5%) upon (a) 8S7/2→ 6GJ excitation on Gd3+ (202

nm) and (b) 8S7/2→ 6IJ excitation on Gd3+ (273 nm), both at 300 K. The spectra are scaled on

141
the 5D1→ 7FJ emission intensity [81]. Reproduced from J. Lumin., 1999, 82, 93. Copyright ©

1999, Elsevier Science B.V.

Fig. 24. Emission spectra of LiGdF4:0.5%Eu3+ upon 8S7/2→ 6GJ excitation on Gd3+ (202 nm)

at (a) 7 K, (b) 50 K and (c) 300 K [81]. Reproduced from J. Lumin., 1999, 82, 93. Copyright

© 1999, Elsevier Science B.V.

Fig. 25. VUV excitation spectra of GdPO4:Eu3+ x% (x= 1, 7 and 10), monitored by a low-pass

filter λ≥580 nm. C.T.B stands for charge transfer band [227]. Reproduced from Opt. Mater.,

2009, 31, 678. Copyright © 2009, Elsevier B.V.

Fig. 26. Comparison of PL spectra for K2GdF5:Eu3+,Pr3+ with λexc=274 nm (no QC effect),,

210 nm, and 172 nm (showing QC effect). The spectra were scaled to the 5D2→ 7F3 at 509 nm

excitation intensity [336]. Reproduced from Appl. Phys. Lett., 2008, 92, 081106. Copyright ©

2008, American Institute of Physics.

Fig. 27. Schematic energy levels of K2GdF5:Eu3+,Pr3+ showing possible mechanisms for

visible QC with excitation in the VUV with λexc= (a) 274 and (b) 210 nm (① and ② denote

the CR and direct ET, respectively) [336]. Reproduced from Appl. Phys. Lett., 2008, 92,

081106. Copyright © 2008, American Institute of Physics.

Fig. 28. Emission spectra of K2GdF5:Tb3+ (5%) excited at λexc= (a) 274, (b) 212, and (c) 172

nm. The spectra are scaled to the 5D3→ 7F5 excitation intensity (*) [91]. Reproduced from

Appl. Phys. Lett., 2006, 89, 131121. Copyright © 2006, American Institute of Physics.

Fig. 29. Exciation spectra of K2GdF5:Tb3+ (5%) monitored at (a) λemi=542 nm (5D4→ 7F5

emission of Tb3+) and (b) λemi=415 nm (5D3→ 7F6 emission of Tb3+). The spectra are scaled to

142
the 8S7/2→ 6IJ excitation intensity (*) [91]. Reproduced from Appl. Phys. Lett., 2006, 89,

131121. Copyright © 2006, American Institute of Physics.

Fig. 30. Schematic energy levels of K2GdF5:Tb3+ showing possible mechanisms for visible

QC under excitation of VUV with λexc=(a) 274, (b) 212, and (c) 172 nm; 1 and 2 denote cross

relaxation and direct energy transfer, respectively [91]. Reproduced from Appl. Phys. Lett.,

2006, 89, 131121. Copyright © 2006, American Institute of Physics.

Fig. 31. Energy level diagram for an Er3+–Gd3+–Tb3+ system, showing the possibility for

visible QC by making use of ET upon excitation in the 4ƒ105d state of Er3+. For the sake of

clarity, the energy level schemes of the ions are simplified [82]. Reproduced from J. Lumin.,

2000, 90, 111. Copyright © 2000, Elsevier Science B.V.

Fig. 32. Emission spectra of LiGdF4:Er3+,Tb3+ (1.5%, 0.3%) upon (a) 4ƒ11(4I15/2)→ 4ƒ105d

excitation on Er3+ (145 nm) and (b) 8S7/2 → 6IJ excitation on Gd3+ (273 nm), both at 300 K.

The spectra are scaled on the Tb3+ 5D3→ 7FJ emission intensity [82]. Reproduced from J.

Lumin., 2000, 90, 111. Copyright © 2000, Elsevier Science B.V.

Fig. 33. Excitation spectrum of YF3:1%Eu3+ (solid line) and emission spectrum of

YF3:1%Pr3+ (dashed line). The excitation and emission wavelengths are indicated in the figure.

Both spectra were recorded at RT. Note the spectral overlap between the Pr3+ 1S0→ 1I6, 3PJ

emissions and the Eu3+ 7F0,1→ 5D3, 5L6 absorptions [100]. Reproduced from J. Lumin., 2005,

114, 267. Copyright © 2005, Elsevier Science B.V.

Fig. 34. RT emission spectra of YF3 doped with 1% Pr3+ and codoped with Eu3+ plotted on a

relative intensity scale. The concentrations of Eu3+ are indicated in the figure. Excitation is

into the Pr3+ 3H4→ 4ƒ5d band at 190 nm. The Pr3+ 1S0→ 1I6, 3PJ and 3P0→ 3H4 transitions are

143
indicated in the figure [100]. Reproduced from J. Lumin., 2005, 114, 267. Copyright © 2005,

Elsevier Science B.V.

Fig. 35. The excitation spectrum (dashed line) of 4S3/2→ 4I15/2 emission of Er3+ in

CaAl12O19:Pr3+, Er3+ and the emission spectrum (solid line) of CaAl12O19:Pr3+, Er3+ under 205

nm excitation [101]. Reproduced from Opt. Commun., 2001, 195, 405. Copyright © 2001,

Elsevier Science B.V.

Fig. 36. Schematic of the Pr3+ cascade emission (left hand side) and the desired ET process

between Pr3+(1S0) and Mn2+ (right hand side). Solid lines denote radiative transitions. CT

stands for charge transfer. The ET process is depicted by dotted lines. The zigzag line denotes

non-radiative relaxation. After the ET process both Pr3+ and Mn2+ each emit one photon [106].

Reproduced from Opt. Mater., 2006, 28, 575. Copyright © 2006, Elsevier B.V.

Fig. 37. Emission spectra of SrB4O7:Pr3+ (a), SrB4O7:Pr3+,Mn2+ (b), and SrB4O7:Mn2+ (d) at

206 nm excitation, and the emission spectra of SrB4O7:Mn2+ excited by 410 nm (c, curve 2)

along with its excitation spectra monitoring 640 nm emission (c, curve 1) [107]. Reproduced

from Appl. Phys. Lett., 2006, 88, 061906. Copyright © 2006, American Institute of Physics.

Fig. 38. Part of the energy level scheme of Pr3+ and Mn2+ showing the transitions involved in

the ET from Pr3+ to Mn2+ and the emission processes [107]. Reproduced from Appl. Phys.

Lett., 2006, 88, 061906. Copyright © 2006, American Institute of Physics.

Fig. 39. Part of the energy level diagram of the Pr3+–Cr3+ pair in the SrAl12O19 (or CaAl12O19)

system. The dashed lines indicate possible nonradiative ET processes from Pr3+ to Cr3+. For

the sake of clarity, the energy level schemes of the ions are simplified [113]. eproduced from

J. Phys.: Condens. Matter., 2007, 19, 076204. Copyright © 2007, IOP Publishing Ltd.

144
Fig. 40. (a) VUV excitation spectrum (solid line) of CaAl12O19:Pr3+ monitoring the Pr3+

emission at 402 nm and the emission spectrum (dashed line) upon 4ƒ5d excitation at 205 nm.

(b) UV and VUV excitation spectra (solid lines) of the Cr3+ 686 nm emission (dashed line) in

CaAl12O19:Cr3+. The UV and VUV excitation spectra are scaled to the 4A2–4T1(P) excitation

intensity [114]. Reproduced from J. Solid State Chem., 2007, 180, 2933. Copyright © 2007,

Elsevier Inc.

Fig. 41. (a) Normalized emission spectra of Pr3+ in CaAl12O19:1%Pr3+,x%Cr3+ (x =0, 1, 2, 3, 5),

excited at 205 nm. (b) Emission spectra of Pr3+ 3


P0 → 3
H4 transition in

CaAl12O19:1%Pr3+,x%Cr3+ (x =0, 1, 2, 3, 5), excited at 465 nm [114]. Reproduced from J.

Solid State Chem., 2007, 180, 2933. Copyright © 2007, Elsevier Inc.

Fig. 42. VUV excitation/emission (λemi = 545 nm, λexc = 147 nm) spectra of CaSO4:Tb3+,Na+

(made with NaCl) and YBO3:Tb3+ phosphors. A and A′ denote excitation and emission

spectra (red lines) of CaSO4:Tb(4%), Na(12%), whereas B and B′ represent those (black lines)

for the commercial green PDP phosphor, YBO3:Tb3+ [120]. Reproduced from Adv. Funct.

Mater., 2007, 17, 212. Copyright © 2007, Wiley-VCH.

Fig. 43. An energy-level diagram of Tb3+ in a CaSO4:Tb, Na phosphor. The diagram shows

various energy levels of two Tb3+ ions in the neighborhood of an SO42– anion complex in

CaSO4:Tb3+,Na+, participating in the second-order DC process following the VUV absorption

at an SO42– anion complex. Downward arrows show the emission transitions from 5D3 and 5D4

to various 7FJ energy levels. The green arrow corresponds to the 5D4→ 7F5 transition that

results in the intense 545 nm green emission. Other transitions are: 5D3→ 7F6 = 381 nm, 5D3→
7
F5 = 413 nm, 5D3→ 7F4 = 435 nm, 5D4→ 7F6 = 480 nm, 5D4→ 7F5 = 545 nm, 5D4→ 7F4 = 588

nm, and 5D4→ 7F3 = 621 nm [120]. Reproduced from Adv. Funct. Mater., 2007, 17, 212.

145
Copyright © 2007, Wiley-VCH.

Fig. 44. AM1.5G spectrum showing the fraction that is currently absorbed by a thick silicon

device and the additional regions of the spectrum that can contribute towards UC and DC

[129]. Reproduced from Sol. Energy Mater. Sol. Cells, 2006, 90, 2329. Copyright © 2006,

Elsevier B. V.

Fig. 45. Schematic diagram of DC system. The phosphor converter is located on the front

surface of a solar cell, which has a band-gap energy Eg. High-energy photons with ћω>2Eg

are absorbed by the converter and efficiently down converted into two lower energy photons

with ћω>2Eg, which can both be absorbed by the solar cell.

Fig. 46. Schematic representation of the cooperative and accretive pathways for ET from Tb3+

to Yb3+. The bold arrows indicate excitation of Tb3+ into the 5D4 state, after which energy

transfer may occur. The ET processes are depicted by the dotted lines. In both mechanisms a

virtual state is involved. For the cooperative mechanism, the virtual state is located on Tb3+.

For the accretive mechanism, the virtual state is located on Yb3+. Since the total amount of

energy is unchanged after the energy transfer, the resonance condition implies that the

transition energy of Tb3+ balances the sum of the transition energies of the Yb3+ ions [122].

Reproduced from Phys. Rev. B, 2005, 71, 014119. Copyright © 2005, The American Physical

Society.

Fig. 47. Schematic energy level diagrams of GdAl3(BO3)4:Tb,Yb, GdAl3(BO3)4:Pr,Yb, and

GdAl3(BO3)4:Tm,Yb, showing the concept of NIR QC with visible excitation at 489, 485, and

475 nm, respectively [133]. Reproduced from Appl. Phys. Lett., 2007, 91, 051903. Copyright

© 2007, American Institute of Physics.

146
Fig. 48. Emission spectra of Tb3+-doped KYb(WO4)2 excited by 308 nm laser line [126].

Reproduced from J. Lumin., 2001, 92, 229. Copyright © 2001, Elsevier Science B.V.

Fig. 49. Energy level schemes of Yb3+ and Tb3+ ions and UC and DC processes [126].

Reproduced from J. Lumin., 2001, 92, 229. Copyright © 2001, Elsevier Science B.V.

Fig. 50. Visible/NIR emission spectrum of Y0.74Yb0.25Tb0.01PO4 upon Tb3+ 7F6→ 5D4

excitation (489 nm). The spectral region between 900 and 1100 nm is amplified by a factor of

50 [122]. Reproduced from Phys. Rev. B, 2005, 71, 014119. Copyright © 2005, The

American Physical Society.

Fig. 51. Excitation spectra of the Tb3+ 5D4→ 7F3 emission (615 nm, solid line) and the Yb3+
2
F5/2→ 2F7/2 emission (1000 nm, dotted line) in Y0.74Yb0.25Tb0.01PO4 [122]. Reproduced from

Phys. Rev. B, 2005, 71, 014119. Copyright © 2005, The American Physical Society.

Fig. 52. (a) Luminescence decay curves of the Tb3+ 5D4 emission for various concentrations of

Yb3+ plotted with simulation results. The dots are the experimental results. The solid lines are

simulated curves using a cooperative dipole-dipole model. Dashed lines are simulated curves

using an accretive dipole-dipole model. (b) Simulated decay curves for the phonon-assisted

dipole-dipole model. For both figures the Yb3+ concentrations are indicated in the figure [122].

Reproduced from Phys. Rev. B, 2005, 71, 014119. Copyright © 2005, The American Physical

Society.

Fig. 53. (a) PLE spectra of the Tb3+ 5D4→ 7F4 emission (543 nm, solid line) and the

Yb3+:2F5/2→ 2F7/2 emission (970 nm, dotted line), and (b) Visible-NIR PL spectrum upon

excitation of 486 nm (Tb3+: 7F6→ 5D4) of GdBO3:Tb3+,Yb3+ [132]. Reproduced from Appl.

Phys. Lett., 2007, 90, 061914. Copyright © 2007, American Institute of Physics.

147
Fig. 54. The dependence of the green- and NIR-emission intensity on the Yb3+

doping-concentration in GdBO3:Tb3+,Yb3+ nanophosphors. Inset shows schematic energy

levels of the nanophosphors showing possible mechanisms for an NIR QC under excitation of

visible with λex=486 nm [132]. Reproduced from Appl. Phys. Lett., 2007, 90, 061914.

Copyright © 2007, American Institute of Physics.

Fig. 55. Decay lifetimes of the Tb3+ 5D4→ 7F4 luminescence under excitation of 486 nm. The

different fractions x of Yb3+ in the samples are indicated in the figure. Inset shows the transfer

efficiency (TE), QE, and decay lifetime as a function of the Yb3+ doping-concentration [132].

Reproduced from Appl. Phys. Lett., 2007, 90, 061914. Copyright © 2007, American Institute

of Physics.

Fig. 56. Transfer efficiency and decay lifetime as a function of the Yb3+ concentration in

GdBO3:Tb3+,Yb3+. The calculated transfer efficiencies beyond the concentration quenching

are also shown (asterisks) [132]. Reproduced from Appl. Phys. Lett., 2007, 90, 061914.

Copyright © 2007, American Institute of Physics.

Fig. 57. Visible-NIR excitation spectra of (a) GdAl3(BO3)4:1%Pr,x%Yb (x=0, 2, 10, 20, and

30) (λexc=489 nm), (b) GdAl3(BO3)4:1%Tb,2%Yb (λexc=485 nm), and (c)

GdAl3(BO3)4:1%Tm,2%Yb (λexc=475 nm). Emission spectra of RE3+ (RE=Pr, Tb, and Tm)

(solid line) and Yb3+ (dotted line) are also given [133]. Reproduced from Appl. Phys. Lett.,

2007, 90, 061914. Copyright © 2007, American Institute of Physics.

148
Figures

Figure 1.

Figure 2.

149
Figure 3.

Figure 4.

150
Figure 5.

Figure 6.

151
Figure 7.

Figure 8.

152
Figure 9
.

Figure 10.

153
Figure 11.

Figure 12.

154
Figure 13.

155
Figure 14.

Figure 15.

156
Figure 16.

157
Figure 17.

158
Figure 18.

159
Figure 19.

Figure 20.

160
Figure 21.

161
Figure 22.

162
Figure 23.

Figure 24

163
Figure 25.

Figure 26.

164
Figure 27.

Figure 28.

165
Figure 29.

Figure 30.

166
Figure 31.

Figure 32.

167
Figure 33.

Figure 34.

168
Figure 35.

Figure 36.

169
Figure 37.

170
Figure 38.

Figure 39.

171
Figure 40.

172
Figure 41.

Figure 42.

173
Figure 43.

174
Figure 44.

Figure 45.

175
Figure 46.

1
5 D2
D3
24 3
1
G4 P 3 1
5 3 2 P1, I6
D4 P0
20

1
D2
Energy (103cm-1)

16 3
F2
3
12 H4
2 2
F5/2 1 F5/2
G4
3
8 H5 3
F2
980 nm
980 nm

980 nm
980 nm

3
7
F0 F4 3
4 H6
7
F5 3
2 3 2
0
7
F6 F7/2 H6 H4 F7/2
3+ 3+ 3+ 3+ 3+ 3+ 3+
Tb Yb Yb Tm Pr Yb Yb

Figure 47.

176
Figure 48.

Figure 49.

177
Figure 50.

Figure 51.

178
Figure 52.

179
3+ 2 2
Emission Yb : F5/2-> F7/2
Excitation (λex=486 nm)

3+ 5 7
Tb : D4-> Fj
Intensity (a.u.)

5 4 3 210
3+
%Yb in
3+
(YbxGd1-x)BO3:Tb
0
1
10
20
30
50
75

400 500 600 700 800 900 1000 1100


Wavelength (nm)
Figure 53.

5 5
D3, G6
24
970-nm intensity (a.u.)
543-nm intensity (a.u.)

5
D4
20
Energy (10 3cm -1 )

16
Ex. 486 nm

12 2
F5/2

8
7
F0,1,2
4
7
F4
7 2
0 F6 F7/2
3+ 3+ 3+
Tb Yb Yb

0.0 0.2 0.4 0.6 0.8 1.0


3+
Yb concentration (mol.)
Figure 54.

180
3+ 3+
%Yb in (YbxGd1-x)BO3:Tb

0
1
Intensity (a.u.)
10
20
30
50
75

0.00 0.01 0.02 0.03 0.04

Decay lifetime (s)

Figure 55.

100

4
Transfer efficiency (%)
Decay lifetime (ms)

80

60
3

40

2
20

1 0
0 10 20 30 40 50 60 70 80
3+
Yb concentration (mol.%)
Figure 56.

181
(a) Emission 3+ 2 2
Yb : F5/2-> F7/2
(λex=489 nm)
Excitation
x%Yb
Pr : P0→ H6
3+ 3 3
0
2
10
20
30
Emission intensity (a.u.)

400 500 600 700 800 900 1000 1100


(b) Emission
3+ 5 7
(λex=485 nm)
Tb : D4-> F5
Excitation
3+ 2 2
Yb : F5/2-> F7/2
5 7
D4-> F4,3,2

450 500 600 700 800 900 1000 1100


(c) Emission
(λex=475 nm)
Tm : G4→ F4
3+ 1 3

Excitation 3+ 2 2
Yb : F5/2-> F7/2

G4→ H5
1 3

400 500 600 700 800 900 1000 1100


Wavelength (nm)

Figure 57.

182

View publication stats

You might also like