Download as pdf or txt
Download as pdf or txt
You are on page 1of 26

Chem Soc Rev

View Article Online


REVIEW ARTICLE View Journal

Recent progress in luminescence tuning of Ce3+


Published on 30 September 2015. Downloaded by University of Tasmania on 30/09/2015 10:15:45.

and Eu2+-activated phosphors for pc-WLEDs


Cite this: DOI: 10.1039/c4cs00446a
Guogang Li,*a Ying Tian,a Yun Zhaoa and Jun Lin*b

Nowadays, phosphor converted white light-emitting diodes (pc-WLEDs) have been widely used in solid-
state lighting and display areas due to their superior lifetime, efficiency, and reliability as well as
significant reduction in power consumption. Phosphors are indispensable components of pc-WLED
devices, and their luminescence properties determine the quality of WLED lighting and displays. In order
to further achieve high luminous efficacy, chromatic stability, and color-rending properties in pc-WLEDs,
much effort has been focused on improving current pc-WLED phosphors and developing novel
pc-WLED phosphors recently. This review article concerns commonly used rare earth ion (Eu2+ and Ce3+)
Received 2nd December 2014 activated inorganic phosphors, highlighting the important effect of spectral tuning via local structural variations
DOI: 10.1039/c4cs00446a on improving the luminescence performance of phosphors. The main spectral tuning strategies are discussed
in detail and summarized, including (1) doping level control; (2) cationic substitution; (3) anionic substitution;
www.rsc.org/chemsocrev (4) cationic–anionic substitution; (5) the crystal-site engineering approach; (6) mixing of nanophases.

1. Introduction sources owing to their extraordinary luminous efficiency and


brightness, low power consumption, reliability, long operation
Nowadays, phosphor-converted white light-emitting diodes life and environmental friendliness compared to other lighting
(pc-WLEDs) are burgeoning and are extensively considered as technologies, like incandescent, halogen, xenon, fluorescent
the most promising new generation solid state lighting (SSL) lamps and so on.1 Especially in terms of energy saving,
pc-WLEDs have tremendous advantages compared to the above
other light sources. According to the statistics of the US
a
Faculty of Materials Science and Chemistry, China University of Geosciences,
Department of Energy (DOE) in 2012, in the next two decades
Wuhan 430074, P. R. China. E-mail: ggli8312@gmail.com
b
State Key Laboratory of Rare Earth Resource Utilization, Changchun Institute of
(2010–2030), WLED lighting technology will save at least 20% of
Applied Chemistry, Chinese Academy of Sciences, Changchun 130022, P. R. China. total electricity consumption per year if a nationwide move
E-mail: jlin@ciac.ac.cn; Fax: +86-431-85698041 toward SSL for general illumination.1 No other lighting

Guogang Li is currently an Ying Tian is currently studying


associate professor in the China for her master’s degree in the
University of Geosciences (Wuhan). China University of Geosciences
He received his BSc degree from (Wuhan). She received her BSc
the Hebei Normal University of degree from the College of
Science & Technology in 2007 Technology, Hubei Engineering
and was a graduate student at University in 2014 and entered
the Changchun Institute of Applied into Prof. Guogang Li’s group in
Chemistry (CIAC), Chinese Academy the same year. Her current research
of Sciences, and received a PhD interests focus on the synthesis of
degree in 2012. After graduation, pc-WLED phosphors for white light
he worked as a postdoctor in emitting diodes (W-LEDs) for solid-
Guogang Li Taiwan University for one year. Ying Tian state lighting and displays via solid
His current research interests state reaction and soft chemistry
mainly include synthesis of luminescent materials for field methods.
emission displays (FEDs) and white light emitting diodes
(W-LEDs) via solid state reaction and soft chemistry methods.

This journal is © The Royal Society of Chemistry 2015 Chem. Soc. Rev.
View Article Online

Review Article Chem Soc Rev


Published on 30 September 2015. Downloaded by University of Tasmania on 30/09/2015 10:15:45.

Fig. 1 Schematics of the two principal white-lighting strategies of pc-WLED devices. (a) A blue chip with a yellow downconverting phosphor. (b) A n-UV
chip with red, green, and blue (RGB) downconverting phosphors. In order to obtain warm white light in the strategy (a), it can be modified as follows:
(c) mixing the blue chip with yellow and red downconverting (RY) phosphors; or (d) mixing the blue chip with green and red downconverting (RG)
phosphors.

technology can offer so much potential to save energy. Therefore, light (n-UV) LEDs to generate white light, and two main strategies
they are very important for the reduction of global power require- are exploited as follows: one is to combine highly efficient
ments and the use of fossil fuels. On the other hand, pc-WLEDs (Ga,In)N blue LEDs (420–480 nm) with yellow-emitting phos-
are excellent backlight sources for use in display areas such as phors, as in a device with a schematic structure shown in
liquid crystal displays (LCDs).2 To build highly efficient WLED Fig. 1a.2 For example, the most common and easiest way is the
devices, many aspects should be considered and optimized, combination of a 460 nm blue InGaN LED chip and a yellow
including semiconducting components (generally InGaN semi- phosphor of a cerium(III)-doped yttrium aluminum garnet
conductor chips), phosphor conversion components, and (YAG:Ce3+), which is still in widespread use today. However,
luminaires as well as packaging technologies. Phosphors are such a device has a poor color rendering index (Ra = 70–80) and
indispensable components of pc-WLED devices, and the explora- a high correlated color temperature (CCT = 4000–7500 K)
tion and development of pc-WLED phosphors for use as lighting because of the lack of a red component. The modified ways
and display backlight sources is one of the most important and for achieving white light with a high Ra and a suitable CCT are
urgent challenges, which can be met by advanced science and to coat yellow-emitting phosphors with a more red component
technology.3,4 Generally, in pc-WLEDs, downconverting phos- (Fig. 1c, right) or to mix green and red multi-phased phosphors
phors are used to connect with bright blue or near ultra violet on 460 nm blue LEDs (Fig. 1d, right).5–15

Yun Zhao is now a postgraduate Jun Lin is a professor at the


student in China University of Changchun Institute of Applied
Geosciences (Wuhan). She received Chemistry (CIAC), Chinese
her BSc from JiangHan University Academy of Sciences (CAS). He
in 2011. She worked as a chemical received BSc and MSc degrees in
laboratory technician for 3 years inorganic chemistry from Jilin
after undergraduate graduation. University, in 1989 and 1992,
Now, Prof. Guogang Li is her super- respectively, and a PhD degree
visor for her Master’s degree. Her (inorganic chemistry) from
current research interests focus CIACCAS, in 1995. He has
on the synthesis of pc-WLED phos- worked as a postdoctor for more
phors for white light emitting than 4 years in CityU (HK, 1996),
Yun Zhao diodes (W-LEDs) for solid-state Jun Lin INM (Germany, 1997) and VCU/
lighting and displays via solid UNO (USA, 1998–1999). His
state reaction and soft chemistry research interests include luminescent materials and multifunc-
methods. tional composite materials together with their applications in
display, lighting and biomedical fields. He has published more
than 400 peer-reviewed journal articles in related fields.

Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2015
View Article Online

Chem Soc Rev Review Article

It is noticed that multi-phased phosphor converted systems phosphors for pc-WLEDs based on the available LED chips.39,41–47
frequently result in the addition of the manufacture cost, and The emphasis was also be placed on the relationships among
low luminescence efficiency owing to the strong reabsorption crystal structures, luminescence properties, and device perfor-
of the emission of blue or green phosphors by yellow and red- mances. He et al. summarized the dependence of spectroscopic
emitting phosphors. The second approach involves mixing of properties on composition for Eu2+, Ce3+ or Yb2+-activated
the emission from red, green, and blue (RGB) phosphors with nitride and oxynitride phosphors based on a-SiAlON, b-SiAlON,
n-UV LED chips (380–420 nm), as in a device with a schematic oxonitridosilicates, nitridoalumosilicates and nitridosilicates.40
structure shown in Fig. 1b.16–24 It is similar to the first strategy Generally, all possible variations in host composition usually
that low efficiency due to reabsorption, and problems of sedi- alter the symmetry, coordination environment, crystal field
Published on 30 September 2015. Downloaded by University of Tasmania on 30/09/2015 10:15:45.

mentation and uniformity distribution of phosphors in silicon strength, covalence, and the energy transfer, thus influencing
resin cannot be avoided. In return, several benefits including luminescence properties. In fact, the focus on the tunable
an extremely high color rendering index (Ra 4 90), a wide color luminescence of phosphors in the above reviews is mainly based
gamut coverage and a stable light color output at different drive on the average change in crystal structures of phosphors. Though
currents can be owned.16–24 In addition, in order to decrease the variation of the local crystal field also have extremely impor-
the reabsorption effect among phosphors and enhance their tant effect on the luminescence properties of Ce3+ and Eu2+-
luminescence efficiency and color rendering index, WLEDs activated pc-WLED phosphors, which has become a research
based on single-phased white light emitting phosphors have hotspot. In recent years, many famous research groups have
been extensively investigated and developed.25–30 In general, focused their attention on this topic. For example, Liu’s group
phosphors are vital to the development of the WLED lighting have proposed some new luminescence mechanisms in Ce3+ and
and backlight display industry. Highly efficient and emission- Eu2+-doped (oxy)nitrides based on the variation of the local crystal
tunable phosphors not only can enhance the energy efficiencies structure including the ‘‘Cation-Size-Mismatch effect’’, the
of WLEDs but also present potential advantages in the color ‘‘Neighboring-Cation Substitution effect’’, and the ‘‘Nano-
rendering index, correlated color temperature and color gamut, segregation and Neighbor-Cation Control effect’’ to clarify
which enlarge their use in vehicle lights, architecture decorations, some abnormal luminescence phenomena.48–51 Ram’s group
streetlights, directional lighting, clinical medical lighting, etc. also confirmed that some small variations in the local structure
Therefore, novel and acceptable fluorescent materials for improving of La3xCexSi6N11, SrxBa2xSiO4:Eu2+ and so on phosphor systems
the performances of solid-state lighting must be developed.31,32 might generate a big improvement on their luminescence perfor-
So far, there have been a number of useful reviews summarizing mances.1,53,54 Cheetham’s group found that the large spectral
on the pc-WLED phosphors for solid-state lighting or backlight shift from blue to yellow light in Ca2SiO4:Ce should be attributed
displays by Jüstel et al.,33 Lin & Liu,3 Xie,34 Höppe,35 Schnick,36 to the distortion of the local crystal structure.55 In summary, more
and Shang & Lin37 et al. Especially, some recent reviews have and more researchers have recently focused their attention on the
systematically summarized the design, optical properties and the adjustment and improvement of luminescence properties of
application of pc-WLED phosphors with the emphasis on the phosphors through the variation of the local crystal structure.
dependence of luminescence properties on composition.38–40 However, so far there is no related review to mainly concern and
For example, Smet and co-authors presented, identified and systematically discuss the effect of the local crystal field on the
discussed six main criteria dealing with the shape and position luminescence performances of rare earth activated pc-WLED
of the emission and excitation spectra, the thermal quenching phosphors. Therefore, a related review article is necessary. In view
behaviour, the quantum efficiency, the chemical and thermal of this blank, we summarize the current approaches to realize
stability and finally with the occurrence of saturation effects.38 spectral tuning by the variation of the local crystal structure
Based on these criteria, the most common dopant ions (broad- including doping level control, cationic substitution, anionic
band emitting Eu2+, Ce3+ and Mn2+, and light-emitting rare earth substitution, cationic–anionic substitution, the crystal-site engineer-
ions) and host compounds (garnets, sulfides, and (oxy)nitrides) ing approach, mixing of nanophases, and explain some abnormal
are evaluated. Ye et al. reviewed the most recent advances in the luminescence mechanisms based on the actual variation of the local
synthesis and application of phosphors for pc-WLEDs with the bond length, covalence, coordination environment, crystal field
emphasis specifically on: (a) principles to tune the excitation and strength and so on for spectral tuning in rare earth ion activated
emission spectra of phosphors, predictions according to crystal pc-WLED phosphors through the variation of the local crystal field.
field theory, and structural chemistry characteristics (e.g. cova- Therefore, it presents a new angle of view to concern the improve-
lence of chemical bonds, electronegativity, and polarization ment of pc-WLED phosphors, which is different from the previous
effects of elements); (b) pc-WLEDs with phosphors excited by reviews involving pc-WLED phosphors. This review does not intend
blue-LED chips: phosphor characteristics, structure, and acti- to comprehensively understand all aspects of pc-WLED phosphors
vated ions (i.e. Ce3+ and Eu2+), including YAG:Ce, other garnets, and many of these aspects will not be discussed here such as Eu3+,
non-garnets, sulfides, and (oxy)nitrides; (c) pc-WLEDs with Mn2+ and Mn4+-doped materials, semiconducting quantum dots,
phosphors excited by near ultraviolet LED chips: single-phased sulfide-based materials and defect-related luminescent materials.
white-emitting phosphors (e.g. Eu2+–Mn2+ activated phosphors), The concerns are primarily focused on several commonly used rare
red–green–blue phosphors, energy transfer, and mechanisms earth ion (Eu2+ and Ce3+) activated inorganic phosphors [mainly
involved; and (d) new clues for designing novel high-performance including garnets, silicate oxides, (oxo)nitrido(alumino)silicates],

This journal is © The Royal Society of Chemistry 2015 Chem. Soc. Rev.
View Article Online

Review Article Chem Soc Rev

emphasizing the important effect of spectral tuning via local


structural variations on improving the luminescence performances
of phosphors.

2. Improving the photoluminescence


performances by spectral tuning
It is known that the colors of light from pc-WLED lighting and
display devices based on rare earth ion (Eu2+ or Ce3+) activated
Published on 30 September 2015. Downloaded by University of Tasmania on 30/09/2015 10:15:45.

phosphors are usually recorded through luminescence spectra,


namely, the excitation and emission spectra. While the lumi-
nescence spectra are closely related to several vital lumines-
cence performance parameters in pc-WLED devices such as
correlated color temperature (CCT), Commission Internationale
de I’Eclairage (CIE) color coordinates, color rendering index (CRI
or Ra), luminous efficacy, quantum yield (QY) and so on.1,3,37 By
adjusting the position (blue shift or red shift) and full width at
half maximum (FWHM, more broad or more narrow) of photo-
luminescence emission spectra of rare earth activated phosphors,
these parameters such as CCT, CIE color coordinates, and Ra can
be precisely controlled and thus further improve the lumines-
cence performances of WLED devices. For example, the color
coordinates of white light from single pc-WLED devices based on
multivariate phosphors could span the whole while light region Fig. 2 (a) A schematic luminescence process from activator ions (Ce3+ or
with various CCTs (cold or warm white light) and Ras through Eu2+) in a host crystal. (b) Schematic explanation of the effects of the
host crystal on the luminescence of Ce3+ or Eu2+ ions, including the
shifting of the emission position of phosphors. In addition,
nephelauxetic effect (NE, centroid shift), crystal field splitting (ecfs) and
changing the excitation spectra (enlarging the absorption range) the Stokes shift (SS) and so on. Such effects will lead to an increase or
of pc-WLED phosphors might enhance the luminescence effi- decrease in the energy difference between the 5d and 4f levels, known as
ciency of phosphors and the luminescence efficiency of WLED the blueshift or redshift, respectively.
devices. Generally, the above mentioned performance parameters
(CIE color coordinates, CCT, Ra, luminous efficacy, and QY) of
type.37 The most widely used activators are Ce3+ or Eu2+ ions,
rare earth activated phosphors can be reasonably optimized by
whose function in phosphors is to convert n-UV or blue light
spectral tuning including the change of spectral position, shape
radiation to visible light, as shown in Fig. 2a. Due to the 5d–4f
and FWHM to improve the luminescence performances of WLEDs,
transition, the luminescence of Ce3+ and Eu2+ doped phosphors
which can be summarized by the following several aspects: tuning
is related to the intrinsic feature of activator ions and the host
the spectral position to approach the eye sensitivity curve can
crystal structure. Especially the latter has a greater effect on
improve the luminous efficiency of WLEDs; matching the spectral
the resulting photoluminescence properties. Herein, we mainly
position with the standard illuminators can optimizing the color
discuss the effects of host lattices on the 5d–4f transition
rendering properties of WLEDs; adjusting the spectral position can
luminescence of Ce3+ and Eu2+ activated phosphors, highlighting
achieve the desired color temperature; matching spectral position
the relationship between local crystal structural variations around
and shape with the color filters can enhance the color gamut of
activator ions and spectral tuning. According to energetics, the 4f
WLED backlights.3,34 Accordingly, reasonably tuning the lumines-
and 5d energy levels of Ce3+ and Eu2+ ions in phosphors determine
cence spectra of rare earth ion activated phosphors are very
their excitation and emission properties. The Ce3+ ion with the 4f1
important and necessary for the rapid development of WLED
configuration in solids shows efficient broadband luminescence
lighting and backlight display undertakings. For the researchers
due to the 4f–5d parity allowed electric dipole transition. While the
in the phosphor area, great effort is needed to focus on the modi-
Eu2+ ion has the 4f7 electron configuration and its emission is
fication of phosphors’ structures and components for tuning their
partly forbidden by the spin selection role, which whether present
luminescence spectra.
broadband emission depends on the crystal field strength. The
above reasons result in different photoluminescence properties of
3. The spectral tuning approaches Ce3+ and Eu2+ ions in some same hosts. For free Ce3+ or Eu2+ ions,
in pc-WLED phosphors they have large energy difference between the 4f ground state and
the 5d excited state, which are B6.2 eV (50 000 cm1) for Ce3+ and
Commonly, phosphors used in WLEDs consist of an inactive B4.2 eV (34 000 cm1) for Eu2+ ions.56–58 Since the 5d orbitals are
host like aluminates, silicates, (oxo)nitrido(alumino)silicates, strongly affected by the surrounding crystal field, symmetry, anion
etc., and activators, which are defined as the ‘‘host + activator’’ polarizability, and covalency of the host crystal, the excitation and

Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2015
View Article Online

Chem Soc Rev Review Article

emission energies can be modulated through changing the be used as an approximation for describing crystal field splitting
composition and structure of the host crystal. Dorenbos has trends with the bond distance. The shorter the bond length and
proposed a series of equations to preliminarily determine the the stronger the crystal field splitting, and then the longer
f–d excitation and d–f emission of rare earth ions in inorganic the emission wavelength. As the other important aspect, the
compounds.59 For free lanthanide ions and the ions in any nephelauxetic effect (literally means ‘‘charge cloud expansion’’)
inorganic compound, the energy difference E(n,Q,A) between has been studied for the ff-transitions in lanthanides,73,74
the lowest 4f n level and the first 4f n15d level in trivalent the dd-transitions in transition metal elements,70 and the
(Q = 3+) and divalent (Q = 2+) lanthanide ions always shows sp-transitions in Tl+; Pb2+; and Bi3+.75,76 It leads to a displacement
the same characteristic variation with the number of electrons of levels to lower energy depending on the type of compound. The
Published on 30 September 2015. Downloaded by University of Tasmania on 30/09/2015 10:15:45.

n in the 4f shell.59 The effect of the host crystal, denoted by ideas developed for the nephelauxetic effect in the above metal
the parameter A on the energy difference is expressed by the elements may also be used to describe the energy difference
redshift D(Q,A) and the Stokes shift S(Q,A): for the energy of between the 5d and 4f levels.59–69 The decrease of 5d levels
fd-absorption one may write relative to the free ion situation is known as the centroid shift.
It could lead to the centroid shift to lower energy depending on
Eabs(n,Q,A) = EA,free(Q,A)  D(Q,A) (1)
the type of compound due to a reduction of interelectron repul-
and for the df-emission sion.70 The size of the nephelauxetic effect can be characterized
by the nephelauxetic ratio b as
Eem(n,Q,A) = EA,free(Q,A)  D(Q,A)  DS(Q,A) (2)
1  b = hk (4)
where EA,free(Q,A) is a constant for each lanthanide ion with a
value equal or close to the f–d transition energies for the free where 1  b is written as a product of two functions, each
(gaseous) lanthanide ions.60–69 When the first 5d level energy is variable.70 h characterizes the anion ligands and k the metal.
known for a specific lanthanide ion in compound A; one may When anions or ionic complexes are put in the order of
determine the redshift D(Q,A). From that, the first 5d level increasing h-parameter the well-known nephelauxetic series is
energies for all other lanthanides in the same compound can obtained, i.e., no ligands o F o H2O o NH3 o Cl o Br o
be predicted. The values for D(Q,A) and DS(Q,A) (Q = 3+, 2+) of N3 o I2 o O2 o S2 o Se2 and H2O o SO42 o CO32 o
several hundred different compounds including fluorides, PO43 o BO33 o SiO44 o AlO45 o O2. The reduction of
chlorides, bromides, iodides, oxides, sulfides, selenides, (oxy)- interelectron repulsion and the increase of the h-parameter are
nitrides and so on were compiled in ref. 60–69. The above often ascribed to the covalence,70,74 i.e., the sharing of electrons
contributions are very important, which not only guide the between the metal and the surrounding ligands. Therefore,
researchers to develop desired rare earth ion doped phosphors the centroid shift originated from the nephelauxetic effect is
for WLEDs, but also help to deeply study the structure-related also related to the covalency between the activator ions and
luminescence properties. Simply, when accommodating Ce3+ or the surrounding anion ligands.71,72 The higher the degree of
Eu2+ ions into a host, this energy difference can range from UV covalency, the larger centroid downshift will be generated.
across visible wavelengths, depending on two main aspects: Except for the above two main aspects, some other factors such
crystal field splitting (CFS) and the nephelauxetic effect (NE). as the Stokes shift (SS), the coordination geometry, the d orbital
The big 5d orbital crystal field splitting and the nephelauxetic hybridization of the activator ion due to the interaction between
effect (centroid shift) of Ce3+ or Eu2+ ions can substantially molecular orbitals of neighboring anions, the distortion of the
decrease their energy difference, leading to obvious n-UV or activator ion site and so on, can also affect the energy difference
blue light absorption and tunable emission all over the visible between the 5d and 4f levels and thus give rise to a spectral shift.1
light range. These features of Ce3+ and Eu2+ activated phosphors In general, the position of the 5d energy levels of Ce3+ and Eu2+
perfectly match with n-UV or blue LEDs, and thus are extremely ions can change by tens of thousands of wavenumbers from one
conducive to their application in WLEDs. For the crystal field compound to another, for example the 5d configuration of Ce3+ is
splitting effect, it describes the energy difference between the known to lower the energy by 5000–25 000 cm1 under different
highest and lowest 5d energy levels, which is the complex results crystal environments, and thus their emission spectra can be
of some effects like the bond lengths from the activator ion to the adjusted in the whole visible light region.58,59 In other words, this
coordinating anions, the degree of covalency between the acti- essence of spectral tuning in Ce3+ or Eu2+ doped phosphors is
vator ion and its ligands, the coordination environment, and the ascribed to the variation of the microstructural environment
symmetry of the activator site. Generally, the crystal field splitting surrounding Ce3+ or Eu2+ ions. Herein, we summarized several
(Dq) can be determined by the following equation:37,59,70–72 main aspects to realize spectral tuning of rare earth activated
phosphors based on the local crystal structural change of
1 r4
Dq ¼ Ze2 5 (3) phosphors including doping level control, cationic substitu-
6 R
tion, anionic substitution, cationic–anionic substitution, the
where Dq is the magnitude of the 5d energy level separation, Z is crystal-site engineering approach, and mixing of nanophases.
the anion charge or valence, e is the electron charge, r is the The detailed explanations are presented in Sections 3.1–3.4.
radius of the d wavefunction, and R is the bond length. Although The summary of spectral tuning in some representative phosphor
this equation was derived by using a point charge model, it can systems by compositional adjustment based on the variation of

This journal is © The Royal Society of Chemistry 2015 Chem. Soc. Rev.
View Article Online

Review Article Chem Soc Rev

Table 1 Summary of the spectral tuning mechanism in some representative phosphor systems by compositional adjustments based on the variation of
the local crystal structure

Spectral tuning/emission
Type Representative examples Blueshift Redshift Ref.
Doping level control La3xCexSi6N11 (0 o x r 3) B30 nm 53
Ba1xEuxSi2O2N2 (0 o x r 0.1) B10 nm 80
Sr1xEuxSi2O2N2 (0 o x r 0.16) B19 nm 81
Ca-a-SiAlON:xEu2+ (0 o x r 0.25) B20 nm 82
Y3xCexAl5O12 (0 o x r 0.0333) B22 nm 83
Ca12xCexLixAlSiN3 (0 o x r 0.05) B33
Published on 30 September 2015. Downloaded by University of Tasmania on 30/09/2015 10:15:45.

nm 84
Ca2xEuxSiO4 (0 o x r 0.08) B70 nm 86

Cationic substitution SrxBa2xSiO4:Eu2+ (0 r x r 0.2) B63 nm 54


(Sr1xCax)Si2O2N2:Eu2+ (0 r x r 1) B10 nm 81
(Sr0.98xBaxEu0.02)Si2O2N2 (0 r x r 0.63) B25 nm 48, 100
(Sr2xBax)Ga2SiO7:Eu2+ (0 r x r 1) B48 nm 88
a-(Y,Gd)FS:Ce3+ (0 r y r 0.3) B8 nm intensitym 42
Y1.98Ce0.02(Ca1ySry)F4S2 (0 r y r 1) B40 nm 43
(Sr1xCax)2Si5N8:Eu2+ (0 r y r 0.3) B50 nm 93
Ba2xSrxSi5N8:Eu2+ (0 r y r 0.5) 14–33 nm 95
(Ca0.993xyMgxSry)9Y(PO4)7:Eu0.007 (0 r x r 0.5) B51 nm 90
(Ca0.993xyMgxSry)9Y(PO4)7:Eu0.007 (0 r x r 0.5) B22 nm
Sr3xBaxSiO5:Eu2+ (0 r x r 0.2) B15 nm 101
(Na1xCax)(Sc1xMgx)Si2O6:Eu2+ (0 r x r 1) B83 nm 105
(Ce0.005Y0.995)3Sc2Al3xGaxO12 (0 r x r 3) B24 nm 108
CaAlxSi(73x)/4N3:Eu2+ (0.7 r x r 1.3) FWHMm 110
Y3Al52xMgxSixO12:Ce3+ (0 r x r 2) B41 nm 111
Y3Al52xMgxGexO12:Ce3+ (0 r x r 1.6) B19 nm
(Ca1xLax)(Al1+xSi1x)N3:Eu (0 r x r 0.15) B20 nm 50
(Ca1xLix)(Al1xSi1+x)N3:Eu (0 r x r 0.15) FWHMm
La1x0.025Ce0.025Sr2+xAl1xSixO5 (0 r x r 0.975) B29 nm 112
(Ca1xCex)6Ba(P1xSix)4O17 (0 r x r 0.07) FWHMm 113

Anionic substitution Ca2.97Sc2Si3O126xN4x:0.03Ce3+ (0 r x r 0.8) FWHMm 116


Sr1.984Si(N,O)4:0.016Eu2+ (0.7 o z o 1.2) B30 nm 118
Ba5SiO4(FxCl6x):Eu2+ (0 r z r 0.15) B61 nm 119

Cationic–anionic substitution Ca1.95Eu0.05Si5xAlxN8xOx (0 r x r 1) B20 nm 49


Sr1.95Eu0.05Si5xAlxN8xOx (0 r x r 1) B15 nm
Ba1.95Eu0.05Si5xAlxN8xOx (0 r x r 1) B40 nm
Y2.94Ce0.06Al(5x)SixO(12x)Nx (0.20 r x r 0.3) B31 nm 123
Y2.91Ce0.09Al5xSixO12xNx (0 r x r 0.3) B45 nm 117
Sr2.975Ce0.025Al1xSixO4+xF1x (0 r x r 1) B63 nm 136
Gd1xSr2+xAlO5xFx:Ce3+ (0 r x r 1) B100 nm 137
Sr2Ba(AlO4F)1x(SiO5)x:Ce3+ (0 r x r 0.9) B29 nm 138
Ca12Al14zSizO32+zF2z:Eu (0 r z r 0.5) B173 nm 139
(SiC)x–(AlN)1x:Eu2+ (0 r x r 0.5) B35 nm 140
Sr1xY0.98+xCe0.02Si4N7xCx (0 r x r 1) B75 nm 51
(KSrPO4)1x(Ba2SiO4)x (0 r x r 1) B50 nm 149
(KSrPO4)1x(Sr2SiO4)x (0 r x r 1) B95 nm

Crystal-site engineering approach b-YFS:Ce3+ B45 nm 41


Sr32xCexNaxAlO4F (0 o x r 0.2) B66 nm 87
(Ca1xySrxEuy)7(SiO3)6Cl2 (0 r x r 0.5) B5 nm 150
Ca3La3(1x)Ce3x(BO3)5 (0 r x r 0.5) B50 nm 151

the local crystal structure is shown in Table 1. It is noted that visible light region. Thus they are usually used in specific UV-LED
in addition to the 5d–4f transition luminescence, some 4f–4f based WLED lighting and display devices. Some transition metal
transition luminescence such as Eu3+ or Tb3+-activated phos- ions Mn2+ (3d5 electronic configuration) and Mn4+ (3d3 electronic
phors has also been developed as potential phosphors for use configuration) activated phosphors have also been developed as
in pc-WLEDs.31,32,40,77 However, due to the narrow excitation potential pc-WLED phosphors.78,79 The Mn2+ ion can give a broad
(Eu3+: 393 nm 7F0–5L6, 463 nm 7F0–5D2; Tb3+: 4f–4f, 378 nm) band emission in the visible range owing to the 4T1 - 6A1
and emission lines (Eu3+: 5D0–7F2 612 nm; Tb3+: 5D4–7F5 543 nm), transition, and the emission color of Mn2+ can vary from green
these phosphors commonly cannot be efficiently excited by the to red depending on the crystal field.78 If the crystal field around
wide UV-to-blue light region, and have poor color rendering the Mn2+ ion is weak, the splitting of the excited energy levels in d
indices when they are complexed with other phosphors for orbits will be small resulting in Mn2+ emission with higher energy,
generating white light due to the incomplete cover over the whole whereas it will give lower energy emission. However, the d–d

Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2015
View Article Online

Chem Soc Rev Review Article

transition of Mn2+ is forbidden and difficult to pump, the


Mn2+ ion activated phosphors have low luminescence efficiency.
Mn4+ doped materials, for example oxides and fluorides, are
commonly used as red emitting phosphors. Because they present
broad adsorption bands at 380–490 nm and sharp emission
peaks at 610–760 nm due to their distinct 2E–4A2 transition in
the symmetrically octahedral crystal field. However, some draw-
backs of Mn4+ doped materials including a lot of use of HF in the
synthesis process, so far redshift for efficient warm WLEDs with
Published on 30 September 2015. Downloaded by University of Tasmania on 30/09/2015 10:15:45.

a high luminous efficiency, low resistance to moisture, and


unclear structure-related luminescence mechanisms, limit their
rapid progress in scientific investigation and application of
WLED lighting.79 In addition, the long decay lifetime at the
millisecond order of magnitude of Mn2+ and Mn4+-activated
phosphors also tremendously hinders for their application in
WLED lighting. Based on the above reasons Eu3+, Mn2+ and
Mn4+-doped materials are not discussed here temporarily.
On the other hand, the thermal stability of luminescence pro-
perties of phosphors is an important parameter to be considered
in practical WLED applications, which influences the lumines- Fig. 3 Room-temperature (a) absorption, (b) excitation, (c) emission
cence efficiency and lighting quality of the WLED devices.1,34,48–52 spectra of La3xCexSi6N11 (0.18 r x r 3). (Reproduced with permission
from ref. 53, copyright 2013, American Chemical Society.)
Because the emission intensity decreases (namely thermal
quenching) and the wavelength shifts clearly with increasing
temperature due to nonradiative transitions from the excited the decrease in the Ce–N bond length, as shown by eqn (3). The
state to the ground state, which could be explained by hole-trapping increase in CFS strength can be confirmed in the excitation/
or electron transition mechanisms.48–52 Thus an improvement in absorption spectra. There is a significant shift toward lower
the thermoluminescence stability of phosphors used in WLEDs is energies in the position of the lowest energy excitation/
indispensable. Generally, the thermal stability of phosphors is absorption band with increasing Ce content, as exhibited in
closely related to the covalency of activator–anions, the rigidity of Fig. 3a and b. A larger CFS thus means a reduction in the energy
the crystal lattice, and the Stokes shift. An excellent thermal stability difference between the ground states and the lowest 5d-state of
could be expected in some phosphors with high covalency, high Ce3+, shifting the emission peak to the lower energy part of the
rigidity and small Stokes shift. Otherwise, a poor thermal stability spectrum. Except for the CFS effect, a reduction in the Ce–N
will be observed. Because the variation of local crystal structure of distance also increases the interaction of Ce3+ and N3, which
phosphors could influence the covalency of activator–anions, the enlarges the nephelauxetic effect. Therefore, a redshift in the
rigidity of the crystal lattice, and the spectral shift. Therefore, it emission spectrum appears. In addition, the self-absorption is
could be used to improve the thermoluminescence stability also one possible reason for the red shift due to the relatively
of phosphors, and some specific discussion is shown in the small stokes shift of Eu2+ and Ce3+ emission.3,31,32 More examples
following section. for the spectral shift originating from the increase of activator
concentration can be found in Ce3+ or Eu2+ ion doped phosphor
3.1 Doping level control systems like Ba1xEuxSi2O2N2 (10 nm),80 Sr1xEuxSi2O2N2 (19 nm),81
As mentioned in the above section, the ‘‘host + activator’’ typed Ca-a-SiAlON:xEu2+ (20 nm),82 Y3xCexAl5O12 (22 nm),83 Ca12xCex-
luminescent materials are the most widely used phosphors LixAlSiN3 (33 nm),84 Sr3yCeyAlO4F (50 nm),85 etc. The values in the
in pc-WLEDs, and the doped amount of activator ions (Ce3+ above brackets show the redshift amplitude of emission peaks. In
or Eu2+) needs to be reasonably selected in the host (usually addition, George et al. also observed that the different increase
less than 10%). By controlling the doping levels, not only the variance for (La/Ce)–N bonds (0 o x r 3) with increasing Ce
luminescence intensity and efficiency of phosphors can reach a content (for La(1)/Ce(1)–N and La(2)/Ce(2)–N bond lengths by 5%,
maximum value but also the position of luminescence spectra and by 21%, respectively) in La3xCexSi6N11 can contribute to an
can be tuned to some extent. To our knowledge, changing the increase in CFS. Therefore, they concluded that if multi-cation sites
doping levels of Ce3+ or Eu2+ ions in phosphors is the easiest existed, the increase in distortion of the bond length distribution of
and most common way to realize spectral tuning, namely, changing different activator–anion bonds could also cause an increase in CFS
the energy difference between the 5d and 4f levels of Ce3+ or Eu2+ to generate redshift.53 Upon changing the doping level of activators,
ions among all available approaches. For example, George et al. the spectral tuning can also be realized by the crystal-site engineer-
reported a redshift of B30 nm of the position of the maximum ing approach, which involves the replacement of activator ions into
emission wavelength in the La3xCexSi6N11 (0 o x r 3) system with different cation sites owning to different coordination environments
increasing Ce3+ doping concentration, as shown in Fig. 3c.53 This is and energy transfer of activator ions at different crystal sites.86,87
mainly ascribed to the increased crystal field splitting (CFS) due to This section will be elaborately introduced in Section 3.5.

This journal is © The Royal Society of Chemistry 2015 Chem. Soc. Rev.
View Article Online

Review Article Chem Soc Rev

substitutions in the host are classified into three kind patterns:


Ch substitution, Cic substitution and Ch–Cic substitution.
3.2.1 Host cation (Ch) substitution. For Ch substitution, a
large amount of cases such as Ce3+ or Eu2+ ion doped alkaline
earth metal silicates, aluminates, and (oxy)nitrides have been
investigated and reported, as shown in Table 1. It commonly
involves two mechanisms to explain spectral tuning: the average
lattice variation and the local lattice variation of the crystal
structure. On the basis of the average lattice variation of the
Published on 30 September 2015. Downloaded by University of Tasmania on 30/09/2015 10:15:45.

crystal structure, Denault et al. recently revealed a tunable


excitation and emission in SrxBa2xSiO4:Eu2+ by adjusting the
Sr/Ba ratio.54 With increasing Sr content, the excitation spectrum
exhibits a slight redshift (Fig. 5a), while the emission spectrum
shows a large redshift with a maximum emission wavelength of
574 nm (yellow-orange) for Sr2SiO4:Eu2+ and of 511 nm (green) for
Fig. 4 A schematic explanation of (a) the variation of the host lattice. Ba2SiO4:Eu2+ (Fig. 5b), increasing the Stokes shift. Because
Green and gray spheres represent host cations and anion ligands, respec- Ba2SiO4 and Sr2SiO4 all can crystallize in the orthorhombic
tively. Blue and red spheres are activator ions. The ionic radius is blue 4
b-K2SO4 structure, and the addition of Eu in Sr2SiO4 can’t change
green 4 red spheres. The replacement of green spheres by blue or red
spheres will increase or decrease the activator–anion bond length, this structure. Therefore, the intermediate compositions of the
respectively. (b) The spectral red shift and blue shift mechanisms induced SrxBa2xSiO4:Eu2+ phosphor maintain the same structure for all
by the variation of the crystal lattice. The increase of the activator–anion Sr concentrations. The unit cell of SrxBa2xSiO4 contains two
bond length reduces crystal field splitting and the nephelauxetic effect, cation sites, the 10-coordinate Wyckoff site 4c (Sr/Ba1) with a
resulting in the spectral blue shift. On the contrary, a spectral red shift will
larger polyhedral volume and the 9-coordinate Wyckoff site 4c
appear.
(Sr/Ba2) with a smaller polyhedral volume, as shown in Fig. 5c.
The comparison of the ionic radius for Sr2+, Ba2+, and Eu2+ in
In general, phosphors frequently involve in the substitution Sr/Ba1 and Sr/Ba2 indicates possible site mixing of Sr2+ and Ba2+
of cations in the host lattice with activator ions (Ce3+ or Eu2+) to in the intermediate compositions. Moreover, Eu2+ substitution
produce luminescence, which easily results in a mismatch of can occur on both sites. Fig. 5d shows that the unit cell para-
the crystal lattice due to different ion radii. These mismatches meters a, b, c and unit cell volume (V) present a linear decrease
lead to lattice expansion or contraction, which have a substan- following Vegard’s law with increasing Sr content, indicating the
tial effect on CFS and thus further affect the energy difference formation of SrxBa2xSiO4:Eu2+ solid solution and a continuous
between the 5d and 4f states of Ce3+ or Eu2+ ions. A simple decrease in the Sr/Ba–O bond lengths. When Eu2+ ions enter into
model to illustrate the spectral tuning mechanism of phosphors Sr/Ba sites, the average Eu–O bond length also continuously
by controlling the doping level of Ce3+ or Eu2+ ions is shown in decreases and thus leads to an increasing CFS, generating red-
Fig. 4. During the process of doping Ce3+ or Eu2+ ions into the shift emission. Another reason to induce redshift emission is
host lattice, if its ionic radii are smaller than that of host cations, a more Eu2+ ions into the Sr/Ba1 sites with increasing Sr content,
lattice contraction will occur (Fig. 4a). Subsequently, the activator– which offers a larger nephelauxetic effect on the Eu2+ ions.
anion bond lengths will decrease and the interactions between More interestingly, the thermal stability of SrxBa2xSiO4:Eu2+
activator ions and anion ligands will strengthen (Fig. 4b), which phosphors were improved by changing the Sr/Ba ratio. Denault
brings about a larger CFS and a stronger nephelauxetic effect of et al. observed that the intermediate composition with 46% Sr
5d energy levels of Ce3+ or Eu2+ ions. Finally, a red shift emission had the highest resistance to thermal quenching of lumines-
can be observed in this system upon gradually increasing the cence, remaining stable up to 413 K.54 The possible reason is that
activator concentration, whereas a blue shift emission occurs the intermediate compositions possess optimal bonding that
(Fig. 4b). creates a more rigid crystal structure compared to the end-
member compositions. Consequently, they concluded that this
3.2 Cationic substitution result provides an impetus to the search for newly efficient
Cationic substitution is another common route to realize phosphor hosts that possess highly connected structures and
spectral tuning of Ce3+ or Eu2+ doped phosphors by altering optimized bonding interactions. Based on the same mechanism,
their coordination environments. It is well known that there are Wu et al. reported a spectral blue shift from 553–590 nm (yellowish
two kinds of cations in the host including host cations (Ch) and orange light) to 440–470 nm (blue light) in Y1.98Ce0.02(Ca1ySry)F4S2
cations in ionic complexes (Cic).49–51,71 For example, Y3+ ions (0 r y r 1) when gradually replacing Ca2+ with Sr2+ (Fig. 5e),
and Al3+ ions are Ch and Cic in Y3Al5O12, respectively. Both the which is related to the changed crystal field strength.43 Because the
substitutions of Ch and Cic by other cations in the host can above substitution leads to the expansion of the lattice volume due
affect the CFS and NE of Ce3+ or Eu2+ ions, inducing the change to the larger Sr2+ ions, thus changing four Ce–S and four Ce–F
in their energy difference between the lowest 5d energy level bonds to 286.94 and 262.41 pm within the internal YF4S5 poly-
and the ground state. To facilitate discussion here, the cationic hedra. The bond lengths of apical Ce–S, 284.63 and 284.71 pm are

Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2015
View Article Online

Chem Soc Rev Review Article


Published on 30 September 2015. Downloaded by University of Tasmania on 30/09/2015 10:15:45.

Fig. 5 (a) Excitation and (b) emission spectra collected at the maximum emission/excitation wavelength for each sample. (c) Crystal structure of
SrxBa2xSiO4 with SiO4 polyhedra shown as blue, O2 shown as orange, the Sr/Ba1 10-coordinate site shown as dark gray, and the Sr/Ba2 9-coordinate
site shown as light gray. (d) Refined unit cell parameters a, b, c, and V of (Sr,Ba)2SiO4:Eu2+ with increasing Sr content. (e) Excitation (monitored at
550–590 nm) and emission (lex = 440–470 nm) spectra of Y1.98Ce0.02(Ca1ySry)F4S2 (y = 0, 0.1, 0.25, 0.5, 0.75 and 1). The insets are the corresponding
photos taken under 365 nm excitation. (f) Schematic explanation of the spectral tuning mechanism based on the average lattice contraction of the crystal
structure. C1–3 represent host cations and the order of ionic radius is C3 o C1 o C2 (Reproduced with permission from ref. 43 and 54, copyright 2012,
2014, American Chemical Society, Royal Society of Chemistry.)

almost the same in Y1.98Ce0.02SrF4S2 and Y1.98Ce0.02CaF4S2. In such was formed at x = 0–0.63 and a average lattice expansion could be
cases, the Ce3+ ion experiences a weaker crystal field splitting observed with the increase of Ba concentration due to a larger
due to the expansion of YF4S5 polyhedra in the Y1.98Ce0.02SrF4S2 ionic radius of Ba2+ than Sr2+.48 Based on the previous mechanism,
system; therefore, it is reasonable to observe the blue-shifted a blueshift of the maximum emission would be expected. However,
excitation and emission spectra. More evidence can be found an abnormal redshift emission about 35 nm was observed with
about spectral tuning and the improvement of thermal stability increasing Ba content in the (Sr0.98xBaxEu0.02)Si2O2N2 (x = 0 to
based on the average lattice variation of the crystal structure in 0.63) system, as shown in Fig. 6a. The corresponding color
(Sr2xBax)Ga2SiO7:Eu2+,88 (Sr1xCax)Si2O2N2:Eu2+,89 Sr4xMgxSi3- coordinates that moved from the green to the yellow region and
O8Cl4:Eu2+,90 (Ba1xSrx)9Sc2Si6O24:Ce3+,Li+,91 A32xRECexNax(BO3)3 the corresponding digital photos (Fig. 6b) also verified this red-
(A = Sr, Ca; RE = Y, La),92 a-(Y,Gd)FS:Ce3+,42 and (Sr1xyCax- shifted luminescence. Li et al. investigated the local fine structure
Bay)2Si5N8:Eu2+.93–95 A schematic explanation is shown in Fig. 5f. of (Sr0.98xBaxEu0.02)Si2O2N2 (x = 0 to 0.63) phosphors by the
Upon using a larger host cation (C2) to replace a smaller one (C1), extended X-ray absorption spectra (EXAFS) and found that there
an average lattice expansion occurs, which brings out a linear is no a linear increase in the Sr/Ba–O distance, whereas the big
increase in the bond length of activator–anions followed by doping difference in end-member crystal structures (triclinic: P1 for
of the activator ions (Ce3+ or Eu2+) into C1/C2 sites. This decreases SrSi2O2N2 and orthorhombic: Pbcn for BaSi2O2N2).48 On the
the interaction between activator ions and anion ligands, and contrary, the Sr–O and Sr–O/N bond lengths gradually decrease
weaken the CFS strength and the nephelauxetic effect, and thus with the increase in x in (Sr0.98xBaxEu0.02)Si2O2N2 (x = 0 to
shifts the maximum emission position to the shorter wavelength, 0.63) series, as shown in Fig. 6c.
as shown by the (i) process in Fig. 5f. On the contrary, the redshift Because Eu2+ preferentially replace Sr2+ sites in (Sr0.98xBax-
emission will appear if the Ce3+ or Eu2+ ions enter the smaller Eu0.02)Si2O2N2 (x = 0 to 0.63) due to the isotypic structure of
C3/C1 sites, as shown by the (ii) process in Fig. 5f. It is noted that EuSi2O2N2 and SrSi2O2N2, and the similar ionic radius of Eu2+
although a linear decrease/increase of the average activator–anion and Sr2+. Therefore, they proposed a possible mechanism for
distance appears, abnormal blueshift/redshift might be observed the unexpected redshift emission based on the local lattice
in some systems, e.g., (Sr0.98xBaxEu0.02)Si2O2N2,96–100 Sr3xBax- variation,48 as shown in Fig. 6d: first, in a local structure, one
SiO5:Eu2+,101,102 (Ba1xSrx)Al2Si2O8:Eu2+,103 (Ca0.993xyMgxSry)9- Eu2+ occupies one Sr2+ site in (Sr0.98xBaxEu0.02)Si2O2N2 (x = 0 to
Y(PO4)7:Eu0.0072+,104 (Na1xCax)(Sc1xMgx)Si2O6:Eu2+,105 and 0.63) series. The strain of the Ba–O bond increases when the
(Y,Gd,Lu)3Al5O12:Ce.106 The possible reason is that the local lattice neighboring Sr2+ ion is replaced by one larger Ba2+ ion (i).
variation originating from the neighboring cation exerts a more To release lattice strain, the Eu–O bond length decreases (i).
important effect on the luminescence properties of Ce3+ or Eu2+. For With the increase of Ba2+ content, other neighboring Sr2+ sites
instance, in the (Sr0.98xBaxEu0.02)Si2O2N2 system, a solid solution are successively replaced by Ba2+ ions (ii and iii), thereby gradually

This journal is © The Royal Society of Chemistry 2015 Chem. Soc. Rev.
View Article Online

Review Article Chem Soc Rev

increases the strength of the crystal field. As a result, the 5d


energy levels of Eu2+ are lowered, and the emission wavelength
is redshifted from 486 to 508 nm with Sr2+ content increasing.
Except for the main CFS effect, the nephelauxetic effect also
exerts an import effect on the resulting spectral shift. In fact,
the final luminescence spectra of many phosphor systems
are the complex results of several effects including the CFS
effect, the NE effect and so on. Sometimes, the CFS effect and
the NE effect play a conflicting role in spectral tuning, and the
Published on 30 September 2015. Downloaded by University of Tasmania on 30/09/2015 10:15:45.

ultimate results depend on the bigger effect. For example, Jang


et al. observed that the emission spectra of (Sr1xBax)3SiO5:Eu2+
appeared to be first redshifted (x = 0–0.2) and then blue shifted
(x = 0.2–1.0) with the increasing x value from 0 to 1.0, as shown
in Fig. 7a.101 They thought that there exist the competitive CFS
effect and the NE effect in (Sr1xBax)3SiO5:Eu2+ systems. Upon
increasing the Ba content in (Sr1xBax)3SiO5:Eu2+, the former
enables the CFS effect of the 5d level of the Eu2+ ion to increase
in terms of the local lattice variation mechanism, while the
latter results in an reduced CFS effect due to the increased
Eu–O covalency. At x r 0.2, the crystal field splitting effect
dominates and thus appears to be redshifted (Fig. 7b).
However, after x Z 0.2, the NE effect surpasses the CFS
effect in (Sr1xBax)3SiO5:Eu2+, leading to a blue shift (Fig. 7b). In
addition, Xia et al. presented a double Ch substitution model
in (Na1xCax)(Sc1xMgx)Si2O6:0.03Eu2+, by which they can tune

Fig. 6 (a) Typical PL emission spectra of (Sr0.98xBaxEu0.02)Si2O2N2 (0 r


x r 0.63) phosphors with the increase of Ba contents (lex = 405 nm).
(b) The corresponding CIE color coordinate diagram of the above samples
and digital photos for x = 0, 0.30, 0.63 samples. (c) Extended X-ray
absorption spectra of (Sr0.98xBaxEu0.02)Si2O2N2 (x = 0 to 0.63) samples.
(d) Schematic explanation of the redshift luminescence mechanism in
(Sr0.98xBaxEu0.02)Si2O2N2 (0 r x r 0.63) series. (e) Dependence of
emission spectra of (Ca0.993xyMgxSry)9Y(PO4)7:Eu2+0.007 phosphors on
concentrations of Mg2+ and Sr2+ (Reproduced with permission from ref. 48
and 104, copyright 2011, 2014, American Chemical Society, Royal Society
of Chemistry.)

decreasing the Eu–O bond length. This result increases 5d-orbital


crystal field splitting of Eu2+ ions, causing a continuous redshift in
the emission spectra. Similar situation occurs in (Ca0.993xy-
MgxSry)9Y(PO4)7:Eu0.0072+ systems. There is a continuous blue-
shift emission from 486 to 435 nm with increasing Mg2+ content
(x r 0.5) and a redshift from 486 to 508 nm with increasing Sr2+
concentration (y r 0.5), as indicated in Fig. 6e.104 According to
the local lattice variation mechanism, the above phenomenon can
be explained as follows: when Ca2+ is substituted by a smaller
Mg2+ ion, the distance between Eu2+ and O2 becomes longer,
Fig. 7 (a) PL spectra of (Sr1xBax)3SiO5:Eu2+ samples with various Ba con-
resulting in smaller crystal field splitting, so that there is a
tent: x = 0, 0.2, 0.4, 0.6, 0.8 and 1.0. (b) Schematic diagram of the energy
continuous blue-shift with the doped Mg2+ concentration. In band change of Eu2+ in (Sr1xBax)3SiO5:Eu2+ phosphors. D represents the
contrast, when Ca2+ is substituted with a larger Sr2+ ion, the crystal field splitting of the 5d level of the Eu2+ ion (Reproduced with
distance between Eu2+ and O2 becomes shorter, which in turn permission from ref. 101, copyright 2009, The Electrochemistry Society.)

Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2015
View Article Online

Chem Soc Rev Review Article

the emission from yellow light (537 nm) to blue light (454 nm) with other effects. Therefore, it is widely used in the modification of Ce3+
the x value from 0 to 1, as shown in Fig. 8b.105 The tunable yellow- or Eu2+ activated phosphors for promoting their application in
blue emission in this system is related to the local variation of the WLED lighting and display areas.
crystal structure. The crystal structure of NaScSi2O6 is isomorphous 3.2.2 Cations in ionic complex (Cic) substitution. As men-
with that of CaMgSi2O6, and the octahedral Mg site is occupied by tioned in the previous section, there are two kinds of cations in
Sc and the Ca site is occupied by Na (Fig. 8a). Therefore, they can most inorganic phosphor hosts: the host cation (Ch) and the
form a complete series of solid-solutions. In this solid-solution, cation of ionic complexes (Cic).49–51,70 The role of Cic is to
the replacement of Mg2+ by Sc3+ is strongly accompanied by the connect anion ligands for forming stable frames of inorganic
replacement of Ca2+ by Na+, to ensure the charge balance. The Na crystals and cation vacancies. Ch and doping cations can
Published on 30 September 2015. Downloaded by University of Tasmania on 30/09/2015 10:15:45.

environment of the Eu2+ center leads to the spectrum with maxi- occupy these cation vacancies to form host materials and the
mum emission in the yellow region, while the Ca environment resulting phosphors, respectively. The common Ch in inorganic
produce another emission center in the blue region. There is a phosphors contains alkali metal ions (Li+, Na+, and K+), alkaline
random distribution of these small NaSc units and CaMg units. earth metal ions (Mg2+, Ca2+, Sr2+, and Ba2+), rare earth metal
However, there are local clusters of NaSc and CaMg that would ions (Y3+, Sc3+, La3+, Gd3+, and Lu3+) etc., and the usual Cic is
result in two different Eu2+ sites with a difference in the covalency. focused on the third and fifth main groups of the periodic
Therefore both bands (blue and yellow emission) are visible in the table, e.g. B3+, Al3+, Ga3+, Si4+, Ge4+, P5+ and so on. The Cic
solid solution. Moreover, the ratio of yellow emission to blue substitution can be divided into the equivalent valence and
emission varied with x values due to the different ratios of NaSc nonequivalent valence substitution based on the replacement
local clusters to CaMg clusters, realizing a yellow-blue tunable of Cic by metal ions in the same group and different main
emission. In other words, Ch substitution is an effective route to groups, respectively. For example, Ga3+ - Al3+, and Ge4+ - Si4+
realize spectral tuning by the competition of CFS, NE and some are equivalent valence replacements, while Si4+ - Al3+ is a
nonequivalent valence replacement. The spectral tuning
mechanism by two Cic substitutions can also change the local
coordination environments of activator ions, and thus influences
their energy difference between the lowest 5d state and the
ground state. The most representative examples are to tune the
luminescence spectra of garnet-structured phosphors by modify-
ing the composition of Al with Ga and Si.107 Ueda et al. reported
that the Ce3+ luminescence was blue-shifted with increasing Ga
content in (Ce0.005Y0.995)3Sc2Al3xGaxO12 (x = 0, 1, 2, 3), as shown
in Fig. 9a.108 In the photoluminescence excitation (PLE) spectra,
two excitation bands at approximately 340 nm and 430 nm were
attributed to transitions from the 4f level to the 5d1 and 5d2 levels
(5d1 and 5d2 are the lowest and the second lowest 5d levels),
respectively. When doping the larger Ga into the (Ce0.005Y0.995)3-
Sc2Al3xGaxO12 system, a linear increase of unit cell parameters
was observed (Fig. 9b), indicating the formation of solid solution.
With increasing lattice constants, the excitation energy of the
4f–5d1 peak shifts to higher energies and the 5d1–4f emission
peak also shifts to higher energies. However, the excitation
energy of the 4f–5d2 peak shifts to lower energies. Moreover,
the energy difference between the 5d1 and 5d2 levels decreased
with increasing lattice constants, as shown in Fig. 9c. The
decrease in the splitting energy between the 5d1 and 5d2 levels
shows that the crystal field strength becomes weaker with
increasing Ga content at the tetrahedral site. However, if the 5d
levels were modified by the nephelauxetic effect, both the 5d1 and
the 5d2 level would shift to lower or higher energies and would
maintain a constant splitting energy between the 5d1 and 5d2
level. Therefore, they concluded that the blue shift of the 5d1 level
with increasing Ga content at the tetrahedral site was caused
Fig. 8 (a) The polyhedral representation of the structure of the mainly by crystal field splitting not by the nephelauxetic effect.
(Na1xCax)(Sc1xMgx)Si2O6:Eu2+ solid-solution sample. (b) PL spectra (lex =
Briefly, the substitution of smaller Al3+ ions by larger Ga3+ ions
365 nm) of (Na1xCax)(Sc1xMgx)Si2O6:Eu2+ (x = 0–1) phosphors with different
x values, and the upper inset gives the digital images of some selected
enables the average crystal structure to expand and the rigidity
phosphors under a 365 nm UV lamp. (Reproduced with permission from of the unit cell decreases, leading to a loosed environment
ref. 105, copyright 2013, Nature Publishing Group.) surrounding the Ce3+ ions. Accordingly, the crystal field splitting

This journal is © The Royal Society of Chemistry 2015 Chem. Soc. Rev.
View Article Online

Review Article Chem Soc Rev

the local structure of phosphors. Recently, Shang et al. realized


the redshift emission in the Y3Al52xMgxSixO12:Ce3+ (41 nm)
and the Y3Al52xMgxGexO12:Ce3+ (19 nm) series by replacing
Al(1)3+–Al(2)3+ pairs with Mg2+–Si4+ and Mg2+–Ge4+ pairs, respec-
tively (Fig. 11a and b).111 In general, this should be mainly
assigned to the lattice distortion in local crystal structures. The
detailed mechanism can be explained as follows: the ionic radius
of Mg2+ with 6-fold coordinated O is larger than that of Al3+. When
the Al3+(1) site is occupied by the Mg2+ ion, the Mg2+–O2 bond
Published on 30 September 2015. Downloaded by University of Tasmania on 30/09/2015 10:15:45.

length increases, while the adjacent Ce3+–O2 bond length (RCe–O)


decreases, resulting in the increase of the crystal field strength. So
the emission peak shifts to a longer wavelength. On the other
hand, the smaller Si4+ or Ge4+ ion substituted for Al(2)3+ would lead
to a longer RCe–O bond length and decrease the crystal field
strength around the Ce3+ ion, so that the emission spectra would
be blue-shifted. In the Y3Al52xMgxSixO12:Ce3+ and Y3Al52xMgxGex-
O12:Ce3+ solid solutions, Mg2+–Si4+/Ge4+ as ion pairs substituted for
Al(1)3+–Al(2)3+ are synchronously incorporated into the Y3Al5O12:Ce3+
host lattice. According to the crystal structure shown in Fig. 11c,
Fig. 9 (a) The photoluminescence excitation (PLE) and emission (PL) YO8 dodecahedra share the vertex with SiO4 tetrahedra, while YO8
spectra of the (Ce0.005Y0.995)3Sc2Al3xGaxO12 (x = 0, 1, 2, 3) (Ce:YSAGG) dodecahedra share the edge with MgO6 octahedra. The augmenta-
solid solutions. (b) Lattice constants of Ce:YSAGG. (c) Lattice constant tion of the crystal field strength arising from Mg2+ substitution
dependence of transition energies of the Ce3+ ions in Ce:YSAGG. (d) The
would dominate, namely, DE3 4 DE2 4 DE1, resulting in the
PL of Y2.9Ce0.1Al5xSixO12 as a function of x values. (Reproduced with
permission from ref. 108 and 109, copyright 2013, 2014, Elsevier.)
redshift emission in the above two phosphor systems. Moreover,
there is a larger redshift in the Mg2+–Si4+ systems than that in
Mg2+–Ge4+ systems.
of the 5d levels of Ce3+ ions in Ce:YSAGG decreased and induced
a blueshift of the emission spectra. Unfortunately, a decrease in
the rigidity of phosphors with increasing Ga or Si contents also
increased the nonradiative rate and decreased the distance
between the 5d1 level and the conduction band. The former
damaged the quantum yields of phosphors, and the latter
increased luminescence quenching. Therefore, the researchers
need to search for suitable composition to balance the various
luminescence performances of phosphors. Such phosphors
can produce the maximum effect in WLED lighting and display
applications. Based on a similar mechanism, Zhang et al. found a
blueshift in Y3yCeyAl5xSixO12 by the Si4+ - Al3+ nonequivalent
valence replacement, as shown in Fig. 9d.109 Jung et al. also
reported a tune emission from orange to red in CaAlxSi(73x)/4N3:Eu2+
phosphors by varying the Si/Al ratio, as shown in Fig. 10a.110 Fig. 10b
revealed that the emission spectrum of the representative
CaAl0.7Si1.225N3 sample could be deconvoluted into two Gaussian
components due to the presence of two Ca2+ sites (Ca1 and Ca2)
in the host. Moreover, the deconvoluted Gaussian components
of all non-stoichiometric CaAlxSi(73x)/4N3 phosphors depended
on their corresponding local environments, namely, the Al/Si
molar ratio around the Eu2+ activator. The high Al/Si ratio near
the Eu2+ activator abolished the high energy peak. Therefore, the
Eu2+ activator site in an Al-rich environment with a smaller
average Ca–N distance was assigned to the lower energy emission
peak due to the increased crystal field splitting and nephelauxetic
effect, while that in a Si-rich environment with a longer average
Fig. 10 (a) The emission spectra of CaAlxSi(73x)/4N3:Eu2+ phosphors
Ca–N distance was assigned to the higher energy emission peak. with x = 0.7–1.3 and images of each sample taken under an 365 nm UV
Except for the single substitution of Al3+ by Si4+, some researchers lamp. (b) Gaussian deconvolution of CaAl0.7Si1.225N3. (Reproduced with
exploited double substitution approaches to form the distortion of permission from ref. 110, copyright 2010, Optic Society of America.)

Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2015
View Article Online

Chem Soc Rev Review Article


Published on 30 September 2015. Downloaded by University of Tasmania on 30/09/2015 10:15:45.

Fig. 11 The photoluminescence emission spectra of (a) Y3Al52xMgxSixO12:Ce3+ (x = 0–2) and (b) Y3Al52xMgxGexO12:Ce3+ (x = 0–1.6) with the excitation
monitoring wavelengths of 456 nm. (c) The crystal structure of Y3Al5O12 and the schematic illustrates the replacement of Al(1)3+–Al(2)3+ pairs by
Mg2+–Si(Ge)4+ pairs. (Reproduced with permission from ref. 111, copyright 2014, American Chemical Society.)

3.2.3 Ch–Cic substitution. Since both Ch and Cic substitutions thermal quenching behaviour based on the (2) and (3) Ch–Cic
can distort the microenvironments of the local structure surroun- substitution types that substituted Li+/Si4+ for Ca2+/Al3+ (Li series)
ding Ce3+ or Eu2+ ions in phosphor compounds, namely, by and La3+/Al3+ for Ca2+/Si4+ (La series) in the CaAlSiN3:Eu2+ lattice.50
changing their CFS and NE effects, the simultaneous Ch–Cic Linear shifts in the lattice parameters in both series with increas-
substitutions are predicted to possibly further intensify the ing x, as plotted in Fig. 13a, which demonstrates the formation of
above effects and thus have been developed to tune the lumi- solid solutions in both materials. The sum of the cation radii of
nescence properties of phosphors recently.112,113 Considering a the CaAlSiN3 parent structure is smaller than that of the La lattice,
charge balance in phosphors that is used to reduce nonradia- but larger than that of the Li lattice. Therefore, the lattice
tive transitions, the Ch–Cic substitutions mainly contained the expansion with x in the La series occurs, while a lattice contrac-
following several types: (1) A+ + C3+ 2 2B2+ or B12+ + B22+, (2) A+ tion in the Li series is observed. The maximum emission peak in
+ D4+ 2 B2+ + C3+, (3) B2+ + D4+ 2 2C3+ or C13+ + C23+, (4) B2+ + the La series shifts to the shorter wavelength region with x due to
E5+ 2 C3+ + D4+, (5) B2+ + D4+ 2 A+ + D5+ 2 2C3+ or C13+ + a decrease in the CFS of 5d orbitals caused by lattice expansion
C23+, as shown in Fig. 12; The above 5 types are the predictable (Fig. 13bi). However, an unexpected broadening of the emission
Ch–Cic substitution types; ‘‘2’’ means that the substitution not peak in the shorter wavelength region with increasing x instead
only can be executed from left to right but also can be of shifting to the longer wavelength region appears despite the
substituted from right to left; A, B, C, D, and E in these lattice contraction in the Li series (Fig. 13bii), which is signi-
inorganic phosphors generally represent the alkali metal, the ficantly correlated with the local structural change of activators.
alkali earth metal, the IIIA and IIIB metal, the IVA metal, the VA In addition, with increasing x values in the two systems, the
metal elements in the periodic table, respectively, for example, thermal quenching of the La series increased while the thermal
A are Li, Na, and K; B are Mg, Ca, Sr, and Ba; C are Sc, Y, La, Gd, quenching of the Li series decreased, as depicted in Fig. 13c.
Lu or B, Al, Ga, and In; D are Si and Ge; E are P, Mo, and W. B1 Wang et al. proposed a possible mechanism that the accom-
and B2 are different alkali earth metal elements, while C1 and modation of the Eu2+ activator is controlled by variable neighboring
C2 are different IIIA and IIIB metal elements. In addition, there is a cations to elucidate the variations in the photoluminescence and
basic requirement that phosphor compounds should form a quenching behaviour in both series. Schematic models for the
solid-solution or a single phase after Ch–Cic substitution. Wang neighboring-cation induced effect are shown in Fig. 13d. This
et al. systematically investigated spectral tuning properties and model explains the site occupation of Eu2+ activators in both series,

This journal is © The Royal Society of Chemistry 2015 Chem. Soc. Rev.
View Article Online

Review Article Chem Soc Rev


Published on 30 September 2015. Downloaded by University of Tasmania on 30/09/2015 10:15:45.

Fig. 12 (left) Schematic illustration of the Ch–Cic substitutions in Ce3+ or Eu2+ activated phosphors. (right) The typical types of Ch–Cic substitutions.
(5) are the predictable types. ‘‘2’’ means that the substitution can be executed from left to right or from right to left; A, B, C, D, and E in these inorganic
phosphors generally represent the alkali metal, the alkali earth metal, the IIIA and IIIB metal, the IVA metal, the VA metal elements in the periodic table,
respectively. B1 and B2 are different alkali earth metal elements, while C1 and C2 are different IIIA and IIIB metal elements. The blue and green spheres are
Ch, the red and yellow are Cic, and the gray spheres are anion ligands.

view of the larger ion size of Eu2+ than Ca2+, thereby providing
a weak covalent coordination environment for Eu2+. A reverse
argument can predict the accommodation of Eu2+ ions in the Li
series. Therefore, the Eu2+ activator is controlled to locate at the
relatively loose outer Ca2+ site, which is connected to the nitride
with Li+ and SiSi8+ pair branches, namely, the Si-rich site. The
Eu2+ activators in both series locate at the relatively loose site for
minimizing the lattice strain. Thus, the photoluminescence in the
La series is expected to blueshift because there is a progressive
decrease in the covalency and crystal field strength with x. By
contrast, the nitrides connecting with excess Si4+ neighbors in the
Li series possess a smaller nephelauxetic effect than that connecting
with Si4+/Al3+-equivalent neighbors for Eu2+ ions. This result creates
the other high-energy emission as fitted in Fig. 13bii. In general, a
remote control effect that guides Eu2+ activators into different
Ca2+ sites is proposed for neighboring-cation substitution when
the Si4+/Al3+ ratio varies with the valence of the Mn+ (n = 1–3) cation.
According to the remote control effect, the Eu2+ activators are
Fig. 13 (a) Relative changes in the lattice parameters with x in (Ca1xLax)- surrounded by nitride anions that neighboring with M3+-
(Al1+xSi1x)N3:Eu (La series) and (Ca1xLix)(Al1xSi1+x)N3:Eu (Li series). (b) PLE dominant and Si4+/Al3+-equivalent coordination when M is a
and PL spectra of (i): La series and (ii): Li series (lex = 460 nm). The arrows trivalent. However, when M is a monovalent, the Eu2+ ions
present the energy shift for the emission spectra of the La series and the
will be surrounded by nitride anions neighboring with M+-
relative intensity change for the Li series with increasing x. The PL spectra
of the Li series are decomposed into two Gaussian peaks from different dominant and Si-rich coordination.50 This mechanism not only
Eu2+ environments. (c) Thermal quenching behavior for La series and can reveal the underlying mechanisms in optical changes by
Li series. (d) Local coordination of CaAlSiN3, La and Li series lattices. The adjusting the cation/anion composition of phosphors but also
dashed line represents the longer and looser bonds. The broad solid line can efficiently tune optical properties, especially thermal stability.
represents the shorter and tenser bonds. (Reproduced with permission
Therefore, it could be general to luminescent materials that are
from ref. 50, copyright 2013, American Chemical Society.)
sensitive to valence variation in local environments. Other types of
simultaneous Ch–Cic substitutions such as La1x0.025Ce0.025Sr2+x-
as presented in Fig. 13dii and diii. Since the bonding distance of Al1xSixO5 (x = 0–0.975) and (Ca1xCex)6Ba(P1xSix)4O17 (x =
the nitride and Ca2+ that connects with the La3+ and AlSi7+ pair 0–0.07) have been reported to tune spectral positions or improve
becomes longer due to the neighboring-cation induced effect. emission intensity.112,113 The above results obviously confirm that
When Eu2+ activator is introduced, it would preferably be accom- the Ch–Cic substitutions based on the variation of the local
modated at the looser Ca2+ site to minimize the lattice strain in environment of activators can efficiently tune the luminescence

Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2015
View Article Online

Chem Soc Rev Review Article

properties of phosphors, and thus could be greatly developed. Ca3Sc2Si3O12:Ce3+ resulting in a great enhancement in the
Moreover, on the basis of the summarized (1)–(4) types of longer wavelength (610 nm) side of the typical green band,
Ch–Cic substitution, some new Ch–Cic substitution such as correspondingly leading to the change of luminescent colors
B2+ + D4+ 2 A+ + D5+ 2 2C3+ or C13+ + C23+, as shown by (5) from green to yellow-orange, as shown in Fig. 14a.116 This is
in Fig. 12, can be predicted. This can help improve the assigned to the presence of N3 around Ce3+ ions in Ca3Sc2-
application of phosphors in WLED lighting and display areas. Si3O126xN4x:Ce3+ (x = 0–0.8) systems. Because N3 can offer
However, the underlying luminescence mechanisms in many the stronger nephelauxetic effect on Ce3+ than O2, the incor-
Ch–Cic substituted systems are still unclear, and need to be poration of N3 anions leads to a decrease in the energy
further investigated and explored. difference between the lowest 5d excited state and the ground
Published on 30 September 2015. Downloaded by University of Tasmania on 30/09/2015 10:15:45.

state of Ce3+ ions, which promotes a redshift emission.


3.3 Anionic substitution However, a decreased emission intensity of the representative
In inorganic phosphor compounds, the first coordination layer Ca2.97Sc2Si3O126xN4x:0.03Ce3+ phosphor was also observed
of host cations and activator ions consists of various anions, with the increase of x values in the inset of Fig. 14a. The
and the interaction of activator ions with anions directly influ- possible explanation is that the substitution of O2 by N3 in
ences the energy difference between 5d levels and the ground Ca3Sc2Si3O126xN4x generates oxygen vacancy defects, resulting
state, which is described by the nephelauxetic effect.58–69 in the enhancement of the non-radiative transition host, and
Commonly, the nephelauxetic effect is linked to the covalency thus quenches the luminescence. Based on the similar mechanism,
of the crystal.59 As the degree of covalency between the activator Setlur et al. realized a distinct low-energy Ce3+ absorption and
ion and the surrounding anions increases, the stronger inter- emission bands in Ce3+ doped RE3Al5O12:Ce3+ (RE = Lu3+, Y3+,
action causes the energy difference between the lowest 5d state or Tb3+) garnet phosphors through the introduction of N3 into
and the ground state of activator ions to decrease. Thus the
covalency of the crystal is related to the electronegativity of
anions and the cation—anion distance. For the same cation,
the higher negative charge and the smaller ionic radius of the
coordinated anion usually lead to the stronger covalency
between the cation and the anion. Thus, the nephelauxetic effect
(centroid shift) in the descending order is selenides 4 sulphides 4
nitrides 4 oxides 4 fluorides, namely, Se2 4 S2 4 N3 4
O2 4 F. Typically, Ce3+ or Eu2+ doped sulfide and nitride
phosphors have a longer emission wavelength than oxide phos-
phors and commonly give yellow-to-red emission.41–47,93,95 If the
activator ions were surrounded by anion groups, the higher
charge density on the anion groups would exert the stronger
nephelauxetic effect on activator ions. The charge density of some
common anion groups is oxides 4 aluminates 4 silicates 4
borates 4 phosphates.71 On the other hand, the interaction
between the molecular orbitals of neighboring anions causes
the 5d orbitals of the activator ion to hybridize, creating bonding
orbitals at lower energies and antibonding orbitals at higher
energies. With the increase of the degree of covalency, both the
bonding and antibonding orbitals move to lower energies, which
causes a red shift. Thus a blue shift occurs with a decrease in the
degree of covalency. In addition, the coordination geometry also
affects the nephelauxetic effect and further changes the crystal
field splitting of Ce3+ or Eu2+ 5d levels. By comparing five different
coordination geometries, Dorenbos determined that octahedral
coordination generates the largest crystal field splitting, followed
by cubic coordination and dodecahedral coordination, and the
smallest crystal field splitting originates from tricapped trigonal
prisms and cuboctahedral coordination.1,71 In general, the varia-
tion of anion coordination circumstances around Ce3+ or Eu2+
ions, namely, anionic substitution, can affect the optical properties
Fig. 14 (a) PL spectra and relative intensity (inset) for Ca2.97Sc2-
of phosphors due to the change in crystal field splitting strength
Si3O126xN4x:0.03Ce3+ (x = 0–0.8). (b) Excitation and emission spectra of
and the nephelauxetic effect. One of the most commonly used Sr1.984Si(N,O)4:0.016Eu2+ (0.7 o z o 1.2). Green lines are two Gaussian
anion substitutions is to change the N/O ratio of phosphor fitting components. (Reproduced with permission from ref. 116 and 118,
compounds.114,115 Recently, Liu et al. introduced N3 into copyright 2011, 2013, Royal Society of Chemistry.)

This journal is © The Royal Society of Chemistry 2015 Chem. Soc. Rev.
View Article Online

Review Article Chem Soc Rev

phosphor hosts.117 In addition, changing the nephelauxetic effect the physical properties of functional materials that are sensitive
in some common phosphors by designing anionic substitution to the coordination environment and the valence change at specific
could obtain unusual luminescence properties. For example, by sites such as magnetic, electrical and mechanics properties.49,50
increasing the N/O ratio of the host, Zhao et al. obtained novel Recently, it has also been extended to tune the luminescence
red-emitting Sr2SiNzO41.5z:Eu2+ (0.7 o z o 1.2) phosphors with properties of Ce3+ or Eu2+ ion activated phosphors.49 The simulta-
the maximum emission wavelength at 617 nm (Fig. 14b), which neous C–A substitutions can controllably change the crystal field
has a obvious redshift than that of Sr2SiO4:Eu2+ phosphors splitting and nephelauxetic effects around Ce3+ or Eu2+ ions, and
(530–560 nm).118 This might originate from the formation of thus can realize a controllable adjustment of spectra and thermal
new Eu2+–N sites in Sr2SiNzO41.5z:Eu2+ (0.7 o z o 1.2) systems. quenching behaviour of phosphors. Based on the electrically
Published on 30 September 2015. Downloaded by University of Tasmania on 30/09/2015 10:15:45.

Since the nephelauxetic effect of N3 is larger than O2, it leads neutral principle in compounds, the C–A substitutions are generally
to an enhanced covalency between Eu2+ and anions, finally classified into the following types: (1) C3+ + Y2 2 D4+ + Z3
resulting in redshift emission. Generally, the increase of the (Al3+ + O2 2 Si4+ + N3), (2) D4+ + Y2 2 C3+ + X (Si4+ + O2 2
N/O ratio will enable the excitation and emission spectra of Al3+ + F), (3) C3+ + U4 2 B2+ + C3 (Y3+ + C4 2 Sr2+ + N3), (4)
phosphors to shift the lower energy position, while a reverse D4+ + U4 2 C3+ + Z3 (Si4+ + C4 2 Al3+ + N3), (5) D4+ + Y2 2
change will generate blue-shifted luminescence. Of course, in E5+ + Z3 2 C3+ + X (Si4+ + O2 2 P5+ + N3 2 Al3+ + F), and
addition to changing the N/O ratio, other anionic substitutions (6) E5+ + Y2 2 D4+ + X (P5+ + N3 2 Si4+ + F), as shown in
with different electronegativities can also designed to tune the Fig. 15. The expressions in brackets are representative C–A
luminescence spectra of phosphors. Xia et al. reported an asym- substitution examples. The (5) and (6) substitution types are the
metric blue emission band from 420 nm to 550 nm with a peak predictable Ch–Cag substitution types. ‘‘2’’ means that the sub-
at 442 nm in Ba5SiO4Cl6:Eu2+ under UV light excitation.119 Eu2+ stitution can be substituted from left to right or from right to left.
ions in Ba5SiO4Cl6 formed two emission centers, namely Eu(1) B, C, D, and E in these inorganic phosphors generally represent
and Eu(2), with maxima at 442 and 470 nm, respectively. some of the alkali earth metal, the IIIA,B metal, the IVA metal, the
Moreover, Eu(1) centers act as the main luminescent centers VA metal elements in the periodic table, respectively, for example,
in the Ba5SiO4Cl6:Eu2+ phosphor. By replacing Cl by F, the B are Mg, Ca, Sr, and Ba; C are Sc, Y, La, Gd, Lu or Al, Ga, and In;
emission peaks of Ba5SiO4(FxCl6x):Eu2+ (x = 1, 3, 4, 6) phos- D are Si and Ge; E is P. X, Y, Z, and U are F, O, N and C elements,
phors presented a green emission band from 450 nm to 575 nm respectively. Similarly, the solid-solution should be formed after
with centers at 503 nm. Because the emission band on the long C–A substitution in phosphor compounds. These C–A substitu-
wavelength side of the Ba5SiO4Cl6:Eu2+ phosphor was centered tions will be illustrated through some representative examples as
at 470 nm, with a full width at half maximum (FWHM) of follows.
50.5 nm, while the main emission band of the Ba5SiO4(FxCl6x):Eu2+ Many oxides and nitrides, particularly silicon nitride-based
(x = 1, 3, 4, 6) phosphor was centered at 503 nm, with a FWHM materials with good thermal stabilities, such as M2Si5N8 and
of 48.5 nm. It indicates that the two emissions belong to the MSi2O2N2 (M = Ca, Sr, Ba), CaAlSiN3, and SiAlON, are useful
same luminescent center, which should be the Eu(2) centers. phosphor hosts.34,36,49,50,121,122 When doping Ce3+ or Eu2+ ions
Therefore, the spectral red shift from 442 nm to 503 nm in into silicon nitride-based materials, it generates excellent
Ba5SiO4(FxCl6x):Eu2+ systems should be attributed to two yellow to red emission under the excitation of UV or blue light.
reasons. The one is that the addition of F ions generates the Therefore, they are widely combined with a blue light-emitting
strong polarization effect, which promotes the preferred occupa- diode to emit white light for everyday applications. Sometimes,
tion of the loose environment of the Eu(2) sites, generating an in order to satisfy the diverse WLED lighting requirements
intense green emission.119 The other reason is that the F element (for example, different correlated color temperatures and color
has a larger electronegativity than the Cl element, while the F ion rendering indices), the shape, FWHM and position of lumines-
has a smaller ionic radius than Cl. The introduction of F ions cence spectra need to be tuned. Chen et al. showed that the size-
increases the crystal field strength and covalency of Eu2+ ions, mismatch between the host and doping cations based on the
generating a red shift from 470 nm to 503 nm at Eu(2) sites. Except substitution of Si4+–N3 pairs by Al3+–O2 pairs can systematically
for single anion substitution, some anion group substitution, tune photoluminescent shifts in M1.95Eu0.05Si5xAlxN8xOx, leading
for example, substituting SiO44 by PO43 in apatite-structured to a red shift when the M = Ba and Sr host cations are larger than
phosphor systems can also tune the luminescence properties the Eu2+ dopant (Fig. 16a and 17b), but a blue shift when the
of phosphors and improve their luminescence performance.120 M = Ca host cations are smaller than the Eu2+ dopant (Fig. 16c).49
In summary, anionic substitution can realize an efficient and In addition, the size-mismatch tuning of thermal quenching was
controllable adjustment of the optical properties of phosphors, also observed. A local anion clustering mechanism in which Eu2+
which has enlarged the phosphor family used in the pc-WLEDs, gained excess N coordination in the M = Ba and Sr structures, but
and improved their luminescence performances. excess O coordination in the Ca analogues, was proposed for
explaining these mismatch effects, as shown in Fig. 16d. The local
3.4 Cationic–anionic substitution structure of Ca1.95Eu0.05Si4AlN7O (top), where Eu2+ is bigger than
Because cationic–anionic (C–A) substitution can efficiently and the Ca2+ ions and lattice strain are minimized by clustering oxide
systematically control the local structures of inorganic com- anions around Eu2+ cations (and hence more nitrides around Ca2+
pounds to some extent, it has been extensively used to adjust ions), resulting in a blue shift emission. Ba1.95Eu0.05Si4AlN7O

Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2015
View Article Online

Chem Soc Rev Review Article


Published on 30 September 2015. Downloaded by University of Tasmania on 30/09/2015 10:15:45.

Fig. 15 (left) Schematic illustration of the cation–anion (C–A) substitutions in Ce3+ or Eu2+ activated phosphors. (right) (1)–(4) are the typical types of
C–A substitutions. (5) and (6) are the predictable types. ‘‘2’’ means that the substitution can be executed from left to right or from right to left; C, D, and E
in these inorganic phosphors generally represent some of the IIIA,B, IVA and VA metal elements in the periodic table, respectively. The blue and green
spheres are cations in anion groups; the red and gray spheres are anion ligands.

Fig. 17 (a) Excitation and (b) emission spectra of Sr2.975Ce0.025Al1x-


SixO4+xF1x (SASF:Ce3+; x = 0–1). Sr1 and Sr2 polyhedra looking down
the [100] direction are shown to the right of (a); the upper two polyhedra
are from the Sr3AlO4F (SAF) structure and the lower two are from the
Fig. 16 Normalized excitation and emission spectra of M1.95Eu0.05-
Sr3SiO5 (SSO) structure. Light gray, black, orange, and green spheres
Si5xAlxN8xOx:M = (a) Ca, (b) Sr, (c) Ba. The emission spectra are decom-
represent Sr1, Sr2, O, and F atoms, respectively. (c) VUV emission spectra
posed into Gaussian contributions. The arrows illustrate the energy shift
(lex = 254 nm) and dependence of the CIE chromaticity coordinates on
and the change in relative intensity of the spectra with increasing x.
varying z values in Ca11.9Al14zSizO32+zF2z:Eu0.1 (z = 0–0.5). The inset
(d) Schematic local anion distributions driven by the cation size-
shows the schematic variation of the activator site driven by Si4+–O2
mismatch in M1.95Eu0.05Si5xAlxN8xOx phosphors. (e) Emission spectra
incorporation and the irradiated phosphor images of each composition.
of Y2.91Ce0.09Al5xSixO12xNx for x = 0, 0.10, 0.20; (g) emission spectra
(Reproduced with permission from ref. 136 and 139, copyright 2011, 2012,
of Y2.94Ce0.06Al(5x)SixO(12x)Nx for x = 0.2–0.3. (Reproduced with
American Chemical Society, Wiley.)
permission from ref. 49, 117 and 123, copyright 2008, 2012, American
Chemical Society, Wiley.)

redshift was observed in Sr1.95Eu0.05Si5xAlxN8xOx. They found


that the size-mismatch effect originated from the substitution of
(bottom), where Eu2+ is smaller than the Ba2+ ions, tends to Si4+–N3 pairs by Al3+–O2 pairs in M1.95Eu0.05Si5xAlxN8xOx
be coordinated by nitride, while the introduced oxide anions materials could also be used to tune photoluminescence in some
preferentially coordinate Ba2+. Since N3 ions can induce a phosphors having (oxo)nitridosilicate frameworks such as a- and
stronger nephelauxetic effect than O2 ions, Ba1.95Eu0.05- b-SiAlONs and the MSi2O2N2 (M = Ca, Sr, and Ba) family. There-
Si5xAlxN8xOx presents a redshift emission with x. A similar fore, this mechanism was predicted to be general to oxynitride

This journal is © The Royal Society of Chemistry 2015 Chem. Soc. Rev.
View Article Online

Review Article Chem Soc Rev

materials and would be useful in tuning optical and other properties bonding of Ce3+ ions in the structure. Unfortunately, SASF:Ce3+
that are sensitive to local coordination environments. A reverse have a more serious thermal quenching for luminescence, where
substitution of Al3+–O2 by Si4+–N3 was also reported in the maximum PL intensity of the x = 0.5 sample can only retain
Y3Al5O12:Ce3+ garnet phosphors by Setlur and Michalik et al., 55% of the initial value at 170 1C with respect to the 64% of
respectively, and they obtained a similar redshift phenomenon by commercial YAG:Ce3+. This serious thermal quenching arises from
this reverse substitution, as described in Fig. 16e and f.117,123 This softer phonon modes associated with fluorine atoms in the host.
is assigned to the occurrence of N3 in their local coordination However, the SSO:Ce3+ (x = 1.0) phosphor had improved thermal
of Ce3+ ions. In view of the larger covalency of Ce–N than that of quenching properties compared to the x = 0.7 or x = 0.9 materials.
Ce–O, indicating the stronger nephelauxetic effect surrounding Disordering of Al3+/Si4+ and F/O2 provides an alternate quench-
Published on 30 September 2015. Downloaded by University of Tasmania on 30/09/2015 10:15:45.

the Ce3+ ions compared to typical garnets. It is noted that there is ing path in the host lattice which outweighs the effect of softer
also a stronger luminescence quenching in these systems versus phonon modes of F. On the basis of the same substitution type,
typical Ce3+ ions in garnets. This might result from the damage Liu’s group developed an new approach that can tune the valency
due to the rigidity of the Y3Al5O12:Ce3+ lattice due to the size- of Eu by controlling the size of the activator site in phosphors for
mismatch between Al3+–O2 pairs and Si4+–N3 pairs. In general, designing phosphors used in pc-WLEDs.139 They have revealed
the substitution between Al3+–O2 pairs and Si4+–N3 pairs in that by replacing Al3+–F by the appreciate dopant Si4+–O2,
phosphors can change the local environment around the activator the valence state of Eu3+ can be tuned to Eu2+ in Ca12Al14z-
ions, and thus realize a controllable tuning of luminescence SizO32+zF2z:Eu (z = 0–0.5, CASOF:Eu) phosphors due to the
properties of phosphors. By the above substitutions, some promising enlargement of the activator site, which leads to a rise in broad-
phosphor systems including Si–N doped BaMgAl10O17:Eu2+,124 band emission at 440 nm of Eu2+ (Fig. 17c). Moreover, the
Ba(Si,Al)5(O,N)8:Eu2+,125 Sr5Al5+xSi21xN35xO2+x:Eu2+ (x E 0),126 emission intensity of Eu2+ and Eu3+ alternatively increases and
Sr2Al2xSi1+xO7xNx,127 Mval+
m/val+Si12(m+n)Alm+nOnN16n (M = Mg, Ca, decreases with z. The chromaticity coordinates of CASOF:Eu
and Lu (0.5 r m r 2.0, 1 r n r 1.8),128 La1xCexAl(Si6zAlz)- phosphors could be tuned from (0.6101, 0.3513) for z = 0 to
(N10zOz) (z B 1),129 and a-SiAlON series,130–135 have been modified (0.1629, 0.0649) for z = 0.5, as shown in Fig. 17d, suggesting their
or developed. potential for use in WLEDs. They thought that this approach could
Another common C–A substitution type is D4+ + Y2 2 C3+ + not only be limited to present materials but could be general to

X and the representative example is the double substitution other related systems, which may lead to more opportunities
of Si4+–O2 by Al3+–F. According to this strategy, Seshadri’ for developing novel phosphors for WLEDs. As mentioned in
group have made much effort to modify and develop newly SASF:Ce3+, double substitutions of cations and anions may
oxyfluoride-based pc-WLED phosphors recently.136–138 The optimize the Ce3+ ions bonding with the surrounding anion
replacement of Al3+ by Si4+ and of F by O2 not only compensate ligands, improving the quantum efficiency of phosphors. Similar
charge, but also more importantly change the local environment work was reported by Shi et al. and Xie et al. on the (SiC)x–
of the lattice. For example, in the Sr2.975Ce0.025Al1xSixO4+xF1x (AlN)1x:Eu2+ solid-solution phosphors by replacing Al3+–N3 by
(SASF:Ce3+) system, abnormal spectral tuning were obtained by Si4+–C4.140,141 The (SiC)x–(AlN)1x:Eu2+ (x = 0–1) phosphor shows
substituting Si4+–O2 by Al3+–F across the solid solution series broad absorption bands in the 250–425 nm range, and bright
of Sr3AlO4F (SAF) and Sr3SiO5 (SSO) (Fig. 17a and b).136 A blue emissions peaking at 470 nm, which has a more high color
conventional d-orbital splitting model predicts a blueshift in purity (0.135 and 0.167) for the x = 0.5 sample than that of the
the excitation and emission spectra as x increases in SASF:Ce3+, AlN:Eu2+ sample.140 Interestingly, the luminescence intensity
based on the increasing average Sr–O/F distance. However, with increases with the SiC content and reaches its maximum at
the increasing x value in SASF:Ce3+, the emission peak positions x = 0.20.140 The reason is that the formation of (SiC)x–(AlN)1x
gradually move toward longer wavelengths, in the range from solid solution promotes the doping of Eu2+ ions into Al sites
474 nm for SAF:Ce3+ (x = 0) to 537 nm for SSO:Ce3+ (x = 1.0) of the host. Moreover, the composition-optimized (SiC)0.20–
(Fig. 17b). The modified SASF:Ce3+ phosphors have wider emission (AlN)0.80:0.006Eu2+ phosphor shows a small thermal quenching
peaks than SAF:Ce3+, which allows SASF:Ce3+ phosphors to provide effect than AlN:Eu2+. This is because the incorporation of SiC into
a higher color rendering properties when incorporated into an the AlN host increases the rigidity of the crystal lattice.
LED chip. The emission red-shift is partially attributed to increas- Generally, the (1), (2) and (4) C–A substitutions only involve
ing distortion of the Ce3+ sites, which causes an increased crystal the substitutions between the Cic and anion ligands. In fact, the
field splitting. Other contributions originate from the increasing C–A substitution could also occur between the Ch and anion
nephelauxetic effect offered by O2 than F in the lattice (different ligands. Huang et al. have confirmed this situation in the
Sr sites) as x increases. Note that there is a non-monotonic increase Sr1xY0.98+xCe0.02Si4N7xCx (x = 0–1) system, in which Sr2+–
of the excitation wavelength maximum with varying x. The above N3 was successfully substituted by Y3+–C4. By this substitution,
inconsistency in the shift of excitation and emission wavelengths the emission peak of Sr1xY0.98+xCe0.02Si4N7xCx phosphors can
with x should be attributed to the increasingly anisotropic of the be widely and controllably tuned from blue light to yellow light
coordination around both Sr sites, as shown in Fig. 17a. In with the increase of Y3+–C4 content; blue for x = 0–0.07, blue-
addition, the modified Sr2.975Ce0.025Al1xSixO4+xF1x phosphors green for x = 0.1, green for x = 0.2–0.6, yellow-green for x = 0.7 and
had a high quantum efficiency of 85% at room temperature, close yellow for x = 0.8 and 1.0 samples, as shown in Fig. 18a.51
to that of commercial YAG:Ce3+, which may be due to the optimal However, the overall trend of increasing thermal quenching

Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2015
View Article Online

Chem Soc Rev Review Article

behaviour with increasing x is surprising, as the progressive shape, quantum efficiency, thermal quenching behaviour and
replacement of Sr2+ and N3 by Y3+ and C4 was expected to so on through controlling the local structure around activator
introduce more covalency into the materials and hence ions. Therefore, modifying the current Pc-WLED phosphors or
increase lattice rigidity and the quenching barrier height. They developing newly efficient phosphors for application in WLEDs
proposed a dominant neighboring-cation effect to explain this by this strategy is necessary. Furthermore, some newly C–A
color-tunable emission and the thermal quenching luminescence. substitution types such as (5) and (6) in Fig. 15 (right) can be
The local structures around Ce3+ activator ions when occupying predicted to realize spectral tuning based on a similar principle
Y3+ sites in the x = 0 and x = 1 materials are shown in Fig. 18b and to that in (1)–(4) types. Schnick and coworkers have developed a
c, respectively. Briefly, the Sr2+/Y3+ substitution in the second series of oxonitridophosphate (MgSrP3N5O2, CaMg2P6O3N10,
Published on 30 September 2015. Downloaded by University of Tasmania on 30/09/2015 10:15:45.

coordination sphere has a strong effect on the coordination of SrP3N5O, Ba3P6O6N8, (Ca, Sr, Ba)P2N4, etc.) recently, which offer
Ce3+. The degree of covalency in the Ce–N bonding is controlled more phosphor host materials and also offer a possibility to
by the bonding of the nitride with other neighbors. Sr2+ ions form tune their luminescence properties by the Si4+ + O2 2 P5+ +
more strong ionic bonds with N3, and so the Ce–N covalency is N3 2 Al3+ + F and P5+ + N3 2 Si4+ + F substitutions.142–148
enhanced, whereas Y–N bonds are more covalent, and so the Thus, more highly efficient phosphors might be developed for
Ce–N covalency is reduced. Thus the Ce–N bonds become less WLED applications. Certainly, the above-mentioned several
covalent as the second-neighbor Sr2+ ions around Ce3+ are gradu- types in Section 3.4 do not contain all C–A substitution types
ally replaced by Y3+, namely, the nephelauxetic effect continuously due to the limitation of classification conditions. For example,
decreases. This results in a progressive increase of the thermal in the (KSrPO4)1x(Ba,Sr)2SiO4)x system, it exhibits tunable
quenching luminescence of phosphors. While the redshift of emission from blue to green for (KSrPO4)1x(Ba2SiO4)x (Fig. 19a)
emission can be explained, the neighboring-cation effect of and from blue to yellow for (KSrPO4)1x(Sr2SiO4)x (Fig. 19b) by
replacing Sr2+ by Y3+ is suppressed by the overall effect of the substituting of K+–Sr2+ by 2Ba2+/2Sr2+ and of PO43 groups by
lattice contraction (a larger crystal field splitting). SiO44 groups simultaneously.149 Both emission peaks of Eu2+
According to the previous discussion, it obviously indicates have a redshift as x increases due to the increasing crystal field
that the C–A substitutions can efficiently tune the optical pro- effect and an anomalous occupation at some cation sites with the
perties of pc-WLED phosphors including the spectral position, bigger nephelauxetic effect.

3.5 Crystal-site engineering approach


It is known that the activator ions randomly occupy the cation
sites in the phosphor host according to the balance of the ionic
radius and valency. Nevertheless, there is still some strategies
to change the sites of activator ions in the phosphor matrix
when multi-cation sites are available, including the control of
the doping level of activator ions, cationic substitution and so
on, which is called the crystal-site engineering approach.86,87
By these crystal-site engineering approaches, one can expect
spectral tuning of phosphors owning to different coordination
environments and energy transfer of activator ions between
different crystal sites.150–153 According to the cationic substitution
approach, Daicho et al. reported a yellow-emitting (Ca1xySrxEuy)7-
(SiO3)6Cl2 (Cl_MS:Eu2+) with high luminescence efficiency.150 The
as-prepared Cl_MS:Eu2+ can be efficiently excited by UV light or blue
light, exhibiting a broad emission from 475 nm to 725 nm with the
maximum emission at 580 nm under the excitation of 400 nm UV
light, as shown in Fig. 20a. This broad emission band is attributed
to two different Eu2+-occupied sites, the M(2) and M(3) sites in
Fig. 20c and d, respectively. Moreover, they found that the dominant
wavelength was linearly red-shifted with the increase of the Ca/Sr
ratio (Fig. 20b), namely, the emission color of could be tuned by
varying the Ca/Sr atomic ratio. The reason can be explained as
follows: the increase of the Ca/Sr ratio drives more Eu2+ ions into the
M(3) sites; then, anisotropic lattice contraction occurred along the
a axis and this contraction induced substantial deformation of
Fig. 18 (a) Excitation and emission spectra of Sr1xY0.98+xCe0.02Si4N7xCx
the M(3) site; finally, such deformation strengthened the crystal-
samples. The representation of the local first- and second-coordination
around Ce3+ activator ions occupying Y3+ sites in Sr1xY0.98+xCe0.02-
field splitting, and the 580 nm luminescence band was red-
Si4N7xCx phosphors for (c) x = 0 and (d) x = 1. (Reproduced with shifted.150 By the doping level control, Liang’s group has successfully
permission from ref. 51, copyright 2013, Wiley.) tuned the photoluminescence properties of many phosphor systems

This journal is © The Royal Society of Chemistry 2015 Chem. Soc. Rev.
View Article Online

Review Article Chem Soc Rev

to Ce(2)3+. This energy transfer will quench the emission of Ce(1)3+


and strengthen the emission of Ce(2)3+. Certainly, different nephel-
auxetic effects under different sites also generate an obvious
influence on the resulting luminescence of Ce3+ and Eu2+ ions.
Through the doping level control, Sato et al. reported a novel deep-
red light (B650 nm) emission, when doped a large amount of Eu2+
ions in Ca2xEuxSiO4 phosphors, as shown in Fig. 20e. This should
be attributed to the substitution of Eu2+ ions in the small Ca sites
that possessed strong interactions between Eu2+ and O2.86 Chen’s
Published on 30 September 2015. Downloaded by University of Tasmania on 30/09/2015 10:15:45.

group found that this approach is available in sulfide-based


phosphor systems.41–43 For example, the b-YFS lattice consists of
two distinct coordination polyhedra. The Y(1) site in the YS6
polyhedron has a stronger bond covalency than the Y(2) site in
the YS2F6 polyhedron. This suggests that the green emission at
around 495 nm originates from the substitution of Ce3+ ions for
Y3+ ions in the Y(2) site whereas the yellow component at around
547 nm is the emission of Ce3+ in the Y(1) site. By increasing the
Ce3+ doping concentration from 1% to 5% in the b-YFS:Ce3+
system, the emission colors of phosphors were adjusted from
green light (495 nm) to yellow light (540 nm).41 This should be
attributed to energy transfers between two different Ce3+ emission
centers with increasing Ce3+ concentration. A simple model to
illustrate the spectral tuning mechanism of Ce3+ or Eu2+ ion doped
phosphors by the crystal-site engineering approach is shown in
Fig. 21. Commonly, when doping Ce3+ or Eu2+ ions into the host
that have several cation sites, it preferentially enter into a part of
cation sites at a low doping concentration, while other sites are
Fig. 19 Photoluminescence emission spectra of Eu2+-activated (a)
(KSrPO4)1x(Ba2SiO4)x and (b) (KSrPO4)1x(Sr2SiO4)x at various x values
subsequently seized with increasing the doping content. Because of
under 380 nm excitation. The insets are photos showing colors of different coordination environments in different crystal sites that
the corresponding samples, which were taken with a 380 nm emitting result in a difference in CFS and the nephelauxetic effect, the
UV-LED. (Reproduced with permission from ref. 149, copyright 2013, Wiley.) corresponding shapes and positions of emission spectra are
different. Moreover, the energy transfer easily occur between
different luminescent centers with gradually enhancing the
such as Ca3La3(1x)Ce3x(BO3)5, La2CaB10O19:Ce3+, Sr3yCeyAlO4F and concentration of activator ions in this host, which also leads
so on. For example, in the Ca3La3(1x)Ce3x(BO3)5 system, Liu et al. to spectral shift of phosphors. It is noted that although the
reported spectral tuning from UV light to blue-green light.151 Fig. 20f crystal-site engineering approach is an efficient route to tune
show the normalized emission spectra upon excitation of the Ce(1)3+ the spectra of rare earth ion activated pc-WLED phosphors,
site at 311 nm in the wavelength range of 340–550 nm. It is shown the tunable range for spectra strongly depends on different
that the emission from Ce(2)3+ gradually increases with the cation-coordination environments that contain a sophisticated
increase of the x value, whereas the emission from Ce(1)3+ complex of crystal field splitting, the nephelauxetic effect and
decreases regularly. Meanwhile, emission intensity from so on. Sometimes, it is difficult to realize a continuous and
Ce(2)3+ always keeps increasing when excitation is on the controllable luminescence adjustment for Ce3+ or Eu2+-acti-
Ce(2)3+ site at 368 nm (Fig. 20g). This might be attributed to vated phosphors. Moreover, this spectral tuning approach
the different occupancy probability between Ce(1)3+ and Ce(2)3+ commonly needs a high level of cationic substitution or a
and the energy transfer from Ce(1)3+ to Ce(2)3+. The Ce3+ ions doping level, which easily sacrifices the emission intensity
occupy two sites in the above systems, Ce(1)3+ for the 10-fold and thermal stability of phosphors. That is because that an
coordination La3+ site and Ce(2)3+ for the 8-fold coordination excessive doping level may result in the concentration quenching
Ca2+ site, and the occupations of both the Ce(1)3+ site and the effect and self-absorption.85 Accordingly, the reasonable doping
Ce(2)3+ site increase with the total doping concentration. Before level should be employed to give simultaneous consideration of
the occurrence of concentration-quenching, both the emission luminescence efficiency, thermal stability and emission position
of the Ce(1)3+ site and the Ce(2)3+ site will regularly increase. desired by WLEDs’ requirements.
The lowest 5d absorption band for the Ce(2)3+ site (at about
370 nm) has obvious overlap with the emission bands from the 3.6 Mixing of the nanophase
Ce(1)3+ site, which results in the Ce(2)3+ ions absorbing the Commonly, phosphors used in WLEDs should retain a single
emission from the Ce(1)3+ site efficiently. As a consequence, an phase in order to achieve excellent luminescence efficiency
effective resonance-type energy transfer may occur from Ce(1)3+ and thermal stability. This single phase is a macroscopical

Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2015
View Article Online

Chem Soc Rev Review Article


Published on 30 September 2015. Downloaded by University of Tasmania on 30/09/2015 10:15:45.

Fig. 20 (a) Emission (dotted curves) and excitation spectra (solid curves) of (Ca0.37Sr0.53Eu0.10)7(SiO3)6Cl2 (Cl_MS:Eu2+) phosphors. The inset shows its
IQE and absorption (Ab) values. (b) Dominant wavelengths of luminescence as a function of the Ca fraction in Cl_MS:Eu2+. (c) and (d) are the coordination
geometries of the M(2) and M(3) sites. Pink and orange spheres represent oxide and chloride ions, respectively. (e) Emission spectra of Ca2xEuxSiO4 with
x = 0.2, 0.4, 0.6, and 0.8. Inset: Photographs of these samples under 365 nm UV light. The emission spectra of Ca3La3(1x)Ce3x(BO3)5 with different x
values at RT: (f) lex = 311 nm and (g) lex = 368 nm. (Reproduced with the permission from ref. 86, 150 and 151, copyright 2012, 2014, American Chemical
Society, Nature Publishing Group, Wiley.)

statistical analysis of phosphor compounds. Therefore, this


result can’t guarantee a single phase or a single crystal in all
regions of phosphor compounds. Actually, phosphor powders
commonly consist of polycrystalline particles, which may result
in some mixing of the nanophase in the local region, which
means mixing of different crystal phases in the nano-domain
region. While these mixings of the nanophase in the local
crystal structure of phosphors offer a possibility for spectral
tuning.48,51 For example, Li et al. reported tunable emission
in (Sr0.98xBaxEu0.02)Si2O2N2 systems by the mixing of nano-
phases.21 Fig. 22a and b show the typical PL emission spectra
of (Sr0.98xBaxEu0.02)Si2O2N2 phosphors at x = 0.65 to 0.77 and
x = 0.78 to 0.98, respectively. Obviously, there are two emission
bands with centers at 494 nm and 565 nm in (Sr0.98xBaxEu0.02)-
Si2O2N2 (x = 0.78 to 0.98) series, which respectively originate
from the Ba0.98Eu0.02Si2O2N2 phase and the (Sr0.63Ba0.35Eu0.02)-
Si2O2N2 phase. Similarly, the two emission bands with centers
at 494 nm and 565 nm in the (Sr0.98xBaxEu0.02)Si2O2N2 (x = 0.65 to
0.77) series are attributed to the (Sr0.30Ba0.68Eu0.02)Si2O2N2
phase and the (Sr0.63Ba0.35Eu0.02)Si2O2N2 phase, respectively.
Moreover, the relative emission intensity of two emission bands
Fig. 21 A schematic illustration of the red shift and blue shift mechanisms
in the above two series can be adjusted by the Sr/Ba ratio,
induced by occupying different crystal sites. Green and gray spheres namely, the relative content of two phases in the local region,
represent host cations and anion ligands, respectively. H1n are cation which generates a single-host white light emission. This result
sites with different sizes. The size is H1 4 H2 4 Hn. The crystal field splitting can be further confirmed by HRTEM results, as shown in
strength of activator ions in H1n sites are H1 o H2 o Hn. A spectral red
Fig. 22c, which shows a domain structure of (Sr0.30Ba0.68Eu0.02)-
shift will appear when activator ions entering into cation sites from H1 to
Hn. On the contrary, a spectral blue shift will appear. The energy transfers
Si2O2N2. The d-spacings of two domains are marked. For the
of activator ions among the H1n sites also contribute to a spectral top domain, A = 2.91 Å, B = 5.14 Å, and the interplanar angle is
red shift. 76.51. For the bottom domain, C = 2.94 Å, D = 3.53 Å, and the

This journal is © The Royal Society of Chemistry 2015 Chem. Soc. Rev.
View Article Online

Review Article Chem Soc Rev

effort has been made to improve and develop highly efficient


pc-WLED phosphors. The 5d - 4f transitions of rare-earth ions
(i.e., Eu2+ and Ce3+) depends greatly on the surrounding environ-
ments including symmetry, coordination, covalence, bond length,
site size, and crystal field strength in which they reside, making it
possible to tune the excitation and emission wavelength over a
wide range by varying the composition and local crystal structure.
In this review, we have given several approaches to realize the
spectral tuning of pc-WLED phosphors based on local structural
Published on 30 September 2015. Downloaded by University of Tasmania on 30/09/2015 10:15:45.

variation of phosphor host lattices, including doping level control,


cationic substitution, anionic substitution, cationic–anionic sub-
stitution, the crystal-site engineering approach, and mixing of
nanophases. The spectral tuning based on the above approaches
can improve the luminescence performance of pc-WLED phos-
phors such as improving the luminous efficiency, optimizing the
color rendering properties, achieving the desired color tempera-
ture, enhancing the color gamut and so on. Moreover, by the
variation of local crystal structure, one can also expect to improve
the thermal stability of phosphors due to the optimization of
lattice rigidity. Although spectral tuning originating from local
Fig. 22 Typical PL emission spectra of (Sr0.98xBaxEu0.02)Si2O2N2 samples:
(a) x = 0.65 to 0.77 and (b) x = 0.78 to 0.98. The relative emission intensities structural variation of phosphor hosts can effectively improve the
at 470, 494, and 565 nm as functions of x are shown as insets in (c) and (d). luminescence performances of pc-WLED phosphors, the under-
Black dotted lines in parts (c) and (d) show Gaussian fitting curves of x = 0.73 lying luminescence mechanisms in many systems are unclear due
and x = 0.93 samples, respectively. (c) Typical HRTEM image of to the complex effects of crystal field splitting, the nephelauxetic
(Sr0.30Ba0.68Eu0.02)Si2O2N2 showing a domain structure. (d) HRTEM image
effect, the Stokes shift, the distortion of the host lattice and so on.
of the Sr1xY0.98+xCe0.02Si4N7xCx sample, showing multidomain nanostruc-
tures in the two-phase x = 0.4 sample. (Reproduced with permission from Therefore, there is no a general law to be suitable for all phosphor
ref. 48 and 51, copyright 2013, American Chemical Society, Wiley.) systems at current, and more concerns should be focused on
exploring the structured-related luminescence mechanisms.
On the other hand, some non-rare-earth-based luminescent
interplanar angle is 110.01. The A and B originate from the materials like semiconducting quantum dots,154–159 Mn4+-
Ba0.98Eu0.02Si2O2N2 phase, while C and D are identified as the activated red-emitting materials (aluminates, silicates, fluoro-
(Sr0.63Ba0.35Eu0.02)Si2O2N2 phase. Similar spectral tuning was silicate, fluorotitanate)160–168 and defect-related luminescent
reported by mixing of nanophases in the local structure in materials show excellent luminescence properties under n-UV
Sr1xY0.98+xCe0.02Si4N7xCx (Fig. 18a and 22d) and (Na1xCax)- and blue light excitation. For example, defect-related lumines-
(Sc1xMgx)Si2O6:Eu2+ systems (Fig. 8).51,105 In general, the mix- cent materials can also emit white light by controlling the
ing of nanophase in LED phosphor compounds can be used to concentration of the defect and reaction conditions.169 This
tune spectra. However, this approach is closely related to the kind of luminescent material has been greatly accelerated to
synthesis conditions such as calcination temperature, calcina- develop due to the advantages of non-toxic, environment
tion time, synthesis procedure, the use of flux, which make it friendly and cheap. Newly modified Mn4+-activated red-emitting
has bad controllability. Moreover, it also easily suffers from materials such as (Na,K)2(Si,Ge,Ti)F6:Mn4+ and Ba(Si,Ge)F6:Mn4+
reduced thermal stability and efficiency. In terms of real prepared by modified soft chemical methods show high lumines-
applicability (stable emission color, temperature stable, high cence efficiency and low thermal quenching under blue light
quantum efficiency,. . .) of phosphors in WLEDs, it needs to be excitation, which are efficient red-emitting phosphors for the
further investigated and optimized. WLED backlight source.79 Therefore, these non-rare-earth-based
phosphors should also be actively explored to open the door to
4. Summary and outlook novel WLED lighting and display applications. Finally, the search
for new host materials for the efficient doping of rare-earth
Currently, white LED technology is becoming more and more ions (i.e., Eu2+, Ce3+) is also a huge challenge. Some methods
mature with the development of modern science and technology. for exploration and discovery of novel pc-WLED phosphors,
Except for the wide use in various outdoor lighting and display especially, (oxy)nitride phosphors, including the single crystal
backlight areas, pc-WLEDs have gradually entered into home growth method (by Schinick’s group),36,170–173 the single-particle-
lighting, which will bring enormous energy saving and environ- diagnosis approach (by Xie and Hirosaki’s group),174,175 the
mental protection. In order to meet the needs of different lighting heuristics optimization based solid state combinatorial chemistry
and display, and further enhance the luminescence performances method (by Sohn’s group)176,177 and so on should be concerned
of pc-WLED devices, namely, avoiding phosphor conversion to be and encouraged, which might readily be scaled up for industrial
the performance-limiting factor in pc-WLEDs devices, significant WLED applications. In summary, photoluminescence tuning by

Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2015
View Article Online

Chem Soc Rev Review Article

compositional optimization based on the variation of the local 17 C. H. Huang, Y. C. Chiu, Y. T. Yeh, T. S. Chan and T. M.
crystal structure can be realized, which can improve the lumines- Chen, ACS Appl. Mater. Interfaces, 2012, 4, 6661–6667.
cence performance of rare earth ion activated phosphors in 18 C. H. Huang, Y. C. Chiu, Y. T. Yeh and T. M. Chen, Mater.
pc-WLED devices. However, a large number of novel pc-WLED Express, 2012, 2, 303–310.
phosphors and emerging methods for host materials are still in 19 C. H. Huang, Y. C. Chiu, Y. T. Yeh and T. M. Chen, Mater.
the research stage, and need to be further optimized and Express, 2012, 2, 303–310.
improved for practical WLED applications. Therefore, there is 20 C. H. Huang, D. Y. Wang, Y. C. Chiu, Y. T. Yeh and
a long way in achieving the goal, and much effort is needed by T. M. Chen, RSC Adv., 2012, 2, 9130–9134.
scientists and industrialists. 21 C. H. Huang and T. M. Chen, Inorg. Chem., 2011, 50,
Published on 30 September 2015. Downloaded by University of Tasmania on 30/09/2015 10:15:45.

5725–5730.
22 D. Y. Wang, Y. C. Wu and T. M. Chen, J. Mater. Chem.,
Acknowledgements 2011, 21, 18261–18265.
23 D. Y. Wang, C. H. Huang, Y. C. Wu and T. M. Chen,
This project is financially supported by National Basic Research
J. Mater. Chem., 2011, 21, 10818–10822.
Program of China (Grants No. 2010CB327704), the National
24 C. H. Huang, Y. C. Chen, T. M. Chen, T. S. Chan and
Natural Science Foundation of China (Grants No. NSFC
H. S. Sheu, J. Mater. Chem., 2011, 21, 5645–5649.
21301162, 60977013, 91433110, U1301242, 21221061) and the
25 C. H. Huang, T. S. Chan, W. R. Liu, D. Y. Wang, Y. C. Chiu,
Fundamental Research Founds for National University, China
Y. T. Yeh and T. M. Chen, J. Mater. Chem., 2012, 22,
University of Geosciences (Wuhan) (Grant No. CUG130402,
20210–20216.
CUG130614, CUG130624, CUG120864, GBL31510).
26 C. H. Huang and T. M. Chen, J. Phys. Chem. C, 2011, 115,
2349–2355.
Notes and references 27 C. H. Huang, P. J. Wu, J. F. Lee and T. M. Chen, J. Mater.
Chem., 2011, 21, 10489–10495.
1 N. C. George, K. A. Denault and R. Seshadri, Annu. Rev. 28 C. H. Huang, W. R. Liu and T. M. Chen, J. Phys. Chem. C,
Mater. Res., 2013, 43, 481–501. 2010, 114, 18698–18701.
2 G. G. Li and J. Lin, Chem. Soc. Rev., 2014, 43, 7099–7131. 29 C. H. Huang, T. M. Chen, W. R. Liu, Y. C. Chiu, Y. T. Yeh
3 C. C. Lin and R. S. Liu, J. Phys. Chem. Lett., 2011, 2, and S. M. Jang, ACS Appl. Mater. Interfaces, 2010, 2,
1268–1277. 259–264.
4 C. Feldmann, T. Jüstel, C. R. Ronda and P. J. Schmidt, 30 C. C. Lin, Y. P. Liu, Z. R. Xiao, Y. K. Wang, B. M. Cheng and
Adv. Funct. Mater., 2003, 13, 511–516. R. S. Liu, ACS Appl. Mater. Interfaces, 2014, 6, 9160–9172.
5 K. A. Denault, J. Brgoch, S. D. Kloss, M. W. Gaultois, 31 G. J. Gao, J. X. Wei, Y. Shen, M. Y. Peng and L. Wondraczek,
J. Siewenie, K. Page and R. Seshadri, ACS Appl. Mater. J. Mater. Chem. C, 2014, 2, 8678–8682.
Interfaces, 2015, 7, 7264–7272. 32 G. J. Gao and L. Wondraczek, J. Mater. Chem. C, 2014, 2,
6 C. Wang, Z. Y. Zhao, Q. S. Wu, S. Y. Xin and Y. H. Wang, 691–695.
CrystEngComm, 2014, 16, 9651–9656. 33 T. Jüstel, H. Nikol and C. Ronda, Angew. Chem., Int. Ed.,
7 Y. H. Song, T. Y. Choi, K. Senthil, T. Masaki and D. H. 1998, 37, 3084–3103.
Yoon, Mater. Lett., 2011, 65, 3399–3401. 34 R. J. Xie and N. Hirosaki, Sci. Technol. Adv. Mater., 2007, 8,
8 W. Y. Li, R. J. Xie, T. L. Zhou, L. H. Liu and Y. J. Zhu, Dalton 588–600.
Trans., 2014, 43, 6132–6138. 35 H. A. Höppe, Angew. Chem., Int. Ed., 2009, 48, 3572–3582.
9 W. B. Park, S. P. Singh, C. Yoon and K. S. Sohn, J. Mater. 36 M. Zeuner, S. Pagano and W. Schnick, Angew. Chem., Int.
Chem., 2012, 22, 14068–14075. Ed., 2011, 50, 7754–7775.
10 M. Zeuner, P. J. Schmidt and W. Schnick, Chem. Mater., 37 M. M. Shang and J. Lin, Chem. Soc. Rev., 2014, 43, 1372–1386.
2009, 21, 2467–2473. 38 P. F. Smet, A. B. Parmentier and D. Poelman, J. Electrochem.
11 R. J. Xie, N. Hirosaki, T. Suehiro, F. F. Xu and M. Mitomo, Soc., 2011, 158, R37–R54.
Chem. Mater., 2006, 18, 5578–5583. 39 X. H. He, N. Lian, J. H. Sun and M. Y. Guan, J. Mater. Sci.,
12 H. L. Li, R. J. Xie, N. Hirosaki, T. Takeda and G. H. Zhou, 2009, 44, 4763–4775.
Int. J. Appl. Ceram. Technol., 2009, 6, 459–464. 40 A. Mayr and V. Kahlenberg, J. Am. Ceram. Soc., 2013, 96,
13 Y. Q. Li, G. de With and H. T. Hintzen, J. Lumin., 2006, 116, 318–322.
107–116. 41 Y. C. Wu, D. Y. Wang, T. M. Chen, C. S. Lee, K. J. Chen and
14 Y. Q. Li, J. E. J. van Steen, J. W. H. van Krevel, G. Botty, H. C. Kuo, ACS Appl. Mater. Interfaces, 2011, 3, 3195–3199.
A. C. A. Delsing, F. J. DiSalvo, G. de With and H. T. 42 Y. C. Wu, Y. C. Chen, D. Y. Wang, C. S. Lee, C. C. Sun and
Hintzen, J. Alloys Compd., 2006, 417, 273–279. T. M. Chen, J. Mater. Chem. C, 2011, 21, 15163–15166.
15 X. Q. Piao, T. Horikawa, H. Hanzawa and K. Machida, Appl. 43 Y. C. Wu, Y. C. Chen, T. M. Chen, C. S. Lee, K. J. Chen and
Phys. Lett., 2006, 88, 161908. H. C. Kuo, J. Mater. Chem., 2012, 22, 8048–8056.
16 S. Ray, Y. C. Fang and T. M. Chen, RSC Adv., 2013, 3, 44 S. N. Chen and T. M. Chen, J. Chin. Chem. Soc., 2013, 60,
16387–16391. 961–964.

This journal is © The Royal Society of Chemistry 2015 Chem. Soc. Rev.
View Article Online

Review Article Chem Soc Rev

45 S. P. Lee, C. H. Huang and T. M. Chen, J. Mater. Chem. C, 74 G. Blasse and B. C. Grabmaier, Luminescent Materials,
2014, 2, 8925–8931. Springer, Berlin, 1994.
46 S. P. Lee, C. H. Huang, T. S. Chan and T. M. Chen, 75 J. A. Duffy and M. D. Ingram, J. Chem. Phys., 1971, 54, 443.
ACS Appl. Mater. Interfaces, 2014, 6, 7260–7267. 76 A. J. Bruce and J. A. Duffy, J. Chem. Soc., Faraday Trans.,
47 S. P. Lee, T. S. Chan and T. M. Chen, ACS Appl. Mater. 1982, I78, 907–914.
Interfaces, 2015, 7, 40–44. 77 C. H. Huang, T. W. Kuo and T. M. Chen, Opt. Lett., 2011,
48 G. G. Li, C. C. Lin, W. T. Chen, M. S. Molokeev, V. V. 19, A1–A6.
Atuchin, C. Y. Chiang, W. Z. Zhou, C. W. Wang, W. S. Li, 78 L. Shi, Y. L. Huang and H. J. Seo, J. Phys. Chem. A, 2010,
H. S. Sheu, T. S. Chan, C. G. Ma and R. S. Liu, Chem. Mater., 114, 6927–6934.
Published on 30 September 2015. Downloaded by University of Tasmania on 30/09/2015 10:15:45.

2014, 26, 2991–3001. 79 H. M. Zhu, C. C. Lin, W. Q. Luo, S. T. Shu, Z. G. Liu,


49 W. T. Chen, H. S. Sheu, R. S. Liu and J. P. Attfield, J. Am. Y. S. Liu, J. T. Kong, E. Ma, Y. G. Cao, R. S. Liu and
Chem. Soc., 2012, 134, 8022–8025. X. Y. Chen, Nat. Commun., 2014, 5, 4312.
50 S. S. Wang, W. T. Che, Y. Li, J. Wang, H. S. Sheu and 80 Y. Q. Li, A. C. A. Delsing, G. de With and H. T. Hintzen,
R. S. Liu, J. Am. Chem. Soc., 2013, 135, 12504–12507. Chem. Mater., 2005, 17, 3242–3248.
51 W. Y. Huang, F. Yoshimura, K. Ueda, Y. Shimomura, 81 V. Bachmann, C. Ronda, O. Oeckler, W. Schnick and
H. S. Sheu, T. S. Chan, H. F. Greer, W. Z. Zhou, S. F. Hu, A. Meijerink, Chem. Mater., 2009, 21, 316–325.
R. S. Liu and J. P. Attfield, Angew. Chem., Int. Ed., 2013, 52, 82 R. J. Xie, N. Hirosaki, K. Sakuma, Y. Yamamoto and
8102–8106. M. Mitomo, Appl. Phys. Lett., 2004, 84, 5404–5406.
52 H. Y. Huang, F. Yoshimura, K. Ueda, Y. Shimomura, 83 V. Bachmann, C. Ronda and A. Meijerink, Chem. Mater.,
H. S. Sheu, T. S. Chan, C. Y. Chiang, W. Z. Zhou and 2009, 21, 2077–2084.
R. S. Liu, Chem. Mater., 2014, 26, 2075–2085. 84 Y. Q. Li, N. Hirosaki, R. J. Xie, T. Takeda and M. Mitomo,
53 N. C. George, A. Birkel, J. Brgoch, B. C. Hong, A. A. Chem. Mater., 2008, 20, 6704–6714.
Mikhailovsky, K. Page, A. Llobet and R. Seshadri, Inorg. 85 W. B. Im, S. Brinkley, J. Hu, A. Mikhailovsky, S. P. DenBaars
Chem., 2013, 52, 13730–13741. and R. Seshadri, Chem. Mater., 2010, 22, 2842–2849.
54 K. A. Denault, J. Brgoch, M. W. Gaultois, A. Mikhailovsky, 86 Y. Sato, H. Kato, M. Kobayashi, T. Masaki, D. H. Yoon and
R. Petry, H. Winkler, S. P. DenBaars and R. Seshadri, Chem. M. Kakihana, Angew. Chem., Int. Ed., 2014, 53, 7756–7759.
Mater., 2014, 26, 2275–2282. 87 W. P. Chen, H. B. Liang, H. Y. Ni, P. He and Q. Su,
55 A. Kalaji, M. Mikami and A. K. Cheetham, Chem. Mater., J. Electrochem. Soc., 2010, 157, J159–J163.
2014, 26, 3966–3975. 88 S. J. Lee, J. K. Park, K. N. Kim and C. H. Kim, Electrochem.
56 C. M. Zhang, S. S. Huang, D. M. Yang, X. J. Kang, Solid-State Lett., 2007, 10, J45–J48.
M. M. Shang, C. Peng and J. Lin, J. Mater. Chem., 2010, 89 V. Bachmann, C. Ronda, O. Oeckler, W. Schnick and
20, 6674–6680. A. Meijerink, Chem. Mater., 2009, 21, 316–325.
57 P. Dorenbos, J. Phys.: Condens. Matter, 2003, 15, 4797–4807. 90 Z. G. Xia, J. Y. Sun, H. Y. Du and W. Zhou, Opt. Mater.,
58 P. Dorenbos, Phys. Rev. B: Condens. Matter Mater. Phys., 2006, 28, 524–529.
2000, 62, 15640–15649. 91 K. A. Denault, Z. Y. Cheng, J. Brgoch, S. P. DenBaars and
59 P. Dorenbos, ECS Solid State Lett., 2013, 2, R3001–R3011. R. Seshadri, J. Mater. Chem. C, 2013, 1, 7339–7345.
60 P. Dorenbos, J. Lumin., 2000, 91, 91–106. 92 J. Brgoch, C. K. H. Borg, K. A. Denault, S. P. DenBaar and
61 P. Dorenbos, J. Phys.: Condens. Matter, 2003, 15, 575. R. Seshadri, Solid State Sci., 2013, 18, 149–154.
62 P. Dorenbos, J. Lumin., 2000, 91, 155–176. 93 X. M. Teng, Y. H. Liu, Y. Z. Liu, Y. S. Hu, H. Q. He and
63 P. Dorenbos, J. Lumin., 2003, 104, 239–260. W. D. Zhuang, J. Rare Earths, 2009, 27, 58–61.
64 P. Dorenbos, Phys. Rev. B: Condens. Matter Mater. Phys., 94 O. M. ten Kate, Z. Zhang, P. Dorenbos, H. T. Hintzen and
2000, 62, 15640–15649. E. van der Kolk, J. Solid State Chem., 2013, 197, 209–217.
65 P. Dorenbos, Phys. Rev. B: Condens. Matter Mater. Phys., 95 T. Suehiro, R. J. Xie and N. Hirosaki, Ind. Eng. Chem. Res.,
2000, 62, 15650–15659. 2013, 52, 7453–7456.
66 P. Dorenbos, Phys. Rev. B: Condens. Matter Mater. Phys., 96 Y. Q. Li, G. De With and H. T. Hintzen, J. Solid State Chem.,
2001, 64, 125117. 2008, 181, 515–524.
67 P. Dorenbos, J. Lumin., 2002, 99, 283–299. 97 A. A. Setlur, W. J. Heward, M. E. Hannah and U. Happek,
68 P. Dorenbos, J. Alloys Compd., 2002, 341, 156–159. Chem. Mater., 2008, 20, 6277–6283.
69 P. Dorenbos, J. Lumin., 2007, 122–123, 315–317. 98 M. Seibald, O. Oeckler, V. R. Celinski, P. J. Schmidt,
70 C. K. Jørgensen, Modern Aspects of Ligand Field Theory, A. Tücks and W. Schnick, Solid State Sci., 2011, 13,
North-Holland, Amsterdam, 1971. 1769–1778.
71 P. Dorenbos, J. Andriessen and C. W. E. van Eijk, J. Solid 99 M. Seibald, T. Rosenthal, O. Oeckler, F. Fahrnbauer,
State Chem., 2003, 171, 133–136. A. Tücks, P. J. Schmidt and W. Schnick, Chem. – Eur. J.,
72 P. Dorenbos, J. Lumin., 2013, 135, 93–104. 2012, 18, 13446–13452.
73 E. Antic-Fidancev, M. Lemaitre-Blaise and P. Caro, New 100 M. Seibald, T. Rosenthal, O. Oeckler and W. Schnick, Crit.
J. Chem., 1987, 11, 467–472. Rev. Solid State Mater. Sci., 2014, 39, 215–229.

Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2015
View Article Online

Chem Soc Rev Review Article

101 H. S. Jang, Y. H. Won, S. Vaidyanathan, D. H. Kim and 128 L. H. Liu, R. J. Xie, N. Hirosaki, T. Takeda, J. G. Li and
D. Y. Jeon, J. Electrochem. Soc., 2009, 156, J138–J142. X. D. Sun, J. Am. Ceram. Soc., 2009, 92, 2668–2673.
102 J. K. Park, K. J. Choi, J. H. Yeon, S. J. Lee and C. H. Kim, 129 K. Takahashi, N. Hirosaki, R. J. Xie, M. Harada, K. I. Yoshimura
Appl. Phys. Lett., 2006, 88, 043511. and Y. Tomomura, Appl. Phys. Lett., 2007, 91, 091923.
103 J. Brgoch, S. D. Kloß, K. A. Denault and R. Seshadri, 130 R. J. Xie, N. Hirosaki, K. Sakuma, Y. Yamamoto and
Z. Anorg. Allg. Chem., 2014, 640, 1182–1189. M. Mitomo, Appl. Phys. Lett., 2004, 84, 5404–5406.
104 C. H. Huang, P. J. Wu, J. F. Lee and T. M. Chen, J. Mater. 131 R. J. Xie, N. Hirosaki, M. Mitomo, K. Takahashi and
Chem., 2011, 21, 10489–10495. K. Sakuma, Appl. Phys. Lett., 2006, 88, 101104.
105 Z. G. Xia, Y. Y. Zhang, M. S. Molokeev, V. V. Atuchin and 132 K. Sakuma, N. Hirosaki, R. J. Xie, Y. Yamamoto and
Published on 30 September 2015. Downloaded by University of Tasmania on 30/09/2015 10:15:45.

Y. Luo, Sci. Rep., 2013, 3, 3310. T. Suehiro, Mater. Lett., 2007, 61, 547–550.
106 J. K. Li, J. G. Li, S. H. Liu, X. D. Li, X. D. Sun and Y. Sakka, 133 R. J. Xie, N. Hirosaki, M. Mitomo, Y. Yamamoto and
Sci. Technol. Adv. Mater., 2013, 14, 054201. T. Suehiro, J. Phys. Chem. B, 2004, 108, 12027–12031.
107 J. L. Wu, G. Gundiah and A. Cheetham, Chem. Phys. Lett., 134 K. Sakumaa, N. Hirosaki and R. J. Xie, J. Lumin., 2007, 126,
2007, 441, 250–254. 843–852.
108 J. Ueda, K. Aishima and S. Tanabe, Opt. Mater., 2013, 35, 135 R. J. Xie, N. Hirosaki, M. Mitomo, K. Sakuma and
1952–1957. N. Kimura, Appl. Phys. Lett., 2006, 89, 241103.
109 F. F. Zhang, K. X. Song, J. Jiang, S. Wu, P. Zheng, Q. M. 136 W. B. Im, N. George, J. Kurzman, S. Brinkley, A. Mikhailovsky,
Huang, J. M. Xu and H. B. Qin, J. Alloys Compd., 2014, 615, J. Hu, B. F. Chmelka, S. P. DenBaars and R. Seshadri,
588–593. Adv. Mater., 2011, 23, 2300–2305.
110 Y. W. Jung, B. Lee, S. P. Singh and K. S. Sohn, Opt. Express, 137 W. B. Im, Y. Fourré, S. Brinkley, J. Sonoda, S. Nakamura,
2010, 18, 17805–17818. S. P. DenBaars and R. Seshadri, Opt. Express, 2009, 17,
111 M. M. Shang, J. Fan, H. Z. Lian, Y. Zhang, D. L. Geng and 22673–22679.
J. Lin, Inorg. Chem., 2014, 53, 7748–7755. 138 K. A. Denault, N. C. George, S. R. Paden, S. Brinkley,
112 W. B. Im, N. N. Fellows, S. P. DenBaars and R. Seshadri, A. A. Mikhailovsky, J. Neuefeind, S. P. DenBaars and
J. Mater. Chem., 2009, 19, 1325–1330. R. Seshadria, J. Mater. Chem., 2012, 22, 18204–18213.
113 N. Komuro, M. Mikami, Y. Shimomura, E. G. Bithell and 139 K. W. Huang, W. T. Chen, C. I. Chu, S. F. Hu, H. S. Sheu,
A. K. Cheetham, J. Mater. Chem. C, 2015, 3, 204–210. B. M. Cheng, J. M. Chen and R. S. Liu, Chem. Mater., 2012,
114 Y. H. Song, T. Y. Choi, K. Senthil, T. Masaki and D. H. 24, 2220–2227.
Yoon, Mater. Lett., 2012, 67, 184–186. 140 Z. Q. Shi, Z. L. Wei, C. He, R. F. Jing, H. J. Wang and
115 Y. X. Gu, Q. H. Zhang, Y. G. Li and H. Z. Wang, J. Alloys G. J. Qiao, RSC Adv., 2014, 4, 62926–62934.
Compd., 2011, 509, L109–L112. 141 R. J. Xie, H. Naoto, B. Dierre, T. Takeda and T. Sekiguchi,
116 Y. F. Liu, X. Zhang, Z. D. Hao, X. J. Wang and J. H. Zhang, J. Soc. Inf. Disp., 2011, 19, 627–630.
J. Mater. Chem., 2011, 21, 6354–6358. 142 F. J. Pucher and W. Schnick, Z. Anorg. Allg. Chem., 2014,
117 A. A. Setlur, W. J. Heward, M. E. Hannah and U. Happek, 640, 2708–2713.
Chem. Mater., 2008, 20, 6277–6283. 143 D. Baumann and W. Schnick, Eur. J. Inorg. Chem., 2015,
118 Z. Y. Zhao, Z. G. Yang, Y. R. Shi, C. Wang, B. T. Liu, G. Zhu 617–621.
and Y. H. Wang, J. Mater. Chem. C, 2013, 1, 1407–1412. 144 S. J. Sedlmaier, D. Weber and W. Schnick, Z. Kristallogr. -
119 Z. G. Xia, Q. Li and J. Y. Sun, Mater. Lett., 2007, 61, New Cryst. Struct., 2012, 227, 1–2.
1885–1888. 145 F. J. Pucher, A. Marchuk, P. J. Schmidt, D. Wiechert and
120 Y. Zhang, G. G. Li, D. L. Geng, M. M. Shang, C. Peng and W. Schnick, Chem. – Eur. J., 2015, 21, 6443–6449.
J. Lin, Inorg. Chem., 2012, 51, 11655–11664. 146 A. Marchuk, V. R. Celinski, J. Schmedt auf der Günne and
121 K. S. Sohn, B. Lee, R. J. Xie and N. Hirosaki, Opt. Lett., 2009, W. Schnick, Chem. – Eur. J., 2015, 21, 5836–5842.
34, 3427–3429. 147 A. Marchuk and W. Schnick, Angew. Chem., Int. Ed., 2015,
122 K. S. Sohn, S. J. Lee, R. J. Xie and N. Hirosaki, Appl. Phys. 54, 2383–2387.
Lett., 2009, 95, 121903. 148 S. J. Sedlmaier, E. Mugnaioli, O. Oeckler, U. Kolb and
123 M. Sopicka-Lizer, D. Michalik, J. Plew, T. Juestel, W. Schnick, Chem. – Eur. J., 2011, 17, 11258–11265.
H. Winkler and T. Pawlik, J. Eur. Ceram. Soc., 2012, 32, 149 J. K. Han, M. E. Hannah, A. Piquette, J. B. Talbot,
1383–1387. K. C. Mishra and J. Mckittrick, J. Am. Ceram. Soc., 2013,
124 Y. F. Wang, Q. Q. Zhu, L. Y. Hao, X. Xu, R. J. Xie and 96, 1526–1532.
S. Agathopoulos, J. Am. Ceram. Soc., 2013, 96, 2562–2569. 150 H. Daicho, T. Iwasaki, K. Enomoto, Y. Sasaki, Y. Maeno,
125 W. B. Park, S. P. Singh and K. S. Sohn, J. Am. Chem. Soc., Y. Shinomiya, S. Aoyagi, E. Nishibori, M. Sakata, H. Sawa,
2014, 136, 2363–2373. S. Matsuishi and H. Hosono, Nat. Commun., 2012, 3, 1132.
126 O. Oeckler, J. A. Kechele, H. Koss, P. J. Schmidt and 151 C. M. Liu, H. B. Liang, X. J. Kuang, J. P. Zhong, S. S. Sun
W. Schnick, Chem. – Eur. J., 2009, 15, 5311–5319. and Y. Tao, Inorg. Chem., 2012, 51, 8802–8809.
127 Y. Q. Li, N. Hirosaki, R. J. Xie and M. Mitomo, Sci. Technol. 152 L. Li, H. B. Liang, Z. F. Tian, H. H. Lin, Q. Su and
Adv. Mater., 2007, 8, 607–616. G. B. Zhang, J. Phys. Chem. C, 2008, 112, 13763–13768.

This journal is © The Royal Society of Chemistry 2015 Chem. Soc. Rev.
View Article Online

Review Article Chem Soc Rev

153 K. A. Denault, Z. Y. Cheng, J. Brgoch, S. P. DenBaars and 166 H. D. Nguyen, C. C. Lin, M. H. Fang and R. S. Liu, J. Mater.
R. Seshadri, J. Mater. Chem. C, 2013, 1, 7339–7345. Chem. C, 2014, 2, 10268–10272.
154 J. Ziegler, S. Xu, E. Kucur, F. Meister, M. Batentschuk, 167 L. L. Wei, C. C. Lin, M. H. Fang, M. G. Brik, S. F. Hu, H. Jiao
F. Gindele and T. Nann, Adv. Mater., 2008, 20, 4068–4073. and R. S. Liu, J. Mater. Chem. C, 2015, 3, 1655–1660.
155 C. Li, M. Ando, H. Enomoto and N. Murase, J. Phys. 168 Q. Zhou, Y. Zhou, Y. Liu, L. J. Luo, Z. L. Wang, J. H. Peng,
Chem. C, 2008, 112, 20190–20199. J. Yan and M. M. Wu, J. Mater. Chem. C, 2015, 3, 3055–3059.
156 J. Lee, V. C. Sundar, J. R. Heine, M. G. Bawendi and 169 C. M. Zhang and J. Lin, Chem. Soc. Rev., 2012, 41, 7938–7961.
K. F. Jensen, Adv. Mater., 2000, 12, 1102–1105. 170 P. Pust, F. Hintze, C. Hecht, V. Weiler, A. Locher,
157 E. Jang, S. Jun, H. Jang, J. Lim, B. Kim and Y. Kim, Adv. D. Zitnanska, S. Harm, D. Wiechert, P. J. Schmidt and
Published on 30 September 2015. Downloaded by University of Tasmania on 30/09/2015 10:15:45.

Mater., 2010, 22, 3076–3080. W. Schnick, Chem. Mater., 2014, 26, 6113–6119.
158 J. S. Steckel, P. Snee, S. Coe-Sullivan, J. P. Zimmer, J. E. 171 P. Pust, V. Weiler, C. Hecht, A. Tücks, A. S. Wochnik, A.-K.
Halpert, P. Anikeeva, L. A. Kim, V. Bulovic and M. G. Henß, D. Wiechert, C. Scheu, P. J. Schmidt and W. Schnick,
Bawendi, Angew. Chem., Int. Ed., 2006, 45, 5796–5799. Nat. Mater., 2014, 13, 891–896.
159 R. Xie, D. Battaglia and X. Peng, J. Am. Chem. Soc., 2007, 172 P. Pust, A. Wochnik, E. Baumann, P. J. Schmidt, D. Wiechert,
129, 15432–15433. C. Scheu and W. Schnick, Chem. Mater., 2014, 26, 3544–3549.
160 D. Sekiguchi, J. Nara and S. Adachi, J. Appl. Phys., 2013, 173 S. Schmiechen, H. Schneider, P. Wagatha, C. Hecht, P. J.
113, 183516. Schmidt and W. Schnick, Chem. Mater., 2014, 26,
161 R. Kasa and S. Adachi, J. Appl. Phys., 2012, 112, 013506. 2712–2719.
162 W. Lü, W. Z. Lv, Q. Zhao, M. M. Jiao, B. Q. Shao and 174 N. Hirosaki, T. Takeda, S. Funahashi and R. J. Xie, Chem.
H. P. You, Inorg. Chem., 2014, 53, 11985–11990. Mater., 2014, 26, 4280–4288.
163 L. F. Lv, X. Y. Jiang, S. M. Huang, X. Chen and Y. X. Pan, 175 R. J. Xie, Y. Q. Li, N. Hirosaki and H. Yamamoto, Nitride
J. Mater. Chem. C, 2014, 2, 3879–3884. Phosphors and Solid State Lighting, Taylor & Francis, 2011.
164 X. Y. Jiang, Y. X. Pan, S. M. Huang, X. Chen, J. G. Wang and 176 B. Park, S. P. Singh, C. Yoon and K. S. Sohn, J. Mater. Chem.
G. K. Liu, J. Mater. Chem. C, 2014, 2, 2301–2306. C, 2013, 1, 1832–1839.
165 J. H. Oh, H. Kang, Y. J. Eo, H. K. Park and Y. Rag Do, 177 W. B. Park, S. P. Singh, C. Yoon and K. S. Sohn, J. Mater.
J. Mater. Chem. C, 2015, 3, 607–615. Chem., 2012, 22, 14068–14075.

Chem. Soc. Rev. This journal is © The Royal Society of Chemistry 2015

You might also like