Next Article in Journal
A Voltage-Tuned Terahertz Absorber Based on MoS2/Graphene Nanoribbon Structure
Previous Article in Journal
MIL-100(Fe)-Based Composite Films for Food Packaging
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Light-Conversion Phosphors for Plant Growth and Strategies for the Modulation of Photoluminescence Properties

1
Institute for Smart Materials & Engineering, School of Materials Science and Engineering, University of Jinan, No. 336 Nanxinzhuang West Road, Jinan 250022, China
2
School of Chemistry and Chemical Engineering, University of Jinan, No. 336 Nanxinzhuang West Road, Jinan 250022, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(11), 1715; https://doi.org/10.3390/nano13111715
Submission received: 30 April 2023 / Revised: 20 May 2023 / Accepted: 22 May 2023 / Published: 23 May 2023

Abstract

:
The advent of greenhouses greatly promoted the development of modern agriculture, which freed plants from regional and seasonal constraints. In plant growth, light plays a key role in plant photosynthesis. The photosynthesis of plants can selectively absorb light, and different light wavelengths result in different plant growth reactions. Currently, light-conversion films and plant-growth LEDs have become two effective ways to improve the efficiency of plant photosynthesis, among which phosphors are the most critical materials. This review begins with a brief introduction of the effects of light on plant growth and the various techniques for promoting plant growth. Next, we review the up-to-date development of phosphors for plant growth and discussed the luminescence centers commonly used in blue, red and far-red phosphors, as well as their photophysical properties. Then, we summarize the advantages of red and blue composite phosphors and their designing strategies. Finally, we describe several strategies for regulating the spectral position of phosphors, broadening the emission spectrum, and improving quantum efficiency and thermal stability. This review may offer a good reference for researchers improving phosphors to become more suitable for plant growth.

1. Introduction

With the booming development of modern agriculture, the techniques used to regulate plant growth have been increasingly well-established. Light, as the key source of all lives, plays an irreplaceable role in plant growth. However, it has been found that the plant growth is selective in light absorption, mainly due to the differences in light absorption by four main pigments in plants (chlorophyll A, chlorophyll B, photosensitive pigment PR and photosensitive pigment PFR). Blue light (400–500 nm), red light (600–700 nm) and far-red light (700–780 nm) play a crucial role in phototropism, photosynthesis and light morphology during plant growth, respectively.
In contrast, ultraviolet (UV) and yellow-green light contribute little to plant photosynthesis [1]. Therefore, the means to increase the light content of blue light, red light and far-red light in the surroundings of plants to enhance plant photosynthesis, and hence promote plant growth has become a research hotspot in this field. Plant-growth LEDs and light-conversion films have been considered as two effective ways to increase the amount of beneficial light (blue, red and far-red) in the plant’s surroundings due to the modulation capabilities of their spectral components. A plant-growth LED is a phosphor-covered light-emitting diode. Due to the advantages of low energy consumption, high luminous efficiency, long life and adjustable wavelength, it has become the primary artificial light source in the process of plant growth and is widely used in indoor plant cultivation (IPC) so as to remove the limitation of insufficient light on plant cultivation [2].
In comparison, light-conversion films are mainly prepared by adding light-conversion agents to agricultural plastic films. It has the function of converting ultraviolet light and yellow-green light into blue light, red light and far-red light, which are strongly absorbed by plants so as to improve the light quality in the greenhouse and improve the comprehensive utilization rate of solar light energy. They are clean and environmentally friendly forms of a “light energy fertilizer” [3].
In both LEDs and light-conversion films, phosphors determine the light-conversion capacity, spectral range and stability of final products, which are key to effectively promote plant growth. Phosphors are mainly composed of the matrix and activator, whereby activator ions can enter the crystal lattice of the matrix as a luminescent center. After photoexcitation, they transit from the ground state to the excited state. Since the electrons in the excited state will transition from the excited state to the ground state, there will be a radiative transition in this process, resulting in luminescence.
In recent years, a large number of phosphors with excellent performance have been reported by researchers, such as SrMgAl10−yGayO17: Mn4+ (red phosphor) [1], Ca1.1Sr0.9SiO4: Ce3+ (blue phosphor) [4] and Ca3Al2Ge3O12: Eu3+ (far-red phosphor) [5]. However, phosphors with high luminous efficiency, good thermal stability and a high degree of match between the emission spectrum and plant pigment absorption spectrum are relatively rare. This is especially true for phosphors that can meet the requirements of different plants or the different growth periods of the same plant, as different kinds of light absorption are almost absent. Therefore, in this review, on the basis of reviewing the latest progress of phosphors used for plant growth, we discuss in detail the current regulatory strategies for the luminescent properties of phosphors used for plant growth. We believe that this review work will be helpful to explore more luminescent materials with excellent properties for plant growth (Figure 1).

2. Influence of Light on Plant Growth

Light, as an environmental signal and energy for photosynthesis, plays a crucial role in the process of plant growth. Only under the condition of light can plants carry out photosynthesis to synthesize and produce the required organic substances. In a review paper published in 2018, Pattison and colleagues suggested that plants might use several photoreceptors to direct their growth through light and control photosynthesis. These include phytochrome, carotenoid, chlorophyll, phototropic pigment and cryptochromes. Light is used as a fuel for photosynthesis and to control a plant’s form, along with delivering signals to guide fruit shape, leaf expansion and even color [6].
The main absorption range of plant photosynthesis is 300–800 nm, but not all light wavelengths can be used by plants, which are mainly determined by the effective characteristic absorption of plant photosynthetic pigments. The photosensitive pigment is mainly affected by red (660 nm) and far-red (730 nm) light. It dominates plant growth, development and differentiation processes, and includes PR and PFR. These two photosensitive pigments with different absorption spectra can switch states between each other. The PR is the biologically inactive state, which converts to the PFR subtype by a strong absorption peak of 660 nm for red light, while the PFR is the biologically active state, which reverts to the PR state by capturing far-red light at a center of approximately 730 nm [7].
In contrast, cryptochromes and photochromes mainly require light from the blue region (400–490 nm), which affects the accumulation of plant biomass, is conducive to the growth and development of roots, and also has an important influence on the growth of plant stems [8]. Chlorophylls and carotenoids respond sensitively to blue (400–500 nm) and red (600–700 nm) light and are essential for plant growth and development [9]. Therefore, in the solar spectrum, blue, red and far-red light are particularly important for the growth and development of plants, which can effectively accelerate the growth rate of plants and improve the yield and quality of plants. Moreover, a small amount of UV light can improve plant resistance. Yellow-green light (490–550 nm) is also a form of photosynthetically active radiation, but it is generally believed that this light band contributes little to plant growth due to its low absorption by chloroplasts.

3. Fundamental Theory of Phosphors Luminescence

Inorganic phosphor materials are a type of photoluminescent material that can convert absorbed excitation light energy into emitted light for radiation. The basic composition of a typical inorganic phosphor material involves a matrix and an activator. In some cases, traces of other impurity ions (e.g., sensitizers, charge compensators) are introduced to enhance the luminescent properties of the materials [10]. The matrix, which is the main body of the phosphor, provides a suitable substitution site and lattice environment for the activating ion and consists mainly of inorganic compounds such as oxides, nitrides and sulfides. The activator is an optically active impurity doped in the matrix, usually in a small amount. The electron transition of the activator ion is crucial for producing luminescence. Activator ions that can be used as a luminescence center generally include rare earth elements, transition metal elements, etc. [11,12,13].
The fundamentals regarding the luminescence process of phosphor are commonly divided into two types: (1) the activator absorbing energy from the ground state to the excited state, and then through the form of light radiation back to the ground state, during which there is also a part of the form of heat generation through the non-radiative transition back to the ground state [14] and (2) the participation of a sensitizer, which absorbs energy from the ground state to the excited state and then transfers the absorbed energy to the activator through a form of energy transfer. The activator transfers the energy from the high energy level state to the low energy level state and then releases the energy back to the ground state in the same form as the previous class. This type of luminescence is often found in phosphors where the excitation energy absorption of the activator is relatively weak, and the energy transfer between the sensitizer and the activator is used to enhance the luminescence properties [15].
Quantum theory states that the electrons in an atom are in a state of rotation around the nucleus. The electrons can only rotate in a specific orbit, where no energy is produced. Atoms usually have many orbitals, and electrons in different orbitals have different energy states and form different energy levels. With a certain amount of energy, the electrons transition between different energy levels to achieve a change in luminescence [16].

4. Brief Description of Techniques for Improving Plant Growth

As the source of all lives, light plays an irreplaceable role in plant growth, such as germination, growth, flowering and fruiting [17]. Different wavelengths of light can lead to different effects on plant growth. In general, blue light favors photosynthesis, red light affects the phototropism of plants, while far-red light controls the entire life cycle of plants from germination to maturity. Therefore, increasing the light content of blue, red and far-red light in the plant’s surroundings will enhance plant photosynthesis, thereby promoting plant growth [18]. The selective absorption of light from plants makes it possible to control plant growth artificially. With the increasing demand for crop yield, the traditional methods of enhancing crop yields through chemical fertilizers and pesticides are no longer compatible due to their polluting nature. The technology of promoting plant growth and development by artificially increasing the content of beneficial light has been developed rapidly and has become one of the hot research fields in facility agriculture [19]. This can be roughly divided into plant-growth LEDs and light-conversion plastic films.

4.1. Indoor Plant-Growth LED

The traditional agricultural production methods are often affected by extreme weather conditions such as frost, hail, drought, fog and floods, making it difficult ensure crop yields and quality, and severely constraining economic and social development. With the improvement of the human living standard, the demand for a healthy life and a green environment is urgent day by day. However, the traditional method of increasing crop yields through fertilizers and pesticides will bring serious long-term pollution, which is no longer suitable for the needs of plant growth development. Therefore, with the development of modern agriculture, indoor plant cultivation has attracted more and more attention [20,21].
As the “engine” of plant growth, light plays a vital role in indoor crop cultivation [22]. The light sources used in traditional plant cultivation are mainly fluorescent, high-pressure sodium, incandescent and metal halide lamps. However, these light sources not only mismatch the photosynthetic spectra of plants, but also have high costs and a short lifespan. In comparison, phosphor-converting LEDs (pc-LEDs) are a new light source with high energy efficiency and environmental protection advantages. More importantly, the spectral composition of the pc-LEDs can be adjusted to better match the photosynthetic spectra of plants, which makes LEDs develop rapidly in the field of plant lighting [23]. LED lighting has been found to affect morphological or physiological changes in plants [24].
Figure 2 shows the configuration of the pc-LED. A pc-LED is a phosphor-covered light-emitting diode that emits light by combining blue or near ultraviolet (NUV) chips and phosphor [25]. Therefore, the final spectral range can be adjusted through the design of fluorescent powders to provide a better suit for plant photosynthesis. Table 1 shows recent research on phosphors for LEDs for indoor plant cultivation. The table shows that most of the reported plant LEDs are red phosphors and far-red phosphors, with relatively little research on blue phosphors. Nowadays, the high requirements for phosphor luminous intensity, thermal stability and spectral match have led to the rapid development of research into phosphors for LEDs on plant growth.

4.2. Light-Conversion Films

Light-conversion film is a functional film fabricated by adding light-conversion materials to a plastic film. It converts ultraviolet and yellow-green light from sunlight into blue, red and far-red light, which are strongly absorbed by plants. It can improve the quality of light in greenhouses and enhance the comprehensive utilization of solar energy, being a clean and environmentally friendly “light energy fertilizer” [37].
As a result of the loss of available arable land, increasing environmental pollution and the decline in food crops, functional films with specific functions are taking over the film industry. The light-conversion film adds the function of light-conversion in comparison to traditional agricultural film. It can effectively improve the photosynthetic efficiency of plants, increase their resistance to pests and reduce the use of pesticides. Compared with the plant-growth LED, it does not require any high-cost equipment, power and other inputs, in line well with the actual situation of Chinese agricultural development, and has excellent potential for development. Light-conversion agents are the core materials for light-conversion films to realize their light-conversion function. The properties of light-conversion films vary with the type of light-conversion agents. According to the material properties, it can be roughly divided into: rare earth inorganic compounds; rare earth organic complexes; inorganic luminous materials and organic fluorescent dyes [38].
Light-conversion film, a new-fangled type of agricultural material, has garnered significant interests in recent times owing to its potentials for enhancing light utilization efficacy, ensuring adequate light conditions for plants, and augmenting crop yield and quality. Consequently, the film has become the focal point of considerable research and investigations. However, there are still many deficiencies in the efficiency, cost, compatibility with agricultural film, light match with plant pigment absorption and light transmittance, which significantly limit the application of multifunctional agricultural films. Therefore, it is particularly important to develop new light-conversion agents with more excellent properties to promote the development of functional agricultural films.

5. Development of Phosphors for Plant Growth

5.1. Red Phosphor

Chlorophyll and photosensitive pigments are the primary photosynthetic pigments of plants, which have strong absorption of red light. Red light had significant effects on fresh weight, dry weight, stem length, leaf and leaf area [39]. Karimi et al. investigated the effects of red light on the physiological, morphological and phytochemical properties of goldenrod and found that with the increase in red light content, the leaf area and root elongation of plants increased, and the total fresh weight and dry weight of plants increased gradually. This is because red light induces cell division and cell expansion [40]. The principal activator ions commonly used in red phosphors for plant growth are Eu3+, Eu2+, Mn4+, etc. The luminescent characteristics and luminescent principles of these activator ions are described below.

5.1.1. Eu3+-Doped Phosphor

The luminescence of Eu3+ belongs to the 4f–4f transition. Because its 4f orbital is shielded by 5s25p6 orbital in space, the matrix crystal field has little influence on the spectrum of Eu3+-activated phosphor, so the luminescence is relatively stable. Since both the ground state energy level 7F0 and the excited state energy level 5D0 of Eu3+ are nondegenerate (J = 0), the emission and excitation transitions between the 5D0 and 7F0 energy levels can be used to determine the symmetry of the lattice sites occupied by Eu3+ ions [41]. Figure 3 shows a schematic diagram of the energy levels of Eu3+. Under near-ultraviolet or blue light irradiation, the excited electron transits from the ground state to excited state. Then, it returns to the 5D0 energy level after a non-radiative transition and achieves luminescence by a radiative transition from the 5D0 energy level to the ground state 7FJ energy level (J = 0–4) [42,43]. The 5D07FJ (J = 2,3,4) transition produces emissions in the 614 nm, 654 nm and 704 nm bands, in the range of red light (600–700 nm) strongly absorbed by plants. In addition, the synthesis process of Eu3+-activated phosphors is relatively simple as they are prepared by calcination directly under air without a reducing atmosphere. Therefore, Eu3+-activated red phosphors are widely used in plant growth.
Wang et al. prepared Ca9MY0.667(PO4)7(M = Li, Na):Eu3+ phosphors by a conventional solid-state reaction process. The matrix belonged to the R3c-type space group structure, a typical β-Ca3(PO4)2-derived structure, where the Ca2+ ions can be distributed in five crystal degree points with rich and diverse symmetries [42]. As shown in Figure 4a, the sample produces characteristic red light emissions under near-ultraviolet and blue light excitation. The 614 nm emission from 5D07F2 and the 700 nm emission from the 5D07F4 transition have the highest intensity, mainly attributed to the fact that Eu3+ occupies the lattice position at the non-inverted center of low symmetry. As shown in Figure 4b, the fluorescence emission spectra of Ca9NaY0.667(PO4)7:0.12Eu3+ matched with the absorption wavelengths of plant photosensitive pigments (PR and PFR), which could effectively promote plant growth [43]. Sivakumar’s team reported a Li3BaSrLa3(MoO4)8: Eu3+ phosphor. Figure 4c indicates the excitation spectrum of the phosphor, which has strong absorption in the 200–430 nm region due to the overlap of the MoO42− group and the O2−→Eu3+ charge transfer band. Additionally, the large distance between the sites occupied by Eu3+ makes it possible to achieve high-concentration doping of the phosphor. Therefore, the phosphor has a high quantum yield (92.6%). Figure 4d shows that the spectral profile of the prepared red LED closely matches the absorption of the photosensitive pigment (PR), which indicates that the red phosphor can be applied in plant-growth LEDs [44].
In addition to the above inorganic luminescent materials, Eu3+-doped rare earth organic complex materials also have good luminescent properties. Shoji et al. prepared transparent films with UV to red wavelength conversion by coating [Eu(hfa)3(TPPO)2] luminophores mixed with TDMPPO on commercially available plastic films (as shown in Figure 5a,b). The film significantly promotes the growth of plants. Compared to plants developed by uncoated films, plants developed by coated films showed an increase in height by 1.2 times and overall biomass by 1.4 times [45,46].

5.1.2. Eu2+-Doped Phosphor

Eu2+ has a 4f7 electronic configuration with a spectral term of 8S7/2 in the ground state. The lowest excited state may consist of the inner layer of the 4f7 group or the 4f65d1 group. Nevertheless, the energy of the 4f65d1 group of Eu2+ ions are generally lower than those of the 4f7 group at room temperature. Thus, most of the Eu2+ ion-activated materials are characterized by f–d transitions, which exhibit broad spectra and high-intensity emissions [47,48]. In addition, the matrix lattice field can significantly affect the Eu2+ activation of phosphors because of crystal field effects and nephelauxetic effects, resulting in an energy variation of the 5d energy level. The energy of the 5d can be reduced by tuning the matrix lattice field so that the Eu2+ transition occurs over a wide wavelength range, achieving a color shift from the ultraviolet to the red light-emitting region [49,50].
Eu2+-activated phosphors have been extensively studied in plant growth. For instance, Xia’s team changed the crystal field strength by adjusting the doping amount of Rb, thus modifying the emission peak of (RbxK1−x)3LuSi2O7: Eu2+ to make it emit more suitable for the red light range required for plant growth (as shown in Figure 6a) [51]. Lei’s team has prepared red phosphors (Li2Ca2Mg2Si2N6: Eu2+) for the first time under atmospheric pressure using a simple solid-state method. It shows a red emission peak at 638 nm and a full width at a half peak (FWHM) of 62 nm under blue light irradiation. The phosphor emission spectrum remains almost constant at different excitation wavelengths except for the luminous intensity, indicating the presence of only one Eu2+ crystallization site in Li2Ca2Mg2Si2N6. The authors coated Li2Ca2Mg2Si2N6: Eu2+ red phosphor onto blue LED chips (λem = 455 nm) to fabricate plant-growth LEDs, and the device’s emission spectrum matches well with chlorophyll’s absorption range (Figure 6b). The cabbages treated by the plant-growth LEDs had better quality by comparing the growth of cabbages after 10 days under the same conditions with white daylight LED irradiation (Figure 6c), demonstrating that plant-growth LEDs based on this red phosphor can promote plant growth [52]. Furthermore, Lei’s team mixed Sr2Si5N8:2%Eu2+ phosphor with polyethylene to produce light-conversion films by extrusion, pelletizing and blow molding. SEM results images show that the film structure is dense after the addition of the light-conversion agents, leading to an increase in the mechanical properties of the film. The fluorescence spectrum shows that the light-conversion films convert blue-violet light to red light. Through simulation experiments, the biomass and quality of cabbage covered by the light-conversion films were found to be enhanced by comparison with ordinary agricultural films (Figure 6d), effectively promoting plant growth [53].

5.1.3. Mn4+-Doped Phosphor

Mn4+, a transition metal ion, is frequently doped in various matrices as a red emission activator ion. The Mn4+ ion belongs to the d3 electron configuration, similar to the 5d electrons of Eu2+. Electrons in the d3 orbitals are also exposed, so the optical properties of Mn4+-doped materials are also strongly influenced by the crystal field environment [54]. The Mn4+ ion is normally stable in the octahedral position of the solid and emits red light between 600 and 700 nm when excited by two strong and broad excitation bands (4A2g4T1g(4F) and 4A2g4T2g). The emission is produced by the energy level transition of 2Eg4A2g. Since the lowest energy state (2E(t23)) hardly changes in different crystal fields, the luminescence properties are very similar between different bodies but strongly depend on the covalency of the Mn4+ ligand bond [55,56]. The red light produced by the Mn4+-doped phosphor is an ideal match for the red light required for plant growth, which has potential applications in the field of plant-growth LEDs.
As shown in Figure 7a, Deng et al. reported a red light-emitting phosphor (BaZrGe3O9: Mn4+ (BZGO)). The main body of BZGO is a hexagonal structure which belongs to the P6(_)c2 space group. Its crystal structure is composed of two octahedrons of BaO6 and ZrO6. Mn4+ will replace Zr4+ due to the similar ionic radius and valence state. Under UV excitation, the red emission is in the wavelength range of 600–800 nm, which matches the absorption range with chlorophylls A, chlorophylls B and photosensitive pigments (PR) (Figure 7b). The emission intensity of BaZrGe3O9:Mn4+ at 150 °C is 81.7% of that at room temperature, showing excellent thermal stability. A red plant-growth LED was prepared using the phosphor, and the growth rate of garlic and corn was significantly improved under its irradiation. Since 3d→3d belongs to a cosmically forbidden transition, most Mn4+-doped phosphors have a relatively low excitation emission intensity [22]. In order to solve this problem, Wang et al. found that the red emission of Sr4Al14O25:0.01Mn4+ increased by 60% through the incorporation of Ga3+. This is mainly because the Ga–O bond is strongly covalent and can significantly alter the overlap of the 2p orbitals in the M–O–Mn bond, distorting the MnO6 octahedron and disrupting the cosmically forbidden d–d transition in Mn4+. They mixed the phosphors with polydimethylsiloxane to prepare the light-conversion film, and the subsequent chlorella growth experiment proved that the light-conversion film has great potential in promoting plant growth (Figure 7c) [21]. Wang et al. synthesized a red-luminescent ceramic (Mg2TiO4:Mn4+) using a high temperature solid-state reaction method. Compared with phosphor powder, luminescent ceramics have a higher thermal conductivity, which can effectively dissipate heat and reduce the influence of temperature on luminescence intensity. Under the excitation of 465 nm, luminous ceramics of Mg2TiO4:Mn4+ emits red light with a peak wavelength at 658 nm, which matches well with the wavelength required for photosynthesis [57].

5.2. Far-Red Phosphor

Light quality has a regulatory effect on plant photosynthesis. Plants perceive changes in light quality through different types of photoreceptors, among which photosensitivities mainly absorb red and far-red light. When R:FR is low, plants will produce a series of shade avoidance reactions, thus affecting plant growth [58,59,60]. Kalaitzoglo et al. found that increasing R:FR in plant-growth LEDs can negatively affect the growth and early fruit yield of tomato seedlings. The main reason for this is that the lack of far-red light reduces the total leaf area, which in turn leads to a reduction in total plant light absorption and plant dry mass [61]. Far-red light has an important influence on plant growth. In recent years, research on far-red phosphors for plant growth mainly doped with oxides of Mn4+ and Cr3+ have developed rapidly. Under UV excitation, Mn4+-doped red phosphors emit primarily red light around 660 nm. Since its luminescence properties are closely related to the crystal field environment of the matrix, the researchers have prepared a series of Mn4+-doped phosphors with far-red light emission by modulating the crystal environment. For example, Ye’s group has successfully synthesized a series of Mn4+-activated SrLa2Al2O7 phosphors using a high-temperature solid-phase method. Excited by 365 nm, the phosphor of SrLa2Al2O7: Mn4+ emits a far-red light with a peak at 731 nm, which matches well with the absorption band of the photosensitive pigment PFR (Figure 8a) [29]. In addition to finding suitable substrates for Mn4+-doped far-red emission phosphors, cationic substitution can also be used to adjust the luminescence spectrum. Zhou et al. substituted Ga3+ for the Al3+ in the red phosphor of SrMgAl10-yGayO17 doped with Mn4+, as shown in Figure 8b. The luminescence spectrum produced an obvious red shift, which was mainly related to the energy reduction of the 2E level caused by the weak electron cloud rearrangement effect brought by the substitution [1,62].
Recently, the transition group metal Cr3+, along with Mn4+, has also been the focus research on far-red phosphor-activated ions. The electronic configuration of Cr3+ is 3d3, with wide absorption in its transition energy levels of 4T24A2 and 4T14A2, which can be well-matched to the UV chip. The characteristic emission of Cr3+ exhibits spin-forbidden narrow-band red emission (2E→4A2) and spin-allowed broadband far-red emission (4T24A2), making Cr3+-doped phosphors a potential application for plant growth [63]. The strength of the crystal field determines the dominance of the narrow-band red light emission and broadband far-red light emission of Cr3+ phosphors, so the choice of matrix is particularly important for Cr3+-doped phosphors [64]. Huy’s team has successfully synthesized Cr3+-doped BaMgAl10O17 (BAM) narrow-band red phosphors with a hexagonal β-aluminum oxide structure using the sol–gel method. Its strong narrow-band emission (FWHM ~ 4 nm) peaked at 695 nm. Cr3+ doping replaces Al3+ in the matrix at a distorted octahedral position with a strong crystal field (Dq/B > 2.3), resulting in a narrow-band red emission from the phosphor of 2E→4A2 (Figure 8c) [65]. In contrast, Xia’s team produced a broadband near-infrared emission at 650–1350 nm by doping Cr3+ into the weak octahedral coordination crystal field of the lead-free metal halide bis-chalcogenide of Cs2Ag1−xNaxInCl6 [66]. Based on the ability of the crystal field strength to modulate the luminescence of Cr3+-doped phosphors, Li’s team has achieved a modulation of the crystal field strength at the center of the luminescence by a co-substitution of Gd3+ in Gd3Ga5O12 with Y3+ and In3+. As the co-substitution proceeds, Dq/B gradually decreases, resulting in a redshift of the spectrum and a broadening of the FWHM. The electroluminescence (EL) spectrum of a pc-LED made using this phosphor is shown in Figure 8d. The EL spectrum covers well with the light required for PR, PFR, chlorophyll B and phycocyanin during plant growth. Light tests on foliage green plants revealed that the best quality foliage green plants were grown in a R+NIR light environment, thus resulting in the positive application of Gd3Y0.5In0.5Ga5O12: Cr3+ phosphors to plant growth [64].
Figure 8. (a) Comparison of the absorption spectrum of phytochrome PFR and EL spectrum of pc-LED by using SrLa2Al1.994O7:0.006Mn4+ phosphor. (Reprinted from ref. [29]. Copyright of Elsevier, 2023). (b) The normalized PL spectra under 340 nm excitation and the insets are an enlarged part of the peaks in the wavelength ranging from 658 to 668 nm. (Reprinted from ref. [1]. Copyright of Elsevier, 2020). (c) PL spectra measured under excitation wavelength of 405 nm of the BAM:1%Cr3+ phosphor. (Reprinted from ref. [65]. Copyright of Elsevier, 2020). (d) Infrared part of the LED package spectrum with the absorption curve of chlorophyll B, PR, PFR and phycoerythrin. (Reprinted from ref. [64]. Copyright of John Wiley and Sons, 2022).
Figure 8. (a) Comparison of the absorption spectrum of phytochrome PFR and EL spectrum of pc-LED by using SrLa2Al1.994O7:0.006Mn4+ phosphor. (Reprinted from ref. [29]. Copyright of Elsevier, 2023). (b) The normalized PL spectra under 340 nm excitation and the insets are an enlarged part of the peaks in the wavelength ranging from 658 to 668 nm. (Reprinted from ref. [1]. Copyright of Elsevier, 2020). (c) PL spectra measured under excitation wavelength of 405 nm of the BAM:1%Cr3+ phosphor. (Reprinted from ref. [65]. Copyright of Elsevier, 2020). (d) Infrared part of the LED package spectrum with the absorption curve of chlorophyll B, PR, PFR and phycoerythrin. (Reprinted from ref. [64]. Copyright of John Wiley and Sons, 2022).
Nanomaterials 13 01715 g008

5.3. Blue Phosphor

The whole life cycle of plants is strongly influenced by the light environment, especially the light involved in photosynthesis and photophore reaction. Blue light is widely involved in plant growth and development processes, such as photomorphogenesis and leaf expansion. In a study by Li et al., the chlorophyll A, chlorophyll B, carotenoids and total photosynthetic pigments in the leaves of plants irradiated with blue light were 1.3 times higher than in control plants. The growth of the leaves, as well as the rootstock of the plant, is remarkable [23,67,68]. Currently, there are relatively few studies on blue phosphors for plant growth, with only a few Eu2+- and Ce3+-doped phosphors. Regarding Eu2+-doped luminescent materials, their use in red phosphors for plant growth has been described above as having a wide range of applications. Since the 5d energy level of Eu2+ is susceptible to the external crystal field environment, the wavelength position of its emission band is very matrix-dependent, allowing for luminescence modulation from the near-UV to the far-IR [69]. Regarding Eu2+-doped blue phosphors for plant growth, Guo’s team used a high-temperature solid-phase method to synthesize an Eu2+-doped luminescent material (Na2BaSr(PO4)2). The phosphor emits a blue light at 400–500 nm, with a main peak at 428 nm when excited by UV light. The blue region has a spectral similarity index (SR) of 92% with chlorophyll A, which is highly compatible with the absorption spectrum of plant chlorophyll A in the blue light region [36].
Among the rare earth luminescent ions, Ce3+ has the simplest 4f1 electronic configuration. The luminescent characteristics of Ce3+ are caused by an optical transition between the 4f1 ground state and the 5d1 excited state. The 4f1 ground state configuration splits into two energy levels (2F5/2 and 2F7/2) with the number of 5d energy levels depending on the symmetry of the crystal field. The most significant effects on the interaction of Ce3+ with the matrix lattice are the crystal field splitting and nephelauxetic effect, so the position of the excitation and emission bands is closely related to the matrix lattice (i.e., crystal structure, crystal field environment and nephelauxetic effect). Ce3+ emits near-UV to yellow light, and is mostly used as a blue light source in plant growth [70,71]. For example, Lian’s team developed the Ce3+-doped alkaline earth silicate phosphor (CamSr2-m-nBanSiO4: Ce3+, Li+ (CSBS: Ce3+)). The spectral modulation was achieved by varying the composition of Sr and Ba as a proportion of the matrix. A redshift of the spectrum from 417 nm to 438 nm was produced from Ca2SiO4 to Sr0.3Ca1.6SiO4, and a blueshift of 401 nm was produced when the matrix was Ba2SiO4. Different plants have slightly different absorption spectra. This spectrally tunable phosphor has great potential for use in plant-growth LEDs and light-conversion films [72].

5.4. Red and Blue Composite Phosphors

Although specific wavelengths of red and blue light are absorbed by plant photoreceptors and have a positive effect on plant growth, individual blue or red light irradiation not only cannot meet their needs in the actual process of plant growth, but also has some negative effects [73]. Research has found that plants exposed to monochromatic red light can develop “red light syndrome.” This is mainly due to dysfunctional photosynthesis, such as low photosynthetic capacity and slow stomatal response, resulting in reduced plant biomass, leaf area, leaf number, chlorophyll content and stem elongation [74,75,76]. This “red light syndrome” can be reversed by adding blue light. Miao et al. studied the effects of monochromatic red and red-blue composite light on chloroplast ultrastructure, photosynthesis and nutrient accumulation in cucumber seedlings. The results show that by adding blue light, an increase in chloroplasts, larger starch grains and thicker leaves can be obtained, which will directly benefit leaf photosynthetic performance and thus maintain dry matter accumulation [77]. Therefore, the study of red and blue composite phosphor is of great significance for plant growth.

5.4.1. Single-Ion-Doped Red and Blue Composite Phosphors

Eu2+ is the best-known and most widely used of a single-ion-activated multi-color phosphor. The 5d→4f energy transition of Eu2+ is closely related to the crystal environment of the occupied sites, so single-phase-doped multi-color phosphors can be obtained by combining multiple cation sites in a single matrix with different crystal field environments [78,79]. Guo’s group has studied the Eu2+-activated phosphors Ba3GdNa(PO4)3F (BGNPF). As shown in Figure 9a, the matrix Ba3GdNa(PO4)3F has three different Ba2+ sites, including Ba1 with six O coordination, Ba2 with two F coordination to five oxygen atoms, and Ba3 with two F atoms coordination to six oxygen atoms, which are partially substituted by Eu2+. Emission spectra and liquid helium temperature lifetimes demonstrate that Eu2+ enters the three Ba2+ crystal sites and produces emissions at 436 nm, 480 nm and 640 nm, originating from the 4f65d1→4f7 transition. Electroluminescence spectra of an LED fabricated by combining this phosphor with a near-UV chip include two broadband emission peaks in blue (472 nm) and red (608 nm), respectively. It has a high degree of match with the absorption spectra of chlorophyll and carotenoid (Figure 9b) [3]. In addition to the blue-red dual emission produced by Eu2+ at different sites, the red light emission of Eu3+ is also present in some Eu2+-doped phosphors. A single-ion-realized red-blue dual-emission phosphor can also be realized by combining it with the blue-emitting Eu2+ in this system. For example, Xia et al. synthesized a blue-red dual-emission dual-phase phosphor (Na3La2(PO4)3:xEu (Na3La(PO4)2/LaPO4:xEu)) by a high-temperature solid-state reaction method. The main body of the phosphor is a composite material composed of Na3La(PO4)2 and LaPO4 crystals. As seen in Figure 9c, the phosphor shows blue light emission at 422 nm and red light emission at 621 nm under UV excitation. This is mainly due to the presence of both Eu2+ and Eu3+ in the phosphor [80].

5.4.2. Double-Ion Co-Doped Red and Blue Composite Phosphors

In addition to the doping of a single luminescent center, the simultaneous doping of the matrix with different activated ions enables the production of red and blue dual-emission phosphors. For example, Zhang et al. doped two luminescent centers (Ce3+ and Mn2+) into the matrix Na4CaSi3O9 (NSCO). The phosphor exhibits a blue-red double emission with the excitation of 336 nm. The blue emission originates from the 5d→4f transition of the Ce3+, and the red emission originates from the 4T1(4G)-6A1(6S) forbidden spin transition of Mn2+. The emission spectrum of the phosphor has a superior overlap with the absorption spectrum of chlorophyll, as shown in Figure 9d. It can be seen that the light emitted by the phosphors can be well-utilized by plants and has potential applications in the field of plant growth [81]. Moreover, Gong et al. investigated the co-doping of Eu2+ and Mn2+ to achieve red and blue dual emissions. (Na2Mg2Si6O15 was chosen as the matrix). Eu2+ randomly occupies the 6-ligand and 7-ligand of Na+ lattice sites in the matrix, while Mn2+ occupies one of the octahedral Mg2 sites to produce the red light emission. On the other hand, a significant spectral overlap can be observed between the Eu2+ emission band and the Mn2+ excitation transition. This demonstrates that Eu2+ can also transfer a part of its energy to Mn2+, effectively enhancing its forbidden energy level (4T1-6A1) transition. Figure 9e shows the emission spectrum of the phosphor under 365 nm UV excitation, with the blue and red emission peaks matching the photosynthetic action spectrum (PAS) of the plant chlorophyll [82].

6. Luminescence Regulation Strategy of Phosphors

To better cater to the light requirements of plant growth, phosphors for plant growth fulfill the following criteria. (1) The excitation spectrum of the phosphor utilized in light-conversion films should exhibit a significant degree overlap and strong absorption with the ultraviolet or yellow-green light in the solar spectrum. The excitation spectrum of the phosphor applied in plant-growth LED should demonstrate a strong correlation with n-UV light or blue light as well as exhibit high absorption intensity. (2) The emission spectrum of the phosphor should have a high spectral match to the absorption spectrum of the plant. The phosphor must not only emit the blue, red and far-red light required for plant growth, but also possess a sufficiently high emission intensity. (3) Possessing a high quantum efficiency results in efficient luminescence with minimal energy loss. (4) The phosphor must exhibit chemical stability in humid air, bright light and other environmental conditions. Furthermore, the phosphor must exhibit exceptional thermal stability to ensure optimal performance in plant-growth LED applications at temperatures of up to 150 °C. According to the aforementioned, it is essential to adjust and improve the luminescence performance of phosphors. Benefiting from the tunability of the rare earth ion/transition metal ion, specific strategies can be employed to optimize the luminescence properties of the phosphor, such as spectral position, half-peak width and thermal stability. The relevant modulation strategies are described below.

6.1. Adjustment of the Spectral Position

The light absorption of plants is dependent on the type of photoreceptors they possess. These photoreceptors absorb various wavelengths of light, resulting in plants having distinct requirements for specific wavelengths of light. Moreover, plants have different light requirements at different stages of growth, making it crucial to develop phosphors with adjustable spectra for optimal plant development. As mentioned above, benefiting from the close link between the energy levels of rare earth ions/transition metal ions and the crystal environment, the crystal field strength can be modified by structural and compositional design to modulate the excitation and emission wavelengths of the activator. The specific strategies are summarized below.

6.1.1. Strategy for Group Substitution

Substituting an element in the matrix composition by another element of a different ionic radius alters the lattice environment surrounding the activating ion, resulting in a shift of its emission peak. Matrix replacement can be divided into three types: cationic substitution, anionic substitution and chemical unit co-substitution. Among these, cation substitution can be further classified into matrix cation substitution and cation substitution within anionic coordination polyhedra. For example, Wang’s group found that the emission spectrum of RbxK2-xCaPO4F: Eu2+ changed from the red emission (665 nm) of K2CaPO4F: Eu2+ to the cyan-blue emission (487 nm) of Rb2CaPO4F: Eu2+ by replacing K+ with Rb+ (as shown in Figure 10a). The blue shift in the emission spectrum can be attributed to the large lattice expansion of the RbxK2-xCaPO4F: Eu2+ as the concentration of Rb+ with a large radius increases. The large lattice expansion results in an increase in the average bond length of the Eu–O bond, a decrease in the crystal field strength and a reduction in the cleavage of the 5d electron, ultimately leading to an increase in the emission energy [83]. However, Gu et al. discovered that a spectral shift was also observed in Mn4+-doped phosphors (SrMgAl10−yGayO17) when Al3+ was replaced by Ga3+ in the anionic group. The nephelauxetic effect is the primary factor contributing to the spectral shift. The introduction of Ga3+ resulted in a weakened electron cloud expansion effect between the bonds, leading to a reduced Racah parameter B value, which subsequently caused a decrease in the energy of 2Eg4A2 transition, thus resulting in a redshift phenomenon [1]. In summary, the crystalline field strength of the activating ion and the nephelauxetic effect are simultaneously altered by cation substitution, with the final change in spectral position guided by the stronger of these effects.
Anion substitution is another strategy for modulating the luminescent properties of phosphors. The covalence of the crystal is greatly influenced by electronegativity difference between the coordination anions, and the nephelauxetic effect is closely related to the covalence of the crystal. Therefore, the substitution of different electronegativity anions can alter the nephelauxetic effect, modify the energy gap between the lowest energy level of 5d orbitals, ultimately resulting in an adjustment of emission energies and spectral position. The order of electronegativity of several common anions is C4− <N3− <Cl <O2− <F. Gong et al. achieved controlled green-to-blue regulation in the Ba3Ca2(PO4)3F: Eu2+ system using an anion substitution strategy of Cl for F [84]. Similarly, the introduction of the N3− anion in the Ce3+-doped Ca3Sc2Si3O12 phosphor resulted in an emission redshift due to the reduction in energy difference between the lowest 5d excited state and ground state of Ce3+ [85]. In addition, luminescence can be regulated by the co-substitution of chemical units. As shown in Figure 10b, Han Tao’s group substituted [Ca2+–Ge4+] with [Lu3+–Ga3+] in the phosphor of Ca3−xLuxGa2+xGe3−xO12: Cr3+, resulting in a shift of the PL peak position from 766 to 803 nm. On the one hand, the co-substitution of [Lu3+–Ga3+] leads to a decrease in lattice parameters and cell volume, and the contraction of the lattice usually leads to a strong crystal field splitting. On the other hand, the co-substitution reduces the band gap, leading to a downward shift in the energy level of the lowest position of Cr3+ (2E→4T2) and the redshift of the luminescence spectrum [86].

6.1.2. Strategy for Activating Ion Concentration and Phosphors Concentration Regulation

For matrix materials containing multiple cation sites, differences in the lattice environment among these sites cause the same luminescent center to exhibit different luminescence. Additionally, changing the doping concentration in certain systems can impact the lattice position occupied by the activator ion, resulting in the change of luminescent color. In the example of Sr3Sc4O9:xEu phosphor, the matrix Sr3Sc4O9 contains three types of Sr2+ sites, Sr1O6, Sr2O9 and Sr3O1, and an Sc site that can be replaced by Eu2+. The selective site occupation of Eu2+ ions in the matrix occurs as the Eu2+ doping concentration increases. The least energy is required to enter the 6-ligand cation site, so the Eu2+ ion preferentially occupies the 6-ligand Sr and Sc sites and exhibits only one red light emission band (Figure 10c). As the x value increases, the Eu2+ ions begin to occupy the Sr2 and Sr3 sites, showing full spectral emission. In higher doping concentrations, the Eu2+ cation only occupies the Sr1–3 sites, and the long-wave NIR emission peak from the Sc3+ site disappears. This is because Eu3+ is a better match to Sc3+ in terms of radius and charge, so at higher concentrations, the Eu ions that enter the Sc3+ site are in the trivalent state to maintain the stability of the overall crystal structure [87]. Furthermore, similar phenomena can be observed in phosphor systems such as Sr3Ga2Ge4O14: Cr3+ [88], LiYGeO4: Tb3+ [89] and Li2Mg3TiO6: Cr3+ [90]. Phosphors must ultimately be dispersed in specific types of polymers, such as poly(methyl methacrylate) (PMMA), poly(dimethylsiloxane) (PDMS), poly(vinylpyrrolidone) (PVP), poly(ethylene oxide) (PEO), etc., to make light-conversion films or plant-growth LEDs. The combination of phosphor and polymer can have a positive effect on the phosphor [91]. For example, by protecting the phosphor from environmental conditions or preventing aggregation, the fluorescent properties of the powder can be improved, which would otherwise lead to sudden extinction. It has been discovered that altering the phosphor concentration in the polymer affects the phosphor’s dispersion and interaction with the polymer, modifying the composite’s luminous color. As an illustration, Maria Luisa Saladino et al. discovered that the emission peaks of Ce3+: Y3Al5O12 (Ce:YAG)-PMMA composites shifted by 15 nm as the amount of Ce:YAG increased and that the clusters of Ce:YAG particles were more evenly distributed in composites with higher doping concentrations. The xrd diffraction pattern of the samples with different concentrations showed no change in the Ce:YAG peaks compared to the powder, proving that it was not the change in phosphor structure that caused the shift in the luminescence position. Further solid-state NMR tests showed that the addition of Ce:YAG particles strongly influenced the T1𝜌(C) values. The change in relaxation time demonstrates that electron donor interactions occur between the carboxy-oxygen lone pair and the surface yttrium or cerium ions, resulting in a shift in the emission spectrum and a more homogeneous composite [92]. The same phenomenon was observed in composites formed by the doping of the rare earth organic complex Eu(TTA)3(H2O)2 (TTA: thenoyltrifluoroacetone) in poly(ethylene oxide) (PEO), where the luminescence position of the composite shifted with the change in the molar ratio of the complex to the polymer [93].

6.1.3. Crystal Phase Engineering

In some special cases, activator doping or component substitution can cause a change in the crystalline phase of the matrix. Different crystalline phases will naturally have different lattice environments, which results in luminescence modulation in the phosphor. For example, Prof. Lian’s team formed an additional K2BaCa(PO4)2 phase by introducing Ba2+ into K2Ca(PO4)F: Eu2+, and its content increased with Ba2+. The luminescence spectrum also gradually shifts from red emissions in the K2Ca(PO4)F matrix to cyan emissions in K2Ca(PO4)F, achieving tunable emission colors (Figure 10d) [94].
Figure 10. (a) Normalized emission spectra of RbxK2−xCaPO4F: Eu2+ (0 ≤ x ≤ 2) excited at 380 nm. (Reprinted from ref. [83]. Copyright of John Wiley and Sons, 2021). (b) Schematic illustration of the first and second coordination environments. (Reprinted from ref. [86]. Copyright of John Wiley and Sons, 2021). (c) The PL spectra and of Sr3Sc4O9: xEu excited at 350 nm. (Reprinted from ref. [87]. Copyright of Royal Society of Chemistry, 2022). (d) PL spectra of K2BaCa(PO4): Eu2+ phosphors and the photographs of the samples under a 365 nm UV lamp. (Reprinted from ref. [94]. Copyright of Elsevier, 2023).
Figure 10. (a) Normalized emission spectra of RbxK2−xCaPO4F: Eu2+ (0 ≤ x ≤ 2) excited at 380 nm. (Reprinted from ref. [83]. Copyright of John Wiley and Sons, 2021). (b) Schematic illustration of the first and second coordination environments. (Reprinted from ref. [86]. Copyright of John Wiley and Sons, 2021). (c) The PL spectra and of Sr3Sc4O9: xEu excited at 350 nm. (Reprinted from ref. [87]. Copyright of Royal Society of Chemistry, 2022). (d) PL spectra of K2BaCa(PO4): Eu2+ phosphors and the photographs of the samples under a 365 nm UV lamp. (Reprinted from ref. [94]. Copyright of Elsevier, 2023).
Nanomaterials 13 01715 g010

6.2. Adjustment of the Spectral Width

The broad blue and red light spectra have a better match for the absorption spectrum of the plant, providing a more effective support for plant growth and development. There are two main strategies for broadening the emission spectrum. One method is to dope the activating ion into multiple lattice sites in the matrix, allowing for different crystal field environments and resulting in a broad-spectrum emission from the combination of luminescence at different sites.
As shown in Figure 11a, the BaIn2(P2O7)2 (BIP) structure has two unequal [InO6] octahedral positions, both of which share angles with six [PO4] tetrahedra. The average bond lengths of In1-O and In2-O are 2.1311 Å and 2.2317 Å, respectively. BaIn2(P2O7)2:0.02Cr3+ exhibits a broad near-infrared emission in the range of 700–1400 nm when excitation occurs at 480 nm. The broadband NIR emission arises from the overlap of two spectral bands centered at 879 and 958 nm, originating from Cr3+ located in distinct [InO6] octahedra [95]. Similarly, broadband emission resulting from multi-location occupation was observed in Bi3+-doped phosphors. There are three types of Sr sites in the matrix Sr3Al2O5Cl2 that can be occupied by Bi3+ ions. It was determined that the Bi3+ ion exclusively occupied Sr1 and Sr3 positions. Due to the longer average bond length of the Sr1 site, broadband emission at 490 nm dominated by the Sr1 position is generated under excitation at 345 nm excitation, and 556 nm dominated by the Sr3 position is produced upon excitation at 376 nm [96].
Another strategy is to co-dope the activator with a distinct emission originating from other ions. The emission spectrum can be broadened by activating the energy transfer between the ions and other luminescent centers. In the process of energy transfer, other luminescent centers not only generate their own emissions, but also act as sensitizers to transfer energy to the activating ion. The prerequisite for achieving efficient energy transfer is that the emission spectrum of the sensitizer must have a certain degree of overlap with the absorption spectrum of the activator. For example, the emission spectrum of Eu2+ overlaps with the excitation spectrum of Mn2+ in the phosphor of KMg4(PO4)3: Eu2+, Mn2+, which offers the possibility of energy transfer from Eu2+ to Mn2+. Under excitation of 365 nm, the KMg4(PO4)3: Eu2+, Mn2+ phosphors exhibit a broad excitation band ranging from 250 to 425 nm and two broad emission bands peaking at 450 nm and 625 nm, which are attributed to the 4f-5d transition of Eu2+ and the 4T16A1 transition of the Mn2+ ion. As shown in Figure 11b, the FWHM value of the Mn2+ ion emission peak band in KMg4(PO4)3 gradually increases with increasing Mn2+ concentration. The energy transfer from Eu2+ to Mn2+ effectively broadens the emission spectrum within the red spectral region [97]. In addition, the Cr3+ ion has a broad absorption band throughout the UV–Vis region, and its excitation spectrum overlaps with most of the luminescent ions, such as Eu2+, Ce3+, Bi3+, Yb3+, etc. Lei’s group has studied Eu2+-, Cr3+-doped Mg2Al4Si5O18 phosphors. Mg2Al4Si5O18 provides sufficient sites for the incorporation of Eu2+ and Cr3+. The diffuse reflectance spectra (DRS) of Mg2Al4Si5O18: Cr3+ and the emission spectra of Mg2Al4Si5O18: Eu2+ exhibit spectral overlap in the range of 500 nm to 850 nm (Figure 11c), which indicates the possibility of energy transfer between the Eu2+ and Cr3+ ions [98]. The phosphor spectrum of Mg2Al4Si5O18: Eu2+ shows the characteristic 4f65d→4f7 transition red emission of Eu2+ under 450 nm excitation. Excitation emission spectra with characteristic excitation/emission bands of Eu2+ and Cr3+ were clearly identified in Mg2Al4Si5O18: Eu2+, 0.01Cr3+, implying the successful construction of ultra-wide vis–NIR phosphors. Except for co-doping with other luminescent ions as sensitizers to increase the spectral emission range, doping activating ions in matrix with self-luminescent properties is also an effective means of achieving a broad-spectrum emission. For instance, Xie’s group synthesized Cr3+-doped SrHfO3 (SHO) by high-temperature solid-phase method. SHO emits near-infrared light in the range of 700–800 nm due to the presence of self-trapping excitons. Considering the similarity to an ionic radius, Cr3+ ions are preferred to replace Hf4+ ions in the octahedral position. As shown in Figure 11d, the emission band of SHO:0.005Cr consists of two parts when excited at 460 nm. The spectral emission at 700–850 nm originates from the autofluorescence of the SHO host, the emission at a peak half-height width of 190 nm at 1000 nm results from the transitions of 4T24A2 allowed by the Cr3+ spin [99].

6.3. Adjustments of Other Performance

Quantum efficiency is an essential indicator of a luminescent material’s light-conversion capability, reflecting the phosphor’s ability to convert absorbed energy into emitted light. The higher the quantum efficiency, the more efficient the light-conversion, resulting in less energy loss. There are several strategies to improve the quantum efficiency of phosphors, and all of the methods mentioned above have been reported to optimize quantum efficiency. For example, Wang et al. achieved a 16.67% increase in internal quantum efficiency (IQE) by a cation substitution strategy of Al to In in the Cs2KInF6: Cr3+ system [100]. Yan et al. reported Mn4+ and Bi3+ co-doped Gd2SrAl2O7 phosphors with a significant increase in quantum efficiency (79.9%) by energy transfer from Bi3+ to Mn4+ [101]. The main focus will be to introduce a different strategy: the flux strategy which aims to improve the quantum efficiency of the phosphors. Li et al. successfully prepared a series of yellow light-emitting samples (Ca2MgWO6: Bi3+) using a high-temperature solid-phase method. The addition of MgF2 as a flux in the preparation resulted in the enhancement of the luminescence properties by adjusting the lattice and morphology of the model. The X-ray diffraction pattern is shown in Figure 12a. With the increase in MgF2 concentration, the impurity phase of CaWO4 initially present gradually decreased. When the addition amount reached 2 wt%, the impurity phase disappeared, and the emission intensity of Bi3+ ions increased by 21.4%. The phosphor has a high quantum efficiency and absorption efficiency with maximum values of 42.1% and 89.1%, respectively [102]. Park et al. investigated the effect of different fluoride fluxes on the photoluminescence properties of Lu2.94Al5O12:0.06Ce3+, such as LiF, NaF and BaF2. The melting point of the flux is lower than the solid-state reaction temperature (1400 °C). This liquid phase will be present as a liquid phase during the reaction and acts as a facilitator of ion diffusion, thereby facilitating the formation of highly crystalline phosphors. The excitation-emission spectra show that the addition of the flux does not alter the shape and position of the phosphor excitation peak but only increases the intensity. However, the emission peak was blue-shifted, which was attributed to the addition of a flux. The addition of a flux increased the crystallinity of the phosphor and decreased the lattice defects around the Ce3+ ions, thus reducing the crystal field splitting [103].
Apart from quantum efficiency, excellent thermal stability is essential for phosphors used in plant-growth LEDs because the operating temperature can reach up to 150 °C. Nevertheless, the luminescence intensity of fluorescent materials generally decreases with the increase in temperature, which is called the thermal quenching phenomenon. Therefore, improving the thermal quenching performance of phosphors remains a major research focus. The thermal stability of luminescent materials is closely related to the matrix lattice. The high structural stiffness of the matrix lattice represents a weak lattice vibration, which essentially limits the non-radiative transition within the phosphor and enhances the thermal burst resistance of the phosphor. For example, You’s group has synthesized a series of Cr3+-activated LiABO4 (a = Al, Ga; B = Ge, Si) with a similar benzene ring structure. As shown in Figure 12b, the crystal structures of LiGaGeO4 and LiAlSiO4 are rigid and similar to that of a benzene ring. Due to the extensive structural stiffness, the emission intensity of LiGaGeO4:002Cr3+, LiAl0.1Ga0.9GeO4:0.002Cr3+ and LiAlSiO4:0.02Cr3+ at 423 K (150 °C) remains at 85%, 90% and 94% compare to that at 298 K (25 °C) (Figure 12c,d) [104]. In addition, some trap energy levels can be created in matrix through non-equivalent substitution. These trap energy levels can act as electron-capture centers to store the energy and release electrons at a high temperature. These electrons which return to the ground state can compensate for the loss of thermal bursts, thus improving the thermal stability of the phosphor. Bai et al. reported a new red-emitting phosphor K2MgGeO4: Eu3+ which exhibits zero thermal quenching. As shown in Figure 12e, With the increase in temperature, the emission intensity of the phosphor gradually increased. When the temperature rises to 300 °C, the intensity remains at 107.22% of the initial value at room temperature. The ICP-OES results revealed that the actual content of the potassium element was lower than the theoretical content, indicating the presence of potassium vacancies in the lattice of the doped sample. In addition, the increasing trend of oxygen concentration suggests the presence of interstitial oxygen atoms in the lattice. Oi is also demonstrated by the deconvolution peak at 533.76 eV in the high-resolution XPS spectrum. Due to the non-equivalent substitution of Eu3+ ions for K+, two kinds of defects are generated in the lattice. Under thermal activation, the electrons trapped by the defects are re-released to participate in the radiative transition, thus compensating for the energy loss caused by the non-radiative transition at high temperature, and finally achieving a zero thermal quenching performance of the phosphor [105].

7. Summary and Outlook

In this review, we systematically summarize the research progress of phosphors for plant growth. The main classifications are based on the central light-absorbing regions of plant photosynthesis, including red, blue, far-red and red-blue composite phosphors. The luminescent properties of the relevant phosphors are described from the perspective of the luminescent ions. In particular, several strategies to meet the requirements for phosphors are summarized in the direction of light-conversion films and plant-growth LEDs by adjusting the spectral wavelength, broadening the emission spectrum and improving quantum efficiency and thermal stability. Modifying the spectral position through component substitution in the matrix structure is a very effective method. Additionally, adjusting the concentration of the activator and altering the crystal phase to modify the crystal environment around the luminescent center is an alternative strategy for achieving spectral shifts. The selection of matrixes with multiple occupancy sites for activating ions is often employed to broaden the spectral range, thereby generating a wide range of spectral emissions. Moreover, co-doping with other luminescent ions can be used to complement the spectral emission. Quantum efficiency and thermal stability are crucial factors in the applications of phosphors. Here, the improvement of phosphor quantum efficiency through flux strategies and the mechanism of defect engineering to enhance thermal stability are highlighted.
Despite extensive research on regulating the luminescence properties of phosphors used for plant growth, there remains many numerous unresolved issues. (1) For light-conversion films, the compatibility between inorganic luminescent materials and resin is poor, which significantly impacts the uniformity and light transmittance of the film. (2) Although the co-doping strategy can broaden the luminescence spectrum of phosphors, it often results in a reduction in luminescence efficiency due to the energy transfer between luminescent ions. (3) The formation of defects in the matrix lattice can enhance the thermal stability of phosphors by trapping electrons, but it can also lead to the intrinsic luminescence reduction in phosphors. (4) The rigid structure of phosphors results in weak thermal vibrations, which are beneficial for reducing non-radiative transitions, but results in a narrow-band emission.

Author Contributions

Conceptualization, X.J.; methodology, C.Y., W.L. and X.Z.; resources, X.J. and C.Y.; writing—original draft preparation, C.Y.; writing—review and editing, X.J., C.Y., X.Z., W.L., S.L., L.X. and J.S.; visualization, X.J., Q.Y. and C.Y.; supervision, W.L. and L.X.; funding acquisition, X.J. All authors have read and agreed to the published version of the manuscript.

Funding

Supported by the Science and Technology Program of University of Jinan (No. XKY2068-140200568) and The Major Subject of the University of Jinan (1420702).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gu, S.; Xia, M.; Zhou, C.; Kong, Z.; Molokeev, M.S.; Liu, L.; Wong, W.-Y.; Zhou, Z. Red shift properties, crystal field theory and nephelauxetic effect on Mn4+-doped SrMgAl10-yGayO17 red phosphor for plant growth LED light. Chem. Eng. J. 2020, 396, 125208. [Google Scholar] [CrossRef]
  2. Ma, N.; Li, W.; Devakumar, B.; Zhang, Z.; Huang, X. Finding an efficient far-red-emitting CaMg2La2W2O12:Mn4+ phosphor toward indoor plant cultivation LED lighting. Mater. Today Chem. 2021, 21, 100512. [Google Scholar] [CrossRef]
  3. Chen, J.; Zhang, N.; Guo, C.; Pan, F.; Zhou, X.; Suo, H.; Zhao, X.; Goldys, E.M. Site-Dependent Luminescence and Thermal Stability of Eu2+ Doped Fluorophosphate toward White LEDs for Plant Growth. ACS Appl. Mater. Interfaces 2016, 8, 20856–20864. [Google Scholar] [CrossRef] [PubMed]
  4. Chen, Y.; Zheng, X.; He, Y.; Han, Y.; Zhang, J.; Qiu, Z.; Zhou, W.; Yu, L.; Lian, S. Precisely control the ultraviolet to blue light conversion for plant growth: Rigid crystal structure, lattice substitution and flux effect in the Ca1.1Sr0.9SiO4:Ce3+, Li+ phosphor. Mater. Res. Bull. 2022, 150, 111760. [Google Scholar] [CrossRef]
  5. Miao, S.; Shi, R.; Zhang, Y.; Chen, D.; Liang, Y. Deep-Red Ca3Al2Ge3O12:Eu3+ Garnet Phosphor with Near-Unity Internal Quantum Efficiency and High Thermal Stability for Plant Growth Application. Adv. Mater. Technol. 2023, 2023, 2202103. [Google Scholar] [CrossRef]
  6. Pattison, P.M.; Tsao, J.Y.; Brainard, G.C.; Bugbee, B. LEDs for photons, physiology and food. Nature 2018, 563, 493–500. [Google Scholar] [CrossRef]
  7. Zhou, Z.; Zheng, J.; Shi, R.; Zhang, N.; Chen, J.; Zhang, R.; Suo, H.; Goldys, E.M.; Guo, C. Ab Initio Site Occupancy and Far-Red Emission of Mn4+ in Cubic-Phase La(MgTi)1/2O3 for Plant Cultivation. ACS Appl. Mater. Interfaces 2017, 9, 6177–6185. [Google Scholar] [CrossRef]
  8. Pedmale, U.V.; Huang, S.C.; Zander, M.; Cole, B.J.; Hetzel, J.; Ljung, K.; Reis, P.A.B.; Sridevi, P.; Nito, K.; Nery, J.R.; et al. Cryptochromes Interact Directly with PIFs to Control Plant Growth in Limiting Blue Light. Cell 2016, 164, 233–245. [Google Scholar] [CrossRef]
  9. Chen, J.; Guo, C.; Yang, Z.; Li, T.; Zhao, J. Li2SrSiO4:Ce3+, Pr3+ Phosphor with Blue, Red, and Near-Infrared Emissions Used for Plant Growth LED. J. Am. Ceram. Soc. 2016, 99, 218–225. [Google Scholar] [CrossRef]
  10. Gupta, I.; Singh, S.; Bhagwan, S.; Singh, D. Rare earth (RE) doped phosphors and their emerging applications: A review. Ceram. Int. 2021, 47, 19282–19303. [Google Scholar] [CrossRef]
  11. You, L.; Tian, R.; Zhou, T.; Xie, R.-J. Broadband near-infrared phosphor BaMgAl10O17:Cr3+ realized by crystallographic site engineering. Chem. Eng. J. 2021, 417, 129224. [Google Scholar] [CrossRef]
  12. Huang, M.-H.; Zhu, Q.-Q.; Li, S.; Zhai, Y.; Zhang, H.; Wang, L.; Xie, R.-J. Thermally robust Al2O3–La3Si6N11:Ce3+ composite phosphor-in-glass (PiG) films for high-power and high-brightness laser-driven lighting. J. Mater. Chem. C 2023, 11, 488–496. [Google Scholar] [CrossRef]
  13. Liang, Y.; He, J.; Song, Z.; Han, Y.; Qiu, Z.; Zhou, W.; Zhang, J.; Yu, L.; Lian, S. Novel Dual-Excitation and Dual-Emission Materials: Eu2+,Pb2+ Co-doped Core–Shell-Structured CaS@CaZnOS Phosphors and Their Application for Highly Efficient Photosynthesis of Plants. ACS Appl. Mater. Interfaces 2022, 14, 1413–1422. [Google Scholar] [CrossRef] [PubMed]
  14. Fan, X.; Zhang, W.; Lü, F.; Sui, Y.; Wang, J.; Xu, Z. Research of Fluorescent Properties of a New Type of Phosphor with Mn2+-Doped Ca2SiO4. Sensors 2021, 21, 2788. [Google Scholar] [CrossRef] [PubMed]
  15. Kumar, A.; Manam, J. Color tunable emission and temperature dependent photoluminescence properties of Eu3+ co-doped Gd2Zr2O7:Dy3+ phosphors. Opt. Mater. 2019, 96, 109373. [Google Scholar] [CrossRef]
  16. Wei, Y.; Yang, H.; Gao, Z.; Yun, X.; Xing, G.; Zhou, C.; Li, G. Anti-Thermal-Quenching Bi3+ Luminescence in a Cyan-Emitting Ba2ZnGe2O7:Bi Phosphor Based on Zinc Vacancy. Laser Photonics Rev. 2021, 15, 2000048. [Google Scholar] [CrossRef]
  17. Liu, X.; Shi, Y.; Cheng, K.; Gong, X.; Huang, W.; Deng, C. Na0.2La0.2Sr1.6WO6:Mn4+: An efficient far-red-emitting phosphor via a cation-pair partial substitution. Opt. Mater. 2023, 136, 113478. [Google Scholar] [CrossRef]
  18. Liu, X.; Huang, Q.; Cai, P.; Zhao, H.; Hu, Z.; Lu, L.; Zhu, Z.; Pu, X.; Wang, X.; Ai, Q.; et al. Mn4+ doped Mg2ScSbO6 deep-red-emitting double-perovskite phosphors for plant-cultivation. J. Lumin. 2023, 256, 119624. [Google Scholar] [CrossRef]
  19. Dewir, Y.H.; Al-Ali, A.M.; Rihan, H.Z.; Alshahrani, T.; Alwahibi, M.S.; Almutairi, K.F.; Naidoo, Y.; Fuller, M.P. Effects of Artificial Light Spectra and Sucrose on the Leaf Pigments, Growth, and Rooting of Blackberry (Rubus fruticosus) Microshoots. Agronomy 2023, 13, 89. [Google Scholar] [CrossRef]
  20. Matysiak, B.; Ropelewska, E.; Wrzodak, A.; Kowalski, A.; Kaniszewski, S. Yield and Quality of Romaine Lettuce at Different Daily Light Integral in an Indoor Controlled Environment. Agronomy 2022, 12, 1026. [Google Scholar] [CrossRef]
  21. Wang, S.; Seto, T.; Liu, B.; Wang, Y.; Li, C.; Liu, Z.; Dong, H. Tremendous Acceleration of Plant Growth by Applying a New Sunlight Converter Sr4Al14−xGaxO25:Mn4+ Breaking Parity Forbidden Transition. Adv. Sci. 2023, 10, 2204418. [Google Scholar] [CrossRef] [PubMed]
  22. Deng, K.; Jin, Y.; Yuan, L.; Wu, H.; Hu, Y. Mn4+-activated red pc-LED for precisely matching the spectral absorption and circadian rhythm of photoreceptor toward promoting crop growth. J. Alloys Compd. 2023, 938, 168493. [Google Scholar] [CrossRef]
  23. Dutta Gupta, S.; Jatothu, B. Fundamentals and applications of light-emitting diodes (LEDs) in in vitro plant growth and morphogenesis. Plant Biotechnol. Rep. 2013, 7, 211–220. [Google Scholar] [CrossRef]
  24. Al Murad, M.; Razi, K.; Jeong, B.R.; Samy, P.M.A.; Muneer, S. Light Emitting Diodes (LEDs) as Agricultural Lighting: Impact and Its Potential on Improving Physiology, Flowering, and Secondary Metabolites of Crops. Sustainability 2021, 13, 1985. [Google Scholar] [CrossRef]
  25. Shao, B.; Huo, J.; You, H. Prevailing Strategies to Tune Emission Color of Lanthanide-Activated Phosphors for WLED Applications. Adv. Opt. Mater. 2019, 7, 1900319. [Google Scholar] [CrossRef]
  26. Parale, P.N.; Kadam, A.R.; Dabre, K.V.; Dhoble, S.J. Spectroscopic investigation of KBa2(PO3)5:Mn4+ activated glasses for plant cultivation applications. Mater. Lett. X 2023, 18, 100191. [Google Scholar] [CrossRef]
  27. Jin, C.; Li, R.; Liu, Y.; Zhou, C.; Sun, P.; Luo, Z.; Liu, Z.; Jiang, J. High-performance Gd3Al4GaO12:Cr3+ phosphors for next-generation far-red LEDs. Mater. Res. Bull. 2023, 163, 112234. [Google Scholar] [CrossRef]
  28. Meng, X.; Wang, Z.; Yan, Z.; Guo, D.; Liu, M.; Gong, J.; Feng, X.; Zhang, T.; Li, X.; Li, P. Flux induced highly efficient and stable phosphor Sr2ScSbO6:Mn4+ for plant growth lighting. J. Mater. Chem. C 2023, 11, 5064–5072. [Google Scholar] [CrossRef]
  29. Zou, W.; Nie, W.; Wu, D.; Wu, S.; Wang, W.; Peng, J.; Ye, X. Synthesis, luminescence properties and potential applications for plant growth: A novel Mn4+-activated SrLa2Al2O7 phosphor with far-red emission. J. Lumin. 2023, 257, 119759. [Google Scholar] [CrossRef]
  30. Khan, M.S.; Mehare, M.D.; Parauha, Y.R.; Dhale, S.A.; Dhoble, S.J. Synthesis and novel emission properties of Bi3+-doped Ca2BO3Cl phosphor for plant cultivation. Luminescence 2023, 38, 4–11. [Google Scholar] [CrossRef]
  31. Cheng, K.; Xu, Y.; Liu, X.; Long, J.; Huang, W.; Deng, C. A novel far-red phosphors Li2ZnTi3O8:Cr3+for indoor plant cultivation:Synthesis and luminescence properties. Ceram. Int. 2023, 49, 6343–6350. [Google Scholar] [CrossRef]
  32. Li, L.; Cao, Q.; Xie, J.; Wang, W.; Wang, J.; Pan, Y.; Wei, X.; Li, Y. Novel far-red emitting phosphor Mn4+-activated BaLaLiWO6 with excellent performance for indoor plant cultivation of light-emitting diodes. J. Alloys Compd. 2023, 934, 167927. [Google Scholar] [CrossRef]
  33. Singh, K.; Vaidyanathan, S. Stable and efficient narrow-band red emitters with high colour purity for white LEDs and plant growth applications. Dalton Trans. 2022, 51, 11255–11266. [Google Scholar] [CrossRef] [PubMed]
  34. Bargat, S.R.; Parauha, Y.R.; Shirbhate, N.S.; Mishra, G.; Dhoble, S.J. Novel red colour emitting Ca0.995Mg2(SO4)3:0.5Eu2+ phosphor under ultraviolet, blue, and green excitation for plant growth LEDs. Luminescence 2022, 37, 463–471. [Google Scholar] [CrossRef] [PubMed]
  35. Gao, P.; Li, Q.; Zhou, C.; Chen, K.; Luo, Z.; Zhang, S.; Molokeev, M.S.; Wang, J.; Zhou, Z.; Xia, M. High-Efficiency Continuous-Luminescence-Controllable Performance and Antithermal Quenching in Bi3+-Activated Phosphors. Inorg. Chem. 2022, 61, 13104–13114. [Google Scholar] [CrossRef]
  36. Xiang, J.; Zhao, X.; Suo, H.; Jin, M.; Zhou, X.; Chen, C.; Guo, C. Very stable and high-quantum-yield phosphor Na2BaSr(PO4)2:Eu2+ for plant growth LEDs. Mater. Chem. Front. 2021, 5, 6192–6199. [Google Scholar] [CrossRef]
  37. Zhang, Z.; Zhao, Z.; Lu, Y.; Wang, D.; Wang, C.; Li, J. One-Step Synthesis of Eu3+-Modified Cellulose Acetate Film and Light Conversion Mechanism. Polymers 2021, 13, 113. [Google Scholar] [CrossRef]
  38. Liu, Y.; Gui, Z.; Liu, J. Research Progress of Light Wavelength Conversion Materials and Their Applications in Functional Agricultural Films. Polymers 2022, 14, 851. [Google Scholar] [CrossRef] [PubMed]
  39. Izzo, L.G.; Mickens, M.A.; Aronne, G.; Gómez, C. Spectral effects of blue and red light on growth, anatomy, and physiology of lettuce. Physiol. Plant. 2021, 172, 2191–2202. [Google Scholar] [CrossRef]
  40. Karimi, M.; Ahmadi, N.; Ebrahimi, M. Red LED light promotes biomass, flowering and secondary metabolites accumulation in hydroponically grown Hypericum perforatum L. (cv. Topas). Ind. Crop. Prod. 2022, 175, 114239. [Google Scholar] [CrossRef]
  41. Binnemans, K. Interpretation of europium(III) spectra. Coord. Chem. Rev. 2015, 295, 1–45. [Google Scholar] [CrossRef]
  42. Dickens, B.; Schroeder, L.W.; Brown, W.E. Crystallographic studies of the role of Mg as a stabilizing impurity in β-Ca3(PO4)2. The crystal structure of pure β-Ca3(PO4)2. J. Solid State Chem. 1974, 10, 232–248. [Google Scholar] [CrossRef]
  43. Wang, S.; Han, Y.; Shi, L.; Tong, Y.; Zhao, Q.; Zhang, J.; Mao, Z.; Mu, Z.; Zhang, Z.; Niu, S. A new strategy to the phosphors for plant growth LEDs: Far red emission from the Ca9MY0.667 (PO4)7 (M = Li, Na):Eu3+ phosphors due to the Eu3+: 5D07F4 transition. J. Lumin. 2020, 225, 117404. [Google Scholar] [CrossRef]
  44. Singh, K.; Rajendran, M.; Devi, R.; Vaidyanathan, S. Narrow-band red-emitting phosphor with negligible concentration quenching for hybrid white LEDs and plant growth applications. Dalton Trans. 2021, 50, 4986–5000. [Google Scholar] [CrossRef] [PubMed]
  45. Hasegawa, Y.; Yamamuro, M.; Wada, Y.; Kanehisa, N.; Kai, Y.; Yanagida, S. Luminescent Polymer Containing the Eu(III) Complex Having Fast Radiation Rate and High Emission Quantum Efficiency. J. Phys. Chem. A 2003, 107, 1697–1702. [Google Scholar] [CrossRef]
  46. Shoji, S.; Saito, H.; Jitsuyama, Y.; Tomita, K.; Haoyang, Q.; Sakurai, Y.; Okazaki, Y.; Aikawa, K.; Konishi, Y.; Sasaki, K.; et al. Plant growth acceleration using a transparent Eu3+-painted UV-to-red conversion film. Sci. Rep. 2022, 12, 17155. [Google Scholar] [CrossRef]
  47. George, N.C.; Denault, K.A.; Seshadri, R. Phosphors for Solid-State White Lighting. Annu. Rev. Mater. Res. 2013, 43, 481–501. [Google Scholar] [CrossRef]
  48. Qin, X.; Liu, X.; Huang, W.; Bettinelli, M.; Liu, X. Lanthanide-Activated Phosphors Based on 4f-5d Optical Transitions: Theoretical and Experimental Aspects. Chem. Rev. 2017, 117, 4488–4527. [Google Scholar] [CrossRef]
  49. Hu, T.; Jiang, Z.; Wang, B.; Yu, T.; Wen, D.; Zeng, Q.; Gao, Y. Eu2+ luminescence in CaYGaO4 olivine: A new efficient red phosphor for warm illumination. J. Mater. Chem. C 2023, 11, 2153–2161. [Google Scholar] [CrossRef]
  50. Leaño, J.L., Jr.; Mariano, C.O.M.; Huang, W.-T.; Mahlik, S.; Lesniewski, T.; Grinberg, M.; Sheu, H.-S.; Hu, S.-F.; Liu, R.-S. Thermally Stable and Deep Red Luminescence of Sr1−xBax[Mg2Al2N4]:Eu2+ (x = 0–1) Phosphors for Solid State and Agricultural Lighting Applications. ACS Appl. Mater. Interfaces 2020, 12, 23165–23171. [Google Scholar] [CrossRef]
  51. Lai, S.; Zhao, M.; Qiao, J.; Molokeev, M.S.; Xia, Z. Data-Driven Photoluminescence Tuning in Eu2+-Doped Phosphors. J. Phys. Chem. Lett. 2020, 11, 5680–5685. [Google Scholar] [CrossRef] [PubMed]
  52. Yang, X.; Zhang, Y.; Zhang, X.; Chen, J.; Huang, H.; Wang, D.; Chai, X.; Xie, G.; Molokeev, M.S.; Zhang, H.; et al. Facile synthesis of the desired red phosphor Li2Ca2Mg2Si2N6:Eu2+ for high CRI white LEDs and plant growth LED device. J. Am. Ceram. Soc. 2020, 103, 1773–1781. [Google Scholar] [CrossRef]
  53. Wu, W.; Zhang, Z.; Dong, R.; Xie, G.; Zhou, J.; Wu, K.; Zhang, H.; Cai, Q.; Lei, B. Characterization and properties of a Sr2Si5N8:Eu2+-based light-conversion agricultural film. J. Rare Earths 2020, 38, 539–545. [Google Scholar] [CrossRef]
  54. Adachi, S. Photoluminescence properties of Mn4+-activated oxide phosphors for use in white-LED applications: A review. J. Lumin. 2018, 202, 263–281. [Google Scholar] [CrossRef]
  55. Zhou, Z.; Zhou, N.; Xia, M.; Yokoyama, M.; Hintzen, H.T. (Bert) Research progress and application prospects of transition metal Mn4+-activated luminescent materials. J. Mater. Chem. C 2016, 4, 9143–9161. [Google Scholar] [CrossRef]
  56. Li, Y.; Qi, S.; Li, P.; Wang, Z. Research progress of Mn doped phosphors. RSC Adv. 2017, 7, 38318–38334. [Google Scholar] [CrossRef]
  57. Wang, Z.; Lin, H.; Zhang, D.; Shen, Y.; Li, Y.; Hong, R.; Tao, C.; Han, Z.; Chen, L.; Zhou, S. Deep-red emitting Mg2TiO4:Mn4+ phosphor ceramics for plant lighting. J. Adv. Ceram. 2021, 10, 88–97. [Google Scholar] [CrossRef]
  58. Tan, T.; Li, S.; Fan, Y.; Wang, Z.; Ali Raza, M.; Shafiq, I.; Wang, B.; Wu, X.; Yong, T.; Wang, X.; et al. Far-red light: A regulator of plant morphology and photosynthetic capacity. Crop J. 2022, 10, 300–309. [Google Scholar] [CrossRef]
  59. Park, Y.; Runkle, E.S. Far-red radiation promotes growth of seedlings by increasing leaf expansion and whole-plant net assimilation. Environ. Exp. Bot. 2017, 136, 41–49. [Google Scholar] [CrossRef]
  60. Demotes-Mainard, S.; Péron, T.; Corot, A.; Bertheloot, J.; Le Gourrierec, J.; Pelleschi-Travier, S.; Crespel, L.; Morel, P.; Huché-Thélier, L.; Boumaza, R.; et al. Plant responses to red and far-red lights, applications in horticulture. Environ. Exp. Bot. 2016, 121, 4–21. [Google Scholar] [CrossRef]
  61. Kalaitzoglou, P.; van Ieperen, W.; Harbinson, J.; van der Meer, M.; Martinakos, S.; Weerheim, K.; Nicole, C.C.S.; Marcelis, L.F.M. Effects of Continuous or End-of-Day Far-Red Light on Tomato Plant Growth, Morphology, Light Absorption, and Fruit Production. Front. Plant Sci. 2019, 10, 322. [Google Scholar] [CrossRef] [PubMed]
  62. Ding, F.; Zhou, Y.; He, Y.; Liang, Y.; Luo, P.; Zhou, W.; Zhang, J.; Yu, L.; Qiu, Z.; Lian, S. Broadband UV-Excitation and Red/Far-Red Emission Materials for Plant Growth: Tunable Spectrum Conversion in Eu3+,Mn4+ Co-doped LaAl0.7Ga0.3O3 Phosphors. Inorg. Chem. 2023, 62, 3141–3152. [Google Scholar] [CrossRef] [PubMed]
  63. Dang, P.; Wei, Y.; Liu, D.; Li, G.; Lin, J. Recent Advances in Chromium-Doped Near-Infrared Luminescent Materials: Fundamentals, Optimization Strategies, and Applications. Adv. Opt. Mater. 2023, 11, 2201739. [Google Scholar] [CrossRef]
  64. Wang, Y.; Wang, Z.; Wei, G.; Yang, Y.; He, S.; Li, J.; Shi, Y.; Li, R.; Zhang, J.; Li, P. Highly Efficient and Stable Near-Infrared Broadband Garnet Phosphor for Multifunctional Phosphor-Converted Light-Emitting Diodes. Adv. Opt. Mater. 2022, 10, 2200415. [Google Scholar] [CrossRef]
  65. Huyen, N.T.; Tu, N.; Tung, D.T.; Trung, D.Q.; Anh, D.D.; Duc, T.T.; Nga, T.T.T.; Huy, P.T. Photoluminescent properties of red-emitting phosphor BaMgAl10O17:Cr3+for plant growth LEDs. Opt. Mater. 2020, 108, 110207. [Google Scholar] [CrossRef]
  66. Zhang, A.; Liu, Y.; Liu, G.; Xia, Z. Dopant and Compositional Modulation Triggered Broadband and Tunable Near-Infrared Emission in Cs2Ag1−xNaxInCl6:Cr3+Nanocrystals. Chem. Mater. 2022, 34, 3006–3012. [Google Scholar] [CrossRef]
  67. Kong, L.; Wen, Y.; Jiao, X.; Liu, X.; Xu, Z. Interactive regulation of light quality and temperature on cherry tomato growth and photosynthesis. Environ. Exp. Bot. 2021, 182, 104326. [Google Scholar] [CrossRef]
  68. Li, C.-X.; Xu, Z.-G.; Dong, R.-Q.; Chang, S.-X.; Wang, L.-Z.; Khalil-Ur-Rehman, M.; Tao, J.-M. An RNA-Seq Analysis of Grape Plantlets Grown in vitro Reveals Different Responses to Blue, Green, Red LED Light, and White Fluorescent Light. Front. Plant Sci. 2017, 8, 78. [Google Scholar] [CrossRef]
  69. Zhao, M.; Zhang, Q.; Xia, Z. Structural Engineering of Eu2+-Doped Silicates Phosphors for LED Applications. Acc. Mater. Res. 2020, 1, 137–145. [Google Scholar] [CrossRef]
  70. Xia, Z.; Meijerink, A. Ce3+-Doped garnet phosphors: Composition modification, luminescence properties and applications. Chem. Soc. Rev. 2017, 46, 275–299. [Google Scholar] [CrossRef]
  71. Wang, X.; Zhao, Z.; Wu, Q.; Li, Y.; Wang, Y. Synthesis, structure and photoluminescence properties of Ca2LuHf2(AlO4)3:Ce3+, a novel garnet-based cyan light-emitting phosphor. J. Mater. Chem. C 2016, 4, 11396–11403. [Google Scholar] [CrossRef]
  72. Yang, Y.; Lin, Y.; Han, Y.; Qiu, Z.; Zhou, W.; Zhang, J.; Li, C.; Yu, L.; Lian, S. Fine controllable blue emission and its mechanism in Ce3+-doped orthosilicate solid solution phosphors for different plant growths. J. Rare Earths 2018, 36, 1150–1156. [Google Scholar] [CrossRef]
  73. Barceló-Muñoz, A.; Barceló-Muñoz, M.; Gago-Calderon, A. Effect of LED Lighting on Physical Environment and Microenvironment on In Vitro Plant Growth and Morphogenesis: The Need to Standardize Lighting Conditions and Their Description. Plants 2022, 11, 60. [Google Scholar] [CrossRef]
  74. Appolloni, E.; Paucek, I.; Pennisi, G.; Stringari, G.; Gabarrell Durany, X.; Orsini, F.; Gianquinto, G. Supplemental LED Lighting Improves Fruit Growth and Yield of Tomato Grown under the Sub-Optimal Lighting Condition of a Building Integrated Rooftop Greenhouse (i-RTG). Horticulturae 2022, 8, 771. [Google Scholar] [CrossRef]
  75. Seif, M.; Aliniaeifard, S.; Arab, M.; Mehrjerdi, M.Z.; Shomali, A.; Fanourakis, D.; Li, T.; Woltering, E. Monochromatic red light during plant growth decreases the size and improves the functionality of stomata in chrysanthemum. Funct. Plant Biol. 2021, 48, 515–528. [Google Scholar] [CrossRef] [PubMed]
  76. Gao, Q.; Liao, Q.; Li, Q.; Yang, Q.; Wang, F.; Li, J. Effects of LED Red and Blue Light Component on Growth and Photosynthetic Characteristics of Coriander in Plant Factory. Horticulturae 2022, 8, 1165. [Google Scholar] [CrossRef]
  77. Miao, Y.; Chen, Q.; Qu, M.; Gao, L.; Hou, L. Blue light alleviates ‘red light syndrome’ by regulating chloroplast ultrastructure, photosynthetic traits and nutrient accumulation in cucumber plants. Sci. Hortic. 2019, 257, 108680. [Google Scholar] [CrossRef]
  78. Lu, J.; Su, C.-C.; Hong, C.-S.; Peng, G.; Yang, C.-F. Investigations of Photoluminescence Properties of CaxMg2-xSi2O6:yEu2+ (x = 0.5–1.25, y = 0.015–0.035) Phosphors. Materials 2023, 16, 2032. [Google Scholar] [CrossRef]
  79. Guo, C.; Suo, H. Design of Single-Phased Multicolor-Emission Phosphor for LED. In Phosphors, Up Conversion Nano Particles, Quantum Dots and Their Applications: Volume 1; Liu, R.-S., Ed.; Springer: Berlin/Heidelberg, Germany, 2017; pp. 459–508. ISBN 978-3-662-52771-9. [Google Scholar]
  80. Xia, M.; Wu, X.; Zhong, Y.; Zhou, Z.; Wong, W.-Y. A novel Na3La(PO4)2/LaPO4:Eu blue-red dual-emitting phosphor with high thermal stability for plant growth lighting. J. Mater. Chem. C 2019, 7, 2385–2393. [Google Scholar] [CrossRef]
  81. Zhang, G.; Yin, Y.; Wang, Y.; Yu, S.; Zhang, J.; Wu, J.; Wang, T.; Zhao, L.; Li, Y.; Wang, W. Tunable blue-red dual emission via energy transfer in Na4CaSi3O9: Ce3+, Mn2+ phosphors for plant growth LED. J. Lumin. 2021, 235, 118029. [Google Scholar] [CrossRef]
  82. Gong, W.; Luo, J.; Zhou, W.; Fan, J.; Sun, Z.; Zeng, S.; Pan, H.; Zhu, Z.; Yang, X.; Yu, Z.; et al. Thermal-stable blue-red dual-emitting Na2Mg2Si6O15: Eu2+, Mn2+ phosphor for plant growth lighting. J. Lumin. 2021, 239, 118372. [Google Scholar] [CrossRef]
  83. Kang, Z.; Wang, S.; Seto, T.; Wang, Y. A Highly Efficient Eu2+ Excited Phosphor with Luminescence Tunable in Visible Range and Its Applications for Plant Growth. Adv. Opt. Mater. 2021, 9, 2101173. [Google Scholar] [CrossRef]
  84. Gong, D.; Zhu, D.; Huang, Z.; Wang, C.; Liao, M.; Wu, F.; Mu, Z. Towards high performance cyan phosphor by replacing F− with Cl− in Ba3Ca2(PO4)3(F/Cl): Eu2+. Opt. Mater. 2022, 134, 113123. [Google Scholar] [CrossRef]
  85. Liu, Y.; Zhang, X.; Hao, Z.; Wang, X.; Zhang, J. Generation of broadband emission by incorporating N3− into Ca3Sc2Si3O12: Ce3+ garnet for high rendering white LEDs. J. Mater. Chem. 2011, 21, 6354–6358. [Google Scholar] [CrossRef]
  86. Lang, T.; Cai, M.; Fang, S.; Han, T.; He, S.; Wang, Q.; Ge, G.; Wang, J.; Guo, C.; Peng, L.; et al. Trade-off Lattice Site Occupancy Engineering Strategy for Near-Infrared Phosphors with Ultrabroad and Tunable Emission. Adv. Opt. Mater. 2022, 10, 2101633. [Google Scholar] [CrossRef]
  87. Huang, S.; Shang, M.; Deng, M.; Yan, Y.; Dang, P.; Lin, J. Tunable concentration/excitation-dependent deep-red and white light emission in single-phase Eu2+-activated Sc-based oxide phosphors for blue/UV-LEDs. J. Mater. Chem. C 2022, 10, 14971–14981. [Google Scholar] [CrossRef]
  88. Li, J.; Li, P.; Wang, Y.; Shi, Y.; He, S.; Yang, Y.; Li, R.; Wei, G.; Suo, H.; Wang, Z. Two-stage ultra-broadband luminescence of Cr3+-doped multisite layered phosphor Sr3Ga2Ge4O14 and its application in pc-LEDs. Mater. Today Chem. 2022, 26, 101102. [Google Scholar] [CrossRef]
  89. Wang, H.; Wang, X.; Liu, S.; Mei, T.; Fan, Y.; Kang, L.; Xin, F.; Xing, M.; Fu, Y.; Luo, X.; et al. Color-tunable photoluminescence and persistent luminescence in a single LiYGeO4:Tb3+ phosphor. Mater. Res. Express 2022, 9, 076201. [Google Scholar] [CrossRef]
  90. Tan, T.; Wang, S.; Su, J.; Yuan, W.; Wu, H.; Pang, R.; Wang, J.; Li, C.; Zhang, H. Design of a Novel Near-Infrared Luminescence Material Li2Mg3TiO6:Cr3+with an Ultrawide Tuning Range Applied to Near-Infrared Light-Emitting Diodes. ACS Sustain. Chem. Eng. 2022, 10, 3839–3850. [Google Scholar] [CrossRef]
  91. He, Z.; Zhang, C.; Zhang, J.; Liu, S.; Sun, Y.; Chen, Q.; Chu, Z.; Ye, M.; Zhang, K. Concentration-dependent multi-color humic acid-based carbon dots for luminescent polymer composite films. J. Mater. Sci. 2022, 57, 1069–1083. [Google Scholar] [CrossRef]
  92. Armetta, F.; Sibeko, M.A.; Luyt, A.S.; Chillura Martino, D.F.; Spinella, A.; Saladino, M.L. Influence of the Ce:YAG Amount on Structure and Optical Properties of Ce:YAG-PMMA Composites for White LED. Z. Phys. Chem. 2016, 230, 1219–1231. [Google Scholar] [CrossRef]
  93. Liu, H.-G.; Xiao, F.; Zhang, W.-S.; Chung, Y.; Seo, H.-J.; Jang, K.; Lee, Y.-I. Influences of matrices and concentrations on luminescent characteristics of Eu(TTA)3(H2O)2/polymer composites. J. Lumin. 2005, 114, 187–196. [Google Scholar] [CrossRef]
  94. Zhao, S.; Liao, S.; Shi, R.; Zhang, J.; Han, Y.; Lian, S. Tuning emission color of Eu2+-activated phosphor through phase segregation. Chem. Eng. J. 2023, 452, 139640. [Google Scholar] [CrossRef]
  95. Wang, Q.; Wang, S.; Pang, R.; Tan, T.; Tan, T.; Wen, H.; Zhang, S.; You, H.; Li, C.; Zhang, H. Two-site occupation in Cr3+-activated BaIn2(P2O7)2 phosphor for broadband near-infrared thermometry and LED applications. Mater. Res. Bull. 2023, 163, 112222. [Google Scholar] [CrossRef]
  96. Ma, H.; Guan, L.; Yin, T.; Wu, Y.; Liu, Z.; Wang, D.; Wang, F.; Fu, N.; Li, X. Broadband emission phosphor Sr3Al2O5Cl2:Bi3+: Luminescence modulation and application for a white-light-emitting diode. Ceram. Int. 2022, 48, 33143–33150. [Google Scholar] [CrossRef]
  97. Chen, J.; Liu, Y.; Mei, L.; Wang, Z.; Fang, M.; Huang, Z. Emission red shift and energy transfer behavior of color-tunable KMg4(PO4)3:Eu2+,Mn2+ phosphors. J. Mater. Chem. C 2015, 3, 5516–5523. [Google Scholar] [CrossRef]
  98. Zou, X.; Zhang, H.; Li, W.; Zheng, M.; Molokeev, M.S.; Xia, Z.; Zheng, Y.; Li, Q.; Liu, Y.; Zhang, X.; et al. Ultra-Wide Vis–NIR Mg2Al4Si5O18:Eu2+,Cr3+Phosphor Containing Unusual NIR Luminescence Induced by Cr3+Occupying Tetrahedral Coordination for Hyperspectral Imaging. Adv. Opt. Mater. 2022, 10, 2200882. [Google Scholar] [CrossRef]
  99. Sun, Z.; Zhou, T.; Liu, R.; Tang, X.; Xie, R.-J. Ultrawide near-infrared SrHfO3:Cr3+phosphor with dual emission bands. J. Am. Ceram. Soc. 2023, 106, 3446–3454. [Google Scholar] [CrossRef]
  100. Wang, J.; Han, X.; Zhou, Y.; Wu, Z.; Liu, D.; Zeng, C.; Cao, S.; Zou, B. Ion Substitution Strategy toward High-Efficiency Near-Infrared Photoluminescence of Cs2KIn1−yAlyF6:Cr3+Solid Solutions. J. Phys. Chem. Lett. 2023, 14, 1371–1378. [Google Scholar] [CrossRef]
  101. Yan, Y.; Luo, C.; Ling, S.; Liang, J.; Liao, S.; Huang, Y. Enhancing quantum efficiency and thermal stability in Gd2SrAl2O7: Mn4+, Bi3+, Na+ far-red emitting phosphor by energy transfer and cation substitution strategy for indoor plant growth LED lighting. J. Alloys Compd. 2023, 947, 169609. [Google Scholar] [CrossRef]
  102. Li, L.; Chen, J.; Jiang, Y.; Xu, S.; Guo, H. Bi3+ ions doped double perovskite Ca2MgWO6 phosphor for yellow light emission. J. Lumin. 2023, 258, 119782. [Google Scholar] [CrossRef]
  103. Park, K.; Kim, H.; Gwon, S.Y.; Jung, G.W.; Kim, S.W. Photoluminescence improvement of Lu3Al5O12:Ce3+ phosphors by controlling synthesizing temperature and adding fluoride fluxes. Ceram. Int. 2022, 48, 33848–33860. [Google Scholar] [CrossRef]
  104. Zhang, X.; Dong, L.; Zhang, L.; Xu, Y.; Wu, X.; Yin, S.; Zhong, C.; You, H. High thermal stability phosphors with a rigid structure similar to the benzene ring and application in plant growth. J. Mater. Chem. C 2022, 10, 16857–16864. [Google Scholar] [CrossRef]
  105. Bai, Y.; Jia, Z.; Gao, J.; Wu, L.; Kong, Y.; Zhang, Y.; Xu, J. A novel red-emitting phosphor K2MgGeO4:Eu3+ for WLEDs: Zero-thermal quenching induced by heterovalent substitution. J. Mater. Chem. C 2022, 10, 15957–15966. [Google Scholar] [CrossRef]
Figure 1. Absorption spectra of plant pigments, including: chlorophyll a, chlorophyll b, photosensitive pigment PR, and photosensitive pigment PFR.
Figure 1. Absorption spectra of plant pigments, including: chlorophyll a, chlorophyll b, photosensitive pigment PR, and photosensitive pigment PFR.
Nanomaterials 13 01715 g001
Figure 2. Schematic diagram of the pc-LED structure.
Figure 2. Schematic diagram of the pc-LED structure.
Nanomaterials 13 01715 g002
Figure 3. Schematic diagram of the energy level of Eu3+.
Figure 3. Schematic diagram of the energy level of Eu3+.
Nanomaterials 13 01715 g003
Figure 4. (a) Photoluminescence (PL) spectra for Ca9LiY0.667(PO4)7: xEu3+. (b) Comparison of PL spectrum of CNYPO:0.12Eu3+ and the absorption spectra of phytochrome PR and PFR. (Reprinted from ref. [43]. Copyright of Elsevier, 2020). (c) Photoluminescent excitation (PLE) (under λem = 615 nm) of Li3BaSrLa3−xEux(MoO4)8 where x = 0–3 for selected compositions. (d) The spectral overlap between the red LED and phytochrome (PR) absorption. (Reprinted from ref. [44]. Copyright of Royal Society of Chemistry, 2021).
Figure 4. (a) Photoluminescence (PL) spectra for Ca9LiY0.667(PO4)7: xEu3+. (b) Comparison of PL spectrum of CNYPO:0.12Eu3+ and the absorption spectra of phytochrome PR and PFR. (Reprinted from ref. [43]. Copyright of Elsevier, 2020). (c) Photoluminescent excitation (PLE) (under λem = 615 nm) of Li3BaSrLa3−xEux(MoO4)8 where x = 0–3 for selected compositions. (d) The spectral overlap between the red LED and phytochrome (PR) absorption. (Reprinted from ref. [44]. Copyright of Royal Society of Chemistry, 2021).
Nanomaterials 13 01715 g004
Figure 5. Photographs of agriculture films painted with or without wavelength-converting materials under daylight (a) and ultraviolet irradiation (b). (Reprinted from ref. [46]. Copyright of Springer Nature, 2022).
Figure 5. Photographs of agriculture films painted with or without wavelength-converting materials under daylight (a) and ultraviolet irradiation (b). (Reprinted from ref. [46]. Copyright of Springer Nature, 2022).
Nanomaterials 13 01715 g005
Figure 6. (a) Normalized PL spectra of as-prepared (RbxK1−x)3LuSi2O7:0.01Eu2+ (0 ≤ x ≤ 1). (Reprinted from ref. [51]. Copyright of American Chemical Society, 2020). (b) Absorption spectra of Chlorophyll (A and B) and the emission spectrum of Li2Ca2Mg2Si2N6:1.0%Eu2+-based plant-growth LED device. (c) Indoor pak-choi cultivation irradiated and digital photographs of pak-choi cultivation irradiated by different lighting sources. (Reprinted from ref. [52]. Copyright of John Wiley and Sons, 2019). (d) Biomass of Chinese flowering cabbages obtained under the two films. (Reprinted from ref. [53]. Copyright of Elsevier, 2020).
Figure 6. (a) Normalized PL spectra of as-prepared (RbxK1−x)3LuSi2O7:0.01Eu2+ (0 ≤ x ≤ 1). (Reprinted from ref. [51]. Copyright of American Chemical Society, 2020). (b) Absorption spectra of Chlorophyll (A and B) and the emission spectrum of Li2Ca2Mg2Si2N6:1.0%Eu2+-based plant-growth LED device. (c) Indoor pak-choi cultivation irradiated and digital photographs of pak-choi cultivation irradiated by different lighting sources. (Reprinted from ref. [52]. Copyright of John Wiley and Sons, 2019). (d) Biomass of Chinese flowering cabbages obtained under the two films. (Reprinted from ref. [53]. Copyright of Elsevier, 2020).
Nanomaterials 13 01715 g006
Figure 7. (a) Crystal structure of BZGO. (b) PL spectra upon excitation by 312 nm and PLE spectra of BZGO:xMn4+ (x = 0.5–4%) when monitoring the emission at 666 nm. (Reprinted from ref. [22]. Copyright of Elsevier, 2023). (c) Change of optical density of chlorella in tube in a week and photograph of chlorella tube and nearby film (Sr4Al14O25: Mn, Mg, Ga film and blank). (Reprinted from ref. [21]. Copyright of John Wiley and Sons, 2022).
Figure 7. (a) Crystal structure of BZGO. (b) PL spectra upon excitation by 312 nm and PLE spectra of BZGO:xMn4+ (x = 0.5–4%) when monitoring the emission at 666 nm. (Reprinted from ref. [22]. Copyright of Elsevier, 2023). (c) Change of optical density of chlorella in tube in a week and photograph of chlorella tube and nearby film (Sr4Al14O25: Mn, Mg, Ga film and blank). (Reprinted from ref. [21]. Copyright of John Wiley and Sons, 2022).
Nanomaterials 13 01715 g007
Figure 9. (a) Crystal structure of Ba3GdNa(PO4)3F and three kinds of Ba2+ ions with a different coordination environment. (b) PL and PLE spectra of the phosphor BGNPF:0.9%Eu2+ along with the absorption spectra of chlorophyll A, B and carotenoids. (Reprinted from ref. [3]. Copyright of American Chemical Society, 2016). (c) Emission (λex = 266 nm) spectra of the Na3La2(PO4)3:xEu phosphors, the inset shows the peak changes at 422 nm and 594 nm. (Reprinted from ref. [80]. Copyright of Royal Society of Chemistry, 2019). (d) Normalized PL and PLE spectra of NCSO:0.01Ce3+, Mn2+ and absorption spectra of chlorophyll A and chlorophyll B. (Reprinted from ref. [81]. Copyright of Elsevier, 2021). (e) The emission spectra of Na2-xMg2-ySi6O15:xEu2+, yMn2+ (x = 0.02, y = 0–0.80) upon 365 nm excitation. (Reprinted from ref. [82]. Copyright of Elsevier, 2021).
Figure 9. (a) Crystal structure of Ba3GdNa(PO4)3F and three kinds of Ba2+ ions with a different coordination environment. (b) PL and PLE spectra of the phosphor BGNPF:0.9%Eu2+ along with the absorption spectra of chlorophyll A, B and carotenoids. (Reprinted from ref. [3]. Copyright of American Chemical Society, 2016). (c) Emission (λex = 266 nm) spectra of the Na3La2(PO4)3:xEu phosphors, the inset shows the peak changes at 422 nm and 594 nm. (Reprinted from ref. [80]. Copyright of Royal Society of Chemistry, 2019). (d) Normalized PL and PLE spectra of NCSO:0.01Ce3+, Mn2+ and absorption spectra of chlorophyll A and chlorophyll B. (Reprinted from ref. [81]. Copyright of Elsevier, 2021). (e) The emission spectra of Na2-xMg2-ySi6O15:xEu2+, yMn2+ (x = 0.02, y = 0–0.80) upon 365 nm excitation. (Reprinted from ref. [82]. Copyright of Elsevier, 2021).
Nanomaterials 13 01715 g009
Figure 11. (a) Crystal structure of the BIP sample. (Reprinted from ref. [95]. Copyright of Elsevier, 2023). (b) The emission spectra of KMg4(PO4)3:0.02Eu2+, xMn2+ (x = 0–0.35) phosphors. (inset) The normalized emission spectra for KMg4(PO4)3:0.02Eu2+, xMn2+ (x = 0–0.35) phosphors. (Reprinted from ref. [97]. Copyright of Royal Society of Chemistry, 2022). (c) DR spectrum of Mg2Al4Si5O18:0.01Cr3+, PL and PLE spectra of Mg2Al4Si5O18: Eu2+. (Reprinted from ref. [98]. Copyright of John Wiley and Sons, 2022). (d) Emission and excitation spectra of SHO:0.005Cr. (Reprinted from ref. [99]. Copyright of John Wiley and Sons, 2023).
Figure 11. (a) Crystal structure of the BIP sample. (Reprinted from ref. [95]. Copyright of Elsevier, 2023). (b) The emission spectra of KMg4(PO4)3:0.02Eu2+, xMn2+ (x = 0–0.35) phosphors. (inset) The normalized emission spectra for KMg4(PO4)3:0.02Eu2+, xMn2+ (x = 0–0.35) phosphors. (Reprinted from ref. [97]. Copyright of Royal Society of Chemistry, 2022). (c) DR spectrum of Mg2Al4Si5O18:0.01Cr3+, PL and PLE spectra of Mg2Al4Si5O18: Eu2+. (Reprinted from ref. [98]. Copyright of John Wiley and Sons, 2022). (d) Emission and excitation spectra of SHO:0.005Cr. (Reprinted from ref. [99]. Copyright of John Wiley and Sons, 2023).
Nanomaterials 13 01715 g011
Figure 12. (a) XRD patterns of Ca2MgWO6:1%Bi3+ samples prepared with x MgF2 (x = 0, 1, 2 and 3 wt%) flux and standard card (JCPDS No. 48-0108) for Ca2MgWO6. (Reprinted from ref. [102]. Copyright of Elsevier, 2023). (b) Crystal structure of LiGaGeO4 and LiAlSiO4. (c,d) Temperature-dependent PL spectra of LiGaGeO4:0.002Cr3+ and LiAlSiO4:0.02Cr3+. (Reprinted from ref. [104]. Copyright of Royal Society of Chemistry, 2022). (e) The relationship between luminescence intensity and temperature. (Reprinted from ref. [105]. Copyright of Royal Society of Chemistry, 2022).
Figure 12. (a) XRD patterns of Ca2MgWO6:1%Bi3+ samples prepared with x MgF2 (x = 0, 1, 2 and 3 wt%) flux and standard card (JCPDS No. 48-0108) for Ca2MgWO6. (Reprinted from ref. [102]. Copyright of Elsevier, 2023). (b) Crystal structure of LiGaGeO4 and LiAlSiO4. (c,d) Temperature-dependent PL spectra of LiGaGeO4:0.002Cr3+ and LiAlSiO4:0.02Cr3+. (Reprinted from ref. [104]. Copyright of Royal Society of Chemistry, 2022). (e) The relationship between luminescence intensity and temperature. (Reprinted from ref. [105]. Copyright of Royal Society of Chemistry, 2022).
Nanomaterials 13 01715 g012
Table 1. Excitation and emission wavelengths of phosphors for LEDs on indoor plant cultivation.
Table 1. Excitation and emission wavelengths of phosphors for LEDs on indoor plant cultivation.
Phosphorλex/nmλem/nmEmission ColorRefs.
Ca3Al2Ge3O12: Eu3+393707Red[5]
KBa2(PO3)5: Mn4+438660Red[26]
Gd3Al4GaO12: Cr3+360734Red[27]
Sr2ScSbO6: Mn4+310700Red[28]
SrLa2Al2O7: Mn4+365731Red[29]
Ca2BO3Cl: Bi3+486732Red[30]
Li2ZnTi3O8: Cr3+360/468735Red[31]
BaLaLiWO6: Mn4+332708Red[32]
Li3BaSrY3(WO4)8: Eu3+394615Red[33]
Ca0.995Mg2(SO4)3: Eu2+554635Red[34]
Ca1.1Sr0.9SiO4: Ce3+365425Blue[4]
La3SbO7: Bi3+315520Blue[35]
Na2BaSr(PO4)2: Eu2+325428Blue[36]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yang, C.; Liu, W.; You, Q.; Zhao, X.; Liu, S.; Xue, L.; Sun, J.; Jiang, X. Recent Advances in Light-Conversion Phosphors for Plant Growth and Strategies for the Modulation of Photoluminescence Properties. Nanomaterials 2023, 13, 1715. https://doi.org/10.3390/nano13111715

AMA Style

Yang C, Liu W, You Q, Zhao X, Liu S, Xue L, Sun J, Jiang X. Recent Advances in Light-Conversion Phosphors for Plant Growth and Strategies for the Modulation of Photoluminescence Properties. Nanomaterials. 2023; 13(11):1715. https://doi.org/10.3390/nano13111715

Chicago/Turabian Style

Yang, Chengxiang, Wei Liu, Qi You, Xiuxian Zhao, Shanshan Liu, Liang Xue, Junhua Sun, and Xuchuan Jiang. 2023. "Recent Advances in Light-Conversion Phosphors for Plant Growth and Strategies for the Modulation of Photoluminescence Properties" Nanomaterials 13, no. 11: 1715. https://doi.org/10.3390/nano13111715

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop