Home Polymer nanocomposite sunlight spectrum down-converters made by open-air PLD
Article Open Access

Polymer nanocomposite sunlight spectrum down-converters made by open-air PLD

  • Abdalla M. Darwish EMAIL logo , Sergey S. Sarkisov EMAIL logo , Simeon Wilson , Jamaya Wilson , Eboni Collins , Darayas N. Patel , Kyu Cho , Anit Giri , Lynn Koplitz , Brent Koplitz and David Hui
Published/Copyright: October 30, 2020
Become an author with De Gruyter Brill

Abstract

We report, for the first time to our knowledge, on the polymer nanocomposite sunlight spectrum down-converters made by the concurrent multi-beam multi-target pulsed laser deposition (CMBMT-PLD) of phosphor and polymer in ambient air. Phosphor PLD targets were made of down-converting rare-earth (RE)-doped fluorides NaYF4:Yb3+,Er3+, and NaYF4:Yb3+,Tm3+ with a Stokes shift of 620 nm (from 360 to 980 nm), minimizing the effect of re-absorption. The phosphors were synthesized by the wet method. Polymer target was made of poly (methyl methacrylate) known as PMMA. Target ablation was conducted with 1,064 nm beams from an Nd:YAG Q-switched laser. Beam intensity was 2.8 × 1016 W/cm2 for both targets. The substrate was a microscope glass slide. Phosphor nanoparticles with a size ranging from 10 to 50 nm were evenly distributed in the polymer matrix during deposition. The nanoparticles retained the crystalline structure and the fluorescent properties of the phosphor target. There was no noticeable chemical decomposition of the deposited polymer. The products of laser-induced reaction of the polymer target with atmospheric gases did not reach the substrate during PLD. Post-heating of the substrate at ∼90°C led to fusion of separate polymer droplets into uniform coating. Quantum yield of the down-conversion polymer nanocomposite film was estimated to be not less than ∼5%. The proposed deposition method can find its application in making commercial-size down-converter coatings for photo-voltaic solar power applications.

1 Introduction

There is growing interest in adoption of green-energy technologies that rely heavily on solar power. Alongside the importance of electrical performance come features such as aesthetics, convenience, and safety. These concerns have motivated those working on luminescent solar concentrators (LSCs) for use in buildings [1]. LSCs are large windows filled/coated with luminescent materials that absorb UV/blue spectral components of sunlight and re-emit at near-infrared (NIR) wavelengths. The window acts as a lightguide that redirects the re-emitted radiation to small edge-attached photo-voltaic (PV) cells to generate electricity (Figure 1a). Because LSC area exposed to sunlight is much larger than the edges, LSCs can greatly increase the flux of radiation incident onto the perimeter PV devices and increase the photocurrent, acting as a solar concentrator. In building-integrated applications, LSCs can be introduced without compromising a building’s aesthetic value. Down-converting films can also be deposited directly over PV cells to convert UV/blue sunlight into NIR radiation and thus improve the efficiency [2] (Figure 1b).

Figure 1 
               (a) The concept of LSC. The nanoparticles of a luminescent material in the nanocomposite coating absorb incident UV/blue sunlight (blue arrows; photons with energy hν
                  1, where h is Planck’s constant and ν
                  1 is the frequency of UV/blue radiation) and re-emit at an NIR wavelength (red arrows; photons with energy hν
                  2, where ν
                  2 is the frequency of NIR radiation) in glass window guiding it to a PV cell attached to the edge. (b) The concept of sunlight spectrum down-convertor over a PV cell. A down-converting film over a conventional PV cell absorbs incident UV/blue sunlight and down-converts it into NIR radiation that efficiently generates extra PV power in addition to the one produced by visible and NIR spectral components of sunlight.
Figure 1

(a) The concept of LSC. The nanoparticles of a luminescent material in the nanocomposite coating absorb incident UV/blue sunlight (blue arrows; photons with energy 1, where h is Planck’s constant and ν 1 is the frequency of UV/blue radiation) and re-emit at an NIR wavelength (red arrows; photons with energy 2, where ν 2 is the frequency of NIR radiation) in glass window guiding it to a PV cell attached to the edge. (b) The concept of sunlight spectrum down-convertor over a PV cell. A down-converting film over a conventional PV cell absorbs incident UV/blue sunlight and down-converts it into NIR radiation that efficiently generates extra PV power in addition to the one produced by visible and NIR spectral components of sunlight.

Existing solar spectrum down-converters have numerous problems, such as limited spectral coverage, re-absorption, and toxicity [3,4,5]. The search for better down-converting materials thus continues. One possible alternative is polymer nanocomposites filled with the nanoparticles of down-converting compounds of rare-earth (RE) elements. RE compounds have recently found many applications ranging from up-conversion biomarkers for bio-imaging, agents for cancer diagnostics and treatment, additives in solar power materials, to various sensors [6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25]. The advantages of RE-filled polymer nanocomposites as spectral down-converters are three-fold: their RE-based phosphor constituents can have high down-converting efficiency [26,27,28,29,30], they are environment friendly, and the polymer matrix offers smooth coating/form shaping.

Currently, polymer nanocomposite down-converting films are made by applying phosphor nanocolloids in polymer solutions over LSC panels or PV cells using spin coating [2], spraying, or dip coating [26] followed by drying off the solvent. The major deficiencies include limited concentration of phosphor nanoparticles in the colloids, poor control of film thickness, and possible clustering of the nanoparticles in the initial colloids during storage or when they are drying after deposition. The rationale for this work was to investigate an alternative, new approach for making polymer nanocomposite down-converting films – the concurrent multi-beam multi-target pulsed laser deposition (CMBMT-PLD) in open air. PLD, traditionally conducted in a vacuum chamber, uses a beam of a high-energy pulsed laser to ablate a target and transfer target material on a substrate. The thickness of the deposited film can be maintained with a sub-nanometre precision. Multi-beam multi-target PLD makes films composed of the materials of several targets, which could have quite different physical and chemical properties. The constituents can be combined at any desirable proportion. Accordingly, a polymer nanocomposite film can be loaded with a greater amount of down-conversion phosphor than in the case of traditional techniques of film making. The concentration of the phosphor in the polymer matrix can also be precisely controlled in the direction normal to the film as well as along the film’s plane. Concurrent deposition provides more uniform mixing of the phosphor particles in the polymer matrix as compared to non-concurrent (sequential) PLD. However, the vacuum chamber puts severe restrictions on the size of the substrates that can be coated and negates the benefits of conventional PLD in vacuum. Open-air PLD lacks this restriction. It brings the capability to coat large-size substrates of complex shapes, flexibility, and low cost. The major limitations of open-air PLD are possible degradation of the deposited material due to laser-initiated chemical reaction with atmospheric oxygen and nitrogen and strong scattering of the ejected target matter on its way toward the substrate due to collisions with air molecules. The authors, using their experience with CMBMT-PLD [31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47], close the gap of knowledge on the properties of polymer nanocomposite films deposited in open air and show that, despite limitations, they can potentially function as solar spectrum down-converters.

Section 2 of the article describes deposition setup, preparation of RE-based down-conversion phosphors and phosphor PLD targets, and deposition process. In Section 3, based on optical spectroscopy data, the solar spectrum down-conversion mechanisms in the RE phosphors are discussed. The section also contains the results and discussion of the morphology and chemical integrity of pure polymer films made by open-air PLD. The structure and optical properties of the polymer–phosphor composite films are also discussed. Finally, in Section 4, the conclusions are made on feasibility of the proposed approach of making solar spectrum down-converting coatings.

2 Materials and methods

2.1 Deposition setup

The experimental setup for open-air CMBMT-PLD is schematically presented in Figure 2. Two high-energy focused laser beams from a pulsed laser concurrently ablate polymer and phosphor targets. The targets might rotate to avoid rapid erosion and cracking due to ablation in single spot. The plumes of the ablated material propagate toward the substrate and condensate on it. Since the plumes intersect, the phosphor and the polymer mix in a composite film formed on the substrate. The substrate is moving in X–Y directions to coat over a broad area. The substrate can also be heated to fuse separate melted polymer droplets in uniform film and smoothen its surface by capillary forces.

Figure 2 
                  Schematic of open-air co-deposition of phosphor and polymer using CMBMT-PLD. Parameter t is the distance between the spot of target ablation and the substrate.
Figure 2

Schematic of open-air co-deposition of phosphor and polymer using CMBMT-PLD. Parameter t is the distance between the spot of target ablation and the substrate.

2.2 Targets

The inorganic targets were made of two RE-doped fluorides: NaYF4:Yb3+,Er3+ and NaYF4:Yb3+,Tm3+. Justification of using groups of RE ions (Yb3+, Er3+) and (Yb3+, Tm3+) for efficient down-conversion is given, for instance, in ref. [48]. It will also be briefly discussed in the next section. The selection of fluoride NaYF4 as a host for RE ions is justified by its low phonon energy (∼300 cm−1) that makes multi-phonon-assisted non-radiative relaxation of excited ions negligible [49]. Powders of the phosphors were synthesized using the economical wet process, which in its essence is co-precipitation in the presence of Na2-ethylenediaminetetraacetic acid (EDTA). In a typical procedure, 2.1 g of NaF (0.05 mol) was dissolved in 60 mL of deionized water. Another solution was prepared by mixing together a × 16 mL of 0.2 mol/L aqueous solution of YCl3, b × 16 mL of 0.2 mol/L solution of YbCl3, c × 16 mL (c = 1 (a + b)) of 0.2 mol/L solution of XCl3 (X stands for Er or Tm), and 20 mL of 0.2 mol/L EDTA aqueous stock solution to form the metal–EDTA complex. All the chemicals were acquired from Sigma-Aldrich. The complex solution was injected in the NaF solution quickly, and the mixture was stirred vigorously for 1 h at room temperature. After stirring, the mixture was allowed to stay overnight for the precipitate to settle. The precipitate was collected, washed several times with distilled water and anhydrous ethanol. In the second stage of the process, the precipitate was dried in open air for 48 h at 60°C to remove traces of water. The resulting powder had the doping rate according to the formula NaY a F4:Yb3+ b ,X 3+ c . Four samples of phosphor powders were made. There were two samples of NaYF4:Yb3+,Er3+ type: NaY0.94F4:Yb3+ 0.03,Er3+ 0.03 and NaY0.78F4:Yb3+ 0.2,Er3+ 0.02; and two samples of NaYF4:Yb3+,Tm3+ type: NaY0.88F4:Yb3+ 0.1,Tm3+ 0.02 and NaY0.78F4:Yb3+ 0.2, Tm3+ 0.02. The freshly made crystalline powder had NaYF4 host in the cubic α-phase. It was converted into more efficient hexagonal β-phase using heat treatment in an open-air furnace at a temperature in the range between 400 and 600°C for 1 h. These temperature range and duration of heating were found to be optimal for crystalline phase change in ref. [50]. After that, the phosphor powder was reduced by ball-milling to nano-powder. Finally, the nano-powder was compressed in a solid PLD target using a 25-T manual hydraulic press.

The polymer target was cut from an acrylic microscope slide (UVT-UV Transmittable-pure poly(methyl methacrylate) or PMMA from Ted Pella, Inc.; Catalog # 260226).

2.3 Deposition

During open-air CMBMT-PLD both targets (Figure 2) were ablated with focused beams from a Continuum Surelite SL III-10 Q-switched Nd:YAG laser. Laser pulses had an energy of 0.45 J for each target at a wavelength of 1,064 nm; pulse duration ∼5 ns. Laser beams were focused on the targets with convex lenses (50 mm focal length). Peak beam intensity on the target of ∼2.8 × 1016 W/cm2 was typical for open-air PLD [51]. Corresponding fluence of ∼1.38 × 108 J/cm2 exceeded by many orders of magnitude threshold fluence of PLD of target materials (between 0.4 and 1.25 J/cm2 [34,52]). This resulted in a deposition rate of ∼12 nm/pulse, which was three orders of magnitude faster than for typical vacuum PLD [34]. Ablation of PMMA target at the same fluence as phosphor leads to a higher polymer deposition rate and mixing phosphor particles in the polymer at a proportion of ∼0.024% by weight. The substrate was a glass microscope slide (Ted Pella Catalog #260441). Distance between the spot of target ablation and the substrate t (Figure 2) was ∼2 mm. Incidence angles of the laser beams to the targets were in the range 80–86°. Time of ablation was 5 min. With a pulse repetition rate of 10 Hz, PLD took ∼3,000 laser pulses.

3 Results and discussion

3.1 Down-conversion

3.1.1 Down-conversion mechanisms

The down-conversion mechanisms in the synthesized phosphors NaYF4:Yb3+,Er3+ and NaYF4:Yb3+,Tm3+ can be illustrated by energy-level diagrams of the pairs of coupled RE ions (Er3+, Yb3+) and (Tm3+, Yb3+) proposed in refs. [30,53]. They are presented in Figure 3a and b, respectively. The mechanisms correspond to the concentrations of RE ions and possible impurities and defects like the ones occurring in this work. Ion of Er3+ (Figure 3a) acts as an energy donor and that of Yb3+ as an acceptor. Wavy horizontal arrow marked “Sunlight” indicates optical excitation of Er3+ with UV photons of sunlight at a wavelength of ∼380 nm. The reason to select (Er3+–Yb3+) pair is that, for efficient down-conversion, a donor should have an energy level above 20,000 cm−1 along with an intermediate energy level at approximately 10,000 cm−1, whereas the acceptor should have a single excited-state 2F5/2 above 2F7/2 ground-state corresponding to NIR emission in the range 9,100–11,100 cm−1 (900–1,100 nm wavelength). During exposure to sunlight, the absorbed UV photon excites Er3+ ion (grey arrow up in Figure 3a). Excited Er3+ ion relaxes radiationlessly from level 4G11/2 down to 4F7/2. After that, two cross-relaxation processes (marked by sloping dashed arrows) occur. Er3+ ion returns to the intermediate level 4I11/2 from the 4F7/2 level (dark grey dashed arrow down), resulting in the excitation of neighbouring Yb3+ ion from ground-state 2F7/2 to level 2F5/2 as shown by dashed arrow up. Then Er3+ ion returns to ground-level 4I15/2 from intermediate-level 4I11/2 and its energy is transferred to the second Yb3+ ion. This overall sequence involves a single UV photon and produces two NIR photons. In the case of (Tm3+–Yb3+) pair (Figure 3b), Tm3+ is excited by a blue sunlight photon. The energy of Tm3+ transition 1G43H6 is approximately twice as high as the energy difference between 2F7/2 and 2F5/2 levels of Yb3+. This fact makes the cooperative down-conversion possible. Corresponding mechanism (1) thus results in two NIR photons emitted by Yb3+ ions. Since tetravalent Tm does not exist, the cooperative down-conversion from Tm3+ to Yb3+ would not be influenced by the “metal-to-metal” charge-transfer state. In addition, the low phonon energy of NaYF4 (∼300 cm−1) together with the large energy gap between Tm3+:1G4 and Yb3+:2F5/2 makes the Tm3+:1G4 → Yb3+:2F5/2 multi-phonon-assisted nonradiative relaxation negligible. There is also a cross-relaxation energy-transfer mechanism (2) between transition Tm3+:1G43H5 and Yb3+:2F7/22F5/2.

Figure 3 
                     Energy-level diagram explaining the mechanisms of down-conversion in groups of ions (a) Er3+ and Yb3+ [30] and (b) Tm3+ and Yb3+ [53].
Figure 3

Energy-level diagram explaining the mechanisms of down-conversion in groups of ions (a) Er3+ and Yb3+ [30] and (b) Tm3+ and Yb3+ [53].

3.1.2 Optical absorption

Diffuse reflectance spectra of the phosphor powders (Figure 4) were taken with a Shimadzu UV-2600-ISR-2600 Plus UV-VIS-NIR spectrophotometer with integrating sphere (measurable spectral range 220–1,400 nm). Dynamic light-scattering measurement of the powder suspensions in water with a Malvern Zetasizer revealed that the dominating size of the powder particles was between 5 and 10 µm. The samples were prepared by compressing the micro-powders into pellets with a 25-T manual hydraulic press. Following [54], the absorption peaks in the diffuse reflectance spectra were related to some transitions in pairs of ions (Er3+, Yb3+) (curve A) and (Tm3+, Yb3+) (curve B) as presented in Figure 4 caption. The strongest peaks (6 and 6′) correspond to the optical absorbance of the ion of Yb3+ due to transition 2F7/22F5/2 (∼980 nm). The ion was responsible for down-conversion NIR emission when relaxing back from excited to ground state. The strength of peaks 6 and 6′ is similar in magnitude. The absorption peaks 1 through 5, attributed to Er3+, are less strong than peak 6, but still significant (curve A). But the absorption peaks that could be attributed to Tm3+ are not noticeable as compared to peak 6′ in curve B. This can be explained by different concentrations of Yb3+. In phosphor NaY0.94F4:Yb3+ 0.03,Er3+ 0.03 (curve A), the concentration of Yb3+ is equal to that of Er3+, while in NaY0.78F4:Yb3+ 0.2,Tm3+ 0.02 the concentration of Yb3+ is ten times greater than that of Tm3+. This correlates with findings in ref. [54].

Figure 4 
                     Diffuse reflectance spectra of the powder samples of down-conversion RE phosphors. Curve A corresponds to phosphor NaY0.94F4:Yb3+
                        0.03,Er3+
                        0.03 baked after synthesis for 1 h at 500°C; B – NaY0.78F4:Yb3+
                        0.2,Tm3+
                        0.02 baked for 1 h at 500°C. Absorption peak 1 corresponds to transition in ion of Er3+:4I15/2 → 4G11/2 (380 nm); 2 – Er3+:4I15/2 → 4F7/2 (488 nm); 3 – Er3+:4I15/2 → 2H11/2 (520 nm); 4 – Er3+:4I15/
                        2 → 4S3/2 (540 nm); 5 – Er3+:4I15/2 → 4F9/2 (653 nm); and 6, 6′ – Yb3+:2F7/2 → 2F5/2 (980 nm).
Figure 4

Diffuse reflectance spectra of the powder samples of down-conversion RE phosphors. Curve A corresponds to phosphor NaY0.94F4:Yb3+ 0.03,Er3+ 0.03 baked after synthesis for 1 h at 500°C; B – NaY0.78F4:Yb3+ 0.2,Tm3+ 0.02 baked for 1 h at 500°C. Absorption peak 1 corresponds to transition in ion of Er3+:4I15/24G11/2 (380 nm); 2 – Er3+:4I15/24F7/2 (488 nm); 3 – Er3+:4I15/22H11/2 (520 nm); 4 – Er3+:4I15/ 24S3/2 (540 nm); 5 – Er3+:4I15/24F9/2 (653nm); and 6, 6′ – Yb3+:2F7/22F5/2 (980 nm).

3.1.3 Light sources simulating sunlight

Down-conversion luminescence spectroscopy of the phosphors was conducted using three light sources imitating UV-blue components of sunlight: (a) Insect Killer Bulb (40 W UV black light; fits UVB45 and UV40 units; intensity in close vicinity of the bulb ∼8.8 mW/cm2; spectral peak ∼369 nm; FWHM ∼20 nm; emission spectrum is shown in Figure 5a); (b) UV301 LED flashlight from Shenzhen Lightfe Lighting Co., Ltd, China (intensity ∼25.6 mW/cm2; peak of the spectrum is at 368 nm; FWHM ∼ 20 nm; Figure 5b), and (c) Ar-ion laser. The first UV source produced ∼1.76 times radiation of UV spectral component of sunlight at the air mass coefficient AM1.5 [55]. The second source produced 5.12 times radiation. Maximum power of 80 mW of 488-nm spectral line of the Ar-ion laser was ∼1.6 times the visible sunlight at AM1.5. Sharp peaks in the spectrum of the Insect Killer Bulb (Figure 5a) at 365.4 and 404.7 nm belonged to the mercury-vapour light leaking through the UV fluorescent coating of the bulb.

Figure 5 
                     Emission spectra of UV sources used to simulate UV component of sunlight in the experiments with down-conversion materials: (a) Insect killer UV black light (intensity ∼ 8.8 mW/cm2). (b) UV LED flashlight (intensity ∼ 25.6 mW/cm2).
Figure 5

Emission spectra of UV sources used to simulate UV component of sunlight in the experiments with down-conversion materials: (a) Insect killer UV black light (intensity ∼ 8.8 mW/cm2). (b) UV LED flashlight (intensity ∼ 25.6 mW/cm2).

3.1.4 Down-conversion and down-shifting emission

All the down-conversion spectra presented below were taken with a spectrometer SP2500i from Teledyne Princeton Instruments equipped with uncooled InGaAs photodetector ID-441. Figure 6a presents down-conversion emission spectra of phosphors (1) NaY0.94F4:Yb3+ 0.03,Er3+ 0.03 and (2) NaY0.78F4:Yb3+ 0.2,Er3+ 0.02 pumped with UV LED flashlight (Figure 5b). Apparently, the NIR emission from phosphor (2) with seven-fold greater concentration of Yb3+ ions (its diffusive reflectance spectrum was presented by curve A in Figure 4) was ∼2.5 times stronger than that of phosphor (1). Besides NIR, the phosphors had a visible down-shifting emission. The spectrum of such an emission was taken with the Princeton SP2500i spectrometer equipped with photo-multiplying tube R928. The spectrum (presented for phosphor 2 in Figure 6a) has spectral peaks corresponding to the absorption peaks in Figure 4 (curve A) that were attributed to various energy transitions in the diagram in Figure 3a. These visible emission spectral peaks have green (520 and 540 nm) and red (653 nm) colours.

Figure 6 
                     (a) Down-conversion emission spectra of phosphors (1) NaY0.94F4:Yb3+
                        0.03,Er3+
                        0.03 and (2) NaY0.78F4:Yb3+
                        0.2,Er3+
                        0.02 pumped by UV flashlight (Figure 5b). (b) Visible (down-shifting) emission spectrum of phosphor NaY0.78F4:Yb3+
                        0.2,Er3+
                        0.02 pumped by UV bulb for insect killer (Figure 5a). Group of peaks 1 corresponds to the transition of excited ion Er3+:2H11/2 → 4I15/2 (520 nm); group 2 – Er3+:4S3/2 → 4I15/2 (540 nm); group 3 – Er3+:4F9/2 → 4I15/2 (653 nm). Dashed line is the tail of the spectrum of the UV bulb.
Figure 6

(a) Down-conversion emission spectra of phosphors (1) NaY0.94F4:Yb3+ 0.03,Er3+ 0.03 and (2) NaY0.78F4:Yb3+ 0.2,Er3+ 0.02 pumped by UV flashlight (Figure 5b). (b) Visible (down-shifting) emission spectrum of phosphor NaY0.78F4:Yb3+ 0.2,Er3+ 0.02 pumped by UV bulb for insect killer (Figure 5a). Group of peaks 1 corresponds to the transition of excited ion Er3+:2H11/24I15/2 (520 nm); group 2 – Er3+:4S3/24I15/2 (540 nm); group 3 – Er3+:4F9/24I15/2 (653 nm). Dashed line is the tail of the spectrum of the UV bulb.

Figure 7 presents down-conversion emission spectra of phosphors (a) NaY0.88F4:Yb3+ 0.1,Tm3+ 0.02 and (b) NaY0.78F4:Yb3+ 0.2,Tm3+ 0.02 pumped with Ar-ion laser at 488 nm (power ∼ 75.5 mW). One can see that the intensity of this emission is much stronger than that of the phosphors based on (Er, Yb) pair presented in Figure 6. This can be explained by strong down-shifting radiation (Figure 6b) competing more significantly with the down-conversion process in (Er, Yb) pair as compared to (Tm, Yb). Down-shifting emission from (Tm, Yb) pair pumped with blue light occurs in 650–660 nm and 800–816 nm bands (Figure 3b). As will be shown later (Figure 17), the radiation in the first, red band was much weaker than the radiation in the second, short-wavelength NIR band. Moreover, the second band was also directly involved in the cross-relaxation energy-transfer mechanism (2) between transition Tm3+:1G43H5 and Yb3+:2F7/22F5/2 (Figure 3b) that contributed significantly to down-conversion. In each NIR down-conversion spectrum, Stark splitting peak structure of Yb3+ emission (during transition 2F5/22F7/2, Figure 3b) could not be well distinguished. A dominating spectral peak at 980 nm could be observed for phosphor (a) and 1,100 nm – for phosphor (b) with double content of Yb3+. The latter phosphor had a diffusion reflectance spectrum presented in Figure 4 (curve B) with a strong absorption peak corresponding to Yb3+ transition 2F7/22F5/2. This intensity of down-conversion emission of this phosphor was also ∼5 times greater than that of phosphor (a).

Figure 7 
                     (a) Down-conversion emission spectra of phosphors (a) NaY0.88F4:Yb3+
                        0.1,Tm3+
                        0.02 and (b) NaY0.78F4:Yb3+
                        0.2,Tm3+
                        0.02 pumped with Ar-ion laser at 488 nm (power ∼ 75.5 mW).
Figure 7

(a) Down-conversion emission spectra of phosphors (a) NaY0.88F4:Yb3+ 0.1,Tm3+ 0.02 and (b) NaY0.78F4:Yb3+ 0.2,Tm3+ 0.02 pumped with Ar-ion laser at 488 nm (power ∼ 75.5 mW).

Figure 8a presents the intensity of down-conversion emission at 980 nm of phosphor NaY0.88F4:Yb3+ 0.1,Tm3+ 0.02 plotted versus the wavelength of the pump (different spectral lines of Ar-ion laser). The intensity was re-calibrated to account for different pump powers at various wavelengths. One can see that NIR emission response of the phosphor remains basically “flat” within the spectral band of the pump from 457.9 to 514.5 nm. Quantum yield (QY) of down-conversion emission of the phosphor, defined as the ratio of number N NIR of produced NIR photons to number N 488 of pump blue (488 nm) photons QY = N NIR /N 488 [56], was roughly (neglecting reflectance of pump radiation) estimated in the following way. N 488 was proportional to the power of incident 488-nm laser beam (∼80 mW), N 488 = T × 80 mW (T – time interval). N NIR was computed based on the ratio of the intensity of NIR emission I NIR, as measured by spectrometer SP2500i, times its spectral band Δ NIR to intensity I 980 of the spectral peak of a reference 980 nm laser diode with known power (1.6 mW) times its spectral band Δ 980, N NIR = T × [(I NIR × Δ NIR)/(I 980 × Δ 980)] × 1.6 mW. QY was found to be ∼5% – of the same order as in ref. [56]. As an example of the phosphor response to variable pump power of the Ar-ion laser spectral lines, Figure 8b presents the intensity of down-conversion emission at 980 nm peak plotted versus the power of the 465.8 nm pump from Ar-ion laser. Linear fit of the plot suggests that down-conversion is a linear process [54]. Accordingly, this makes it possible to reach high conversion efficiencies independent of the incident power and allows for the use of non-concentrated sunlight.

Figure 8 
                     Intensity of down-conversion emission at 980 nm of phosphor NaY0.88F4:Yb3+
                        0.1,Tm3+
                        0.02 (a) versus the wavelength of the pump (different spectral lines of Ar-ion laser) and (b) versus the power of pump from Ar-ion laser spectral line at 465.8 nm. Solid red line is the linear fit of the experimental plot.
Figure 8

Intensity of down-conversion emission at 980 nm of phosphor NaY0.88F4:Yb3+ 0.1,Tm3+ 0.02 (a) versus the wavelength of the pump (different spectral lines of Ar-ion laser) and (b) versus the power of pump from Ar-ion laser spectral line at 465.8 nm. Solid red line is the linear fit of the experimental plot.

3.2 Pure polymer film after deposition

3.2.1 Morphology

Since the polymer target was exposed to intense laser radiation in ambient air, our concern was preservation of polymer integrity during PLD. The morphology of pure PMMA coatings was investigated using optical microscopy with an optical AmScope M400 monocular compound microscope (WF10× eyepiece; 40× to 400× magnification; tungsten bulb illuminator; brightfield, Abbe condenser; coarse and fine Focus) with Mu 400 CCD camera. The microscope images of the coatings are presented in Figures 9–11. Images taken on the edge of the coating (Figures 9 and 10) show that the deposited material was in the form of separate polymer islands brought to the substrate more likely by the process of laser-induced catapulting of the droplets of melted target [57]. Glass transition temperature of PMMA is known to be between 85 and 165°C [58]. The substrate with polymer coating was placed on a hot plate heated to a temperature of ∼90°C as measured by a thermo-couple thermometer touching the surface of the hotplate. Heating the substrate for 2 min melted isolated polymer islands and forced some of them to coalesce (Figure 10). In the centre of the coating (Figure 11), coalescence produced a continuous polymer film. The film was separated from the substrate and its thickness was measured with a micometer to be 35 ± 5 µm. Poorly controlled and rather rapid heat influx induced boiling and formed air bubbles in some spots.

Figure 9 
                     Microscope images of PMMA coating on glass substrate right after open-air PLD. The images were taken with (a) ×40 magnification (dimensions: X = 2,257 µm, Y = 1,674 µm) and (b) ×100 magnification (dimensions: X = 903 µm, Y = 670 µm). Dark band on the right in image (a) is the mark made by microscope objective that occasionally touched the coating. The images were taken on the edge of the deposition zone, where deposited material did not cover whole surface.
Figure 9

Microscope images of PMMA coating on glass substrate right after open-air PLD. The images were taken with (a) ×40 magnification (dimensions: X = 2,257 µm, Y = 1,674 µm) and (b) ×100 magnification (dimensions: X = 903 µm, Y = 670 µm). Dark band on the right in image (a) is the mark made by microscope objective that occasionally touched the coating. The images were taken on the edge of the deposition zone, where deposited material did not cover whole surface.

Figure 10 
                     Microscope images of PMMA coating presented in Figure 9 after melting on a hot plate at ∼90°C with (a) ×40 and (b) ×100 magnification. The images were taken on the edge of the deposition zone, where deposited material did not cover whole surface.
Figure 10

Microscope images of PMMA coating presented in Figure 9 after melting on a hot plate at ∼90°C with (a) ×40 and (b) ×100 magnification. The images were taken on the edge of the deposition zone, where deposited material did not cover whole surface.

Figure 11 
                     Microscope image of PMMA coating right after melting on a hot plate at ∼90°C taken with (a) ×40 and (b) ×100 magnification. The images were taken in the centre, where deposited material covered whole surface. Small circles are air bubbles in the boiling melt.
Figure 11

Microscope image of PMMA coating right after melting on a hot plate at ∼90°C taken with (a) ×40 and (b) ×100 magnification. The images were taken in the centre, where deposited material covered whole surface. Small circles are air bubbles in the boiling melt.

3.2.2 Chemical integrity

To verify if the polymer molecules disintegrated after deposition, we compared FTIR absorption spectra of the coating against unexposed target (Figure 12a). FTIR spectra were taken with Nicolet iS5 FTIR spectrometer from Thermo Fisher Scientific with iD7 ATR diamond anvil attachment. Apparent similarity of two FTIR spectra indicated that there was no noticeable decomposition of the deposited polymer. Figure 12b presents FTIR spectra of (1) virgin target (unexposed PMMA slide) and (2) target material on the edge of the cavity made in the target by the laser ablation. Three new absorption peaks can be attributed to the stretch vibration modes of terminal alkynes R–C═C–H (2,140–2,100 cm−1), R–N═C–S (2,140–1,990 cm−1), R–N–C (2,165–2,110 cm−1), or cyanide C═N (2019 cm−1) [59]. Based on the FTIR data, the following assumptions can be made with regard to the sequence of events during open-air PLD of the polymer (Figure 13): (1) laser beam heated and melted the polymer; (2) volatile products of decomposition of the polymer escaped; (3) shock wave pulverized and sputtered (catapulted) droplets of melted polymer on the substrate; (4) polymer target material around the cavity edges underwent pyrolysis and photo-chemical modification; (5) polymer droplets solidified on the substrate; and (6) deposited coating preserved the chemical structure of the polymer target.

Figure 12 
                     (a) FTIR absorption spectra of PMMA slide used as a PLD target and deposited film (Figures 9 through 11). (b) FTIR spectra of (1) PMMA slide (target) and (2) the material on the edge of the cavity created in the PMMA slide by the laser beam during ablation. Three new absorption peaks in the box can be attributed to the stretch vibration modes of terminal alkynes R–C═C–H (2,140–2,100 cm−1), R–N═C–S (2,140–1,990 cm−1), R–N–C (2,165–2,110 cm−1), C═N (cyanides, 2,019 cm−1) [59].
Figure 12

(a) FTIR absorption spectra of PMMA slide used as a PLD target and deposited film (Figures 9 through 11). (b) FTIR spectra of (1) PMMA slide (target) and (2) the material on the edge of the cavity created in the PMMA slide by the laser beam during ablation. Three new absorption peaks in the box can be attributed to the stretch vibration modes of terminal alkynes R–C═C–H (2,140–2,100 cm−1), R–N═C–S (2,140–1,990 cm−1), R–N–C (2,165–2,110 cm−1), C═N (cyanides, 2,019 cm−1) [59].

Figure 13 
                     Schematic illustrating the open-air PLD of polymer coating.
Figure 13

Schematic illustrating the open-air PLD of polymer coating.

3.3 Polymer nanocomposite film

Down-conversion polymer–phosphor composite films made by CMBMT-PLD were further investigated to verify if the phosphor particles retained the properties of the target and to find the average size of the particles and the way they were embedded in the polymer. Figure 14a presents the X-ray diffraction (XRD) spectrum of PMMA + NaYF4:Yb3+,Er3+ film taken with a Bruker D2 Phaser diffractometer. The sample was prepared by separating the film from the substrate and pulverizing it with a mortar. Observed diffraction peaks can be attributed to the hexagonal β-phase of NaYF4 (JCPDS No 28-1192) identical to the crystalline structure of the target. As for the effect of Yb3+ or Er3+ (Tm3+) dopants on the XRD spectrum of the NaYF4 matrix in the phosphors, small variations might be detected in the lattice parameters due to differences in ionic radius size between the Y3+ and the dopants taking its place in random locations in the crystalline matrix [60,61]. Randomness of the dopant locations brings disorder in the matrix and variations of the local crystal field for the second-nearest neighbours of the dopant ions. This imposes a slightly different crystal field environment for each subset of the dopant ions. Such differences imply an inhomogeneous broadening of the bands corresponding to the transitions in the photoluminescence spectra unlike those from ordered crystals that exhibit narrow bands. Detailed investigation of such phenomena was out of scope of this work. High-resolution scanning-electron-microscope (SEM) image of the film taken with a JEOL JSM-7100FA microscope (Figure 14b) shows phosphor particles (white spots) of a size between 10 and 50 nm dispersed in the polymer matrix. Analysis of the SEM images showed that the mean concentration of nanoparticles n was 526 per square micron. The uniformity of their dispersion U, defined as U = (1 SD/n) × 100% [62] (SD is the standard deviation of the concentration of nanoparticles), was 76%. Further improvement of uniformity could be achieved by randomized translation of the substrate during the deposition. Based on average nanoparticle diameter d a ∼ 30 nm and two-dimensional concentration of the nanoparticles in the film, the effective share of sunlight photons captured by a 30 nm-thick nanocomposite sub-layer S c could be estimated as S c = nd 2 /4) = 0.546. The share of sunlight photons passing freely through the sub-layer was thus P f = 1 − S c = 0.454. For a 300 nm-thick nanocomposite film made of ten 30 nm-thick sub-layers the share of passing sunlight photons would be (P f)10 = 3.4 × 10−4. So almost all the photons were captured. QY ∼ 5% of the phosphor determined earlier could thus be applied to a nanocomposite film of a thickness >300 nm as an estimate of its down-conversion efficiency as well.

Figure 14 
                  (a) XRD spectrum of the polymer nanocomposite film made of PMMA and the nanoparticles of NaYF4:Yb3+,Er3+ with the diffraction peaks attributed to the hexagonal β-phase of NaYF4. (b) High-resolution SEM image of the polymer nanocomposite film of PMMA + NaYF4:Yb3+,Er3+ deposited on glass substrate at magnification ×100k.
Figure 14

(a) XRD spectrum of the polymer nanocomposite film made of PMMA and the nanoparticles of NaYF4:Yb3+,Er3+ with the diffraction peaks attributed to the hexagonal β-phase of NaYF4. (b) High-resolution SEM image of the polymer nanocomposite film of PMMA + NaYF4:Yb3+,Er3+ deposited on glass substrate at magnification ×100k.

One convenient way of checking preservation of optical properties of the phosphor target in deposited nanoparticles could be observation of visible up-conversion radiation from a polymer nanocomposite film illuminated with a 980 nm laser diode. Here we consider nanoparticles of NaYF4:Yb3+,Tm3+. Energy-level diagram explaining mechanisms of up-conversion in coupled ions of Yb3+ and Tm3+ is presented in Figure 15 [63]. Solid, dotted, and wavy arrows represent photon absorption/emission, energy transfer, and non-radiative relaxation processes, respectively. Four ions of Yb3+ are excited by NIR pump radiation and transfer energy of their NIR photons via a multi-step excitation process to an ion of Tm3+. During relaxation to the ground state Tm3+ ion can emit blue, green, red, or short-wavelength NIR photon. We illuminated the deposited and melted PMMA + NaY88%F4:Yb3+ 10%Tm3+ 2% film with a 980-nm laser diode and were able to see up-conversion emission from illuminated spots in the form of blue strips having the shape of the laser beam (Figure 16). Furthermore, spectroscopy of up-conversion emission from the film (Figure 17) revealed the presence of blue (488 nm), weak (slightly higher than noise) red (658 nm), and strong 802 nm NIR peak in agreement with the energy diagram (Figure 15) and literature data [63]. This confirmed that the phosphor nanoparticles retained the optical properties of the target after PLD process and were evenly distributed in the deposited polymer film.

Figure 15 
                  Energy-level diagram of Yb3+ and Tm3+ ions explaining the mechanism of up-conversion. Solid, dotted, and wavy arrows represent photon absorption/emission, energy-transfer, and non-radiative relaxation processes, respectively [63].
Figure 15

Energy-level diagram of Yb3+ and Tm3+ ions explaining the mechanism of up-conversion. Solid, dotted, and wavy arrows represent photon absorption/emission, energy-transfer, and non-radiative relaxation processes, respectively [63].

Figure 16 
                  Optical microscopic images of PMMA+ NaY0.88F4:Yb3+
                     0.1Tm3+
                     0.02 nanocomposite film made by open-air CMBMT-PLD and heated at ∼90°C at magnification (a) ×40 and (b) ×100. The film was illuminated with a 200 mW 980 nm laser. Blue strip is produced by up-conversion emission of the embedded phosphor nanoparticles initiated by the infrared pump.
Figure 16

Optical microscopic images of PMMA+ NaY0.88F4:Yb3+ 0.1Tm3+ 0.02 nanocomposite film made by open-air CMBMT-PLD and heated at ∼90°C at magnification (a) ×40 and (b) ×100. The film was illuminated with a 200 mW 980 nm laser. Blue strip is produced by up-conversion emission of the embedded phosphor nanoparticles initiated by the infrared pump.

Figure 17 
                  Upconversion emission spectrum of PMMA + NaY0.88F4:Yb3+
                     0.1Tm3+
                     0.02 nanocomposite film made by open-air CMBMT-PLD. The film was pumped with a 200 mW 980 nm laser.
Figure 17

Upconversion emission spectrum of PMMA + NaY0.88F4:Yb3+ 0.1Tm3+ 0.02 nanocomposite film made by open-air CMBMT-PLD. The film was pumped with a 200 mW 980 nm laser.

4 Conclusions

RE-based phosphors NaYF4:Er3+,Yb3+ and NaYF4:Yb3+,Tm3+, synthesized by the wet method, were found to be suitable for polymer nanocomposite sunlight spectrum down-converters with a Stokes shift between the spectrum of absorption and fluorescence as large as 620 nm (from 360 to 980 nm) minimizing the effect of re-absorption. The latter phosphor was more efficient in down-converting blue light in NIR with a QY of ∼5%. Polymer nanocomposite down-converters were successfully deposited on glass substrates by co-deposition of polymer PMMA and the phosphors using the open-air CMBMT-PLD process. The polymer was found to preserve its chemical integrity during deposition that more likely occurred as catapulting polymer droplets from the melted target. Post-heating of the substrate at ∼90°C led to fusion of melted polymer islands into uniform coating. The phosphor nanoparticles with a size ranging from 10 to 50 nm were evenly embedded in the polymer matrix during the co-deposition. The nanoparticles retained the crystalline structure and optical properties of the phosphor target. QY of the down-conversion polymer nanocomposite film was estimated to be not less than 5%. The proposed deposition method thus demonstrated its feasibility and potential for making commercial-size thin-film sunlight down-converters for PV solar power applications.

Acknowledgments

Dillard team appreciates the financial support from US Air Force Office of Scientific Research Grant FA9550-18-1-0364 AFOSR and Army Research Office Grant No W911-NF-19-1-0451 and the partial support from DU Minority Health and Health Disparity Research Center MHHDRC. Research of Dr. Patel was sponsored by the Army Research Office and was accomplished under Grants Number W911NF-18-1-0446 and W911NF-19-1-0506. The views and conclusions contained in this document are those of the authors and should not be interpreted as representing the official policies, either expressed or implied, of the Army Research Office or the U.S. Government. The U.S. Government is authorized to reproduce and distribute reprints for Government purposes notwithstanding any copyright notation herein. Additionally, Dr. Patel’s research was supported by the grant from UNCF, Henry C. McBay Research Fellowship 2019.

  1. Conflict of interest: The authors declare no conflict of interest regarding the publication of this paper.

References

[1] Debije M. Luminescent solar concentrators. Luminescent solar solutions. Semiconductor solutions. Nat Photonics. 2017;11:143–4.10.1038/nphoton.2017.20Search in Google Scholar

[2] Ho W-J, Yang G-C, Shen Y-T, Deng Y-J. Improving efficiency of silicon solar cells using europium-doped silicate-phosphor layer by spin-on film coating. Appl Surf Sci. 2016;365:120–4.10.1016/j.apsusc.2016.01.061Search in Google Scholar

[3] Krumer Z, Pera SJ, van Dijk – Moes RJA, Zhao Y, de Brouwer S, et al. Tackling self-absorption in luminescent solar concentrators with type-II colloidal quantum dots. Sol Energ Mat Sol C. 2013;111:57–65.10.1364/PV.2012.PW2B.3Search in Google Scholar

[4] Pietryga JM, Park Y-S, Lim J, Fidler AF, Bae WK, Sergio Brovelli S, et al. Spectroscopic and device aspects of nanocrystal quantum dots. Chem Rev. 2016;116:10513–622.10.1021/acs.chemrev.6b00169Search in Google Scholar PubMed

[5] Giebink NC, Wiederrecht GP, Wasielewski MR. Resonance-shifting to circumvent reabsorption loss in luminescent solar concentrators. Nat Photonics. 2011;5:694–701.10.1038/nphoton.2011.236Search in Google Scholar

[6] Bhattacharya S, Alkharfy KM, Janardhanan R, Mukhopadhyay D. Nanomedicine: Pharmacological perspectives. Nanotechnol Rev. 2012;1:235–53.10.1515/ntrev-2011-0010Search in Google Scholar

[7] Huang Y-Y, Sharma SK, Dai T, Chung H, Yaroslavsky A, Garcia-Diaz M, et al. Can nanotechnology potentiate photodynamic therapy? Nanotechnol Rev. 2012;1(2):111–46.10.1515/ntrev-2011-0005Search in Google Scholar PubMed PubMed Central

[8] Le Trequesser Q, Seznec H, Delville M-H. Functionalized nanomaterials: Their use as contrast agents in bioimaging: mono- and multimodal approaches. Nanotechnol Rev. 2013;2(2):125–69.10.1515/ntrev-2012-0080Search in Google Scholar

[9] Youkhanna J, Syoufjy J, Rhorer M, Oladeinde O, Zeineldin R. Toward nanotechnology-based solutions for a disease: Ovarian cancer as an example. Nanotechnol Rev. 2013;2(4):473–84.10.1515/ntrev-2013-0008Search in Google Scholar

[10] Sheeba CJ, Marslin G, Revina AM, Franklin G. Signaling pathways influencing tumor microenvironment and their exploitation for targeted drug delivery. Nanotechnol Rev. 2014;3(2):123–51.10.1515/ntrev-2013-0032Search in Google Scholar

[11] Yao L, Xu S. Detection of magnetic nanomaterials in molecular imaging and diagnosis applications. Nanotechnol Rev. 2014;3(3):247–68.10.1515/ntrev-2013-0044Search in Google Scholar

[12] Rai M, Birla S, Ingle AP, Gupta I, Gade A, Abd-Elsalam K, et al. Nanosilver: An inorganic nanoparticle with myriad potential applications. Nanotechnol Rev. 2014;3(3):281–309.10.1515/ntrev-2014-0001Search in Google Scholar

[13] Li J, Liu H, Deng Y, Liu G, Chen Y, Yang J. Emerging nanostructured materials for the catalytic removal of volatile organic compounds. Nanotechnol Rev. 2016;5(1):147–81.10.1515/ntrev-2015-0051Search in Google Scholar

[14] Zoghbi ME, Altenberg GA. Membrane protein reconstitution in nanodiscs for luminescence spectroscopy studies. Nanotechnol Rev. 2017;6(1):33–46.10.1515/ntrev-2016-0078Search in Google Scholar

[15] Song C, Zhang S, Zhou Q, Hai H, Zhao D, Hui Y. Upconversion nanoparticles for bioimaging. Nanotechnol Rev. 2017;6(2):233–42.10.1515/ntrev-2016-0043Search in Google Scholar

[16] Yang Y, Qina Z, Zenga W, Yanga T, Cao Y, Mei C, et al. Toxicity assessment of nanoparticles in various systems and organs. Nanotechnol Rev. 2017;6(3):279–89.10.1515/ntrev-2016-0047Search in Google Scholar

[17] Malekzad H, Zangabad PS, Mirshekar H, Karimi M, Hamblin MR. Noble metal nanoparticles in biosensors: Recent studies and applications. Nanotechnol Rev. 2017;6(3):301–29.10.1515/ntrev-2016-0014Search in Google Scholar PubMed PubMed Central

[18] Zhoua Q, Zhanga L, Wua H. Nanomaterials for cancer therapies. Nanotechnol Rev. 2017;6(5):473–96.10.1515/ntrev-2016-0102Search in Google Scholar

[19] Fan R, Lu F, Li K. Fabrication and simulation of 1,540 nm transmission by 532 nm excitation in photonic crystal of Er:ZnO film. Nanotechnol Rev. 2017;6(6):497–503.10.1515/ntrev-2017-0143Search in Google Scholar

[20] Li J, Yao M, Shao Y, Yao D. The application of bio-nanotechnology in tumor diagnosis and treatment: A view. Nanotechnol Rev. 2018;7(3):257–66.10.1515/ntrev-2018-0011Search in Google Scholar

[21] Jia F, Li G, Yang B, Yu B, Shen Y, Cong H. Investigation of rare earth upconversion fluorescent nanoparticles in biomedical field. Nanotechnol Rev. 2019;8:1–17.10.1515/ntrev-2019-0001Search in Google Scholar

[22] Sawicki K, Czajka M, Matysiak-Kucharek M, Fal B, Drop B, Męczyńska-Wielgosz S, et al. Toxicity of metallic nanoparticles in the central nervous system. Nanotechnol Rev. 2019;8:175–200.10.1515/ntrev-2019-0017Search in Google Scholar

[23] Miao Y, Lu J, Yin J, Zhou C, Guo Y, Zhou S. Yb3+-containing chitosan hydrogels induce B-16 melanoma cell anoikis via a Fak-dependent pathway. Nanotechnol Rev. 2019;8:645–60.10.1515/ntrev-2019-0056Search in Google Scholar

[24] Godlewski MM, Kaszewski J, Kielbik P, Olszewski J, Lipinski W, Słońska A, et al. New generation of oxide-based nanoparticles for the applications in early cancer detection and diagnostics. Nanotechnol Rev. 2020;9(1):274–302.10.1515/ntrev-2020-0022Search in Google Scholar

[25] Nakamura M. Biomedical applications of organosilica nanoparticles towards theranostics. Nanotechnol Rev. 2012;1:469–91.10.1515/ntrev-2012-0005Search in Google Scholar

[26] Wang T, Zhang J, Maa W, Luo Y, Wang L, Hua Z, et al. Luminescent solar concentrator employing rare earth complex with zero self-absorption loss. Sol Energy. 2011;85:2571–9.10.1016/j.solener.2011.07.014Search in Google Scholar

[27] Park WJ, Oh SJ, Kim JK, Heo J, Wagner T, Strizik L. Down-conversion in Tm3+/Yb3+ doped glasses for multicrystalline silicon photo-voltaic module efficiency enhancement. J Non-Cryst Solids. 2014;83:181–3.10.1016/j.jnoncrysol.2013.03.037Search in Google Scholar

[28] Melvin ten Kate O, Krämer KW, Van, der Kolk E. Efficient luminescent solar concentrators based on self-absorption free, Tm2+ doped halides. Sol Energ Mat Sol C. 2015;140:115–20.10.1016/j.solmat.2015.04.002Search in Google Scholar

[29] Xia W, Jin X, Yang X, Xiao S. Down conversion of Er3+-Yb3+ couple in Y2BaZnO5. Opt Mater. 2016;54:294–9.10.1016/j.optmat.2016.02.037Search in Google Scholar

[30] Verma A, Sharma SK. Down-conversion from Er3+-Yb3+ codoped CaMoO4 phosphor: A spectral conversion to improve solar cell efficiency. Ceram Inter. 2017;43:8879–85.10.1016/j.ceramint.2017.04.023Search in Google Scholar

[31] Darwish A, Sarkisov S. Method and apparatus for open-air pulsed laser deposition. US Patent No. 10316403 B2. June 11, 2019.Search in Google Scholar

[32] Darwish AM, Sagapolutele MT, Sarkisov S, Patel D, Hui D, Koplitz B. Double beam pulsed laser deposition of composite films of poly (methyl methacrylate) and rare earth fluoride upconversion phosphors. Compos Part B - Eng. 2013;55:139–46.10.1016/j.compositesb.2013.06.013Search in Google Scholar

[33] Darwish AM, Wilson S, Sarkisov S, Patel D. Double pulse laser deposition of polymer nanocomposite NaYF4:Tm3+,Yb3+ films for optical sensors and light emitting applications. In: Yin S, Guo R, editors. Photonic fiber and crystal devices: advances in materials and innovations in device applications VII. Proceedings of SPIE 8847; 2014 Sept 28; San Diego, CA. Bellingham, WA: SPIE; 2013. p. 884702.10.1117/12.2027144Search in Google Scholar

[34] Darwish AM, Burkett A, Blackwell A, Taylor K, Sarkisov S, Patel D, et al. Polymer-inorganic nano-composite thin film upconversion light emitters prepared by double-beam matrix assisted pulsed laser evaporation (DB-MAPLE) method. Compos Part B–Eng. 2015;68:355–64.10.1016/j.compositesb.2014.08.044Search in Google Scholar

[35] Darwish AM, Burkett A, Blackwell A, Taylor K, Walker V, Sarkisov S, et al. Efficient upconversion polymer-inorganic nanocomposite emitters prepared by the double beam matrix assisted pulsed laser evaporation (DB-MAPLE). In: Yin S, Guo R, editors. Photonic fiber and crystal devices: advances in materials and innovations in device applications VIII. Proceedings SPIE 9200; 2014 Sept 5; San Diego, CA. Bellingham, WA: SPIE; 2014. p. 92000C.10.1117/12.2063129Search in Google Scholar

[36] Patel D, Lewis A, Wright III D, Lewis D, Sarkisov S, Darwish AM. Optical properties and size distribution of the nano-colloids made of rare-earth ion-doped NaYF4. In: Jiang S, Digonnet MJF, editors. Optical components and materials XII: Proceedings of SPIE 9359; 2015 March 16; San Francisco, CA. Bellingham, WA: SPIE; 2015. p. 93591L.10.1117/12.2077680Search in Google Scholar

[37] Darwish AM, Wilson S, Blackwell A, Taylor K, Sarkisov SS, Patel DN, et al. Ammonia sensor based on polymer-inorganic nano-composite thin film upconversion light emitter prepared by double-beam pulsed laser deposition. Am J Mater Sci. 2015;5(3A):8–15.Search in Google Scholar

[38] Darwish AM, Wilson S, Blackwell A, Taylor K, Sarkisov S, Patel D, et al. Polymer-inorganic nanocomposite thin film emitters, optoelectronic chemical sensors, and energy harvesters produced by multiple-beam pulsed laser deposition. In: Yin S, Guo R, editors. Photonic fiber and crystal devices: advances in materials and innovations in device applications IX: Proceedings of SPIE 9586; 2015 Aug 26; San Diego, CA. Bellingham, WA: SPIE; 2015. p. 958602.10.1117/12.2185498Search in Google Scholar

[39] Darwish AM, Wilson S, Blackwell A, Sarkisov S, Patel D, Mele P. et al. Multi-beam pulsed laser deposition: New method of making nanocomposite coatings. In: Yin S, Guo R, editors. Photonic fiber and crystal devices: advances in materials and innovations in device applications IX: Proceedings of SPIE 9586; 2015 Aug 26; San Diego, CA. Bellingham, WA: SPIE; 2015. p. 958605.10.1117/12.2214633Search in Google Scholar

[40] Darwish AM, Moorea S, Muhammad A, Sarkisov SS, Patel DN, Mele P, et al. Organic-inorganic nano-composite films for photonic applications made by multi-beam multi-target pulsed laser deposition with remote control of the plume directions. In: Yin S, Guo R, editors. Photonic fiber and crystal devices: advances in materials and innovations in device applications X: Proceedings of SPIE 9958; 2016 Sept 15; San Diego, CA. Bellingham, WA: SPIE; 2016. p. 995802.10.1117/12.2237538Search in Google Scholar

[41] Darwish AM, Moore S, Muhammad A, Sarkisov S, Patel D, Mele P, et al. Polymer nano-composite films with inorganic upconversion phosphor and electro-optic additives made by concurrent triple-beam matrix assisted and direct pulsed laser deposition. Compos Part B Eng. 2017;109:82–90.10.1016/j.compositesb.2016.10.053Search in Google Scholar

[42] Darwish AM, Sarkisov SS, Patel DN. Concurrent multi-target laser ablation for making nano-composite films. Chapter 6. In: Yang D, editor. Applications of laser ablation – Thin Film Deposition, Nanomaterial Synthesis and Surface Modification. Rijeka, Croatia: InTech; 2016. p. 129–48.10.5772/64816Search in Google Scholar

[43] Darwish AM, Sarkisov SS, Patel DN, Muhammad A, Thompson K, Johnson M, et al. Polymer nanocomposite luminescent films for solar energy harvesting made by concurrent multi-beam multi-target pulsed laser deposition. In: Yin S, Guo R, editors. Photonic fiber and crystal devices: advances in materials and innovations in device applications XII: Proceedings of SPIE 10755; 2018 Sept 4; San Diego, CA. Bellingham, WA: SPIE; 2018. p. 107552.Search in Google Scholar

[44] Darwish AM, Sarkisov SS, Patel DN. Nanocomposite luminescent solar concentrators: Optics for green energy. Asian J Phys. 2018;27(9–12):625–36.Search in Google Scholar

[45] Darwish AM, Sarkisov SS, Wilson J, Muhammad A, Thompson K, Patel DN, et al. Luminescent solar concentrators based on polymer nanocomposite films made by open-air pulsed laser deposition. In: Yin S, Guo R, editors. Photonic fiber and crystal devices: advances in materials and innovations in device applications XIII: Proceedings of SPIE 11123; 2019 Sept 9; San Diego, CA. Bellingham, WA: SPIE; 2019. p. 1112302.10.1117/12.2530359Search in Google Scholar

[46] Darwish AM, Sarkisov SS, Patel DN. Nanocomposite windows converting solar power into electricity for self-sustaining buildings. In: Bumajdad A, Bouhamra W, Alsayegh OA, Kamal HA, Alhajraf SF, editors. Gulf Conference on Sustainable Built Environment. Cham, Switzerland: Springer Nature Switzerland AG; 2020. p. 367–82.10.1007/978-3-030-39734-0_21Search in Google Scholar

[47] Husainat A, Warsame A, Cofie P, Attia J, Fuller J, Darwish A. Simulation and analysis method of different back metals contact of CH3NH3PbI3 perovskite solar cell along with electron transport layer TiO2 using MBMT-MAPLE/PLD. Am J Opt Photonics. 2020;8(1):6–26.10.11648/j.ajop.20200801.12Search in Google Scholar

[48] Van der Ende BM, Aarts L, Meijerink A. Lanthanide ions as spectral converters for solar cells. Phys Chem Phys. 2009;11:11081–95.10.1039/b913877cSearch in Google Scholar PubMed

[49] Baker CC, Fontana J, Kim W, Bowman SR, Sanghera J, Ballato J, et al. Nanoparticle doping for high power fiber lasers at eye-safer wavelengths. Opt Express. 2017;25(12):13903–15.10.1364/OE.25.013903Search in Google Scholar PubMed

[50] Patel D, Sarkisov SS, Darwish A, Ballato J. Optical gain in capillary light guides filled with NaYF4:Yb3+,Er3+ nanocolloids. Opt Express. 2016;24(18):21147–58.10.1364/OE.24.021147Search in Google Scholar PubMed

[51] Konov VI, Kononenko TV, Loubnin EN, Dausinger F, Breitling D. Pulsed laser deposition of hard coatings in atmospheric air. Appl Phys A. 2004;79:931–6.10.1007/s00339-004-2570-9Search in Google Scholar

[52] Torrisi L, Lorusso A, Nassisi V, Picciotto A. Characterization of laser ablation of polymethylmethacrylate at different laser wavelengths. Rad Eff Defects Solids: Incorp Plasma Sci Plasma Technol. 2008;163(3):179–87.10.1080/10420150701259172Search in Google Scholar

[53] Lin H, Zhou S, Hou X, Li W, Li Y, Teng H, et al. Down-conversion from blue to near infrared in Tm3+ –Yb3+ codoped Y2O3 transparent ceramics. IEEE Photon Tech Lett. 2010;22(12):866–88.10.1109/LPT.2010.2046631Search in Google Scholar

[54] Aarts L, Van der Ende BM, Meijerink A. Downconversion for solar cells in NaYF4:Er, Yb. J Appl Phys. 2009;106:023522.10.1063/1.3177257Search in Google Scholar

[55] Sunlight. Wikipedia [Internet] 2020 July [cited 2020 August 23]. Available from https://en.wikipedia.org/wiki/Sunlight#cite_ref-6.Search in Google Scholar

[56] Zhydachevskyy Y, Tsiurma V, Baran M, Lipinska L, Sybilski P, Suchoki A. Quantum efficiency of the down-conversion process in Bi3+ –Yb3+ co-doped Gd2O3. J Lumin. 2018;196:169–73.10.1016/j.jlumin.2017.12.042Search in Google Scholar

[57] Vogel A, Loretz K, Horneffer V, Huttman G, von Smolinski D, Gebert A. Mechanisms of laser-induced dissection and transport of histologic specimens. Biophys J. 2007;93:4481–500.10.1529/biophysj.106.102277Search in Google Scholar PubMed PubMed Central

[58] Poly(methyl methacrylate). Wikipedia [Internet] 2020 August [cited 2020 August 28]. Available from https://en.wikipedia.org/wiki/Sunlight#cite_ref-6.Search in Google Scholar

[59] Larkin P. Infrared and Raman Spectroscopy: Principles and Spectral Interpretation. Waltman, MA: Elsevier; 2011.10.1016/B978-0-12-386984-5.10002-3Search in Google Scholar

[60] Deng K, Gong T, Hu L, Wei X, Chen Y, Yin M. Efficient near-infrared quantum cutting in NaYF4:Ho3+,Yb3+ for solar photovoltaics. Opt Express. 2011;19(3):1749–54.10.1364/OE.19.001749Search in Google Scholar PubMed

[61] Barranco J, Mendez-Blas A, Calixto ME. Structural, morphology and optical properties of NaYF4 thin films doped with trivalent lanthanide ions. J Mat Sci: Mat Electron. 2019;30:4855–66.10.1007/s10854-019-00780-9Search in Google Scholar

[62] Prokhorov AV, Hanssen LM. Numerical modeling of an integrating sphere radiation source. In: Wooley CB, editor. Modeling and Characterization of Light Sources: Proceedings of SPIE 4775; 2002 August 16; Seattle, WA. Bellingham, WA: SPIE; 2002. p. 106–18.10.1117/12.479642Search in Google Scholar

[63] Arppe R, Hyppänen I, Perälä N, Peltomaa R, Kaiser M, Würth C, et al. NaYF4:Yb3+,Er3+ and NaYF4:Yb3+,Tm3+ nanophosphors by water: The role of the sensitizer Yb3+ in non-radiative relaxation. Nanoscale. 2015;7:11746–57.10.1039/C5NR02100FSearch in Google Scholar PubMed

Received: 2020-08-11
Accepted: 2020-09-16
Published Online: 2020-10-30

© 2020 Abdalla M. Darwish et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Research Articles
  2. Generalized locally-exact homogenization theory for evaluation of electric conductivity and resistance of multiphase materials
  3. Enhancing ultra-early strength of sulphoaluminate cement-based materials by incorporating graphene oxide
  4. Characterization of mechanical properties of epoxy/nanohybrid composites by nanoindentation
  5. Graphene and CNT impact on heat transfer response of nanocomposite cylinders
  6. A facile and simple approach to synthesis and characterization of methacrylated graphene oxide nanostructured polyaniline nanocomposites
  7. Ultrasmall Fe3O4 nanoparticles induce S-phase arrest and inhibit cancer cells proliferation
  8. Effect of aging on properties and nanoscale precipitates of Cu-Ag-Cr alloy
  9. Effect of nano-strengthening on the properties and microstructure of recycled concrete
  10. Stabilizing effect of methylcellulose on the dispersion of multi-walled carbon nanotubes in cementitious composites
  11. Preparation and electromagnetic properties characterization of reduced graphene oxide/strontium hexaferrite nanocomposites
  12. Interfacial characteristics of a carbon nanotube-polyimide nanocomposite by molecular dynamics simulation
  13. Preparation and properties of 3D interconnected CNTs/Cu composites
  14. On factors affecting surface free energy of carbon black for reinforcing rubber
  15. Nano-silica modified phenolic resin film: manufacturing and properties
  16. Experimental study on photocatalytic degradation efficiency of mixed crystal nano-TiO2 concrete
  17. Halloysite nanotubes in polymer science: purification, characterization, modification and applications
  18. Cellulose hydrogel skeleton by extrusion 3D printing of solution
  19. Crack closure and flexural tensile capacity with SMA fibers randomly embedded on tensile side of mortar beams
  20. An experimental study on one-step and two-step foaming of natural rubber/silica nanocomposites
  21. Utilization of red mud for producing a high strength binder by composition optimization and nano strengthening
  22. One-pot synthesis of nano titanium dioxide in supercritical water
  23. Printability of photo-sensitive nanocomposites using two-photon polymerization
  24. In situ synthesis of expanded graphite embedded with amorphous carbon-coated aluminum particles as anode materials for lithium-ion batteries
  25. Effect of nano and micro conductive materials on conductive properties of carbon fiber reinforced concrete
  26. Tribological performance of nano-diamond composites-dispersed lubricants on commercial cylinder liner mating with CrN piston ring
  27. Supramolecular ionic polymer/carbon nanotube composite hydrogels with enhanced electromechanical performance
  28. Genetic mechanisms of deep-water massive sandstones in continental lake basins and their significance in micro–nano reservoir storage systems: A case study of the Yanchang formation in the Ordos Basin
  29. Effects of nanoparticles on engineering performance of cementitious composites reinforced with PVA fibers
  30. Band gap manipulation of viscoelastic functionally graded phononic crystal
  31. Pyrolysis kinetics and mechanical properties of poly(lactic acid)/bamboo particle biocomposites: Effect of particle size distribution
  32. Manipulating conductive network formation via 3D T-ZnO: A facile approach for a CNT-reinforced nanocomposite
  33. Microstructure and mechanical properties of WC–Ni multiphase ceramic materials with NiCl2·6H2O as a binder
  34. Effect of ball milling process on the photocatalytic performance of CdS/TiO2 composite
  35. Berberine/Ag nanoparticle embedded biomimetic calcium phosphate scaffolds for enhancing antibacterial function
  36. Effect of annealing heat treatment on microstructure and mechanical properties of nonequiatomic CoCrFeNiMo medium-entropy alloys prepared by hot isostatic pressing
  37. Corrosion behaviour of multilayer CrN coatings deposited by hybrid HIPIMS after oxidation treatment
  38. Surface hydrophobicity and oleophilicity of hierarchical metal structures fabricated using ink-based selective laser melting of micro/nanoparticles
  39. Research on bond–slip performance between pultruded glass fiber-reinforced polymer tube and nano-CaCO3 concrete
  40. Antibacterial polymer nanofiber-coated and high elastin protein-expressing BMSCs incorporated polypropylene mesh for accelerating healing of female pelvic floor dysfunction
  41. Effects of Ag contents on the microstructure and SERS performance of self-grown Ag nanoparticles/Mo–Ag alloy films
  42. A highly sensitive biosensor based on methacrylated graphene oxide-grafted polyaniline for ascorbic acid determination
  43. Arrangement structure of carbon nanofiber with excellent spectral radiation characteristics
  44. Effect of different particle sizes of nano-SiO2 on the properties and microstructure of cement paste
  45. Superior Fe x N electrocatalyst derived from 1,1′-diacetylferrocene for oxygen reduction reaction in alkaline and acidic media
  46. Facile growth of aluminum oxide thin film by chemical liquid deposition and its application in devices
  47. Liquid crystallinity and thermal properties of polyhedral oligomeric silsesquioxane/side-chain azobenzene hybrid copolymer
  48. Laboratory experiment on the nano-TiO2 photocatalytic degradation effect of road surface oil pollution
  49. Binary carbon-based additives in LiFePO4 cathode with favorable lithium storage
  50. Conversion of sub-µm calcium carbonate (calcite) particles to hollow hydroxyapatite agglomerates in K2HPO4 solutions
  51. Exact solutions of bending deflection for single-walled BNNTs based on the classical Euler–Bernoulli beam theory
  52. Effects of substrate properties and sputtering methods on self-formation of Ag particles on the Ag–Mo(Zr) alloy films
  53. Enhancing carbonation and chloride resistance of autoclaved concrete by incorporating nano-CaCO3
  54. Effect of SiO2 aerogels loading on photocatalytic degradation of nitrobenzene using composites with tetrapod-like ZnO
  55. Radiation-modified wool for adsorption of redox metals and potentially for nanoparticles
  56. Hydration activity, crystal structural, and electronic properties studies of Ba-doped dicalcium silicate
  57. Microstructure and mechanical properties of brazing joint of silver-based composite filler metal
  58. Polymer nanocomposite sunlight spectrum down-converters made by open-air PLD
  59. Cryogenic milling and formation of nanostructured machined surface of AISI 4340
  60. Braided composite stent for peripheral vascular applications
  61. Effect of cinnamon essential oil on morphological, flammability and thermal properties of nanocellulose fibre–reinforced starch biopolymer composites
  62. Study on influencing factors of photocatalytic performance of CdS/TiO2 nanocomposite concrete
  63. Improving flexural and dielectric properties of carbon fiber epoxy composite laminates reinforced with carbon nanotubes interlayer using electrospray deposition
  64. Scalable fabrication of carbon materials based silicon rubber for highly stretchable e-textile sensor
  65. Degradation modeling of poly-l-lactide acid (PLLA) bioresorbable vascular scaffold within a coronary artery
  66. Combining Zn0.76Co0.24S with S-doped graphene as high-performance anode materials for lithium- and sodium-ion batteries
  67. Synthesis of functionalized carbon nanotubes for fluorescent biosensors
  68. Effect of nano-silica slurry on engineering, X-ray, and γ-ray attenuation characteristics of steel slag high-strength heavyweight concrete
  69. Incorporation of redox-active polyimide binder into LiFePO4 cathode for high-rate electrochemical energy storage
  70. Microstructural evolution and properties of Cu–20 wt% Ag alloy wire by multi-pass continuous drawing
  71. Transparent ultraviolet-shielding composite films made from dispersing pristine zinc oxide nanoparticles in low-density polyethylene
  72. Microfluidic-assisted synthesis and modelling of monodispersed magnetic nanocomposites for biomedical applications
  73. Preparation and piezoresistivity of carbon nanotube-coated sand reinforced cement mortar
  74. Vibrational analysis of an irregular single-walled carbon nanotube incorporating initial stress effects
  75. Study of the material engineering properties of high-density poly(ethylene)/perlite nanocomposite materials
  76. Single pulse laser removal of indium tin oxide film on glass and polyethylene terephthalate by nanosecond and femtosecond laser
  77. Mechanical reinforcement with enhanced electrical and heat conduction of epoxy resin by polyaniline and graphene nanoplatelets
  78. High-efficiency method for recycling lithium from spent LiFePO4 cathode
  79. Degradable tough chitosan dressing for skin wound recovery
  80. Static and dynamic analyses of auxetic hybrid FRC/CNTRC laminated plates
  81. Review articles
  82. Carbon nanomaterials enhanced cement-based composites: advances and challenges
  83. Review on the research progress of cement-based and geopolymer materials modified by graphene and graphene oxide
  84. Review on modeling and application of chemical mechanical polishing
  85. Research on the interface properties and strengthening–toughening mechanism of nanocarbon-toughened ceramic matrix composites
  86. Advances in modelling and analysis of nano structures: a review
  87. Mechanical properties of nanomaterials: A review
  88. New generation of oxide-based nanoparticles for the applications in early cancer detection and diagnostics
  89. A review on the properties, reinforcing effects, and commercialization of nanomaterials for cement-based materials
  90. Recent development and applications of nanomaterials for cancer immunotherapy
  91. Advances in biomaterials for adipose tissue reconstruction in plastic surgery
  92. Advances of graphene- and graphene oxide-modified cementitious materials
  93. Theories for triboelectric nanogenerators: A comprehensive review
  94. Nanotechnology of diamondoids for the fabrication of nanostructured systems
  95. Material advancement in technological development for the 5G wireless communications
  96. Nanoengineering in biomedicine: Current development and future perspectives
  97. Recent advances in ocean wave energy harvesting by triboelectric nanogenerator: An overview
  98. Application of nanoscale zero-valent iron in hexavalent chromium-contaminated soil: A review
  99. Carbon nanotube–reinforced polymer composite for electromagnetic interference application: A review
  100. Functionalized layered double hydroxide applied to heavy metal ions absorption: A review
  101. A new classification method of nanotechnology for design integration in biomaterials
  102. Finite element analysis of natural fibers composites: A review
  103. Phase change materials for building construction: An overview of nano-/micro-encapsulation
  104. Recent advance in surface modification for regulating cell adhesion and behaviors
  105. Hyaluronic acid as a bioactive component for bone tissue regeneration: Fabrication, modification, properties, and biological functions
  106. Theoretical calculation of a TiO2-based photocatalyst in the field of water splitting: A review
  107. Two-photon polymerization nanolithography technology for fabrication of stimulus-responsive micro/nano-structures for biomedical applications
  108. A review of passive methods in microchannel heat sink application through advanced geometric structure and nanofluids: Current advancements and challenges
  109. Stress effect on 3D culturing of MC3T3-E1 cells on microporous bovine bone slices
  110. Progress in magnetic Fe3O4 nanomaterials in magnetic resonance imaging
  111. Synthesis of graphene: Potential carbon precursors and approaches
  112. A comprehensive review of the influences of nanoparticles as a fuel additive in an internal combustion engine (ICE)
  113. Advances in layered double hydroxide-based ternary nanocomposites for photocatalysis of contaminants in water
  114. Analysis of functionally graded carbon nanotube-reinforced composite structures: A review
  115. Application of nanomaterials in ultra-high performance concrete: A review
  116. Therapeutic strategies and potential implications of silver nanoparticles in the management of skin cancer
  117. Advanced nickel nanoparticles technology: From synthesis to applications
  118. Cobalt magnetic nanoparticles as theranostics: Conceivable or forgettable?
  119. Progress in construction of bio-inspired physico-antimicrobial surfaces
  120. From materials to devices using fused deposition modeling: A state-of-art review
  121. A review for modified Li composite anode: Principle, preparation and challenge
  122. Naturally or artificially constructed nanocellulose architectures for epoxy composites: A review
Downloaded on 4.11.2025 from https://www.degruyterbrill.com/document/doi/10.1515/ntrev-2020-0079/html?lang=en
Scroll to top button