You are on page 1of 10

Research Collection

Journal Article

Luminescent solar concentrators based on melt-spun polymer


optical fibers

Author(s):
Jakubowski, Konrad; Huang, Chieh-Szu; Gooneie, Ali; Boesel, Luciano F.; Heuberger, Manfred; Hufenus,
Rudolf

Publication Date:
2020-04

Permanent Link:
https://doi.org/10.3929/ethz-b-000396538

Originally published in:


Materials & Design 189, http://doi.org/10.1016/j.matdes.2020.108518

Rights / License:
Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International

This page was generated automatically upon download from the ETH Zurich Research Collection. For more
information please consult the Terms of use.

ETH Library
Materials and Design 189 (2020) 108518

Contents lists available at ScienceDirect

Materials and Design

journal homepage: www.elsevier.com/locate/matdes

Luminescent solar concentrators based on melt-spun polymer


optical fibers
Konrad Jakubowski a,b, Chieh-Szu Huang c,d, Ali Gooneie a, Luciano F. Boesel c,
Manfred Heuberger a,b, Rudolf Hufenus a,⁎
a
Empa, Swiss Federal Laboratories for Materials Science and Technology, Laboratory of Advanced Fibers, Lerchenfeldstrasse 5, St Gallen 9014, Switzerland
b
Department of Materials, ETH Zurich, 8092 Zurich, Switzerland
c
Empa, Swiss Federal Laboratories for Materials Science and Technology, Laboratory of Biomimetic Membranes and Textiles, Lerchenfeldstrasse 5, St Gallen 9014, Switzerland
d
Laboratory of Inorganic Chemistry, ETH Zurich, 8092 Zurich, Switzerland

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• Bi-component melt spinning was ex-


plored as a novel way to manufacture
fiber luminescent solar concentrators
(LSCs).
• Results revealed that solubility and dis-
persion of luminescent dye in polymers
play an important role in LSCs' perfor-
mance.
• Presented LSCs are a cost-effective solu-
tion to enhance the illumination-angle
dependence of solar cell power produc-
tion.
• Due to high industrialization potential
and overall performance, such fibers
may find their way into future
applications.

a r t i c l e i n f o a b s t r a c t

Article history: Luminescent solar concentrators (LSCs) collect incoming sunlight and direct it to a smaller-area photovoltaic cell.
Received 16 December 2019 In the presented work, form factor and illumination angle-dependent performance of LSCs consisting of bi-
Received in revised form 20 January 2020 component melt-spun fibers is demonstrated. Three thermoplastic polymers act as dispersing host material for
Accepted 21 January 2020
the luminescent dye Lumogen Red 305 (LR305). Molecular dynamics simulations provide numerical access to
Available online 22 January 2020
Hildebrand solubility parameters, which are an estimate for the mixing compatibility of dye with polymer matrix.
Keywords:
Actual emission intensity measurements from material samples are compared to Monte Carlo ray tracing simu-
Luminescent solar concentrators lations. Some samples show an increased absorption, which led to the hypothesis that there exist optically pas-
Polymer optical fibers sive dye aggregates if the dispersion is not optimal. The best-performing polymer/dye pair is identified and used
Energy harvesting to melt-spin fibers. Geometrically defined bundles of LSC fibers are studied in a scenario of white light illumina-
Melt spinning tion and variation of illumination-angle. This experiment simulates a theoretical daily course-of-sun illumination
in absence of atmospheric effects. We report optical conversion efficiencies of the prepared LSCs between 2% and
15%, depending on illumination angle and bundle geometry.
© 2020 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://
creativecommons.org/licenses/by-nc-nd/4.0/).

⁎ Corresponding author.
E-mail address: rudolf.hufenus@empa.ch (R. Hufenus).

https://doi.org/10.1016/j.matdes.2020.108518
0264-1275/© 2020 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
2 K. Jakubowski et al. / Materials and Design 189 (2020) 108518

1. Introduction In this study, some key material factors pertaining to the develop-
ment of melt-spun fiber LSCs are considered. Three polymers, namely
Current photovoltaic (PV) cells exhibit efficiencies in the range 10% cycloolefin polymer (COP), polycarbonate (PC) and PMMA were evalu-
to 40% [1]. To increase PV system power, large-area panels are required. ated as possible host materials for the here fixed choice of LR305 dye. To
The costs scale essentially linearly with PV area [2], and the maximum assess luminescent properties in these host polymers, we prepared dye-
power achieved is often limited by the available system integration doped polymer plates. Monte-Carlo ray tracing and molecular dynamics
area (e.g. roof top) [3,4]. The output power of planar solar cells depends simulations were used to link molecular properties to the plate geome-
on the cosine of the illumination angle, manifesting itself as a midday try and results were validated experimentally. Prototype melt-spun bi-
power peak, which has to be embraced by the power grid [5]. Lumines- component LSC fibers were manufactured using a doped COP core and
cent solar concentrators (LSCs) were first proposed in 1976 to enhance low-refractive index sheath material, namely terpolymer of tetrafluoro-
the performance of PV solar cells. In practical use, a LSC must not only be ethylene, hexafluoropropylene and vinylidene fluoride (THVP). The
efficient in collecting photons at fixed illumination angle, but ideally illumination-angle dependent performance in absence of atmospheric
during the course of a whole day. Furthermore, the power gain by effects (such as scattering) was determined for defined bundles of
LSCs should be cost-effective compared to the PV panel's price per such prototype LSC fibers.
area. Polymer optical fibers (POFs) provide a suitable form factor to ad-
dress those requirements [6]. The favorable impact of the fiber geome-
try becomes apparent in terms of the LSC gain factor F(α, λ), which is 2. Materials and methods
defined as the ratio of power generated by a solar cell with/without
LSC [7,8]: 2.1. Materials

P LSC Three polymers were tested as host materials for LR305 in the fiber
F ðα; λÞ ¼ ¼ ηopt ðα; λÞ  Gðα Þ  ηcell ðλÞ ð1Þ core: COP granulate (Zeonor 1020R, refractive index 1.53) was pur-
P cell
chased from Zeon Europe GmbH, Germany; PC granulate (Sabic Lexan
where ηopt(α, λ) signifies the optical conversion efficiency [9,10]. We 103R, refractive index 1.58) was purchased from Lenorplastics AG,
note that for a fixed illumination, ηopt(λ) contains the optical attenua- Switzerland; PMMA granulate (Plexiglas 7 N, refractive index 1.49)
tion in the fiber. The factor ηcell(λ) is the wavelength-dependent effi- was purchased from Evonik Industries AG, Switzerland. In addition, ter-
ciency of the solar cell and G(α) is the angle-dependent geometrical polymer of tetrafluoroethylene, hexafluoropropylene and vinylidene
concentration factor, i.e. the ratio of the illumination angle dependent fluoride granulate (THVP 2030GZ refractive index 1.35) was purchased
collection surface (projected LSC shadow area), divided by the LSC out- from 3 M Company, Germany, and used as a low-refractive-index, me-
put area (out-coupling area facing photovoltaic cell). chanically stable fiber sheath. The dye molecule LR305 was purchased
from BASF SE, Germany. All materials were used without further purifi-
Gðα Þ ¼ Acollection ðα Þ=Aoutput ð2Þ cation. The chemical structures of the materials are shown in Fig. S1 in
the supporting information.
For illumination at a normal angle, the G-factor of a fiber is maximal
and it is also proportional to the fiber aspect ratio, underlining the ad- 2.2. Melt-processing of materials
vantage of the length-scalable fiber form factor [11].
Synthetic fibers can be produced in large quantity, which speaks for Polymer granulates were first dried overnight in a vacuum oven at
the economic benefit and feasibility of the proposed approach. Melt- elevated temperatures. LR305 powder was then added to the dried
spinning is a widely used technique for production of textile fibers granulates of the fiber core polymer. To assure a uniform distribution
[12], and we explore an adaptation for manufacturing fiber LSCs. of the dye, the granulate/powder mixtures were tumbled for 5 h. To pre-
Namely, bi-component melt-spinning enables a continuous, single- pare thin plates, the pre-mixed samples were melt-mixed at their re-
step production of polymer optical fibers (POFs) [13]. The classical spective processing temperatures for 2 min in a torque rheometer
step-index POF consists of a light-conducting core, coaxially embedded (HAAKE™ Rheomix OS Lab Mixer). The solidified extrudate collected
inside a lower-refractive-index cladding. In-situ spinning with the clad- after melt-mixing was grinded, and compression molded into 1 mm-
ding in place, elegantly removes the need for a solution coating process. thick plates with a hot press (Lindenberg Technics).
Bi-component spinning is solvent-free and represents a cost-effective The fiber LSCs was melt-spun in a custom-made pilot plant de-
production method [14,15]. scribed elsewhere [30]. The sheath (cladding) material and the core
The most common polymer used in the design, production and re- polymer melt were fed from two separate extruders. Metering pumps
search of LSCs is polymethyl metacrylate (PMMA). Several other trans- transferred the melts into the spin pack, and the bi-component fiber
parent thermoplastics are available [16–18], which may offer a variation exited the spinneret into the quenching chamber, where it air-cooled
of mechanical or optical properties. Luminescent light converting dyes and solidified. Finally, the POF was taken up, drawn by heated godets,
constitute a fast growing research and application field. Many different and spooled onto a bobbin. Adapted godet speeds defined the optimal
luminophores are being studied as sunlight-converting additives, such tension for each fiber type. The nominal throughput was 150 m of con-
as organic dyes, nanocrystals or quantum dots [19–23]. Lumogen Red tinuous fiber LSC per minute. More details on melt-processing, with
305 (LR305) is a prominent example: this commercially available dye exact processing temperatures and godet speeds, can be found in the
has a close-to-unity quantum yield and an excellent thermal stability Supporting Information.
up to 370 °C – a suitable additive for melt-spinning.
A careful selection of the polymer host-luminescent dye pairing is
crucial to help dispersion of luminescent molecules. The self- 2.3. Optical characterization of plates
absorption parameter [10,24] describes how the light emitted from a
dye molecule is reabsorbed by adjacent dye molecules along the light Absorption of the plates was measured using a UV–Vis spectropho-
path. Likewise, a low solubility of the dye in the host polymer melt tometer (Agilent Cary 4000 UV/Vis). Emission intensity was measured
may lead to dye aggregation, which leads to suppressed light emission from the plates using a spectrofluorometer (Horiba FluoroMax) with a
via luminescence quenching [25,26]. While the dispersion of dye mole- custom-made sample holder. Illumination and detection were set up
cules in liquid solvents is well understood [27–29], solubility of dye in at perpendicular sample faces. Each measurement was repeated three
polymer melts has received less systematic attention. times.
K. Jakubowski et al. / Materials and Design 189 (2020) 108518 3

2.4. Performance of melt-spun bi-component LSCs used in the experiment. For the presented analysis, the histograms
were fitted using Origin 2018.
To study optical behavior of individual fiber LSCs, an established
setup was used [31]. A glass optical fiber (Thorlabs M28L02) coupled 2.7. Molecular dynamics simulations
to a green LED light source (Thorlabs M505F3), illuminating the fiber
from the side, was translated along the fiber axis. At each position, the Fully atomistic molecular dynamics (MD) simulations were per-
emitted light at the fiber's tip was collected by another glass optical formed in order to calculate the Hildebrand solubility parameters of dif-
fiber (Thorlabs M59L01), which was coupled to a mini-spectrometer ferent molecules of interest from the cohesive energy density (CED)
(Hamamatsu C10083MD). [33–35]. Boxes containing 100 molecules were generated using the
Amorphous Cell module of the BIOVIA Materials Studio simulation soft-
ware [36]. The molecules were allowed to interact with each other using
2.5. Performance of LSC bundles
the COMPASS force field of the Forcite module of Materials Studio. Prior
to CED calculations, the simulation box was equilibrated in the NPT en-
Melt-spun fibers were combined into defined bundles (Fig. 5b): each
semble for 2–7 ns to reach equilibrium at 298 K and 1 atm.
bundle consisted of 200 fibers. To define their arrangement, the fibers
were inserted into transparent PMMA tubes (Rohm AG, Switzerland)
3. Results and discussion
of 5 mm inner and 7 mm outer diameter. The fibers covered 85% of
the cross-sectional area of the tube's lumen. The tubed bundles had
3.1. Luminescent properties of the dye-polymer systems
lengths of 20 mm, 40 mm and 80 mm. A calibrated photodetector
head with input optics 11 mm diameter diffusor window and spectral
In order to examine how different polymer hosts influence the lumi-
responsivity between 400 and 1000 nm (RW-3705-2, Gigahertz-Optik
nescence of LR305, the optical emission was measured from dye-doped
GmbH, Germany), attached to an optometer (P 9710, Gigahertz-Optik
polymer plates. PMMA, PC and COP plates were prepared with different
GmbH, Germany) was placed on a rotational stage. The sensitive surface
LR305 concentrations, as summarized in Table 1.
was masked with a black, non-transparent plate, leaving a circular
Absorbance spectra of three host materials (COP, PMMA and PC)
opening of diameter corresponding to that of the inner diameter of
with 0.05 wt% LR305 concentration were measured to determine the
the tubed fiber bunches. The rotational stage with the sensor resided in-
absorption coefficients for each wavelength required for ray-tracing
side a black box to prevent illumination by stray light. A Xenon short-arc
simulations described later in the paper, as well as potential self-
lamp (OSRAM XBO 450 W OFR (6000 K)) was used as a white light
absorption region of the LR305. Fig. 1 displays those results.
source, placed 3 m from the detector to approximate a plane wave
To quantify the emission strength of each of the host materials with
front, reducing square-distance intensity artefacts to b2%. Light inten-
different quantities of LR305, the emission spectra of all prepared plates
sity was measured with the masked photodetector and with perpendic-
were measured and comparatively shown in Fig. 2.
ular tubed fiber bundles on the circular opening. The detector in both
To avoid spectral overlap between emission spectra and excitation
configurations was rotated, and the intensity was recorded for a selec-
beam, the plates were excited by the fixed wavelength 450 nm. This
tion of illumination angles. Each measurement was repeated 3 times.
wavelength lies far from the expected emission and exhibits similar ab-
In addition to these detector measurements, a PV solar cell with dimen-
sorbance for LR305 in the three host materials (Fig. 1). The illumination
sions of 46 mm by 40 mm (Velleman SOL3N Polycrystalline Solar Mod-
was perpendicular to the plate's surface, and measured in transmission
ule 1 V, Conrad Electronic AG, Switzerland) was used for similar
at the same angle. As can be seen in Fig. 1, different plates exhibit a sim-
measurements. The solar cell was also masked with a black, non-
ilar absorbance at that wavelength. For each of the host materials, two
transparent foil, with a circular opening of diameter corresponding to
prominent emission peaks are visible (Fig. 2): namely at ≈590 nm
that of the inner diameter of the tubed fiber bunches. The spectral sen-
and at ≈640 nm. A third, significantly smaller peak is present at
sitivity of the PV solar cell is more closely representing a real application,
≈675 nm, which can be deconvoluted by means of more detailed
i.e. higher sensitivity in red and NIR. The mask was placed identically on
peak analysis. For PMMA and PC samples, both emission and absorption
the cell for all measurements to exclude variations of internal resistance
spectra seem red-shifted with respect to COP samples – this effect oc-
from the shadowed area. The illumination-angle dependent power was
curs due to the polarity of the host material [37]. It can also be observed,
measured, in the same setup as in the previous case, but now the max-
that the position of the first peak shifts towards longer wavelengths.
imum power point was followed on the I(V) curve. Each measurement
To gain additional insight, Fig. 3 displays the integral spectral area for
was repeated 3 times. More details about these measurements can be
all the measured thin-plate samples, calculated from the data in Fig. 2.
found in the SI.
We observe that the LR305 yields a higher emission intensity in the
COP matrix than in PMMA or PC. In addition, the position of the first
2.6. Monte-Carlo ray tracing simulations
Table 1
An open-source, Python-based ray-tracing software PVTRACE (ver- LR305 concentration in the prepared melt-processed thin plates.
sion 2.0.4) [32] was used to track random photons inside each of the
Sample LR305 concentration [wt%] LR305 concentration [μmol/cm3]
prepared polymer plates. The measured absorption and emission spec-
tra of LR305/polymer host systems were used as input data for these COP0 0.002 0.02
calculations. For the simulation, a point light source was placed inside COP1 0.01 0.09
COP2 0.05 0.47
the plate geometry next to the illuminated plate/air interface. The mo- COP3 0.07 0.66
lecular emission spectrum in the simulation corresponded to the previ- COP4 0.1 0.94
ously measured emission spectrum of LR305. For each simulation, COP5 0.25 2.34
250,000 random photons were generated. Each generated photon's ini- PMMA0 0.002 0.02
PMMA1 0.05 0.55
tial wavelength was randomly selected based on the defined emission
PMMA2 0.1 1.10
spectrum. The simulation accounts for photons that undergo different PMMA3 0.25 2.76
optical processes inside the geometry of the LSC plate (like self- PC0 0.002 0.02
absorption, re-emission, absorption by the polymer matrix, etc.). The PC1 0.05 0.56
resulting distribution of emitted photon wavelengths was statistically PC2 0.1 1.11
PC3 0.25 2.78
analyzed in a histogram (Fig. 4), which imitates the spectral analysis
4 K. Jakubowski et al. / Materials and Design 189 (2020) 108518

1.2
2.0 COP2 COP
PMMA1 PMMA
PC1 1.0
PC

Normalized intensity
1.5
Absorbance [a.u]

0.8

1.0 0.6

0.4
0.5

0.2

0.0
400 450 500 550 600 650 700 0.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Wavelength [nm] 3
Concentration [Pmol/cm ]
Fig. 1. Absorption spectra (with standard deviation) of COP2, PMMA1 and PC1 (all with
0.05 wt% LR305). The results indicate good absorption of blue, green and a small portion Fig. 3. Normalized integrated emission intensities of the measured thin-plate samples.
of red light by LR305. (For interpretation of the references to colour in this figure legend,
the reader is referred to the web version of this article.)
simulated and experimental results is thus summarized in Table 2.
peak apparently shifts towards longer wavelength with increasing dye One can readily recognize a trend; namely, that theory predicts higher
concentration (Fig. 2), which is the result of a marked self-absorbance peak ratios—an effect most prominent for the PC, and least prominent
in this wavelength range (Fig. 1). As expected, the position of the second for the COP sample.
and the third peak remain unchanged, in agreement with negligible Table 2 and Fig. 4 indicate the presence of additional attenuation ef-
self-absorbance above 650 nm. fects not accounted for in the simulation; these effects depend on the
We used Monte Carlo ray-tracing simulations to connect the previ- host material.
ously measured emission data to the light path in this plate geometry. Collected results thus suggest there is an excess of self-absorption,
Normalized emission spectra of COP0, PMMA0 and PC0 (0.002 wt% which is more pronounced in PMMA and PC samples. It is reasonable
LR305 each) were used as reference emission spectra in the simulation, to hypothesize that aggregation-induced quenching of the luminescent
since they were measured at minimal self-absorption (i.e. at low LR305 dye molecules is the likely cause; it can explain the observed depen-
concentration). The absorption behavior of the plates with 0.05 wt% dence on the host polymer matrix.
LR305 (COP2, PMMA1 and PC1) was then simulated, assuming the pre- To investigate this hypothesis, we performed molecular dynamics
viously measured absorption behavior (Fig. 1). Fig. 4 shows a compari- (MD) simulations on pure COP and LR305 individually to compute
son between simulated and measured emission spectra of these their respective Hildebrand solubility parameters, δ. This parameter
samples. represents the cohesive energy density for each pure phase. The square
The simulated emission spectra reproduce the peak positions and of the difference between them, Δ = (δA − δB)2, is a simple estimate of
peak shapes rather well. A comparison reveals, however, that the simu- the dispersion quality. A lower parameter Δ would indicate a better dis-
lation predicts a relatively stronger first emission peak. For quantification persion of the dye in the polymer matrix [38]. For the MD computation,
of this effect, peak area ratios are calculated using the following relation: the structures of COP and LR305 were taken from previous publications
[39,40]. Hildebrand solubility parameters of pure PMMA and PC were
Afirst peak taken from literature [41]. The MD simulation is described in more de-
Rpeak area ¼ ð3Þ
Asecond peak þ Athird peak tail in the Experimental Section. Table 3 summarizes the solubility pa-
rameters and Δ values.
This parameter estimates the intensity of the first emission peak (i.e. The predicted best matrix for the dispersion of LR305 is indeed COP.
the one most affected by die self-absorption) relative to the other two Therefore, we would expect less dye aggregates in COP than in PMMA or
peaks (unaffected by self-absorption). The comparison of fitted PC [25,26].

Fig. 2. Luminescence spectra of a) COP plates, b) PMMA plates and c) PC plates with different LR305 concentrations (see: Table 1), measured perpendicular to the plates.
K. Jakubowski et al. / Materials and Design 189 (2020) 108518 5

Fig. 4. Simulated (gray histogram) and measured (red curve) emission spectra of a) COP2 b) PMMA1 c) PC1 (all with 0.05 wt% LR305) in the plate geometry. (For interpretation of the
references to colour in this figure legend, the reader is referred to the web version of this article.)

We think the systematically higher absorption of the first peak ob- 3.3. Characterization of multi-fiber bundle LSCs
served in the previously simulated spectra (Fig. 4, Table 3) can be ex-
plained by aggregate formation, particularly since this 600 nm-peak Fiber bundles were prepared to test their light concentrating perfor-
would be selectively affected by absorption from non-emitting dye ag- mance with respect to the illumination angle. Bundles of three different
gregates. The intensity ratio (“sim/exp” in Table 2) can thus be regarded lengths were used, namely 20 mm, 40 mm and 80 mm (Fig. 5b). The
as a simple measure for dye-aggregate formation in the polymer matrix. maximum length was chosen in accordance to the previously presented
attenuation measurements for a single fiber. To preserve a well-defined
orientation of the 200 fibers as part of a 5 mm diameter bundle, they
3.2. Properties of melt-spun, bi-component fiber LSCs were held together by transparent PMMA tubes. Due to geometric con-
straints, a portion of ~62% of the tube area was covered with emitting
Based on previous results, a continuous bi-component fiber LSC was fiber cores. The photodetector used for intensity measurements was
produced on a pilot melt-spinning plant with COP as a core and THVP as masked with a fixed circular aperture matching the inner diameter of
a cladding material. The production rate was 150 m per minute. The ob- the PMMA tubes. A Xe-arc lamp was used as an excitation light source;
tained fibers exhibit a numerical aperture of 0.72. The LR305 dye con- it exhibits a quasi-white light illumination that approximates the solar
centration was 0.05 wt% in the core – this value was found to have the spectrum. A calibrated photodetector was rotated on a stage to vary
highest emission intensity (Fig. 4). More details on the continuous the relative direction of the incoming light and thus simulate different
fiber production can be found in the Experimental section as well as in sun positions on the sky with respect to the receiver, disregarding atmo-
the Supporting Information. Fig. 5a shows a cross-section of the studied spheric effects such as Rayleigh scattering. We adapt the following con-
melt-spun bi-component fiber LSC – its core radius was 139 ± 1μm, vention: angle α = 0° for the illumination parallel to the surface of the
with an outer radius of 163 ± 2μm, amounting to a core portion of detector and α = 90° for the illumination perpendicular to the surface
73 ± 1vol%. This geometry allowed easy handling and characterization. of the detector. Fig. 7a compares the recorded intensities at different
Melt-spinning provides excellent uniformity of produced fiber LSCs. bundle lengths as well as the masked bare sensor.
The hence obtained fiber LSC was studied with respect to the spectral The highest power was measured for the bare photodetector (PD) at
characteristics of the emitted light. Therefore, a point light source was perpendicular illumination (α = 90 °). We note that the measured in-
translated along the fiber axis, illuminating it from the side and the tensity exhibits a distinctively different angular dependence in presence
light emitted at the tip was recorded using a fiber-coupled spectrome- of the fiber bundles. Namely, at shallower illumination angles, the inten-
ter. The distance dependence of the emitted spectrum in Fig. 6a served sity recorded for bundles is higher, as the incoming light is received by a
as basis to calculate a wavelength-dependent attenuation as displayed larger projected area of the LSC bundle, which is expressed by the fol-
in Fig. 6b, described in more detail the SI. lowing equation:
As seen in Fig. 6a, the spectrum maxima apparently shift towards
longer wavelength as the distance to the illumination point increases; Acollection ðα Þ ¼ L∙2r∙ cosα þ π∙r 2 ∙ sinα ð4Þ
this is, again, the result of dye self-absorption. Light is lost along the
way in an exponential manner, which can be described by an attenua- where L is the length of the cylinder and r is its radius. To study the
tion ratio [dB]. As seen in Fig. 6b, the attenuation spectrum resembles length-dependence of the bundle LSC efficiency from such angle-
the L305 absorption spectrum in COP (Fig. 1). Light in a wavelength dependent measurement, we split the incoming illumination into two
range below 615 nm is attenuated below measurement noise within a orthogonal components:
traveling distance of 1 mm. On the other hand, the light in the range
above 615 nm can travel longer distances. For example, this can be Iðα Þ ¼ A  cosα þ B  sinα ð5Þ
quantified in Fig. 6a (pink curve) where around 16% of the total light
can travel 160 mm to the fiber tip. Additional details about the setup where I(α) is the angle dependent intensity, and A and B are the orthog-
of the optical measurements are available in the SI. onally convoluted contributions containing geometrical and optical

Table 3
Table 2 Hildebrand solubility parameters of luminescent dye and polymer hosts.
Peak area ratios (Rpeak area) for simulated and experimental results.
Material Hildebrand solubility parameter [(J/cm3)0.5] Δ [J/cm3]
Simulation Experiment Ratio (sim/exp)
LR305 17.05 –
COP2 2.37 1.85 1.28 COP 15.47 2.50
PMMA1 1.55 1.14 1.36 PMMA 19.00 3.80
PC1 1.81 1.28 1.42 PC 19.90 8.12
6 K. Jakubowski et al. / Materials and Design 189 (2020) 108518

Fig. 5. a) Microscopic image of a cross-section of the prepared melt-spun bi-component fiber LSC. b) 20 mm, 40 mm and 80 mm PMMA tubes filled with fiber LSCs.

factors influencing the performance of bundle LSC. By fitting the curves Eq. (4). Fig. 8 shows the calculated efficiencies for varying angles α, in
in Fig. 7a to Eq. (5), parameters A and B can be extracted. The hence ob- comparison with previously published references.
tained values are plotted in Fig. 7b for different lengths of fiber bundles. We note that the values obtained for bundle LSCs are comparable to
As expected, the portion A grows with increasing bundle length, due to those reported in other studies [14,15,42–46], underlining the compet-
enhanced collection of lateral light by the LSC. We note that the growth itiveness of the presented bi-component polymer optical fibers. The 80-
of A is sub-linear with bundle length. Fitting the observed asymptotic mm-long bundle, despite yielding the highest absolute intensities
behavior of this component reveals the useful LSC length of around (Fig. 7a), achieves the lowest optical efficiencies among studied config-
80 mm, as imposed by fiber attenuation, presented in Fig. 6b. This max- urations (Fig. 8), while the 20-mm-long bundle performs in the oppo-
imum effective length will depend on dye and agglomerate concentra- site way. The increase of optical efficiency above 80° is presumably
tion. The portion B encompasses a light collection maximum for bare caused by the non-ideal structure of a bundle – as luminescent fibers
detector, and sharply decreases with implementation of bundles. This do not cover the whole cross section, direct light outside the fiber
relates to the attenuated direct transmission of illuminating light cores can reach the detector at α = 90°.
along the bundle axis, i.e. mainly in-coupling at the open fiber tip end. We now assess a more realistic scenario comprising the spectral re-
To study angle-dependent optical conversion within the fiber, we sponse of a polycrystalline silica (pSi) PV solar cell, i.e. additionally
expand the analysis used in literature by, for example, R. Reisfeld et al. benefitting from the conversion of blue into red light. As before, we dis-
[9], E.-H. Banaei et al. [10], S. Correia et al. [42] and I. Parola et al. [43], regard atmospheric effect. Fig. 9 shows the maximal power generated
and calculate the optical efficiency ηopt as a ratio between optical by the pSi cell, again without and with fiber bundles of different lengths,
power produced by bundle LSC, to power received from the illumina- using the same illumination setup as before.
tion, taking into account the effect of varying illumination angle: The power is determined at the maximum power point of the I
(V) curve. The polycrystalline silica solar cells have increased sensitivity
P out ðα Þ Iout ðα Þ∙Aoutput around 760–800 nm. [47] Compared to the white-calibrated photode-
ηopt ðα Þ ¼ ∙100% ¼ ∙100% ð6Þ
P in ðα Þ Iin ∙Acollection ðα Þ tector, the light conversion by LR305 thus explains the excess of
power recorded (viz. Figs. 9 and 7a). This conversion gain is known in
where Pout(α) and Pin(α) are produced and received powers, respec- the LSC literature [48].
tively, Iout(α) and Iin are output and incoming intensities, respectively, For α = 0°, the F-factor (Eq. (1)) reaches values of 2.8, 4.3 and 5.3 for
and Aoutput and Acollection(α) are emitting and collecting areas, respec- bundle lengths of 20 mm, 40 mm, and 80 mm, respectively. To estimate
tively. In the case of cylindrical geometry, such as the bundle LSC pre- the theoretical performance of a LSC-enhanced PV solar cell at varying
sented, the collecting area is changing with the angle following angles (i.e. simulating different arrangements between the sun and

Fig. 6. a) Spectrum of light reaching the fiber LSC's tip for different distances between illumination point and emitting tip. b) Calculated attenuation of light as a function of wavelength. For
wavelengths b 615 nm it was calculated between 0 mm and 2 mm and for wavelengths N 615 nm attenuation was calculated between 0 mm and 160 mm.
K. Jakubowski et al. / Materials and Design 189 (2020) 108518 7

Fig. 7. a) Intensity of light detected by a bare photodetector, as well as a photodetector in contact with LSC bundles of three different lengths. b) Calculated values of A and B deconvolution
parameters for bundles of different length.

cell's surface without considering atmospheric effects), one can inte- polymer melts. The here described methods are suitable to study
grate the measured intensity of Fig. 9 over the full range of angles; other polymer-dye systems. A prototype of an endless melt-spun
Table 4 summarizes the results. bi-component fiber LSC demonstrates the feasibility and up-scaling
We estimate that the pSi cell, enhanced with 40 mm and 80 mm capability of the proposed solution. Fiber-based LSCs could enhance
bundles, can be improved 13% and 26% respectively, compared to the and complement the illumination-angle dependency of commercial
bare PV panel. A beneficial side effect is that power generation is now PV solar panels. More generally, the high values of the F-factors for
distributed more evenly over the angles, as compared to the reference shallow angles suggest that fiber-based LSCs, in form of textiles or
pSi cell without LSC. This angular redistribution property is interesting bundles, could also find photovoltaic applications in more indirect il-
to mitigate the mid-day power peak generated by conventional photo- lumination scenarios. Our results suggest that there is a significant
voltaics with fixed orientation. potential in utilizing industrial melt-spinning as cost-effective
means for fiber-based LSC production.
4. Conclusions

Three transparent thermoplastic polymers, COP, PMMA and PC, CRediT authorship contribution statement
were evaluated as host matrices for LR305. COP was best suited as
host material for LR305, in terms of luminescent properties. Compar- Konrad Jakubowski: Conceptualization, Investigation, Validation,
ison of measured spectra and ray tracing simulations revealed an ex- Visualization, Formal analysis, Writing - original draft. Chieh-Szu
cess absorption in plate model samples that we attribute to the Huang: Investigation, Validation. Ali Gooneie: Methodology, Validation.
formation of dye aggregates in the polymer matrix. Molecular dy- Luciano F. Boesel: Conceptualization, Validation. Manfred Heuberger:
namics simulation were invoked to estimate the interactions be- Conceptualization, Validation, Writing - review & editing, Supervision.
tween luminescent dye and polymer host in melt-processing, Rudolf Hufenus: Conceptualization, Validation, Writing - review &
providing a tool to predict dispersion of (luminescent) additives in editing, Supervision.

100
20mm bundle 1.4
Unassisted cell 8

40mm bundle 20mm bundle power & F-factor


Normalized maximal power

80mm bundle 1.2 40mm bundle power & F-factor 7

Ref. 42 80mm bundle power & F-factor


Calculated F-factor

6
1.0
10
5
Ref. 46
0.8
ηopt [%]

4
0.6
Ref. 15
3
Ref. 14
1
0.4
2

Ref. 44
0.2 1
Ref. 43, Ref. 45
0.0 0
0.1 0 20 40 60 80
0 20 40 60 80 100
Angle D [o]
Angle D [ ] o

Fig. 9. Maximal power registered as a function of the illumination angle α (solid lines) and
Fig. 8. Angle-dependent optical efficiencies of the studied bundle LSCs. Reference values, values of the F-parameter (dashed lines), for a bare PV solar cell and an enhanced cell with
reported previously in the literature, are marked on the graph at α = 0°, as reported. LSC bundles of different lengths, in absence of atmospheric scattering effects.
8 K. Jakubowski et al. / Materials and Design 189 (2020) 108518

Table 4 polymer and structural characteristics revealed by wide angle X-ray diffraction,
Integral of the theoretical maximal power between illumination angles of 0° and 90° of the Polym. (United Kingdom). 55 (2014) 5695–5707, https://doi.org/10.1016/j.
LSC-assisted solar cell, in proportion to the unassisted cell. polymer.2014.08.071.
[14] S.F.H. Correia, A.R. Frias, L. Fu, R. Rondão, E. Pecoraro, S.J.L. Ribeiro, P.S. André, R.A.S.
Measurement Ratio between area under curves (Fig. 9) of Ferreira, L.D. Carlos, Large-area tunable visible-to-near-infrared luminescent solar
configuration LSC-assisted and unassisted cell concentrators, Adv. Sustain. Syst. 1800002 (1–9) (2018) 1800002, https://doi.org/
10.1002/adsu.201800002.
Unassisted solar cell – [15] R.H. Inman, G.V. Shcherbatyuk, D. Medvedko, A. Gopinathan, S. Ghosh, Cylindrical
Cell with 20 mm fiber LSC bundle 0.92 luminescent solar concentrators with near-infrared quantum dots, Opt. Express 19
Cell with 40 mm fiber LSC bundle 1.13 (2011) 24308, https://doi.org/10.1364/OE.19.024308.
Cell with 80 mm fiber LSC bundle 1.26 [16] G. Griffini, Host matrix materials for luminescent solar concentrators: recent
achievements and forthcoming challenges, Front. Mater. 6 (2019) https://doi.org/
10.3389/fmats.2019.00029.
[17] M. Zettl, O. Mayer, E. Klampaftis, B.S. Richards, Investigation of host polymers for lu-
Declaration of competing interest
minescent solar concentrators, Energy Technol 5 (2017) 1037–1044, https://doi.
org/10.1002/ente.201600498.
The authors declare that they have no known competing financial [18] Y. Li, X. Zhang, Y. Zhang, R. Dong, C.K. Luscombe, Review on the role of polymers in
interests or personal relationships that could have appeared to influ- luminescent solar concentrators, J. Polym. Sci. Part A Polym. Chem. 57 (2019)
201–215, https://doi.org/10.1002/pola.29192.
ence the work reported in this paper. [19] K. Wu, H. Li, V.I. Klimov, Tandem luminescent solar concentrators based on
engineered quantum dots, Nat. Photonics 12 (2018) 105–110, https://doi.org/10.
Acknowledgements 1038/s41566-017-0070-7.
[20] A.R. Frias, M.A. Cardoso, A.R.N. Bastos, S.F.H. Correia, P.S. André, L.D. Carlos, V. de Z.
Bermudez, R.A.S. Ferreira, Transparent luminescent solar concentrators using Ln 3
The authors thank Martin Amberg and Urs Schütz for assistance with + -based ionosilicas towards photovoltaic windows, Energies 12 (2019) https://
the solar cell measurement setup, Mathias Lienhard and Benno Wüst for doi.org/10.3390/en12030451.
[21] Y. Zhao, G.A. Meek, B.G. Levine, R.R. Lunt, Near-infrared harvesting transparent lumi-
assistance with polymer melt-processing, Dr. René Rossi for valuable nescent solar concentrators, Adv. Opt. Mater. 2 (2014) 606–611, https://doi.org/10.
discussions, as well as Dr. Sergii Yakunin for assistance with 1002/adom.201400103.
photoluminescent quantum yield measurements. [22] C. Yang, R.R. Lunt, Limits of visibly transparent luminescent solar concentrators, Adv.
Opt. Mater. 5 (2017) 1–10, https://doi.org/10.1002/adom.201600851.
[23] A. Anand, M.L. Zaffalon, G. Gariano, A. Camellini, M. Gandini, R. Brescia, C. Capitani, F.
Data availability Bruni, V. Pinchetti, M. Zavelani-Rossi, F. Meinardi, S.A. Crooker, S. Brovelli, Evidence
for the band-edge exciton of CuInS 2 nanocrystals enables record efficient large-area
luminescent solar concentrators, Adv. Funct. Mater. (2019) 1906629, https://doi.
The raw and processed data required to reproduce these findings
org/10.1002/adfm.201906629.
cannot be shared at this time as the data also forms part of an ongoing [24] L.R. Wilson, B.S. Richards, Measurement method for photoluminescent quantum
study. yields of fluorescent organic dyes in polymethyl methacrylate for luminescent solar
concentrators, Appl. Opt. 48 (2009) 212, https://doi.org/10.1364/AO.48.000212.
[25] Y. Huang, J. Xing, Q. Gong, L.-C. Chen, G. Liu, C. Yao, Z. Wang, H.-L. Zhang, Z. Chen, Q.
Appendix A. Supplementary data Zhang, Reducing aggregation caused quenching effect through co-assembly of PAH
chromophores and molecular barriers, Nat. Commun. 10 (2019) 169, https://doi.
Supplementary data to this article can be found online at https://doi. org/10.1038/s41467-018-08092-y.
[26] R.F. Chen, J.R. Knutson, Mechanism of fluorescence concentration quenching of car-
org/10.1016/j.matdes.2020.108518. boxyfluorescein in liposomes: energy transfer to nonfluorescent dimers, Anal.
Biochem. 172 (1988) 61–77, https://doi.org/10.1016/0003-2697(88)90412-5.
[27] H. Yoo, J. Yang, A. Yousef, M.R. Wasielewski, D. Kim, Excimer formation dynamics of
References intramolecular π-stacked perylenediimides probed by single-molecule fluorescence
spectroscopy, J. Am. Chem. Soc. 132 (2010) 3939–3944, https://doi.org/10.1021/
[1] The National Renewable Energy Laboratory, Best research-cell efficiency chart, ja910724x.
https://www.nrel.gov/pv/cell-efficiency.html 2019, Accessed date: 1 July 2019. [28] R.O. Al-Kaysi, T. Sang Ahn, A.M. Müller, C.J. Bardeen, The photophysical properties of
[2] P.G.V. Sampaio, M.O.A. Gonzalez, Photovoltaic solar energy: conceptual framework, chromophores at high (100 mM and above) concentrations in polymers and as neat
Renew. Sust. Energ. Rev. 74 (2017) 590–601, https://doi.org/10.1016/j.rser.2017.02. solids, Phys. Chem. Chem. Phys. 8 (2006) 3453–3459, https://doi.org/10.1039/
081. b605925b.
[3] F. Meinardi, A. Colombo, K.A. Velizhanin, R. Simonutti, M. Lorenzon, L. Beverina, R. [29] L.R. Wilson, B.C. Rowan, N. Robertson, O. Moudam, A.C. Jones, B.S. Richards, Charac-
Viswanatha, V.I. Klimov, S. Brovelli, Large-area luminescent solar concentrators terization and Reduction of Reabsorption Losses in Luminescent Solar Concentra-
based on stokes-shift-engineered nanocrystals in a mass-polymerized PMMA ma- tors, Vol. 49, 2010 1651–1661.
trix, Nat. Photonics 8 (2014) 392–399, https://doi.org/10.1038/nphoton.2014.54. [30] R. Hufenus, F.A. Reifler, K. Maniura-Weber, A. Spierings, M. Zinn, Biodegradable bi-
[4] M.G. Debije, P.P.C. Verbunt, Thirty years of luminescent solar concentrator research: component fibers from renewable sources: melt-spinning of poly(lactic acid) and
solar energy for the built environment, Adv. Energy Mater. 2 (2012) 12–35, https:// poly[(3-hydroxybutyrate)-co-(3-hydroxyvalerate)], Macromol. Mater. Eng. 297
doi.org/10.1002/aenm.201100554. (2012) 75–84, https://doi.org/10.1002/mame.201100063.
[5] W. Tress, K. Domanski, B. Carlsen, A. Agarwalla, E.A. Alharbi, M. Graetzel, A. Hagfeldt, [31] I. Parola, E. Arrospide, F. Recart, M. Illarramendi, G. Durana, N. Guarrotxena, O.
Performance of perovskite solar cells under simulated temperature-illumination García, J. Zubia, Fabrication and characterization of polymer optical fibers doped
real-world operating conditions, Nat. Energy 4 (2019) 568–574, https://doi.org/10. with perylene-derivatives for fluorescent lighting applications, Fibers 5 (2017) 28,
1038/s41560-019-0400-8. https://doi.org/10.3390/fib5030028.
[6] C.-A. Bunge, T. Gries, M. Beckers, Polymer Optical Fibres, First, Elsevier, 2017https:// [32] D.J. Farrell, D.J. Farrell, pvtrace: optical ray tracing for luminescent materials and
doi.org/10.1016/C2014-0-00562-X. spectral converter photovoltaic devices, https://github.com/danieljfarrell/pvtrace
[7] S.F.H. Correia, P. Lima, E. Pecoraro, S. Ribeiro, P. André, R. Ferreira, L. Carlos, Scale up 2019.
the collection area of luminescent solar concentrators towards metre-length flexible [33] A. Gooneie, P. Simonetti, K.A. Salmeia, S. Gaan, R. Hufenus, M.P. Heuberger, En-
waveguiding photovoltaics, Prog. Photovolt. Res. Appl. 24 (2016) 1178–1193, hanced PET processing with organophosphorus additive: flame retardant products
https://doi.org/10.1002/pip. with added-value for recycling, Polym. Degrad. Stab. 160 (2019) 218–228, https://
[8] R. Reisfeld, D. Shamrakov, C. Jorgensen, Photostable solar concentrators based on doi.org/10.1016/j.polymdegradstab.2018.12.028.
fluorescent glass films, Sol. Energy Mater. Sol. Cells 33 (1994) 417–427, https:// [34] K.A. Salmeia, A. Gooneie, P. Simonetti, R. Nazir, J.-P. Kaiser, A. Rippl, C. Hirsch, S.
doi.org/10.1016/0927-0248(94)90002-7. Lehner, P. Rupper, R. Hufenus, S. Gaan, Comprehensive study on flame retardant
[9] R. Reisfeld, New developments in luminescence for solar energy utilization, Opt. polyesters from phosphorus additives, Polym. Degrad. Stab. 155 (2018) 22–34,
Mater. (Amst). 32 (2010) 850–856, https://doi.org/10.1016/j.optmat.2010.04.034. https://doi.org/10.1016/j.polymdegradstab.2018.07.006.
[10] E.-H. Banaei, A.F. Abouraddy, Design of a polymer optical fiber luminescent solar [35] A. Gooneie, S. Schuschnigg, C. Holzer, A review of multiscale computational methods
concentrator, Prog. Photovolt. Res. Appl. 23 (2015) 403–416, https://doi.org/10. in polymeric materials, Polymers (Basel) 9 (2017) https://doi.org/10.3390/
1002/pip. polym9010016.
[11] K.R. McIntosh, N. Yamada, B.S. Richards, Theoretical comparison of cylindrical and [36] Accelrys Inc, BIOVIA materials studio, http://accelrys.com/, Accessed date: 20 July
square-planar luminescent solar concentrators, Appl. Phys. B Lasers Opt. 88 2018.
(2007) 285–290, https://doi.org/10.1007/s00340-007-2705-8. [37] J.R. Lakowicz, Effects of solvents on fluorescence emission spectra, Princ. Fluoresc.
[12] R. Hufenus, Y. Yan, M. Dauner, D. Yao, T. Kikutani, Biocomponent fibers, Handb. Fi- Spectrosc, Springer US, Boston, MA 1983, pp. 187–215, https://doi.org/10.1007/
brous Mater, 2019. 978-1-4615-7658-7_7.
[13] F.A. Reifler, R. Hufenus, M. Krehel, E. Zgraggen, R.M. Rossi, L.J. Scherer, Polymer op- [38] L.H. Sperling, Introduction to Physical Polymer Science, John Wiley & Sons, Inc., Ho-
tical fibers for textile applications - bicomponent melt spinning from cyclic olefin boken, NJ, USA, 2005https://doi.org/10.1002/0471757128.
K. Jakubowski et al. / Materials and Design 189 (2020) 108518 9

[39] A.A. Leal, J.P. Best, D. Rentsch, J. Michler, R. Hufenus, Spectroscopic elucidation of employing double-doped polymer optical fibers, Sol. Energy Mater. Sol. Cells 178
structure-property relations in filaments melt-spun from amorphous polymers, (2018) 20–28, https://doi.org/10.1016/j.solmat.2018.01.013.
Eur. Polym. J. 89 (2017) 78–87, https://doi.org/10.1016/j.eurpolymj.2017.02.009. [45] R. Rondão, A.R. Frias, S.F.H. Correia, L. Fu, V. De Zea Bermudez, P.S. André, R.A.S.
[40] K.A. Colby, J.J. Burdett, R.F. Frisbee, L. Zhu, R.J. Dillon, C.J. Bardeen, Electronic energy Ferreira, L.D. Carlos, High-performance near-infrared luminescent solar concentra-
migration on different time scales: concentration dependence of the time-resolved tors, ACS Appl. Mater. Interfaces 9 (2017) 12540–12546, https://doi.org/10.1021/
anisotropy and fluorescence quenching of Lumogen red in poly(methyl methacry- acsami.7b02700.
late), J. Phys. Chem. A 114 (2010) 3471–3482, https://doi.org/10.1021/jp910277j. [46] E.-H. Banaei, A.F. Abouraddy, Fiber luminescent solar concentrator with 5.7% conver-
[41] A.F.M. Barton, CRC Handbook of Solubility Parameters and Other Cohesion Parame- sion efficiency, High Low Conc. Syst. Sol. Electr. Appl. VIII. VIII 2013, pp. 1952–1955,
ters, Routledge, 2017https://doi.org/10.1201/9781315140575. https://doi.org/10.1117/12.2022601.
[42] S.F.H. Correia, P.P. Lima, P.S. André, M.R.S. Ferreira, L.A.D. Carlos, High-efficiency lu- [47] A. Pourakbar Saffar, B. Deldadeh Barani, Thermal effects investigation on electrical
minescent solar concentrators for flexible waveguiding photovoltaics, Sol. Energy properties of silicon solar cells treated by laser irradiation, Int. J. Renew. Energy
Mater. Sol. Cells 138 (2015) 51–57, https://doi.org/10.1016/j.solmat.2015.02.032. Dev. 3 (2014) https://doi.org/10.14710/ijred.3.3.184-187.
[43] I. Parola, M.A. Illarramendi, J. Zubia, E. Arrospide, G. Durana, N. Guarrotxena, O. [48] X. Huang, S. Han, W. Huang, X. Liu, Enhancing solar cell efficiency: the search for lu-
García, R. Evert, D. Zaremba, H.-H. Johannes, F. Recart, Polymer optical fibers minescent materials as spectral converters, Chem. Soc. Rev. 42 (2013) 173–201,
doped with organic materials as luminescent solar concentrators, Proc. SPIE - Int. https://doi.org/10.1039/C2CS35288E.
Soc. Opt. Eng. 10101 (2017) 101010Z, https://doi.org/10.1117/12.2252956.
[44] I. Parola, D. Zaremba, R. Evert, J. Kielhorn, F. Jakobs, M.A. Illarramendi, J. Zubia, W.
Kowalsky, H.H. Johannes, High performance fluorescent fiber solar concentrators

You might also like