Li Et Al-2018-Advanced Optical Materials1 PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

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

net/publication/325086742

Optically Transparent Wood: Recent Progress, Opportunities, and Challenges

Article  in  Advanced Optical Materials · May 2018


DOI: 10.1002/adom.201800059

CITATIONS READS

118 10,858

5 authors, including:

Yuanyuan Li Elena Vasileva


KTH Royal Institute of Technology KTH Royal Institute of Technology
98 PUBLICATIONS   6,211 CITATIONS    13 PUBLICATIONS   376 CITATIONS   

SEE PROFILE SEE PROFILE

Ilya Sychugov Sergei Popov


KTH Royal Institute of Technology KTH Royal Institute of Technology
89 PUBLICATIONS   1,966 CITATIONS    344 PUBLICATIONS   3,763 CITATIONS   

SEE PROFILE SEE PROFILE

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

Modeling of Physical Layer Impairments Impact in Optical Fiber Transmission Systems View project

Green Initiative for Future Optical Networks (GRIFFON) View project

All content following this page was uploaded by Yuanyuan Li on 24 January 2019.

The user has requested enhancement of the downloaded file.


PROGRESS REPORT
Transparent Wood www.advopticalmat.de

Optically Transparent Wood: Recent Progress, Opportunities,


and Challenges
Yuanyuan Li,* Elena Vasileva, Ilya Sychugov, Sergei Popov, and Lars Berglund

to improve wood properties for existing


Transparent wood is an emerging load-bearing material reinvented from applications,[8] for instance, reducing
natural wood scaffolds with added light management functionalities. Such moisture sensitivity. However, designing
material shows promising properties for buildings and related structural advanced wood materials combing struc-
tural performance with new functions
applications, including its renewable and abundant origin, interesting optical offers new possibilities to extend the range
properties, outstanding mechanical performance, low density, low thermal of properties and applications of wood.[9,10]
conductivity, and great potential for multifunctionalization. In this study, a Transparent wood is an example of
detailed summary of recent progress on the transparent wood research topic multifunctional wood composites.[11] The
is presented. Remaining questions and challenges related to transparent first report on transparent wood is from
Fink in 1992 to facilitate wood morphology
wood preparation, optical property measurements, and transparent wood
studies.[12] Later on, engineering use of
modification and applications are discussed. transparent wood was suggested by com-
bining mechanical performance considera-
tions with optical transmittance studies.[3,13]
1. Introduction After that, several studies were published along similar engi-
neering use considerations.[14–16] With the addition of optical
With increasing energy demand and requirements for environ- transmittance to basic wood properties, transparent wood facili-
mental conservation, replacement of petroleum-based materials tates wood anatomy studies,[12] and it can be used in light-trans-
with bio-based materials is an interesting opportunity. Wood mitting smart buildings,[3,14,17] electronic devices, and in pho-
from trees is a key resource in the Northern Hemisphere due tonic devices such as photovoltaic cells and light source.[9,18,19]
to its abundance (global number of trees ≈3 trillion),[1] excel- Several works also briefly discussed the transparent wood as a
lent mechanical properties,[2] and potential for functionaliza- part of functionalized wood when reviewing the wood nanotech-
tion.[3–7] The chemical composition with high-strength cellulose nologies.[20,21] In the present study, recent progress and related
nanofibrils and complex hierarchical structure are important challenges about transparent wood are summarized in detail, as
factors contributing to the excellent properties of this native well as mechanisms of importance to optical properties.
composite.[8] This includes sophisticated nanocomposite struc-
ture, characteristic anisotropic organization, complex porosity
combining micro-, meso-, and macropores, and chemical com- 2. Wood Structures
position of three complementary structural components of
high-strength cellulose, soft and hygroscopic hemicelluloses, Wood is of low density due to its polymeric origin and hollow
and the stabilizing phenolic lignin structures. Wood is typically cells functioning as liquid transport channels. Strength and
used as structural materials or as the basis for cellulose/cellu- fracture toughness are related to the composite nature of
losic fibers, for instance, used in buildings, for printing, and wood, including multiscale hierarchical structure and the
as packaging materials. Wood modification is often carried out nanofibrous nature of cellulose (Figure 1a).[22] This has been
discussed in many reviews.[8,23,24] The hierarchical material
structure starts at the millimeter scale of annual rings, via cells
Dr. Y. Y. Li, Prof. L. Berglund in the form of hollow fibers, their layered cell walls, and the
Wallenberg Wood Science Center nanoscale composite structuring of those layers further down
Department of Fiber and Polymer Technology to molecules making up the cellulose nanofibrils (Figure 1a).
KTH Royal Institute of Technology
SE-10044 Stockholm, Sweden
Cellulose is one of the three major structural components in
E-mail: yua@kth.se wood, accounting for around 40–50 wt% of content depending
E. Vasileva, Prof. I. Sychugov, Prof. S. Popov on wood species.[25] Cellulose is present in the form of fibrils
Department of Applied Physics of high axial modulus and strength, embedded in a hydrated
School of Engineering Science matrix of hemicellulose and lignin. Combinations of modeling
KTH Royal Institute of Technology and experimental work reveal the axial modulus and tensile
SE-16440 Kista, Sweden
strength of cellulose crystals to be around 150 and 10 GPa,
The ORCID identification number(s) for the author(s) of this article
can be found under https://doi.org/10.1002/adom.201800059.
respectively.[26,27] In wood, cellulose chains are assembled into
fibrils of 3–4 nm in diameter. The fibrils tend to form larger
DOI: 10.1002/adom.201800059 scale aggregates in the cell wall. The secondary cell wall is the

Adv. Optical Mater. 2018, 6, 1800059 1800059  (1 of 14) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advopticalmat.de

main part of cell wall with thickness in micrometers. It consists


of three major layers, but even those three layers have finer Yuanyuan Li received her
scale layers. The cell wall is organized to form hollow fibers Ph.D. degree in Nanjing
with the total cell wall thickness of a few micrometers, a length Forestry University (China)
of 1–3 mm, and a diameter of 20–50 µm.[23] These hollow fibers in 2014. During her Ph.D.,
are oriented preferably in the longitudinal direction, parallel to she worked as an exchange
the tree stem. In many species of trees, the fiber packing den- student at the Department
sity and fiber cell wall thickness vary during the year, resulting of Materials Science and
in annual growth rings at the millimeter scale. The fiber orien- Engineering, University of
tation leads to the anisotropic properties of wood in terms of Maryland (USA), focusing on
mechanical performance,[28] water transportation, thermal con- nanocellulose-based multi-
ductivity,[29] and optical properties.[30–32] functional composites and
The “matrix” in the native cell wall is a mixture of hemicellu- flexible electronics. Currently,
lose and lignin. Hemicellulose is an amorphous hetero-polysac- she is a postdoc at Kungliga Tekniska Högskolan (KTH)
charide, which is highly hydrated in the cell wall and accounts Royal Institute of Technology (Sweden) working on bioma-
for around 20–40 wt% of wood.[33] It is associated with cellulose terials for functional structural materials.
and may serve to prevent cellulose agglomeration and physi-
cally link cellulose fibrils. Lignin is an aromatic component rich Elena Vasileva holds
in phenols and provides stiffness to cell walls even if there is no M.Sc. degree in Electrical
turgor pressure. It is mainly distributed in the secondary cell Engineering (2013) obtained
wall, but with high concentration in the middle lamella and cell within the M.Sc. double-
wall corners.[34,35] Figure 1b depicts the distribution of polysac- degree program from
charides and lignin in the cell wall. The First Electrotechnical
Porosity and pore size distribution are important param- University (LETI, Russia) and
eters that influence the liquid/gas transportation in wood,[36,37] Lappeenranta University of
thermal conductivity,[29] and optical and mechanical properties Technology (LUT, Finland). Her
of wood.[28] The complex porous structure combining micro-, research interests include laser
meso-, and macropores is a specific characteristic of wood.[38,39] physics, light polarization and
The wood porosity spans over a large range from about 20 to coherence, and novel optical
90 vol%, depending on the wood density. Lumens, pit openings, materials. She is a board member of IEEE Photonics Sweden
and sub-microscale voids in cell walls are included. Lumens are and chairs the Swedish chapter of Women in Engineering
of micrometer scale, and normally of cylinder shape, organ- society (WIE Sweden IEEE). Currently, she is working toward
ized preferably in longitudinal direction. Pit openings can be Ph.D. degree in Photonics at Kungliga Tekniska Högskolan
both meso- and macropores. The presence and organization (KTH) Royal Institute of Technology (Sweden).
of multiscale pores in wood serves to realize water/nutrients
transportation in wood, low thermal conductivity, and influence Lars Berglund studied
mechanical properties. Pores are scattering light due to solid/gas Materials Science and
interface boundaries, and thus hindering light transmittance. Engineering at Luleå University
of Technology (LTU), Luleå,
Sweden, and Polymer Science
and Engineering at University
3. Transparent Wood Fabrication of Massachusetts, Amherst,
Natural wood is not transparent in the visible spectral range due USA. He obtained his Ph.D.
to strong absorption and scattering of light. To make wood trans- from Linköping University
parent, both absorption and scattering need to be eliminated. in 1987. From 1991 to 2002,
The amount of light absorption is strongly related to the chem- he was a Professor at LTU, and
ical composition. Wood is brownish due to the presence of light- from 2002 a Professor of Wood and Wood Composites at
absorbing components such as lignin, chlorophyll, and tannins. the Kungliga Tekniska Högskolan (KTH) Royal Institute of
Among these, lignin is responsible for around 80–95% of light Technology, Stockholm, Sweden. Since 2009 he has been
absorption in wood.[41] This light absorption can be drastically the Director of Wallenberg Wood Science Center at KTH.
reduced by chemical treatment of the wood: either removing
“all” of the lignin (by delignification)[3,12–14,18,42] or deactivating the
chromophores within lignin.[17] Fink treated wood with a 5 wt% wood in boiling solution containing NaOH and Na2SO3, fol-
aqueous solution of sodium hypochlorite for 1–2 days to remove lowed by hydrogen peroxide (H2O2) treatment.[44] The wood tem-
colored substances, including lignin.[12] Berglund and co- plate obtained shows a lignin content of <3 wt%. Yu et al. used
workers did delignification using sodium chlorite under 80 °C the same approach in their work.[15] With decreased lignin con-
for 6 h. The lignin content decreased from around 25 wt% to tent, the wood template shows increased transmittance and min-
less than 3 wt%.[3] This method has also been adapted by several imized light absorption in the visible spectral range (Figure 2a).
groups.[16,42,43] Hu and co-workers removed lignin by treating After chemical treatment, with main focus on delignification,

Adv. Optical Mater. 2018, 6, 1800059 1800059  (2 of 14) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advopticalmat.de

Figure 1.  a) Schematic illustration of hierarchical wood structures from the meter scale of tree to the nanoscale of cellulose chain. Reproduced with per-
mission.[40] Copyright 2011, IOP Publishing Ltd. b) Lignin and polysaccharide map showing their distributions in the cell wall using scanning transmission
X-ray microscopy. Left is the lignin map. Middle is the polysaccharides map. Right image is the map of lignin divided by polysaccharides (image left divided
by image middle) with false color (warmer colors indicate more lignin). Reproduced with permission.[34] Copyright 2014, BioMed Central.

the decolorized, whitish wood (yet not transparent) has a well- wood cell orientation, longitudinal or transverse. As is described
preserved morphological structure at microscale,[3,18] which is in Section 2, the wood consists of hollow fibers with cellular
anisotropic due to the fibrous wood structure. space in the form of “lumen,” forming mesoporous struc-
Although the wood cell structure is not perfectly ordered, tures. Strong light scattering occurs at the boundaries between
rather semiordered, it determines optical properties of the the air-filled cellular void channels and the solid nanocom-
material, depending on the direction of light propagation. A posite matter of the cell walls (e.g., cellulose, residual lignin,
simple coordinate system should, as a minimum, refer to the and hemicellulose) due to refractive index mismatch between

Figure 2.  a) Transmittance of delignified basswood (1 mm thick) with different residual lignin contents. Reproduced with permission.[14] Copyright
2016, Elsevier Ltd. Image of original balsa wood cross sections b) before and c) after delignification. Reproduced with permission.[3] Copyright 2016,
American Chemical Society.

Adv. Optical Mater. 2018, 6, 1800059 1800059  (3 of 14) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advopticalmat.de

Figure 3.  Scheme showing transparent wood preparation by a delignification approach. Scale bars in the scanning electron microscopy (SEM) images
are 40 µm. Reproduced with permission.[3,17] Copyright 2016, American Chemical Society; Copyright 2017, Wiley-VCH.

these components (n ≈1.53 in average) and air (n = 1).[3] Micro- wood. With the “lignin-retaining method,” up to 80 wt% of the
scale and nanoscale voids appear in the cell walls after removal lignin was preserved, leading to better mechanical properties
of lignin. Figure 2b and Figure 2c show the cell wall structure compared with delignified templates.
before and after delignification, respectively. The generated The optical properties of transparent wood based on lignin-
pores lead to increased light scattering inside the wood cell wall. retaining wood templates are comparable to transparent wood
To decrease light scattering from the wood structure, one can based on delignified wood template. Despite the success in
infiltrate the wood template with polymers where the refractive transparent wood preparation and the high transmittance (over
index is matched to the wood template, for example polymethyl 80% at a thickness of 1 mm), transparent wood still demon-
methacrylate (PMMA, n  ≈1.49), epoxy resin (n ≈1.5), polyvi- strated lower transmittance than the neat polymer, and it
nylpyrrolidone (PVP, n  ≈1.53), n-butyl methacrylate (n  ≈1.50), strongly decreases further with increased wood thickness.[3,13] Li
polystyrene (n  ≈1.59), dibutyl phthalate (n  ≈1.52), iso-bornyl et al. proposed that the main reason is interface debond cracks
methacrylate (n  ≈1.48–1.50), diallyl phthalate, and poly vinyl- between the polymer and the wood cell wall.[45] Thus, surface
carbazole (n  ≈1.68).[3,12–14] If the mismatch between refractive acetylation of delignified wood is carried out to increase the
indices of wood and the infiltrated polymer is small, this can interface compatibility between the wood cell wall and PMMA.
reduce the scattering drastically. We generally refer to this com- In this way, transparent wood (balsa) samples (transmittance
posite material, a wood template infiltrated with a suitable poly­ of 60%) with a thickness of 1 cm were successfully made.[45]
­mer, as “transparent wood.” Nevertheless, there is still residual Despite the success in transparent wood fabrication, chal-
scattering due to interfaces, in particular debond gaps caused lenges still exist with respect to tunable optical performance
by incompatibility between the wood cell wall and the polymer and scaled-up production. Effective prevention of polymer
as well as the polymer shrinkage during polymerization.[9,45] shrinkage during polymerization is a remaining question.
Transparent wood was first made by Fink based on the idea of The optical and mechanical stability is also a potential ques-
removing light-absorbing components (mainly lignin) followed tion for industrial applications of the technology. Furthermore,
by infiltration of a polymer with a refractive index matching the transparent wood with increased cellulose content is desirable.
wood substrate (Figure 3).[12] More studies based on the same These questions should be addressed in future studies.
idea, using different treatment methods, have been developed
later on.[3,9,13–16,18,19,42,43] It should be noted that delignification
process typically removes around 30 wt% of wood tissue.[17] 4. Optical Properties
This damages and weakens the wood structure so that han-
dling and fabrication of large substrates become challenging 4.1. Measurement Techniques
and impractical. Li et al. suggested to remove only the chromo-
phoric structures in lignin, while bulk lignin is preserved.[17] Before discussing measurement techniques for optical prop-
The idea has been realized by using alkaline H2O2 treatment of erties, interaction of light with transparent wood requires

Adv. Optical Mater. 2018, 6, 1800059 1800059  (4 of 14) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advopticalmat.de

There are no reports on direct absorbance estimation for


transparent wood, since all measurements are done for trans-
mission, which, in turn, is affected by both absorption and
scattering. The absorption characteristics can be retrieved from
the analysis of reflectance and transmittance (on average 90%).
However, such measurements and data reduction require cau-
tion, since they strongly depend on surface quality.
The most common method to estimate haze and transmit-
tance is described in the ASTM D1003 standard[46] and uti-
lizes an integrating sphere and a spectrometer. However, this
method was mainly developed for materials like plastics or
frosted glasses with haze lower than 30%. Therefore, there may
be some pitfalls for direct implementation of this method for
Figure 4.  Schematic sketch showing light propagation through transparent transparent wood. The instruments commonly used for optical
wood within an input beam of incident intensity I: A—absorption, R—reflec-
measurements on transparent wood are either commercial
tion, BS—back scattering, FS—forward scattering, B—ballistic photons, i.e.,
part of the beam mainly maintaining its original direction of propagation UV–vis spectroscopy with an integrating sphere or homemade
without significant scattering. The surrounding medium is air, and all values setups according to ASTM D1003 standard.[3,13,15] There is no
are given in percentage, whereas incident intensity is taken as 100%. obvious difference between these two types of instruments for
determining the optical properties.
some attention. The processes are schematically illustrated in
Figure 4. As an optical beam travels through transparent wood,
it loses part of the energy and changes its structure due to sev- 4.1.1. Transmittance Measurements
eral factors: 1) back reflection at the front and back planes (R),
2) scattering inside the material (backward and forward scat- An integrating sphere is typically used to measure hemi-
tering, BS and FS in the schematics), and 3) absorption (A) spherical transmittance. Values differ from direct transmittance
inside the material. All these processes depend on physical data because the sphere collects a total cone of scattered light
parameters of the transparent wood. These include quality of consisting of ballistic and randomly scattered photons (FS + B
the sample facets, refractive-index mismatch between wood and in Figure 4). Thus, the measured transmittance value is higher
polymer, the presence of voids, anisotropic structure, to name than direct transmittance values. Certain parameters, such as
a few, which contribute to light scattering. Residual absorbents the setup geometry, sample thickness, light propagation direc-
from original wood and the introduced polymer are responsible tion with respect to the wood fiber orientation, and the size of
for light absorption. the beam in comparison to the integrating sphere port, will
Two quantities are of particular importance in characterizing influence the measured data and must be characterized in
the optical properties of transparent wood: transmittance order to analyze the optical property.
(measure of energy transport losses, including absorption and The semiordered structure of transparent wood causes light
scattering) and haze (information/imaging content loss). Physi- to experience strong anisotropic scattering, as it propagates
cally, the transmittance (T) is the total amount of light energy in transparent wood. Therefore, the beam can diverge signifi-
transported through the sample into forward half-space, both cantly before entering the input port of the integrating sphere
scattered (diffused) light and part of the beam going without (Figure 5). If the sample is thin (Figure 5a) and the beam size is
noticeable disturbance. Effectively, it means the energy of input smaller than the input port diameter, all light passing through
beam subtracted by the reflection, absorption, and scattering in the sample is collected by the sphere. However, if the beam size
the backward half-space. In a simplified way, these values can is larger than or comparable to the input port diameter, part of
be expressed by the following relationships the beam (highlighted in Figure 5b) will be blocked by the inte-
grating sphere. Thus, the measured transmittance value will be
T = B + FS = I − A − R − BS (1) lower than the real value. But even if the incident beam size
is smaller than the port diameter, a sample of a certain thick-
As for haze, phenomenologically it describes the relative ness might still lead to the losses of measured light (Figure 5c).
amount of forward scattered, or diffused, light (subtracted with There may be strong beam scattering inside the medium, so
ballistic photons, ballistic photons is shown as B in Fig. 4), that the light reaching the sphere is reduced. All these exam-
that results in the decrease of image contrast seen through the ples can introduce measurement errors and lead to incorrect
material. In this case, it can be defined as follows values. Therefore, there is a need for common characteristics
of the measurement setups, in order to compare data from dif-
Haze ≈ FS/T (2) ferent samples of transparent wood.
In order to compare results for samples of different thick-
However, there is no strict formal definition, in which ness, a “transmittance coefficient” parameter should be intro-
divergence (solid) angle outgoing photons are still considered duced. This parameter would show the transmittance value
as ballistic. There exists a recommendation document which per unit length. However, the applicability of this parameter
provides technical specifications for a particular measurement is a subject of discussion. It seems that the transmittance
method which we consider later in Section 4.2.1. coefficient could be applied distinctively for the cases when

Adv. Optical Mater. 2018, 6, 1800059 1800059  (5 of 14) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advopticalmat.de

Figure 5.  Sketch of the setup for transparent wood transmittance and haze measurements: a) thin sample on top of input port with a beam size smaller
than the input port diameter; b) thin sample placed on top of input port with a beam size larger than or comparable to the input port diameter; c) thick
sample placed on top of input port with a beam size smaller than the input port diameter; and d) sample with thickness larger than the transport mean
free path (TMFP) so that no ballistic photons are observed.

the thickness of a sample does or does not exceed the trans- detect nondeviating (ballistic) photons, a “bucket”-type detector
port mean free path (TMFP) length.[47] TMFP is typically used with a small numerical aperture is required, in order to clearly
for the description of light propagation inside a scattering distinguish those from scattered photons. Interpretation of the
medium.[48] Its value characterizes the length of propagation results also requires certain care. For the haze measurements,
a photon experiences inside the scattering material, before its the procedure described by the standard is mainly applicable
propagation direction becomes totally random. It means that for materials with haze values lower than 30%. It does not
there are no ballistic photons in the beam after it propagates account for material anisotropy and a possible dependence
a distance longer that TMFP (Figure 5d). TMFP may depend on the TMFP. For example, in the case depicted in Figure 5d,
on the wood type and orientation of the samples relative to the where the thickness of the sample is larger than TMFP, dif-
incident beam. fused transmittance of light will not be measured correctly.
Strong inhomogeneous and anisotropic scattering make
the characterization of the transmittance and haze of trans-
4.1.2. Haze Measurements parent wood significantly different from many other trans-
parent scattering materials, which are homogeneous or have
An important material parameter, along with transmittance, purely random structures. Most results on transparent (semi-
is the haze of transparent wood. In terms of practical utiliza- transparent) materials deal with homogeneous and isotropic
tion, haze is useful to characterize transparency, mostly in the media. However, characterization of unusual transparent mate-
imaging context, i.e., how clearly an ambient image can be rials, such as wood, requires adaptation of classical investiga-
observed through the sample. In principle, haze measurements tion methods. The description of semiordered wood structures
are based on the same setup as shown in Figure 5. Briefly, it cannot be simplified by averaging or by the use of symmetry
can be described as follows:[46,49] properties. They represent the most “complex” case in between
highest and lowest entropies of completely disordered and
1) Measurements of light intensity without specimen and with ordered patterns.[50] During measurements of optical proper-
the white light standard placed (T1—reference transmittance). ties, the mentioned precautions with respect to the ordering of
2) Measurements of light intensity with sample located at the the wood structure are often not taken into account. In addition,
input port of the sphere and white standard placed (T2—light when light propagates perpendicular to the axial wood fiber
transmitted by sample). direction, the scattering introduces a nonsymmetric intensity
3) Measurements of light intensity without the sample and a distribution in two perpendicular planes.[13] One consequence
light trap replacing the white standard (T3—light scattered by is that the measured values of haze will be affected. This leads
instrument). to additional uncertainties in the interpretation of data since the
4) Measurements of light intensity with the sample and the intensity distribution of a beam entering the sphere is assumed
light trap (T4—light scattered by instrument and specimen). to be uniform.[49] The light transmittance should be differen-
tiated for two orthogonal planes relative to the fiber direction
Then the haze value can be calculated as (T⊥ and T||) in order to make the haze definition more rigorous.
In the case of nonuniform intensity distribution, the measured
Td  T4 T3  values of T3 and T4 do not take into account this difference,
Haze = =  −  × 100% (3)
Tt  T2 T1  rather describe “an averaged” spatial value, whereas the haze in
two perpendicular planes is different.
where Tt is total transmittance and Td is transmittance of dif- Besides anisotropy, the material thickness and TMFP factor
fused light. In fact, T3/T1 is the error caused by the experi- are essentially critical parameters for measurement of trans-
mental setup. parent wood haze. To implement the classical approach of
By definition, haze is the percentage of transmitted light haze measurements, the sample thickness should not exceed
that is scattered with a direction deviating more than a speci- TMFP, as discussed before. Otherwise, the light direction will
fied angle (2.5°) from the direction of an incident beam.[46] To be randomized and measured values of T3 and T4 would not

Adv. Optical Mater. 2018, 6, 1800059 1800059  (6 of 14) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advopticalmat.de

bear the original meaning, and depend on the aperture of the when a rather unique combination of high, up to 90%, trans-
detector (Figure 5). With formal treatment in its imaging con- mittance and high (up to 90%) haze can be realized.[3,13,14,45]
text, the haze would attain an infinitely large value, since there This makes transparent wood a great candidate as a structural
is no directly transmitted light (ballistic photons) reaching the “light diffuse” layer for solar cells, lighting systems, and build-
detector plane (Figure 5d). ings (same function as frosted glass). At short distances from
objects, the transparency of relatively thin transparent wood
samples is comparable with glass, plastics, and cellulose-based
4.2. Experimental Findings nanopaper.[52–54] Additionally, in comparison to many other
transparent/translucent materials, transparent wood reveals
In this section, recent results on transmittance and haze meas- strong anisotropic optical behavior, which will be discussed in
urements are reviewed for this unique material. In general, the Section 4.3 in more details.
transmittance of the transparent wood is high in the visible and
near-IR regions. The reason is the very limited absorption of
radiation by residual light-absorbing components in the struc- 4.2.1. Transmittance
ture. The value of transmittance at other wavelengths is lim-
ited by polymer absorption either by molecular vibrations in As mentioned in Section 3, optical properties of transparent
the IR region or electronic states in UV.[51] Besides the amount wood are determined by the inherent cellular structure of the
of residual lignin[14] causing absorption, transmittance is also material. Variation of the diameters, lengths, and cell wall
dependent on factors such as wood type,[17] cellulose volume thicknesses of vessels and fibers results in different material
fraction,[3] cellulose fibers orientation,[13,18] and thickness of the densities and porosities, which also influences transmittance
sample.[3,45] spectra. As was discussed earlier, the cell walls are mesoporous
Transparent wood exhibits high haze due to internal after delignification. Thus, scattering also takes place inside the
light scattering from the refractive-index mismatch between cell wall, in addition to at the cell wall interfaces with the infil-
wood template and infiltrated polymer, and due to the wood/ trated polymer. The increasing number of total refractive-index
polymer interface debond gaps inside the material. Typically, variations at interfaces enhances scattering in the structure.
stronger scattering implies lower transmittance. While this Figure 6a illustrates the dependence of the transmittance
is true for some cases of transparent wood, there are cases on wood species. Optical images of the transparent wood

Figure 6.  a) Transmittance and b) photo of transparent wood (1.5 mm thick) made from balsa (density = 0.16 × 103 kg m−3), pine (density = 0.51 ×
103 kg m−3), birch (density = 0.65 × 103 kg m−3), and ash (density = 0.68 × 103 kg m−3). Reproduced with permission.[17] Copyright 2017, Wiley-VCH.
c) Dependence of transparent balsa wood transmittance (1.2 mm thick) on cellulose volume fraction. Reproduced with permission.[3] Copyright 2016,
American Chemical Society. d) Dependence of transparent wood transmittance on the sample thickness. Transparent wood samples are made from
balsa,[3] beech,[42] and basswood,[13] respectively. ⊥ means light propagates perpendicular to longitudinal direction, and // means light propagates
parallel to longitudinal direction. Transmittance values (c,d) are shown for 550 nm.

Adv. Optical Mater. 2018, 6, 1800059 1800059  (7 of 14) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advopticalmat.de

Figure 7.  a) Dependence of haze on sample thickness. Transparent wood is made from balsa,[3] beech,[42] and basswood,[13] respectively. ⊥ means light
propagates perpendicular to longitudinal direction, and // means light propagates parallel to longitudinal direction. b) Haze spectra of transparent
wood (1.5 mm thick) made from balsa (density = 0.16 × 103 kg m−3), pine (density = 0.51 × 103 kg m−3), birch (density = 0.65 × 103 kg m−3), and ash
(density = 0.68 × 103 kg m−3). c) Dependence of transparent balsa wood haze (1.2 mm thick) on cellulose volume fraction. Reproduced with permis-
sion.[3] Copyright 2016, American Chemical Society. Haze values (a,c) are shown for 550 nm.

made from ash, balsa, pine, and birch are shown in Figure 6b. One would expect to see a dependence of haze values on the
The density of ash is the highest (0.68 × 103 kg m−3), while the density of the structure, which should affect scattering proper-
transmittance is the lowest (≈58% at 1.5 mm sample thickness). ties of the material. Surprisingly, reported haze values are only
As a comparison, the transmittance of transparent wood made slightly affected by the density of the material (different wood
from low-density balsa is 85% for the same thickness (density species)[17] (Figure 7b) or CVF (different degree of compres-
of balsa is 0.16 × 103 kg m−3). Wood density and transmittance sion)[3] (Figure 7c). Such a comparison should be carried out
are correlated, perhaps due to scattering inside the cell wall. at material thicknesses below TMFP, which varies between dif-
This dependency is also observed in Figure 6c, where the trans- ferent wood species.
mittance is modified by changing the cellulose volume fraction
(CVF). The samples made from the same wood species but
with different volume fraction can be obtained by compression 4.3. Anisotropic Scattering and Waveguiding Effect in
of delignified wood templates before polymer impregnation.[3] Transparent Wood
Thus, the through-thickness density of interfaces grows with
CVF (5–65%) while the transmittance is decreased from 85% Based on the anisotropic fiber-based and cellular wood struc-
to 35%.[3] Similar tendency is observed as the wood thickness is ture, light propagation through this material is expected to
increased when light propagates perpendicular to the longitu- reveal strongly anisotropic properties.[13,18] If light travels par-
dinal direction (Figure 6d). It should be pointed out that when allel to the longitudinal and tubular wood cell direction, the
light propagates along the longitudinal direction, the transmit- poly­mer-filled pores can be considered as multimode wave-
tance is almost independent from sample thickness (Figure 6d, guides with relatively high losses.[18] Figure 8a illustrates that
basswood) due to waveguiding effects (Section 4.3). This mode light is guided along the wood fibers independent of the inci-
of orientation is less favorable from a structural mechanics dent beam angle.[18] The guidance effect of wood fibers was also
point of view. demonstrated in ref. [9], where the luminescence of quantum
dots (QDs) inside the transparent wood template is excited by
external radiation from the sample top. Light was collected
4.2.2. Haze from two orthogonal directions of the sample. The intensity
was higher for luminescence collected from the sample edge
The haze value depends on the microstructure of wood.[13,14] perpendicular to the wood fibers’ direction, where light propa-
Some wood fibers and channels contain microcurvatures with gation along the fibers is promoted (Figure 8b).
bumps, microstripes, and microcavities that scatter light[14] Obviously, the waveguiding property also results in different
throughout the whole sample thickness.[13] Literature data for the material transmittance values, depending on the orientation of
haze dependences are sometimes contradictory. Some results an incident beam with respect to wood fiber direction. For the
demonstrate that the thickness of the sample has almost no effect same material thickness, transmittance is higher for light prop-
on the measured haze value (Figure 7a, balsa and basswood).[3,13] agating along the fibers.[13] Moreover, due to light confinement
Other data show strong dependence of haze on wood thickness by the fibers and the semiordered structure (Figure 1a), the scat-
(Figure 7a, beech wood).[42] This is apparently related to the tering properties are also anisotropic (Figure 8c,d). Depending
measurement technique, and how it accounts for TMFP related on the relative orientation of the material direction and the
to material thickness. For thin samples (thickness < TMFP), the incident light, the wood fibers can guide light along the axial
ratio of ballistic and scattered photons would depend on material direction or introduce strong anisotropic scattering. Therefore,
thickness. Thus, it is reasonable to introduce a thickness- the haze value, analogous to transmittance, depends on fiber
dependent parameter for the haze measurements, for example, orientation. Zhu et al. reported that, with the same thickness,
the slope of the curve in Figure 7a in units [mm−1]. Saturation of transparent wood haze is higher when light propagates along
haze values in Figure 7a for thicker samples probably reflects a the longitudinal fiber direction (≈97%) compared with light
transition to the photon transport regime above TMFP. propagating perpendicular to longitudinal direction (≈90%).[13]

Adv. Optical Mater. 2018, 6, 1800059 1800059  (8 of 14) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advopticalmat.de

Figure 8.  Illustration of anisotropic scattering and waveguiding effects in transparent wood. a) Photo shows green light illuminating the transparent
wood surface at different angles (0°, 45°, and 70°). Light propagates mostly along the wood fibers (basswood, 0.5 mm). Reproduced with permis-
sion.[18] Copyright 2016, Wiley-VCH. b) Photoluminescence of QD-filled transparent wood, collected as a function of excitation spot distance from
the sample edge: parallel and perpendicular to the fiber direction (red and black), photoluminescence of QDs in glass collected as a reference (blue).
Reproduced with permission.[9] Copyright 2017, Wiley-VCH. c) The transmittance measurement setup (top), photo image of the scattered light spot
(middle), and the relative intensity distributions in the x- and y-directions (bottom) for R-wood.[13] d) The transmittance measurement setup (top),
photo image of the scattered light spot (middle), and the relative intensity distributions in the x- and y-directions (bottom) for L-wood.[13] In panels
(c) and (d), for R-wood, light propagates along the wood fiber cells, while for L-wood, light propagates perpendicular to the wood fibers. Reproduced
with permission.[13] Copyright 2016, Wiley-VCH.

5. Mechanical Properties of Transparent Wood crystals/cellulose nanofibers are preferably oriented in the
longitudinal direction of the tubular wood cells, further contri­
Mechanical properties of transparent wood are expected to be buting to mechanical anisotropy. In addition, wood properties
improved compared with neat polymers (PMMA). Properties also depend on relative density and morphological details of the
are determined by combined contributions from the wood tem- wood structure.[55] Properties of transparent wood are strongly
plate, the infiltrated polymer phase, and are also influenced by dependent on wood properties. One aspect is the mechanical
interfacial bond properties between the wood template and the anisotropy (Figure 9a,b). Zhu et al. reported that mechanical
infiltrated polymer. properties of transparent wood in the longitudinal direction
As discussed in Section 2, the wood mechanical properties (E  = 2.37 GPa and σ  = 45.38 MPa) are about twice as high as
are dominated by the cellulose content,[22] and the intrinsic geo- those perpendicular to the longitudinal direction (E = 1.22 GPa
metric orientation of the wood fiber cell structure. Cellulose and σ = 23.38 MPa).[13] One may note that this means that the

Figure 9.  Stress–strain curves for the native wood and transparent wood in the a) longitudinal direction and b) transverse direction. Reproduced with
permission.[13] Copyright 2016, Wiley-VCH. c) Stress–strain curves of transparent wood with wood volume fractions of 19% (TW-19) and 5% (TW-5),
delignified wood (DLW) and PMMA. d) SEM image of transparent wood showing the nanofibrous nature of the cell wall region and signs of favorable
interfacial bonding at the PMMA/wood interface. Reproduced with permission.[3] Copyright 2016, American Chemical Society.

Adv. Optical Mater. 2018, 6, 1800059 1800059  (9 of 14) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advopticalmat.de

anisotropy ratio of longitudinal to transverse properties is actu- oriented nanocomposite layers, wood is a sophisticated hier-
ally decreased for transparent wood, as expected. Weak prop- archical material. It is of great interest as a hierarchically struc-
erties in the transverse direction of wood (radial or tangential) tured large-scale template for multifunctional modification. The
are caused by cell wall–bending effects, and this is expected to main modification strategies include modification of the cell
be suppressed by the presence of polymer in the lumen space. wall, cell wall/lumen interface modification, and lumen filling
Another aspect is the dependence of mechanical properties on (Figure 10a).[8] Modification can be performed by chemical reac-
the volume fraction of wood. Li et al. reported that the tensile tions of small molecules,[5,57] physical adsorption,[58,59] physical
properties of transparent wood increase with increased wood/ filling of monomer liquids,[60] thermal treatment,[61,62] or complex
cellulose volume fraction.[3] The modulus was increased from combinations. Successful wood modification examples include
≈2 GPa to ≈4 GPa as the volume fraction increase from 5% to preparation of magnetic wood,[63,64] fire retardant wood,[65] conduc-
19%. With the same PMMA polymer, the mechanical proper- tive wood,[4,66,67] and stimuli-responsive wood.[6] These technolo-
ties of transparent wood will change when wood templates gies can potentially be used for multifunctional transparent wood.
from different wood species are used. Transparent wood from The main approach for transparent wood modification is to
birch shows increased strength (σ  = 140 MPa) compared with fill lumen pore space with monomer of prepolymer liquids,
balsa (σ  = 100.7 ± 8.7 MPa) in bending tests, again reflecting followed by polymerization.[9,15,16,43] Li et al. prepared lumi-
differences in wood/cellulose content.[17,56] Transparent wood nescent transparent wood by including silicon or CdSe QDs in
samples based on balsa, poplar, and beech also show substantial liquid MMA monomer/oligomer mixtures, and impregnating
differences in strength.[3,15,16] A systematic study is required to porous wood templates followed by polymerization.[9] Photolu-
sort out the mechanisms for this. A simple rule of the mixtures’ minescence mapping revealed uniformly distributed emission
approach is expected to predict longitudinal transparent wood centers. The approach is presented in Figure 10b. With similar
modulus, from knowledge of wood and PMMA properties and strategy, magnetic transparent wood (Fe3O4 nanoparticles[16]),
relative volume fractions, but there are several complications. luminescent magnetic wood (γ-Fe2O3@YVO4:Eu3+ nanoparti-
For instance, cell wall properties will differ between wood spe- cles[43]), and heat-shielding transparent wood (CsxWO3 nano-
cies and even between wood samples from the same tree. particles[15]) have later been reported.
Synergies between wood template and polymer properties Another approach is to modify/dope the wood template with
are observed for transparent wood (Figure 9c).[3,15] The load- functional molecules, followed by monomer impregnation and
bearing components in transparent wood are the cell wall and polymerization. Vasileva et al. prepared luminescent wood by
the infiltrated polymer. Favorable interaction (bonding) between impregnation of an organic dye (Rh6G) solution, followed by
wood cellulose nanofibers and PMMA is helpful (Figure 9d). MMA monomer/oligomer impregnation and polymerization
Li et al.[3] reported that the tensile properties of transparent (Figure 10c, left).[19] The optical transmittance is then modified by
wood (Vf = 19%, E = 3.6 GPa, and σ = 90 MPa) are higher than the absorption properties of the dye. A microscopic image of the
for both the delignified wood template (E  = 0.22 GPa and σ  = specimen and the optical transmission curve confirm the pres-
3 MPa) and for PMMA (E = 1.8 GPa and σ = 44 MPa). Yu et al. ence of the dye (Figure 10c, middle). The unique feature of the
reported improved tensile strength for transparent wood com- dyed TW-Rh6G wood material is that its emission spectrum has a
pared with native wood and the neat polymer.[15] Lang et al. narrow peak width, measured as the full width at half maximum
reported high strength in bending tests.[56] One should note (FWHM), of less than 10 nm (Figure 10c, right), compared with
that native wood is not the template for transparent wood, since the typical 50 nm width Rh6G photoluminescence spectrum.[68]
there is a chemical pretreatment step. Native wood has a more Recently, chemical modification of the wood template before
favorable cell wall structure and much better properties than polymer infiltration was demonstrated to improve transparency
delignified wood templates used for transparent wood. of the wood (Figure 10d, left).[45] With the help of acetylation,
One advantage of transparent wood over glass for structural the compatibility between the wood template and the infiltrated
applications is that the ductility and work to fracture are higher. polymer is improved, resulting in higher transmittance and
Transparent wood does not shatter in an unfavorable manner. lower haze (Figure 10d, middle and right). With 1.5 mm thick
The high work to fracture is mainly caused by the intrinsic prop- transparent balsa, the transmittance is increased from 83 ± 2%
erties of the infiltrated polymer, the complex hierarchical wood to 92 ± 1%. Haze is decreased from 70% to 50%. More apparent
structure, and composite material effects in the form of new improvements are obtained with thicker samples and higher
toughening mechanisms. Li et al. performed impact tests on density wood species. This suggests that surface modification is
transparent wood and reported favorable impact toughness.[18] way to control the optical properties.
The three-point bending test on transparent wood showed one- An important aspect is the mechanical performance effect
order of magnitude higher work to fracture of transparent wood from addition of functional particles. Yu et al. reported that
(1.2 MJ m−3) compared with glass (0.1 MJ m−3).[17] This indi- incorporation of functionalized nanoparticles (CsxWO3) in
cates the potential of transparent wood in applications where transparent wood resulted in negligible strength and modulus
impact properties are important. effects. The strength was 60.1 MPa for transparent wood and
59.8 MPa for CsxWO3 /transparent wood. The modulus was
2.67 GPa for transparent wood and 2.72 GPa for CsxWO3 /
6. Transparent Wood Functionalization transparent wood).[15] Gan et al. observed lowered strength
(from 40.6 to 28.7 MPa) and modulus (from 0.79 to 0.57 GPa)
With annual rings, highly ordered porous cells at the scale of when magnetic particles (Fe3O4) were added and the content
10  µm and the cell wall organization in the form of stacked, increased from 0.1 to 0.5 wt%.[16] The strength change is related

Adv. Optical Mater. 2018, 6, 1800059 1800059  (10 of 14) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advopticalmat.de

Figure 10.  Transparent wood modification: a) scheme showing the various types of wood modification. Reproduced with permission.[8] Copyright 2015,
Institute of Materials. b) Sketch of transparent wood modification procedure using quantum dots (QD). Reproduced with permission.[9] Copyright 2017,
Wiley-VCH. c) Microscopic image (left) of illuminated transparent wood with the wood template doped with dye molecules (Rh6G), transmittance
spectrum (middle), and emission spectrum (right) of transparent wood with the wood template doped by dye molecules (TW-Rh6G). Reproduced
with permission.[19] Copyright 2017, Wiley-VCH. d) Photo of transparent wood (left): without (left in the photo) and with (right in the photo) surface
acetylation of wood template; transmittance (middle) and haze (right) spectra of transparent wood without and with surface acetylation of the wood
template. Reproduced with permission.[45] Copyright 2017, The Royal Society of Chemistry.

to embrittlement effects from the particles, whereas the mod- Recently, the mechanical performance of transparent wood
ulus effect is more difficult to understand. was studied.[3] Since the material combines structural perfor-
In general, the added content of nanoparticles or molecular mance with functional (optical) properties (high transmittance
dye is low (less than 1 wt%). Higher content often leads to and haze), this rare combination was emphasized. Most func-
strong aggregation of nanoparticles and lowered strength prop- tional materials do not have load-bearing properties or are suit-
erties. In addition, effects from refractive-index mismatch and able for large-scale structures. Transparent building structure is
intrinsic light absorption characteristics may somewhat com- one demonstration, where light transmittance can be designed
promise optical properties. It is therefore a challenge to increase so that artificial light can be partially replaced by sunlight. High
the content of active components, and yet preserve optical and haze can be used to ensure more uniform and soft nature of
mechanical properties. For improved preparation routes, it is the transmitted light so that indoor privacy is preserved. Trans-
also essential to characterize the distribution of nanoparticles/ parent wood roofs can be designed for some buildings, and
molecules in the material since detailed knowledge is lacking. provide more uniform and comfortable illumination compared
with conventional glass (Figure 11a).[17,18] In addition, building
panels from transparent wood show lower thermal conduc-
7. Applications of Transparent Wood tivity, better impact strength, and lower density compared with
glass.[18] Functionalized transparent wood provides even more
Transparent wood exhibits a combination of high optical opportunities for smart buildings. By inclusion of quantum
transmittance and haze, outstanding toughness, low thermal dots, transparent wood[9,43] attains diffused luminescence prop-
conductivity, low density, anisotropic optical and mechanical erties (Figure 11b). This may be used in planar light sources,
performance, etc.[3,13,17,18] Exploiting and utilizing these proper- luminescent building elements, or design furniture.[9]
ties in practical applications is important for the development Yu et al.[15] demonstrated transparent wood for heat-shielding
of transparent wood research. buildings/windows by the use of NIR absorbing CsxWO3 nanopar-
Early efforts on transparent wood were not motivated by ticles. Strong absorption of light was observed at wavelengths from
engineering considerations. It was first designed to facilitate 780 to 2500 nm. The CsxWO3/transparent wood window resulted
wood anatomy studies.[12] The interior wood structure becomes in lower temperature increase under continuous solar radiation,
visible, and 3D reconstruction of the structure is possible. compared with the glass window reference (Figure 11c). Mag-
Similar concepts have been used for deep imaging of plant and netic nanoparticles have also been incorporated into transparent
animal tissue.[69–71] wood[16,43] to provide electromagnetic interference (EMI) shielding.

Adv. Optical Mater. 2018, 6, 1800059 1800059  (11 of 14) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advopticalmat.de

Figure 11.  Transparent wood applications: a) a schematic of transparent wood roof for comfortable lighting conditions. Reproduced with permission.[18]
Copyright 2016, Wiley-VCH. b) A piece of Si QDs’ transparent wood showing diffused luminescence. Reproduced with permission.[9] Copyright 2017,
Wiley-VCH. c) Temperature changes after 10 min simulated solar radiation of model houses with CsxWO3/transparent wood (left) and glass (right) as
windows. The temperature readings are shown below the image. Reproduced with permission.[15] Copyright 2017, Royal Society of Chemistry. d) Current
density versus voltage characteristics for both bare GaAs cell (black) and the GaAs cell with light management wood coating (red). Reproduced with
permission.[14] Copyright 2016, Elsevier Ltd. e) Stepwise spectral change of the magenta to clear transparent wood ECD upon oxidation from −0.5 to
0.8 V and photographs of the device in its colorless (0.8 V) and colored (−0.5 V) states. Reproduced with permission.[56] Copyright 2018, Wiley-VCH.
f) Optical tunable smart wood window demonstration by lamination of PDLC film on transparent wood substrate. The optical property is tuned by
switching on/off electrical power. Reproduced with permission.[45] Copyright 2017, The Royal Society of Chemistry.

Another direction for transparent wood applications is as Rh6G molecules, and the optical feedback is realized since the
a structural material for photovoltaic devices such as solar aligned, hollow wood fibers in the structure are functioning as
cells and electrochromic devices.[14,45,56] Due to the high haze, small resonators. The emission line width is several nanome-
transparent wood can be designed as a light diffuse layer. ters since the effect is of collective nature and a consequence of
High haze means large scattering angles, which increase the variations in the wood structure. The fine structure of the lasing
length of the light path inside solar cells, so that efficiency is modes is similar to the lasing action of dye molecules in other
improved.[72,73] Based on this idea, Zhu et al. showed improved quasi-random media, for instance, based on sand grains.[74]
solar cell efficiency by positioning transparent wood on top of
solar cells (Figure 11d).[14] Within the transparent wood sub-
strate line, Lang et al. built conjugated polymer-based electro- 8. Conclusions and Outlook
chromic devices (ECD) with transparent wood as the substrate
(Figure 11e).[56] The devices exhibit a vibrant magenta-to-clear Transparent wood is an interesting topic in the emerging field
color change (∆E*  = 43, ∆%T  = 38%) with a high coloration of wood nanotechnology, both in academic and in industrial
efficiency (590 cm2 C−1) and low driving voltage (0.8 V). This context. Noticeable progress has been achieved, including prep-
concept will contribute to smart windows, which save energy. aration of transparent wood with improved mechanical prop-
Incident irradiation into a building is controlled through erties, improvement in making larger and thicker transparent
modulation of transmitted light with the help of ECD. Li et al. wood structures, realization of transparent wood functionaliza-
reported another type of switch on-demand smart window by tion, and the demonstration of applications in smart buildings
attaching polymer-dispersed liquid crystals (PDLC) film on top and photovoltaic devices. More work is required to understand
of transparent wood substrates (Figure 11f).[45] Optical device the light–wood interaction, to tune the optical and mechanical
properties were tuned by changing the alignment of the liquid properties, improve the “green” aspect, and to exploit advanced
crystals through application of an electric field. Without power transparent wood applications. Related questions include: what
supply, the device showed high haze and privacy protection. is the real refractive index of wood? What are the TMFP values
When power was applied, the device was transparent. This par- for different wood species? What are the correct values for trans-
ticular transparent wood substrate is modified to show high mittance and haze per unit length? Can scattering be predicted
transmittance but low haze. from complexity and configurational entropy considerations?
Transparent wood itself does not show intrinsic optical gain. Can we substantially increase the wood content in transparent
The addition of active optical media is needed to facilitate pho- wood? How can the chemical diffusion gradients inside thick
tonic applications. It was demonstrated that transparent wood and large wood specimens be minimized during bleaching and
doped with dye molecules can produce laser action under cer- impregnation processes? How can the content of functional
tain pumping conditions.[19] The active medium of the laser is particles be increased for the purpose of transparent wood

Adv. Optical Mater. 2018, 6, 1800059 1800059  (12 of 14) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advopticalmat.de

functionalization? In order to answer these questions, a system- [13] M. Zhu, J. Song, T. Li, A. Gong, Y. Wang, J. Dai, Y. Yao, W. Luo,
atic development of wood nanotechnologies is needed, with asso- D. Henderson, L. Hu, Adv. Mater. 2016, 28, 5181.
ciated efforts in nanoscience and characterization techniques. [14] M. Zhu, T. Li, C. S. Davis, Y. Yao, J. Dai, Y. Wang, F. AlQatari,
J. W. Gilman, L. Hu, Nano Energy 2016, 26, 332.
[15] Z. Yu, Y. Yao, J. Yao, L. Zhang, Z. Chen, Y. Gao, H. Luo, J. Mater.
Chem. A 2017, 5, 6019.
Acknowledgements [16] W. Gan, L. Gao, S. Xiao, W. Zhang, X. Zhan, J. Li, J. Mater. Sci. 2017,
52, 3321.
Y.Y.L. and E.V. contributed equally to this work. The authors acknowledge
[17] Y. Li, Q. Fu, R. Rojas, M. Yan, M. Lawoko, L. Berglund, Chem-
funding from the European Research Council Advanced Grant project
SusChem 2017, 10, 3445.
No 742733, Wood NanoTech, and Swedish Research Council (VR)
# 621-2012-4421. Min Yan from School of Engineering Science, Kungliga [18] T. Li, M. Zhu, Z. Yang, J. Song, J. Dai, Y. Yao, W. Luo, G. Pastel,
Tekniska Högskolan (KTH) Royal Institute of Technology, is acknowledged B. Yang, L. Hu, Adv. Energy Mater. 2016, 6, 1601122.
for the discussion and help during the preparation of this work. [19] E. Vasileva, Y. Li, I. Sychugov, M. Mensi, L. Berglund, S. Popov, Adv.
Opt. Mater. 2017, 5, 1700057.
[20] L. A. Berglund, I. Burgert, Adv. Mater. 2018, 30, 1704285.
[21] F. Jiang, T. Li, Y. Li, Y. Zhang, A. Gong, J. Dai, E. Hitz, W. Luo, L. Hu,
Conflict of Interest Adv. Mater. 2018, 30, 1703453.
[22] S. J. Eichhorn, A. Dufresne, M. Aranguren, N. E. Marcovich,
The authors declare no conflict of interest.
J. R. Capadona, S. J. Rowan, C. Weder, W. Thielemans,
M. Roman, S. Renneckar, W. Gindl, S. Veigel, J. Keckes, H. Yano,
K. Abe, M. Nogi, A. N. Nakagaito, A. Mangalam, J. Simonsen,
Keywords A. S. Benight, A. Bismarck, L. A. Berglund, T. Peijs, J. Mater. Sci.
2010, 45, 1.
biocomposites, optical property measurement, photonic devices, smart [23] R. J. Moon, A. Martini, J. Nairn, J. Simonsen, J. Youngblood, Chem.
buildings, transparent wood, wood modification Soc. Rev. 2011, 40, 3941.
[24] H. Zhu, Z. Fang, C. Preston, Y. Li, L. Hu, Energy Environ. Sci. 2014,
Received: January 16, 2018 7, 269.
Revised: March 13, 2018 [25] T. Koddenberg, Handbook of Wood Chemistry and Wood Composites,
Published online: May 11, 2018 CRC Press, Boca Raton, FL 2016.
[26] T. Saito, R. Kuramae, J. Wohlert, L. A. Berglund, A. Isogai, Biomacro-
molecules 2013, 14, 248.
[27] I. Sakurada, T. Ito, K. Nakamae, Makromol. Chem. 1964, 75, 1.
[1] T. W. Crowther, H. B. Glick, K. R. Covey, C. Bettigole, D. S. Maynard, [28] D. E. Kretschmann, in Wood Handbook: Wood as an Engineering Material
S. M. Thomas, J. R. Smith, G. Hintler, M. C. Duguid, G. Amatulli, (Ed: R. J. Ross), U.S. Department of Agriculture, Forest Service, Forest
M. N. Tuanmu, W. Jetz, C. Salas, C. Stam, D. Piotto, R. Tavani, Products Laboratory, Madison, WI 2010, p. 1.
S. Green, G. Bruce, S. J. Williams, S. K. Wiser, M. O. Huber, [29] B. M. Suleiman, J. Larfeldt, B. Leckner, M. Gustavsson, Wood Sci.
G. M. Hengeveld, G. J. Nabuurs, E. Tikhonova, P. Borchardt, Technol. 1999, 33, 465.
C. F. Li, L. W. Powrie, M. Fischer, A. Hemp, J. Homeier, P. Cho, [30] S. P. Simonaho, J. Palviainen, Y. Tolonen, R. Silvennoinen, Opt.
A. C. Vibrans, P. M. Umunay, S. L. Piao, C. W. Rowe, M. S. Ashton, Lasers Eng. 2004, 41, 95.
P. R. Crane, M. A. Bradford, Nature 2015, 525, 201. [31] A. Kienle, C. D’Andrea, F. Foschum, P. Taroni, A. Pifferi, Opt. Express
[2] a) U. G. K. Wegst, H. Bai, E. Saiz, A. P. Tomsia, R. O. Ritchie, 2008, 16, 9895.
Nat. Mater. 2015, 14, 23; b) J. Song, C. Chen, S. Zhu, M. Zhu, [32] S. P. Simonaho, R. Silvennoinen, Opt. Rev. 2004, 11, 308.
J. Dai, U. Ray, Y. Li, Y. Kuang, Y. Li, N. Quispe, Y. Yao, A. Gong, [33] P. McKendry, Bioresour. Technol. 2002, 83, 37.
U. H. Leiste, H. A. Bruck, J. Y. Zhu, A. Vellore, H. Li, M. L. Minus, [34] D. Jeremic, R. E. Goacher, R. Yan, C. Karunakaran, E. R. Master,
Z. Jia, A. Martini, T. Li, L. Hu, Nature 2018, 554, 224. Biotechnol. Biofuels 2014, 7, 496.
[3] Y. Li, Q. Fu, S. Yu, M. Yan, L. Berglund, Biomacromolecules 2016, 17, [35] U. P. Agarwal, Planta 2006, 224, 1141.
1358. [36] S. A. Ahmed, S. K. Chun, R. B. Miller, S. H. Chong, A. J. Kim,
[4] C. Chen, Y. Zhang, Y. Li, J. Dai, J. Song, Y. Yao, Y. Gong, J. Wood Sci. 2011, 57, 179.
I. Kierzewski, J. Xie, L. Hu, Energy Environ. Sci. 2017, 10, 538. [37] M. Emaminasab, A. Tarmian, R. Oladi, K. Pourtahmasi,
[5] F. Chen, A. S. Gong, M. Zhu, G. Chen, S. D. Lacey, F. Jiang, Y. Li, S. Avramidis, Wood Sci. Technol. 2017, 51, 261.
Y. Wang, J. Dai, Y. Yao, J. Song, B. Liu, K. Fu, S. Das, L. Hu, ACS [38] M. Plötze, P. Niemz, Eur. J. Wood Wood Prod. 2011, 69, 649.
Nano 2017, 11, 4275. [39] M. Zauer, S. Hempel, A. Pfriem, V. Mechtcherine, A. Wagenführ,
[6] T. Keplinger, E. Cabane, J. K. Berg, J. S. Segmehl, P. Bock, I. Burgert, Wood Sci. Technol. 2014, 48, 1229.
Adv. Mater. Interfaces 2016, 3, 1600233. [40] M. T. Postek, A. Vladár, J. Dagata, N. Farkas, B. Ming, R. Wagner,
[7] E. Cabane, T. Keplinger, T. Künniger, V. Merk, I. Burgert, Sci. Rep. A. Raman, R. J. Moon, R. Sabo, T. H. Wegner, J. Beecher, Meas. Sci.
2016, 6, 31287. Technol. 2011, 22, 024005.
[8] I. Burgert, E. Cabane, C. Zollfrank, L. Berglund, Int. Mater. Rev. [41] U. Müller, M. Rätzsch, M. Schwanninger, M. Steiner, H. Zöbl,
2015, 60, 431. J. Photochem. Photobiol., B 2003, 69, 97.
[9] Y. Li, S. Yu, J. G. C. Veinot, J. Linnros, L. Berglund, I. Sychugov, Adv. [42] H. S. Yaddanapudi, N. Hickerson, S. Saini, A. Tiwari, Vacuum 2017,
Opt. Mater. 2017, 5, 1600834. 146, 649.
[10] W. Gan, L. Gao, Q. Sun, C. Jin, Y. Lu, J. Li, Appl. Surf. Sci. 2015, 332, [43] W. Gan, S. Xiao, L. Gao, R. Gao, J. Li, X. Zhan, ACS Sustainable
565. Chem. Eng. 2017, 5, 3855.
[11] Y. Li, Q. Fu, X. Yang, L. Berglund, Philos. Trans. R. Soc. A 2018, 376, [44] M. Zhu, Y. Wang, S. Zhu, L. Xu, C. Jia, J. Dai, J. Song, Y. Yao,
20170182. Y. Wang, Y. Li, D. Henderson, W. Luo, H. Li, M. L. Minus, T. Li,
[12] S. Fink, Holzforschung 1992, 46, 403. L. Hu, Adv. Mater. 2017, 29, 1606284.

Adv. Optical Mater. 2018, 6, 1800059 1800059  (13 of 14) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.advopticalmat.de

[45] Y. Li, X. Yang, Q. Fu, R. Rojas, M. Yan, L. A. Berglund, J. Mater. [61] G. Chen, T. Chen, K. Hou, W. Ma, M. Tebyetekerwa, Y. Cheng,
Chem. A 2017, 6, 1094. W. Weng, M. Zhu, Carbon 2018, 127, 218.
[46] ASTM D1003-00, in Standard Test Method for Haze and Luminous [62] D. Mohan, C. U. Pittman, P. H. Steele, Energy Fuels 2006, 20,
Transmittance of Transparent Plastics, American Society for Testing 848.
and Materials, West Conshohocken, PA 2000, p. 2. [63] S. Trey, R. T. Olsson, V. Ström, L. Berglund, M. Johansson, RSC Adv.
[47] V. Ntziachristos, Nat. Methods 2010, 7, 603. 2014, 4, 35678.
[48] V. Durán, F. Soldevila, E. Irles, P. Clemente, E. Tajahuerce, [64] V. Merk, M. Chanana, N. Gierlinger, A. M. Hirt, I. Burgert, ACS
P. Andrés, J. Lancis, Opt. Express 2015, 23, 14424. Appl. Mater. Interfaces 2014, 6, 9760.
[49] F. W. Billmeyer, Y. Chen, Color Res. Appl. 1985, 10, 219. [65] Q. Fu, L. Medina, Y. Li, F. Carosio, A. Hajian, L. A. Berglund, ACS
[50] P. Grassberger, Int. J. Theor. Phys. 1986, 25, 907. Appl. Mater. Interfaces 2017, 9, 36154.
[51] T. Kaino, in Encyclopedia of Polymeric Nanomaterials [66] Y. Zhang, W. Luo, C. Wang, Y. Li, C. Chen, J. Song, J. Dai, E. M. Hitz,
(Eds: S. Kobayashi, K. Müllen), Springer, Berlin 2014, p. 1. S. Xu, C. Yang, Y. Wang, L. Hu, Proc. Natl. Acad. Sci. USA 2017, 114,
[52] H. Fukuzumi, T. Saito, T. Iwata, Y. Kumamoto, A. Isogai, Biomacro- 3584.
molecules 2009, 10, 162. [67] B. I. Hassel, S. Trey, S. Leijonmarck, M. Johansson, BioResource
[53] H. Zhu, S. Parvinian, C. Preston, O. Vaaland, Z. Ruan, L. Hu, 2014, 9, 5007.
Nanoscale 2013, 5, 3787. [68] F. J. Duarte, L. Hillman, Dye Laser Principles: with Applications,
[54] I. Siró, D. Plackett, M. Hedenqvist, M. Ankerfors, T. Lindström, Academic Press, San Diego, CA 1990.
J. Appl. Polym. Sci. 2011, 119, 2652. [69] D. Kurihara, Y. Mizuta, Y. Sato, T. Higashiyama, Development 2015,
[55] J. M. Dinwoodie, Timber: its nature and behaviour, E & FN Spon, 142, 4168.
London 2002. [70] C. A. Warner, M. L. Biedrzycki, S. S. Jacobs, R. J. Wisser, J. L. Caplan,
[56] A. W. Lang, Y. Li, M. De Keersmaecker, D. Eric Shen, A. M. Österholm, D. J. Sherrier, Plant Physiol. 2014, 166, 1684.
L. Berglund, J. R. Reynolds, ChemSusChem 2018, 11, 854. [71] H. Hama, H. Kurokawa, H. Kawano, R. Ando, T. Shimogori,
[57] C. M. Popescu, C. A. S. Hill, S. Curling, G. Ormondroyd, Y. Xie, H. Noda, K. Fukami, A. Sakaue-Sawano, A. Miyawaki, Nat. Neurosci.
J. Mater. Sci. 2014, 49, 2362. 2011, 14, 1481.
[58] H. Guo, D. Klose, Y. Hou, G. Jeschke, I. Burgert, ACS Appl. Mater. [72] Y. Chiba, A. Islam, Y. Watanabe, R. Komiya, N. Koide, L. Han, Jpn. J.
Interfaces 2017, 9, 39040. Appl. Phys., Part 2 2006, 45, L638.
[59] H. Guo, P. Fuchs, E. Cabane, B. Michen, H. Hagendorfer, [73] Z. Fang, H. Zhu, Y. Yuan, D. Ha, S. Zhu, C. Preston, Q. Chen, Y. Li,
Y. E. Romanyuk, I. Burgert, Holzforschung 2016, 70, 699. X. Han, S. Lee, G. Chen, T. Li, J. Munday, J. Huang, L. Hu, Nano
[60] J. Li, P. Xue, W. Ding, J. Han, G. Sun, Sol. Energy Mater. Sol. Cells Lett. 2014, 14, 765.
2009, 93, 1761. [74] A. Consoli, C. López, Sci. Rep. 2017, 7, 40141.

Adv. Optical Mater. 2018, 6, 1800059 1800059  (14 of 14) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

View publication stats

You might also like