Download as pdf or txt
Download as pdf or txt
You are on page 1of 29

Progress in Materials Science 132 (2023) 101025

Contents lists available at ScienceDirect

Progress in Materials Science


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

Transparent wood-based functional materials via a


top-down approach
Sailing Zhu a, 1, Subir Kumar Biswas b, 1, Zhe Qiu c, d, Yiying Yue f, Qiliang Fu a, e, *,
Feng Jiang c, *, Jingquan Han a, *
a
Co-Innovation Center of Efficient Processing and Utilization of Forest Resources, International Innovation Center for Forest Chemicals and
Materials, Joint International Research Lab of Lignocellulosic Functional Materials, College of Materials Science and Engineering, Nanjing Forestry
University, Nanjing 210037, China
b
Laboratory of Active Bio-Based Materials, Research Institute for Sustainable Humanosphere (RISH), Kyoto University, Uji, Kyoto 611-0011, Japan
c
Sustainable Functional Biomaterials Laboratory, Department of Wood Science, The University of British Columbia, Vancouver, BC V6T 1Z4,
Canada
d
Key Laboratory of Bio-based Material Science and Technology (Ministry of Education), College of Material Science and Engineering, Northeast
Forestry University, Hexing road 26, Harbin 150040, China
e
Scion, 49 Sala Street, Private Bag 3020, Rotorua 3046, New Zealand
f
College of Biology and Environment, Nanjing Forestry University, Nanjing 210037, China

A R T I C L E I N F O A B S T R A C T

Keywords: Transparent wood-based materials (TWMs), made by either bottom-up or top-down strategies,
Transparent wood have attracted great attention owing to their transparency, sustainability and multifunctionality.
Delignification Compared with materials prepared by a traditional bottom-up method, TWMs fabricated via a
Wood modification
top-down approach are promising as the natural hierarchical structure of wood is preserved with
Top-down approach
a high production efficiency, energy savings and scalability. Wood nanotechnologies, various
Structure and property
Functionalization and application chemical treatments and polymer impregnation or densification techniques have been developed
to fabricate TWMs, including transparent wood and transparent wood films, sharing similar
sustainability and optical transmittance while also preserving some uniqueness in applications
and compositions. Herein, we systematically provide an up-to-date summary on the state-of-the-
art for transparent wood and transparent wood films, with a special highlight on the proc­
ess–structure–property–application perspective, establishing the combination of wood science,
material chemistry and physics with new horizons in TWM science. The integrated topics of

Abbreviations: TWM, transparent wood-based material (fabricated via top-down approach); TW, transparent wood; TWF, transparent wood film;
BWS, bleached wood scaffold; CNF, cellulose nanofibril; e-skin, electronic skin; NW, natural/native wood; UV, ultraviolet; UVA, ultraviolet A; UVB,
ultraviolet B; UVC, ultraviolet C; UPF, UV protection factor; MA, maleic anhydride; IA, itaconic anhydride; SA, succinic anhydride; PMMA, poly­
methyl methacrylate; PVA, poly (vinyl alcohol); PVP, polyvinylpyrrolidone; PLIMA, poly(limonene acrylate); MF, melamine formaldehyde; TEMPO,
2,2,6,6-tetramethylpiperidinyl-1-oxyl; ATO, antimony-doped tin oxide; AgNW, silver nanowire; LED, light-emitting diode; ACEL, alternating current
electroluminescent; PET, polyethylene terephthalate; WLED, white light-emitting diode; FA, formaldehyde; ML, middle lamella; P, primary wall; S1,
outer layer; S2, middle layer; S3, inner layer; L-wood, transparent wood along the longitudinal direction; R-wood, transparent wood in the transverse
direction; PEAG, phosphate ester-polyethylene glycol; PDES, polymerizable deep eutectic solvents; CDs, carbon dots.
* Corresponding authors at: Co-Innovation Center of Efficient Processing and Utilization of Forest Resources, International Innovation Center for
Forest Chemicals and Materials, Joint International Research Lab of Lignocellulosic Functional Materials, College of Materials Science and Engi­
neering, Nanjing Forestry University, Nanjing 210037, China.
E-mail addresses: qiliang.fu@njfu.edu.cn (Q. Fu), feng.jiang@ubc.ca (F. Jiang), hjq@njfu.edu.cn (J. Han).
1
Equal contribution to this work.

https://doi.org/10.1016/j.pmatsci.2022.101025
Received 26 February 2022; Received in revised form 4 September 2022; Accepted 23 September 2022
Available online 3 October 2022
0079-6425/© 2022 Elsevier Ltd. All rights reserved.
S. Zhu et al. Progress in Materials Science 132 (2023) 101025

TWMs emphasize the relationship between nanostructure and properties. The structural and
functional design concepts are described for engineering production purposes. The emerging
applications of TWMs are also discussed for the potential replacement of glass and petroleum-
based plastics. Finally, an outlook and challenges are proposed for the future potential of func­
tional TWMs in engineered production.

1. Introduction

With the rising demand for energy and environmental conservation, eco-friendly materials are desired to replace petroleum-based
materials for sustainable development [1–3]. Natural wood (NW) has been widely used for tools, furniture and buildings over the past
thousands of years, owing to its wide availability, environmentally friendly characteristics, low thermal conductivity, ease of pro­
cessing and high specific strength [4–11]. Recently, nanotechnologies have been developed to modify and functionalize wood
structures at both the nanoscale and molecular levels, as well as to design wood-based functional materials [12–17]. The micro­
structure of wood can be precisely manipulated to introduce additional functionalities by chemical or physical modifications, which
can further extend the applications of wood-based composites in areas including water treatment [18–23], energy harvesting [24–28],
energy storage [29–38], electronic and optical applications [39–48], nanofluidic and hydrogel devices [49–54], and fire retardancy
[55]. Among various wood-based products, transparent wood-based materials are promising bio-based composites with high trans­
parency, low thermal conductivity and multifunctionalities [56–59]. These materials have promise in activities seeking the use of more
sustainable materials in emerging applications, such as energy-saving buildings, smart windows and optoelectronic devices [60–67].
Transparent wood-based materials can be prepared through either bottom-up or top-down approaches [15,68]. In the following
content of this review, all the transparent wood-based materials fabricated via a top-down approach are specially designated as TWMs.
In the bottom-up approach, materials are built from nanocellulose building blocks into thin films or bulk materials (e.g., aerogel and
hydrogel) [69–72]. Nanocellulose is disintegrated from tree, cotton, linen or bacteria cellulose by mechanical, chemical or biological
methods and can be further assembled into a cellulose nanofiber network [73–85]. In 2005, Yano et al. reported the fibrillation of
wood pulp into cellulose nanofibril (CNF) bundles via a mechanical disintegration process, followed by the assembly of nanofibrils to
create a transparent film, which exhibited a promising reinforcement effect for optically transparent composites [86]. On the other
hand, the top-down approach can directly convert NW into transparent wood or transparent thin films with excellent mechanical and
optical properties [87–89]. A porous bleached wood scaffold (BWS) can be prepared by chemically removing lignin and/or hemi­
cellulose from wood cell walls. The BWS is then impregnated by polymers with a refractive index similar to that of cellulose or
densified, fabricating transparent wood (TW) or a transparent wood film (TWF), respectively.
Constructing nanocellulose into transparent thin films based on a bottom-up approach has been thoroughly reviewed
[60,62,64–65,90–94]. Isotropic transparent nanocellulose films can be fabricated by vacuum filtration of nanocellulose colloidal
suspensions followed by resin impregnation, demonstrating excellent properties including high flexibility, superb mechanical and
optical properties, and low thermal expansion efficiency [60,95]. These attributes make nanocellulose-based transparent films
favorable substrates for many emerging applications in optoelectronics, photonics and electronics (e.g., sensors, organic light-emitting
diode displays and energy storage devices) [47–48,60,62,64–65,90,96–104]. In addition, CNFs can be chemically modified or com­
bined with inorganic nanoparticles/organic moieties, generating additional functionalities such as hydrophobicity, magnetic per­
formance, electrical conductivity, fire retardancy and luminescence [105–110]. Although the size of the CNFs and the structure of the
cellulosic films are controllable, the mechanical performances of these isotropic films are suboptimal. To enhance their mechanical
strength, transparent nanocellulose films with anisotropic structures have been intensively studied to uniaxially align CNFs to improve
mechanical properties via mechanical stretching and electric/magnetic field alignment [111–119]. In recent years, these topics have
been summarized in many review articles [120–121]. However, the production of either isotropic or anisotropic nanocellulose films
using the bottom-up method requires relatively complex processes and consumes tremendous amounts of energy, chemicals, water and
time. Moreover, the scalability of functional nanocellulose materials via the bottom-up strategy is challenging.
Alternatively, a top-down approach demonstrates great promise in preparing large-scale transparent wood-based products, which
can facilely overcome the scale-up issue as this strategy starts with NW and ends up with a bulk material [122–124]. The top-down
approach can bypass nanocellulose isolation and reconstruction processes, which are considered the most time- and energy- inten­
sive steps of the bottom-up method. For example, both TW and TWFs have been reported to be directly obtained from bulk wood
[87–89]. The cell walls of these materials with highly oriented cellulose are preserved, resulting in excellent mechanical properties.
More importantly, the preparation of TWMs is much easier, consuming less energy, chemicals, water and time than the bottom-up
approach. Although TW and TWFs are mainly composed of cellulose and directly converted from wood, they are theoretically
different materials in terms of three aspects. 1) Microstructure. TW maintains the honeycomb-like structure of wood cells, while TWFs
are completely densified, with a compact mortar-and-brick structure. 2) Chemical components. TWFs are polymer matrix-free ma­
terials with pure holocellulose. However, the impregnation of refractive-index-matched polymers is necessary in the fabrication of TW,
which means that the main components of TW are holocellulose and infiltrated polymers. 3) Transparency mechanism. The refractive-
index-matched phases of TW allow light transmission through the composite material, whereas photons travel through TWFs as the
defects between air and holocellulose are removed by densification. Therefore, TW and TWFs (both defined as TWMs) are considered
two different wood-based materials in this review.
The top-down method is a promising and straightforward strategy that has prompted scientists to demonstrate multifunctional

2
S. Zhu et al. Progress in Materials Science 132 (2023) 101025

Fig. 1. A brief summary of the inspiration and research progresses of TWMs, it is presented according to the timeline. Inspiration images.
Reproduced with permission [87]. Copyright 2012, Royal Society of Chemistry. Functional & structural TW image (left): Reproduced with
permission [89]. Copyright 2016, Wiley-VCH. Functional & structure TW image (right): Reproduced with permission [88]. Copyright 2016,
American Chemical Society. TW and Light management images: Reproduced with permission [125]. Copyright 2016, Elsevier. Energy efficient TW
image: Reproduced with permission [61]. Copyright 2016, Wiley-VCH. Luminescent TW image: Reproduced with permission [126]. Copyright
2017, Wiley-VCH. Luminescent & magnetic TW image: Reproduced with permission [127]. Copyright 2017, American Chemical Society. Heat-
shielding TW image: Reproduced with permission [128]. Copyright 2017, Royal Society of Chemistry. TWF image: Reproduced with permission
[115]. Copyright 2017, Wiley-VCH. Transparent plywood image: Reproduced with permission [129]. Copyright 2018, Elsevier. Large-thickness TW
image: Reproduced with permission [130]. Copyright 2018, Royal Society of Chemistry. Electrochromic device image: Reproduced with permission
[63]. Copyright 2018, Wiley-VCH. Flexible & conductive TW image: Reproduced with permission [131]. Copyright 2018, Royal Society of
Chemistry. Low haze TW image: Reproduced with permission [66]. Copyright 2019, American Chemical Society. Photochromic TW image:
Reproduced with permission [132]. Copyright 2019, Royal Society of Chemistry. ACEL device image: Reproduced with permission [133]. Copyright
2019, American Chemical Society. Heat-storage TW image: Reproduced with permission [134]. Copyright 2019, American Chemical Society. TW
solar cell image: Reproduced with permission [135]. Copyright 2019, American Chemical Society. Aesthetic TW: Reproduced with permission
[136]. Copyright 2020, Springer Nature. Flame-retardant TW image: Reproduced with permission [137]. Copyright 2020, Elsevier. Flexible TWF
sensor image: Reproduced with permission [138]. Copyright 2020, American Chemical Society. Luminescent TWF image: Reproduced with
permission [139]. Copyright 2020, American Chemical Society. TW gas sensor and TW white light-emitting diode (WLED) images: Reproduced with
permission [140]. Copyright 2020, American Chemical Society. Patternable TW image: Reproduced with permission [124]. Copyright 2021,
American Association for the Advancement of Science. Coloured TW image: Reproduced with permission [141]. Copyright 2021, American
Chemical Society. Shape-memory TW image: Reproduced with permission [142]. Copyright 2021, Elsevier. Fully bio-based TW image: Reproduced
with permission [143]. Copyright 2021, Wiley-VCH. TW e-skin image: Reproduced with permission [144]. Copyright 2021, Elsevier. Photon
diffusion in TW image: Reproduced with permission [145]. Copyright 2022, Wiley-VCH. Smart packaging TW: Reproduced with permission [146].
Copyright 2022, Elsevier.

3
S. Zhu et al. Progress in Materials Science 132 (2023) 101025

TWMs in terms of material design concepts and functionalization routes [12]. Although recent progress on TW has been summarized
[57–58,147–149], TWFs made from top-down approaches have not yet been overviewed. Herein, the state-of-the-art TWMs are
systematically overviewed, including TW and TWFs, as shown in Fig. 1. In particular, this is the first time that TWFs have been
reviewed to the best of our knowledge. The materials science of the process–structure–property–application relationship is discussed.
This consists of applying material chemistry and nanotechnology to tailor the wood chemical structures (lignin removal and modi­
fication), considering nanomorphologies (generation of meso- and/or nanostructures) and adding new functionalities (electrical
conductivity, fire retardancy, thermal insulating and optical properties) to TWMs. TWMs are discussed as potential replacements of
traditional glass and fossil-based plastics in emerging applications. Finally, we highlight some opportunities and challenges in the
engineered production of functional and sustainable TWMs.

2. Transparency mechanism and processing methods

NW is opaque under visible light due to light scattering and reflection at the interface between cell walls and lumens, as well as the
strong light absorption by wood tissue. Approximately 80 ~ 95 % of the light absorption in wood is due to the chromophores associated
with lignin and some extractives [152]. Lignin units have complex chromophores (e.g., carbon–carbon double bonds conjugated with
aromatic rings) that are responsible for the brownish color and light absorption in NW [15,153]. It has been reported that the refractive
indices (n) of lignin, hemicellulose and cellulose are 1.61, 1.53 and 1.53, respectively [154]. The multiscale porous structural com­
bination of micro, meso- and macropores leads to much light scattering because of mismatched refractive indices at the boundaries
between the air-filled (n = 1.0) lumen spaces and cell walls (average n≈1.56) [57,155]. To achieve TWMs, two main steps are
necessary: removing the light-absorbent components (including removing and modifying lignin) and eliminating the mismatched
refractive index in the BWS (average n≈1.53) [88]. In terms of reducing the differences in the refractive index of the TWM, there are
two strategies via the top-down approach (Fig. 2a): 1) impregnating refractive index-matched polymers into BWSs and 2) densifying
BWSs into transparent thin films.

Fig. 2. (a) Schematically illustrate the fabrication of TW and TWFs: Reproduced with permission [150]. Copyright 2019, Elsevier. SEM image of
TW: Reproduced with permission [130]. Copyright 2018, Royal Society of Chemistry. SEM image of TWF: Reproduced with permission [138].
Copyright 2020, American Chemical Society. SEM images of cell wall structures of (b) NW, (c) delignified BWS, and (d) lignin-retaining BWS. The
insets are low-magnification SEM images. Red arrows point to the middle lamella-rich region in lignin. (b-d) Reproduced with permission [151].
Copyright 2017, Wiley-VCH.

4
S. Zhu et al. Progress in Materials Science 132 (2023) 101025

The preparation of a BWS generally starts from removing or modifying the lignin from NW. A range of different chemical treatment
processes, such as delignification (lignin removal) and removal of the chromophores in lignin (lignin modification), are used
[68,156–160]. It is crucial to manipulate nanostructured wood by precisely controlling its nanopores in the cell walls through these
chemical treatments and process conditions (i.e., treatment time, the concentration of chemical solutions, pH value, and temperature).
Acidic sodium chlorite (NaClO2) treatment is commonly used for delignification. NW is mildly treated using 1 ~ 5 wt% NaClO2
aqueous solution in acetate buffer solution with a pH of 4.6 for 6 ~ 18 h (depending on the size and cutting direction of the sample)
[88,133,135]. As a result, most lignin is removed from the cell walls, especially in the lignin-rich middle lamella, resulting in the
creation of microscale and nanoscale pores (Fig. 2c). The original honeycomb structure of NW is preserved in the BWS (Fig. 2b).
NaClO2 generates unstable chlorous acid in acetate buffer solution, which can be further decomposed into chlorine and chlorine di­
oxide to react with lignin by methods including oxidation of the phenolic structure and/or phenoxyl-radical ring opening [161–162].
The lignin content of balsa wood decreased from 24.9 to 2.9 wt% due to the breakdown of lignin carbohydrate complexes [88]. A

Fig. 3. (a) The hierarchical structure of wood from the tree to the cellulose molecule chain. Tree image: Reproduced with permission [215].
Copyright 2021, Wiley-VCH. Transverse section image: Reproduced with permission [155]. Copyright 2016, American Chemical Society. Vascular
tissue, Microfibril bundle, Elementary fibril and Cellulose chain images: Reproduced with permission [122]. Copyright 2020, Springer Nature. Cell
wall image: Reproduced with permission [12]. Copyright 2018, Wiley-VCH. (b) The photograph and morphology of TW. TW image: Reproduced
with permission [89]. Copyright 2016, Wiley-VCH. The center SEM image: Reproduced with permission [143]. Copyright 2021, Wiley-VCH. The
right SEM image: Reproduced with permission [129]. Copyright 2018, Elsevier. (c) The photograph and morphology of TWF. Reproduced with
permission [138]. Copyright 2020, American Chemical Society.

5
Table 1

S. Zhu et al.
Optical properties (550 nm), mechanical performances and thermal properties for TWM.
Sample Wood Thickness Density [g Wood (cellulose)a Optical Mechanical Thermal Ref.
species [mm] cm¡3] [vol%]
trb haze E σ ε [%] toughness [MJ/ tcc [W/m dd
[%] [%] [Gpa] [Mpa] m¡3(¡|-)] K¡1]

Delignified/PMMA Beech 0.1 70 2.5 150 ~10 [218]


0.3 30 18
0.6 18 40
0.7 15 49
Delignified/PMMA/PEG Birch 0.5 84 74 [134]
1.5 68 77 ~14.9 ~70.5 0.3
Delignified/PMMA Balsa 0.7 90 ~48 [88]
3.7 40 ~80
1.2 (5) 85 ~70 ~2.05 38 ~2.2
1.2 (19) 78 ~75 ~3.59 ~90.1 ~3.8
1.2 (65) 34.6 ~78
Delignified/PMMA Birch 1.3 25 70 70 19 270 [219]
Haze wood/epoxy Basswood (0.7) 18.4 (14.7) 90 80 ~76.36 ~2.21 ~0.87 0.24 // [66]
(0.7) 18.4 (14.7) ~14.60 ~1.88 ~0.13 ⊥
Clear wood/epoxy Basswood (0.7) 2.9 (2.5) 92 10 43.39 19.14 ~6.1 0.35 //
(0.7) 2.9 (2.5) 33.29 16.3 ~3.96 ⊥
Delignified/LIMA Balsa 1.2 12 87 46 [143]
SA esterified/ delignified/ Balsa 1.2 12 89 41
LIMA
Delignified/LIMA Balsa 2.0 12 80 62
SA esterified/ delignified/ Balsa 2.0 12 87 45
6

LIMA
Delignified/LIMA Balsa 3.0 12 71 65
SA esterified/ delignified/ Balsa 3.0 12 81 51
LIMA
Delignified/LIMA Birch 0.7 26 88 49 12.6 146.6 1.2
SA esterified/ delignified/ Birch 0.7 26 92 44 17.3 173.6 1.0
LIMA
Delignified/PMMA Balsa 2.0 70.6 76.3 4.8 41.4 // [182]
2.9 11.4 ⊥
MA esterified/ delignified/ Balsa 2.0 79.5 73.6 4.3 46.2 //
PMMA

Progress in Materials Science 132 (2023) 101025


3.1 18 ⊥
IA esterified/ delignified/ Balsa 2.0 76.6 76.2 4.3 54.9 //
PMMA
3.3 33.2 ⊥
SA esterified/ delignified/ Balsa 2.0 83.8 66.9 4.5 61.7 //
PMMA
2.9 12.8 ⊥
Delignified/PMMA Balsa 1.5 83 70 [130]
Acetylated/ delignified/ Balsa 1.5 5 92 50 4 78.9 ~ 2.2 1.0
PMMA
Delignified/epoxy Basswood (3) 90 2.37 45.38 1.2 // [89]
(3) ~80 1.22 23.38 0.59 ⊥
Delignified/PMMA Balsa 0.8 12 ~95 ~60 ~4.3 ~62.5 1.5 // [129]
0.8 12 ~2.4 ~14.6 0.7 ⊥
(continued on next page)
Table 1 (continued )
Sample Wood Thickness Density [g Wood (cellulose)a Optical Mechanical Thermal Ref.

S. Zhu et al.
species [mm] cm¡3] [vol%]
trb haze E σ ε [%] toughness [MJ/ tcc [W/m dd
[%] [%] [Gpa] [Mpa] m¡3(¡|-)] K¡1]

Ply 0/90/0/-90/ Balsa 3.5 10 83 80 ~4.1 ~50.1 1.2 //


0 delignified/PMMA
3.5 10 ~3.9 ~44.9 1.3 ⊥
Ply 0/45/90/-45/ Balsa 3.5 10 75 80 ~3.9 ~45.4 1.2 //
0 delignified/PMMA
3.5 10 ~3.5 ~42 1.4 ⊥
Delignified/PMMA Basswood 1.6 66 ~85 ~14.4 ~131.6 // [220]
1.6 ~3.43 ~9.11 ⊥
Acetylated/ delignified/ Basswood 1.6 72 ~79 ~12.0 ~72.7 //
PMMA
1.6 ~3.75 ~11.49 ⊥
Ply delignified/PMMA Basswood 3.2 ~90 6.7 75
Ply acetylated/ delignified/ Basswood 3.2 ~88 6.8 42
PMMA
Delignified/epoxy Balsa 2 77 ~69.13 // [221]
2 ~6.06 ⊥
Basswood 2 64 ~75.12 //
2 ~8.17 ⊥
Clear wood/PVA Balsa 0.8 91 15 ~3.85 ~143 4.2 ~3.03 ~0.39 // [67]
0.8 ~3.10 ~67 3.7 ~1.06 ~0.19 ⊥
Delignified/PMMA Balsa 1.5 86 68 [151]
Lignin-retaining/PMMA Balsa 1.5 83 75 ~100.7 ~2.18
Delignified/thiol − ene Balsa 1.2 ~5.0 85 63 3.2 53.7 2.1 [184]
7

Lignin-retaining/thiol − Balsa 1.2 ~4.3 90 36 3.4 59 2.5


ene
Lignin-retaining/UV resin Balsa 46.7 95.2 0.678 38.32 [222]
Compressed lignin- 80.8 66.7 2.276 113.75
retaining/UV resin
Delignified/PEAG Balsa 20 93 98 2.2 153.6 // [200]
Delignified/epoxy Poplar 0.09 89.4 // [198]
TWF Basswood 0.05 (0.46) ~1.40 ~90 ~86 [223]
TWF Basswood 0.08 (0.4) 90 350 23.2 7.38 [115]
TWF Balsa 0.055 (1) ~1.20 80 85 49.9 469.9 // [138]
0.1 ~74
0.14 ~70

Progress in Materials Science 132 (2023) 101025


0.2 ~60
TWF Balsa 0.06 (1) ~1.20 81 42.7 393.8 // [139]
0.06 (1) 5.7 76 ⊥
Luminent TWF 0.075 (1) 70 24.4 292 //
Luminent TWF 0.075 (1) 5.3 64 ⊥
TWF Balsa 0.08 (2) ~1.32 80 70 ~51.1 ~449.1 1.75 [191]
TWF Balsa 0.06 ~1.24 77 73.5 426 [189]
TWF Balsa 0.055 ~1.25 24.54 196.8 // [224]
0.055 10.97 37.93 ⊥

The‖symbol represents the light/applied stress/transferred heat along the wood channel direction. The⊥symbol represents the light/applied stress/transferred heat perpendicular to the wood channel
direction.
a
The wood volume fraction denotes the sum of the volume fractions lignin, hemicellulose and cellulose (the cellulose volume fraction refers to the volume fraction of cellulose);
b
tr refers to transmittance;
c
tc refers to the thermal conductivity;
d
d refers to the direction.
S. Zhu et al. Progress in Materials Science 132 (2023) 101025

similar study using sodium hypochlorite (NaClO) was reported by Fink in 1992 [154]. During the delignification treatment, oxidation
and chlorination reactions attacked the benzoquinone structure and the conjugated double bond of the side chain in lignin, leading to
dissolution of lignin in water. A pulping method using alkaline sodium sulfite (Na2SO3 + NaOH) was reported by Hu’s group [89].
Under alkaline cooking conditions, sulfite, bisulfite and hydroxide ions interacted with lignin carbohydrate complexes, enabling the
cleavage of ether bonds [163]. An additional hydrogen peroxide (H2O2) bleaching process was performed, resulting in a lignin content
of less than 3 wt%. In another strategy, peracetic acid or glacial acetic acid/H2O2 solution can selectively remove lignin [55,164–165].
Hydroxonium ions generated from peroxy bonds can attack the electron-rich positions in lignin (olephinic side chain and benzene
ring), resulting in the cleavage of water-soluble lignin [161]. Although several other delignification treatments have also been carried
out (e.g., ionic liquid, deep eutectic solvent or biological enzyme treatments) [166–176], the delignification mechanism of these
treatments is still unclear and needs further investigation.
Unlike the lignin-removal method, the lignin-modification method does not extract lignin but rather changes its structure and
simultaneously removes chromophores. Jungstedt et al. established a suitable analysis method to combine digital image correlation
and a finite element model for the investigation of the fracture toughness of transparent wood [177]. The results show the nonlinear
localized strain and stress development in the cohesive zone. The removal of lignin results in substantially inferior bulk wood structural
properties, since lignin acts as a binder that cross-links holocellulose and the weakened delignified middle lamella facilitates crack
growth [177]. Moreover, during delignification processes, some toxic chemicals (e.g., chlorine, methyl mercaptan, hydrogen sulfide
and dimethyl sulfide) are generated [178]. This is not environmentally friendly in large-scale production. To address these problems, Li
et al. reported a lignin-retaining strategy using a bleaching solution (Na2SO3/NaOH/H2O2), showing that>80 % of the lignin content
was retained [151]. Thus, the cell wall of wood does not contain substantial micro- and nanoscale pores (Fig. 2d), which results in a
better-preserved wood structure and favorable mechanical properties under wet conditions. Subsequently, Xia et al. explored an
improved lignin-modification method that consumed less chemicals and energy than the lignin-removal method [124,179]. This rapid
lignin-modification technique was developed by combining H2O2 brushing and photocatalysis (solar or ultraviolet (UV)) to remove
chromophores rather than immersing NW in the bleaching solution. As immersion treatment is not needed, much larger NW (400 ×
110 × 1 mm3 or even up to ~ 1 m in length) can be used to fabricate a strong and robust BWS. This would allow for the potential

Fig. 4. (a) Optical transmittance of TW with different thicknesses. Reproduced with permission [88]. Copyright 2016, American Chemical Society.
(b) Optical transmittance of TWF with different thicknesses. Reproduced with permission [138]. Copyright 2020, American Chemical Society. (c)
Images of the transverse-cut (left) and longitudinal-cut (right) TW and their scattered light spot, respectively. Reproduced with permission [89].
Copyright 2016, Wiley-VCH. (d) Optical transmittance of acetylated and non-acetylated TW. Reproduced with permission [130]. Copyright 2018,
Royal Society of Chemistry. (e) Optical transmittance of TW without or with esterification. Reproduced with permission [182]. Copyright 2020,
Royal Society of Chemistry. (f) Optical transmittance and haze of low/high haze TW. Reproduced with permission [184]. Copyright 2020, American
Chemical Society.

8
S. Zhu et al. Progress in Materials Science 132 (2023) 101025

scalable manufacturing of larger-scale sustainable wood-based functional materials.


Eliminating light scattering of wood is the second step in fabricating TWMs (Fig. 2a), which can be achieved through two main
approaches: i) the polymer impregnation (backfilling) method and ii) the densification method. For the polymer impregnation method,
the polymers with a refractive index matched with that of the wood cell wall are infiltrated into the BWS under vacuum to obtain TW.
TW was first reported by Fink (in 1992) in a methodology study [154]. Inspired by the transparent crab shell reported by Yano et al.
[87], Berglund’s group developed TW with tunable optical properties and high tensile strength, which was obtained by impregnating
the lumen and cell walls of a balsa BWS with refractive-index-matched polymethyl methacrylate (PMMA, n≈1.49) [88]. At the same
time, Hu’s group backfilled epoxy resin (n≈1.50) into a basswood BWS, resulting in a highly anisotropic and highly transparent wood
composite [89]. Because of the excellent optical transparency (~85 %), haze (~95 %), light-guiding effect, thermal insulation (thermal
conductivity = 0.32 W m− 1 K− 1 along the longitudinal direction) and impact absorption capability, the obtained TW exhibited great
promise as an energy-efficient construction material [61]. Apart from PMMA and epoxy resin, other refractive-index-matched poly­
mers [88,180–183], including poly (vinyl alcohol) (PVA, n≈1.53) [153], polyvinylpyrrolidone (PVP, n≈1.53) [125], polyimide
(precursor, n≈1.45) [137], thiol-ene thermosets (n≈1.56) [184], vitrimers [142,185], poly (limonene acrylate) (PLIMA, n≈1.52)
[143]and melamine formaldehyde (MF, n≈1.50) [186], are also suitable alternatives to infiltrate into BWSs for the preparation of TW.
In some cases, the shrinkage of polymers during the polymerization/curing process inevitably causes cracks or defects along the
interface between infiltrated polymers and BWSs, which has an adverse influence on the optical and mechanical performances of TW
[130]. This could be improved by selecting a thiol-ene thermoset with low shrinkage strains as the backfilling polymer to reduce the
interfacial air gaps [184]. Surface modification (e.g., esterification) of BWSs is also used to enhance the interfacial compatibility
between impregnated polymers and BWSs, resulting in improved transparency [130,143,182,187].
Densification is another strategy that converts BWSs into TWMs with or without external forces to obtain TWFs, which is vital to
achieve TWMs with interfacial defects free and high cellulose volume fractions (80 ~ 90 %) [115,138,188–189]. For example, an
anisotropic TWF was prepared with extensive lignin and hemicellulose removal prior to densification and dehydration processes at
room temperature [138–139]. Water molecule-induced hydrogen bonding (microfibrils and cell walls) is driven with the assistance of
compression [190]. The wood cell walls were condensed and collapsed to form a stacked layered structure (Fig. 2a). Another TWF was
produced by self-densifying 2,2,6,6-tetramethylpiperidinyl-1-oxyl (TEMPO)-oxidized mesoporous wood under ambient conditions
without mechanical pressing or heating [191]. Carboxyl groups were introduced to the TEMPO-oxidized cellulose microfibrils within
the cell walls. Once the two sides of the cell walls were in contact, strong hydrogen bonds formed between the abundant carboxyl and
hydroxyl groups on the cellulose microfibrils, leading to complete collapse of the wood cell walls and the formation of a self-densified
TWF. Although this self-densification method is a more facile than routine densification, the introduction of TEMPO treatment is toxic
and thus leads to an environmental issue.
TWMs have been developed for engineering purposes, combining structural and functional performance. The sustainability and
environmental impacts of TWMs are of significance because of the long service time for many applications, such as buildings, elec­
tronics and photonics. Berglund’s group intended to improve the fabrication processes and environmental sustainability, for instance,
using UV-light assisted polymerization, selection of bio-based polymers, enhancing compatibility and using green processes
[17,130,143,151,182,184]. On the other hand, Hu’s group and most international researchers emphasize functionalization and the
applications of TWMs. Various functional TWMs can be obtained by modifying the BWS, selecting different backfilling polymers or by
adding functional nanoparticles (Fig. 1) to achieve functionalities including luminescence, magnetic and heat-shielding performance,
strain and gas sensing, conductivity, energy conversion, fire resistance, UV blocking, shape memory, thermal response and man­
agement [24,126–128,133–134,138,140,186,189,192–211]. The engineering production cost, scalability and efficiency should be
considered in the future development of TWMs.

3. Structure–property–application

3.1. Structure

Wood has a complex hierarchical structure, from the meter-scale trunk down to nanoscale elementary fibrils and molecular-scale
cellulose chains (Fig. 3a) [212]. Both softwood and hardwood contain different types of cells (for example, vessels and fibers in
hardwood, and tracheids in softwood). The lumens within the cells play an important role in the transportation of water/nutrients in
wood tissue. Cell walls in wood tissue consist of multiple layers, including cell corner middle lamella, middle lamella, primary wall (P),
outer layer (S1), middle layer (S2) and inner layer (S3) of the secondary wall [213]. There are three main components (lignin,
hemicelluloses, and cellulose) and other extractives in the rigid cell walls of wood. The cellulose fibrils are oriented as a skeleton in the
cell wall, embedding in the hemicellulose and lignin matrix [12,214]. At the nanoscale, the strong and stiff cellulose fibril bundles (15
~ 50 nm wide) are constructed from cellulose elementary fibrils (3 ~ 5 nm wide), which are aligned almost parallel along the long
fiberaxis in the S2 layer with small microfibril angles [7]. Various hypotheses have been proposed regarding cellulose biosynthesis in
microfibrils, and it is widely accepted that elementary cellulose fibrils contain a crystalline region where individual cellulose chains are
assembled in an orderly arrangement by intrachain/interchain hydrogen bonding and an amorphous region with a less regular dis­
tribution [213]. At the molecular scale, the linear cellulose chains consist of repeating units of D-glucose connected to each other with a
1,4-β-glycosidic linkage (Fig. 3a). Taking advantage of the hierarchical wood structure, porosity and cell wall assembly, novel func­
tional wood-based materials can be designed, for example, TW and TWFs.
TW and TWFs have a unique anisotropic structure with highly aligned cellulose microfibrils (Fig. 3b and c). Polymers are suc­
cessfully impregnated into the voids and cell walls of the BWS (Fig. 3b center) [143,216–217]. The cellulose microfibrils maintain the

9
S. Zhu et al. Progress in Materials Science 132 (2023) 101025

(caption on next page)

10
S. Zhu et al. Progress in Materials Science 132 (2023) 101025

Fig. 5. (a) Stress–strain curves for TW with different cellulose volume fraction, delignified wood and PMMA. Reproduced with permission [88].
Copyright 2016, American Chemical Society. (b) Stress–strain curves of TWF and other materials. (c) The specific Young’s modulus and specific
strength of TWF and other materials. (b-c) Reproduced with permission [138]. Copyright 2020, American Chemical Society. (d) Stress–strain curves
of the anisotropic films with distinct wood cutting angles of 0◦ , 22.5◦ and 45◦ . Reproduced with permission [115]. Copyright 2017, Wiley-VCH. (e)
Stress–strain curves for NW and TW along and perpendicular to the wood channel direction. Reproduced with permission [89]. Copyright 2016,
Wiley-VCH. (f) Stress–strain curves for the transparent plywood. Reproduced with permission [129]. Copyright 2018, Elsevier. (g) Interactions
between BWS and LIMA (left) and stress–strain curves (right) of TW with and without SA modification. Reproduced with permission [143].
Copyright 2021, Wiley-VCH.

original orientation and are densely packed with each other (Fig. 3b right). Similarly, TWFs present a highly aligned cellulose
microfibril structure without polymer impregnation (Fig. 3c right). The densified cell walls show a mortar-and-brick structure at the
cross-section of the TWF (Fig. 3c center). As they retain the anisotropic structure of the BWS, TW and TWFs possess a range of
anisotropic properties, such as mechanical properties, optical performance and thermal conductivity [122].

3.2. Property

The optical, mechanical and thermal properties of TWMs are significantly impacted by several factors at the micro/nano and
molecular scales, such as the cellulose volume fraction, cellulose microfibril orientation and chemical interactions at the cell wall
interfaces. In addition, these properties are also affected by the thickness, porosity and anisotropic structure of TWMs. The experi­
mental results are referenced from the literature, as presented in Table 1.

3.2.1. Optical properties


In general, the optical properties of TWMs are characterized by measuring transmittance and haze. Haze can be calculated by the
ratio between the intensity of diffused light and the intensity of total transmitted light [57]. TW has a wide range of tunable trans­
mittance (10 ~ 93 %) and adjustable haze (10 ~ 98 %) values, depending on several factors (Table 1), such as wood species, thickness,
cellulose volume fraction and microstructure, which have been summarized in previous reviews [57–58,148]. However, the optical
properties of TWFs have not yet been overviewed. Herein, the effects of wood thickness, anisotropy, lamination and interfacial design
on the optical properties of TW and TWFs (TWMs) are highlighted in this section.
Thickness is one of the main factors governing the optical properties of TWMs. The optical transmittance of TW decreases with
increasing thickness, whereas haze increases [88]. Taking TW (balsa wood) as an example, when the thickness rose from 0.7 mm to 3.7
mm, the transmittance declined from 90 % to 40 %, while the haze increased from 48 % to 80 % (Fig. 4a) [88]. A similar trend was
reported for TW made from beech wood or birch wood [134,218]. This is ascribed to the prolonged light propagation pathway and the
increased interface quantity of polymer/wood, resulting in light attenuation and scattering within TW, respectively. The quantitative
relationship between the total transmittance (Ttotal ) and thickness of TWMs can be expressed by the anisotropic photon diffusion
equation (Equation (1)), showing that Ttotal decays exponentially with increasing TWM thickness [225]. Similarly, the TWF trans­
mittance decreases accordingly with the increasing thickness (Fig. 4b). For example, the transmittance of TWFs was reduced from ~
65 % to ~ 60 % when the thickness was increased from 140 µm to 200 µm [138]. This may be explained by the fact that more air-filled
pores are present at increased thickness, resulting in more light scattering in the TWF.
Ttotal = exp( − ad) (1)
where the constant a is the attenuation coefficient and d is the TWM thickness.
The anisotropic structure of TWMs is expected to result in anisotropic properties in the longitudinal and transverse directions of
TWMs. The light scattering pattern is isotropic for TW in the transverse direction (R-wood), whereas light scattering in TW along the
longitudinal direction (L-wood) is highly anisotropic (Fig. 4c). Interestingly, R-wood showed a higher transmittance (90 %) and haze
(~96 %) than L-wood (80 % transmittance and 90 % haze) due to light following the vertically aligned channels of R-wood with high
propagation scattering [89,226–227]. Similar anisotropic light scattering behavior was observed for TWFs [138,191]. Light propa­
gation was guided along the wood channel direction, resulting in an anisotropic scattering pattern. The wave-guiding property also
leads to an ellipse-shaped scattering pattern with diverse intensities in the × and y directions.
Lamination is an ingenious way of adjusting the optical properties of TWMs from anisotropic to isotropic. Isotropic TWMs can be
achieved by lamination of several layers with cross-ply or quasi-isotropic structures [129,228–229]. Fu et al. investigated transparent
plywood stacked five-layer structures with different fiber orientations. Instead of the squeezed parallelogram-like scattering pattern of
single-ply TW, an almost circularly scattered pattern was projected for the quasi-isotropic transparent plywood with approximately 75
% transmittance and 80 % haze. The light was confined by the laminated fiber in the transparent plywood structure, leading to
isotropic light scattering performance [129]. Similarly, the lamination strategy can also be used to tune the optical performance of
TWFs. For example, anisotropic TWFs can be assembled into a cross-laminated TWF with a double-layer structure and isotropic optical
properties [115]. In another study, isotropic light scattering and uniform illumination of TWFs was achieved by adding CdSe/ZnS
quantum dots [139]. CdSe/ZnS quantum dots with a core–shell structure can be stimulated by UV light to achieve a quantum yield of
up to 35 %. A ZnS shell layer encapsulating a core layer (CdSe) can passivate surface traps, increasing the nonradiative de-excitation
sites for photon-generated charges. The photoluminescence from the quantum dot core layers diffuse new light through the ZnS atomic
crystalline, which can be uniformly spread over the entire surface of the TWF (approximately 70 % transmittance and 81 % haze)
[139].

11
S. Zhu et al. Progress in Materials Science 132 (2023) 101025

Interfacial design of cell walls/polymers is an appealing strategy that allows the optical performance of TWMs to be tailored. To
date, esterification strategies have been reported for the chemical functionalization of cellulose microfibrils to design interfaces be­
tween cell walls and impregnated polymers [130,143,182,184,187,191,230]. The formation of interfacial gaps usually occurs between
the polymers and the inner cell wall due to different shrinkage ratio and poor interface compatibility [145]. The former can be solved
by selecting polymers with smaller shrinkage ratio, such as thiol-ene thermosets [184]. The latter may be addressed by inducing acetyl
groups on the cellulose microfibrils, which can improve the interfacial compatibility between the acylated BWS and the hydrophobic
polymers (e.g., PMMA). For example, acetyl groups were esterified onto cellulose microfibrils using acetate anhydride, resulting in
improved interfacial compatibility [130]. Compared to the nonacetylated TW, acetylated TW with improved interfacial compatibility
led to an increase in transmittance from 83 % to 92 % (Fig. 4d), while the haze decreased from 70 % to 50 % [130]. Another ester­
ification strategy has also been used to modify a BWS to improve interfacial compatibility with impregnated polymers [143,182]. For
example, the chemical modification of a BWS was carried out with renewable cyclic anhydrides, including succinic anhydride (SA),
itaconic anhydride (IA) and maleic anhydride (MA) (Fig. 4e) [182]. The surface modification of the BWS by ring-opening esterification
of anhydrides produces new active sites (e.g., mono- and di-substituted alkenes and terminal carboxylic acid groups), enabling further
interaction with impregnated polymers. Taking the SA-modified BWS as an example, after impregnating PMMA polymer, TW-SA
showed 19 % higher transmittance value and 14 % lower haze value, respectively, than those of unmodified TW due to the
improved interfacial compatibility (Fig. 4e). In a similar report, a fully bio-based TW-esterified material was prepared by impregnation
of a refractive index-matched poly(limonene acrylate) (PLIMA) polymer into an SA-modified BWS with improved interface adhesion,
resulting in 10 % higher optical transmittance and 14 % lower haze than those of unmodified TW [143].

3.2.2. Mechanical properties


The mechanical properties of TW strongly depend on the cellulose volume fraction, wood species, microstructure and wood
densities (Table 1). These have been reviewed by Li et al. but without discussion of TWFs [57–58]. To improve the mechanical
properties of TWMs, both structural manipulation (such as densification and lamination) and chemical modification have been
applied, and these practices are discussed in this section.
Densification of BWSs provides a straightforward strategy to adjust the mechanical properties of TWMs. Cellulose volume fractions
can be controlled by compressing the BWS before the fabrication of TW. With the increase in compression, the cellulose volume
fraction increases, leading to the enhanced tensile strength and modulus of TW. This densification approach was reported by Li et al.,
resulting in an increased Young’s modulus and tensile strength of TW from ~ 2.05 GPa and ~ 38 MPa to ~ 3.59 GPa and ~ 90.1 MPa,
respectively, as the cellulose volume fraction increased from 5 to 19 vol% (Fig. 5a) [88]. This might be related to the synergistic
enhancing effect between the hierarchically/longitudinally oriented cellulose microfibrils and PMMA polymer matrix. In addition, the

Fig. 6. (a) Infrared images of the NaClO-delignified TW (left) and NaClO2-delignified TW (right) indicating the temperature distribution. Repro­
duced with permission [66]. Copyright 2019, American Chemical Society. (b) Thermal conductivities of TW and conventional glass. (c) Infrared
image of the anisotropic heat transfer process in the TW. (d) Infrared images of TW with thermal energy distribution in the axial and radial di­
rections. Reproduced with permission [136]. Copyright 2020, Springer Nature. (e) The thermal conductivity of axial and radial TW heat directions
at different temperature. (b, c, e) Reproduced with permission [67]. Copyright 2019, Wiley-VCH.

12
S. Zhu et al. Progress in Materials Science 132 (2023) 101025

densification of BWSs can not only increase the density of the cellulosic reinforcements but also generate more hydrogen bonds be­
tween cellulose microfibrils of the densified wood cells, resulting in an enhanced mechanical strength [88,222]. A pure TWF was
designed by a fully densified BWS with highly aligned cellulose and a much higher hydrogen bonding density, which had 3-fold
Young’s modulus (50 GPa) and tensile strength (470 MPa) compared to isotropic cellulose nanopaper (Young’s modulus of 15 GPa
and tensile strength of 150 MPa) (Fig. 5b) [138]. Its specific strength was greater than that of most natural fibers, fossil-based plastics
and polymers (Fig. 5c). This positions the material at a prominent location in the Ashby chart, with excellent mechanical performance
but lightweight characteristics (Fig. 5c) [138]. A self-densified film was also reported, yielding a high Young’s modulus (51.1 GPa) and
tensile strength (449.1 MPa) [191]. The mechanical performances of TWFs can be tuned by adjusting the cutting angle (the angle
between the fiber direction and cutting direction) from NW. The lower cutting angle (ϴ = 0◦ ) achieved a TWF with higher cellulose
alignment and anisotropy resulting in a high tensile strength (350 MPa), whereas a higher cutting angle (ϴ = 45◦ ) produced a TWF
with a reduced strength (190 MPa) due to decreased cellulose alignment in the loading direction (Fig. 5d) [115]. The densification
strategy can prevent the density variability caused by different wood species, providing a convenient and feasible strategy for regu­
lating the mechanical properties of TWMs.
Lamination of transparent veneers was found to play a prominent role in overcoming anisotropy and producing an isotropic TWM.
The intrinsic anisotropic structure of TWMs leads to different mechanical properties in the longitudinal and transverse directions
(Fig. 5e). The tensile strength (45.4 MPa) of L-wood is two times as high as that of transverse TW (23.4 MPa, R-wood) [89]. Most
currently reported TWMs are either thin or anisotropic (Table 1) [67,129,139,220]. However, transparent plywood, with a cross-
laminated or quasi-isotropic structure, can be relevant in various emerging applications where isotropic mechanical properties are
crucial. Plies of transparent balsa veneers were laminated by controlling the wood fiber direction and thickness (>3.5 mm), resulting in
an isotropic TWM. The ultimate strength of transparent plywood increased from 15 MPa to 45 MPa under transverse loading (Fig. 5f)
[129]. As a consequence, the transparent plywood fabricated by the laminating method demonstrated isotropic mechanical properties
in both the longitudinal and transverse directions [220]. By applying this lamination technique, the mechanical properties of the TWM
are controllable and predictable based on the mechanical modeling of laminate plate theory.
Chemical modification can functionalize the surface of wood fibers, which can enhance interfacial bonding, thus improving the
mechanical performance of TWMs. For instance, Montanari et al. reported a fully bio-based TW prepared by the diffusion of limonene
acrylate (LIMA) monomer into the SA esterified cell wall [143]. After polymerization, the LIMA polymers were covalently linked to the
functionalized BWS, resulting in a fully bio-based TW with a high mechanical strength of 174 MPa and Young’s modulus of 17 GPa
(Fig. 5g). The chemical modification route is applicable to a variety of wood species, allowing tunable mechanical performance of the
TWM. Other chemical modification methods, e.g., acetylation and TEMPO-oxidation, can potentially be applied to tailor the wood
fiber/polymer interfacial adhesion, achieving mechanically robust TWMs (Table 1) [130,191].

3.2.3. Thermal properties


With the demand for energy-saving building materials in the building sector, thermal insulation properties become essential for
controlling heat exchange. The thermal conductivity is determined by the different wood compositions (the total volume fraction of the
BWS in TW) and anisotropic microstructure of TW. TW with ideal thermal insulation exhibits great potential for application as a
transparent and energy-efficient building material (Table 1).

Fig. 7. The actual performance comparison of between (a) glass and (b) TW. (a) Low thermal energy efficient image: Reproduced with permission
[15]. Copyright 2018, Wiley-VCH. Glaring and highly brittle images: Reproduced with permission [61]. Copyright 2016, Wiley-VCH. (b) Repro­
duced with permission [61]. Copyright 2016, Wiley-VCH.

13
S. Zhu et al. Progress in Materials Science 132 (2023) 101025

Fig. 8. Diagram of TW applied to transparent and smart buildings. (a) Letter A and yin-yang patterned TW. Reproduced with permission [124].
Copyright 2021, American Association for the Advancement of Science. (b) The photo of aesthetic TW, and its demonstration as an aesthetic rooftop
material. Reproduced with permission [136]. Copyright 2020, Springer Nature. (c) Schematic to demonstrate the TW used as clear wood windows,
and its digital images. Reproduced with permission [66]. Copyright 2019, American Chemical Society. (d) Temperature changes in model houses
with glass (right) and CsxWO3/TW (left) as windows after 10 min of simulated solar radiation. Reproduced with permission [128]. Copyright 2017,
Royal Society of Chemistry. (e) Temperature changes inside the box with infrared light irradiation. Insets are images showing the color changes of
ATO/TW with UV irradiation. Reproduced with permission [150]. Copyright 2019, Elsevier. (f) Schematic representation of thermal energy storage
and tunable transmittance by varying the temperature. Reproduced with permission [134]. Copyright 2019, American Chemical Society. (g) For­
mation mechanism illustration, SEM and EDS images of cationic silica-TW. Reproduced with permission [202]. Copyright 2022, Wiley-VCH. (h)
Photograph to illustrate the self-extinguishing property. Reproduced with permission [137]. Copyright 2022, Elsevier. (i) Illustration of a low heat
release value. Reproduced with permission [186]. Copyright 2022, Elsevier.

Thermal conductivity is affected by the volume fraction of the BWS in TW. The impact of different delignification processes (NaClO-
delignified vs NaClO2-delignified) on the thermal conductivity of TW was investigated by Jia et al. [66] NaClO2-delignified TW
possessed a slightly lower thermal conductivity (0.24 W m− 1 K− 1) than NaClO-delignified TW (0.35 W m− 1 K− 1) because NaClO2-
delignified TW had a significantly higher wood component volume (18.4 vol%) than NaClO-delignified TW (2.9 vol%) (Fig. 6a) [66].
The TW had a much lower thermal conductivity of 0.35 W m− 1 K− 1 along the fiber direction, which is only one-third of the thermal
conductivity of glass (~1 W m− 1 K− 1) (Fig. 6b). This is attributed to the complex organization of the hierarchical cell wall structure
(Fig. 3a), which can lead to high resistance to phonon transportation within the cell walls (mainly composed of cellulose and hemi­
cellulose) due to the increased phonon scattering at the numerous interfaces [61]. Therefore, TW with the desired thermal conductivity
can be achieved by controlling the wood composition and microstructure during the delignification process.
Thermal properties are also dominated by the anisotropic microstructure of TW. The thermal energy tends to transfer along TW
with an ellipse shaped distribution under heat radiation (Fig. 6c) [67]. This can be explained by the fact that heat preferentially
propagates along the well-preserved wood cell walls, which yields less phonon scattering along the wood channel direction than in the
vertical direction. Moreover, the irradiated phonons are blocked or confined to travel through the nanoscale pores in the cell walls
[61], resulting in a lower thermal conductivity in the radial direction than in the axial direction (Fig. 6d) [136]. When the heat source
is in the temperature range from 20 ◦ C to 60 ◦ C (Fig. 6e), TW shows a low thermal conductivity (~0.19 W m− 1 K− 1) in the radial
direction and a higher thermal conductivity (~0.4 W m− 1 K− 1) in the axial direction. However, both values are still much lower than
those of traditional glass [67]. In this context, structural manipulation by tuning compositions and/or pore structures provides a
possibility to govern the thermal properties of TWMs. However, the thermal properties of TWFs have not been investigated.

14
S. Zhu et al. Progress in Materials Science 132 (2023) 101025

3.3. Application

3.3.1. Transparent and smart building materials: TWM as an alternative to glass


Developing energy-efficient building technologies is expected reduce residential energy consumption (such as in cooling, heating,
lighting and electrical usage),targeting 50 % or more reduction [231]. Energy-efficient buildings require less energy consumption
while creating a comfortable environment for their occupants. In addition, energy-efficient buildings with low-cost illumination and
thermal insulation can reduce the carbon footprint while contributing to reducing global warming, and providing financial savings. To
this end, the actual performance and economic viability of energy-efficient buildings have gained great attention.
Glass is one of the most common high-transparency and low- haze building materials. However, it has some limitations when used
in buildings (Fig. 7a) [15,61]. (1) Glass has very low energy efficiency due to its high thermal conductivity (~1.0 W m− 1 K− 1), allowing
for the rapid exchange of heat energy between the outdoors and indoors. (2) Reflection effects and discomforting glare from glass may
occur during the daytime, impacting working efficiency and emotional well-being and sometimes leading to traffic accidents due to
‘sun strike’ [61]. (3) Glass has potential safety issues as it can be broken and shatter immediately with a sudden mechanical shock,
long-term creep deformation or earthquakes (Fig. 7a) [15]. (4) The production of glass results in 25 thousand tons of carbon dioxide
every year, contributing to climate change and global warming [15].
Interestingly, TWMs are considered promising candidates as energy-efficient building materials in the substitution of glass, mainly
for the following reasons (Fig. 7b). (1) Excellent thermal properties: TW possesses a lower thermal conductivity (0.19 ~ 0.39 W m− 1
K− 1) than glass, which can effectively reduce energy loss [67]. (2) Well-being in indoor environments: TW can create a comfortable
indoor lighting environment in the daytime. (3) Better security: TW has an improved mechanical toughness (0.56 ~ 6.1 MJ m− 3, glass
= 0.1 MJ m− 3), preventing sudden shattering [66,89,130]. (4) Environmentally friendly: TWMs are relatively ecofriendly because the
BWS derives from biodegradable, renewable and sustainable NW. More promisingly, impregnated resins are moving toward more
environmentally friendly polymers, such as biomass-derived PLIMA and biodegradable PVA. These merits are driving research and

Fig. 9. (a) Schematic of display photochromic TW windows on a building with solar irradiation. Reproduced with permission [132]. Copyright
2019, Royal Society of Chemistry. (b) Images of the electrochromic TW device. Reproduced with permission [63]. Copyright 2018, Wiley-VCH. (c)
Photograph of a luminescent TW stool. Reproduced with permission [151]. Copyright 2017, Wiley-VCH. (d) Illustration of a uniform illumination
using a luminescent TWF as a lighting panel/cover when exposed to a UV light source. Reproduced with permission [139]. Copyright 2020,
American Chemical Society. (e) Optical images and UV spectrum of reconstructed wood. (f) Schematic showing the photothermal conversion device.
(e-f) Reproduced with permission [224]. Copyright 2022, Elsevier.

15
S. Zhu et al. Progress in Materials Science 132 (2023) 101025

development of TWMs as energy-efficient building materials for various applications (Fig. 8).
TW is considered a promising transparent building material for aesthetic and optical design. Xia et al. demonstrated patternable
TW, showing various patterns, such as the letter A and a yin-yang symbol (Fig. 8a) [124]. By applying a solar-assisted chemical (H2O2)
brushing method, the desired patterns of NW turned white. After infiltrating epoxy into the white BWS, the patternable TW can be
precisely designed. In addition, aesthetic TW with natural annual ring features can be fabricated owing to the gradient distribution of
lignin between late and early NW (Fig. 8b). Further aesthetic patterns were achieved by laminating multiple layers of TW. These
aesthetic TW materials were demonstrated as a patterned ceiling in a museum or gallery, where natural light and aesthetic patterns can
be uniformly projected to the indoors (Fig. 8b), which may not happen from glass [136]. Optically designed TW with tunable haze can
be prepared by different delignification processes (NaClO vs NaClO2), as well as tailored interfacial adhesion [66–67,143,184]. Clear
TW with 10 % haze was made by a NaClO delignification process, enabling an appealing substitution of common glass in an energy-
efficient window (Fig. 8c). Translucent TW with approximately 80 % haze can be obtained by the NaClO2 delignification process,
allowing for lightening but with the protection of personal privacy [61,66]. These optical and aesthetic design options with TW
materials provide routes for replacing traditional glass in some engineering applications.
The structural design offers opportunities to achieve large TW as alternatives to glass, such as windows and buildings. TW has been
proposed as a building material that combines structural and functional performances [61]. As a proof-of-concept, a house frame was
demonstrated using a TW roof with more uniform illumination and better thermal insulation performance than a glass rooftop [61]. To
meet the requirements of building applications, TW materials with larger dimensions and higher thicknesses need to be developed. Fu
et al. reported load-bearing and luminescent TW (thickness = 3.5 mm) based on the lamination technique. Transparent plywood can be
designed by cross-stacking layers with an isotropic structure. This improved the mechanical properties in the weakest direction by
controlling fiber orientations and resulted in TW with tunable light scattering behavior [129]. A green steam-modification (H2O2/HAc
steam) method was developed to uniformly modify lignin in bulk wood, resulting in translucent wood with 2 ~ 5 cm thickness that can
be used as a building material [232]. More recently, Xia et al. proposed a scalable technique combining H2O2 lignin modification and
photocatalytic oxidation processes to create large TW (400 × 110 × 1 mm3). This simple, fast and scalable fabrication of TW com­
posites shows great potential for TW to be used as a next-generation building material [124,136]. As a demonstration, these large TW
materials have been showcased as a window or model building (Fig. 8c) that can efficiently harvest sunlight for uniform indoor
illumination, so they can potentially replace conventional glass. Near-infrared light (wavelength λ = 800 ~ 2500 nm)-absorbing
nanoparticles and phase-changed polymers have been encapsulated in TW, resulting in heat-shielding or thermal energy storage
composites that have potential applications in energy-efficient buildings. For example, CsxWO3 or antimony-doped tin oxide (ATO)
nanoparticles were embedded in TW composites, which were demonstrated for thermal-shielding buildings/windows (Fig. 8d)
[128,150]. Due to the near-infrared light being absorbed by the nanoparticles, a much smaller temperature increase was achieved from
these functional TWs compared to that of normal TW and glass [67,128,150]. Functional TW with ATO also demonstrated improved
antiaging performance and prolonged service life (Fig. 8e) [150]. Thermochromic heat-shielding is another important application for
TW materials. Tungsten-doped vanadium dioxide nanoparticles can reflect near-infrared light and reduce the phase transition tem­
perature to room temperature, allowing infrared radiation to be transmitted below room temperature and mostly reflected above room
temperature [205]. In another study, TW for thermal energy storage was designed by adding phase-change materials into a BWS
[134,233]. For example, a polyethylene glycol/PMMA polymer blend was impregnated into a porous BWS, enabling TW to store/
release thermal energy [234–235]. The thermal energy storage/release ability of TW can be adjusted by heating above the melting
temperature or cooling until the crystallization temperature of the polyethylene glycol polymers is reached [134], which cannot be

Fig. 10. Characteristic comparison of plastic and TWF. (a) Reproduced with permission [15]. Copyright 2018, Wiley-VCH. (b) Reproduced with
permission [223]. Copyright 2018, American Chemical Society.

16
S. Zhu et al. Progress in Materials Science 132 (2023) 101025

achieved with glass (Fig. 8f).


The impregnation of flame retardants into a BWS can improve the fire safety of TWMs in the applications of building materials. This
can be achieved by the incorporation of inorganic particles, phosphate- or nitrate-based retardants [137,186,199–200,202,236], such
as MXene nanosheets and silica nanoparticles, polyimide, MFs and phosphate ester-polyethylene glycol (PEAG) polymers. The wood
cell walls were enriched with carboxyl groups, enabling effective permeation of cationic silica nanoparticles and a uniform distribution
of inorganic nanoparticles inside the cell walls (Fig. 8g). An inorganic surface layer formed in the functional TWM, resulting in a
favorable fire retardancy, with limited oxygen permeability and little degradation of volatile gases [199,202]. A phosphate- or nitrate-
based polymer with a refractive index similar to that of the BWS, was impregnated and processed into fire-retardant TW. Due to the
carbonization of polyimide/MF or PEAG chains and the formation of a protective char layer in the combustion test, the resultant TW
exhibited an outstanding self-extinguishing property (Fig. 8h), low heat release (Fig. 8i) and flame retardancy [137,186,200]. These
TW materials also showed higher ignition and lower total smoke release values than NW, suggesting their great potential as a novel
type of fire-retardant transparent composite in building material applications.
Tunable UV-blocking property enables TWM as a smart window, protecting human skin, furniture and interior displays from hazard
and damage. Retaining lignin, adding organic absorber (e. g., 1,3,3-Trimethylindolino-6′ -nitrobenzopyrylospira., 2-(2H-Benzotriazol-
2-yl)-4, 6-di-tert-pentylphenol, CNF, chitosan, lignin) or inorganic UV absorber (e. g., ATO, γ–Fe2O3@YVO4:Eu3+, SrAl2O4:Eu2+,Dy3+)
can realize the UV protective performance of TWMs [127,150,193,209,224,236–240]. Organic UV absorbers can capture the UV
energy and convert it to other forms of energy, which is attributed to the chromophore and auxiliary groups connected with aromatic
groups, such as C– – N, N–N, –NH2, –OH, -SO3H and –COOH. For example, by spatially selective delignification of wood cells or the
addition of organic UV absorbers, TWMs displayed a low transmittance in the UVA (320 ~ 400 nm), UVB (275 ~ 320 nm) and UVC
(200 ~ 275 nm) regions [238]. Different from the light energy conversion, inorganic UV absorbers can connect with wood microfibrils,
followed by increasing the reflection and scattering of UV light on their surfaces, preventing UV rays from passing through TWMs to
achieve the UV protection effect. After embedding the inorganic UV absorbers, the UPF values (a parameter to estimate UV blocking) of
TWMs were 15 ~ 21 times than that of blank sample, indicating an excellent UV protection ability of TWMs [209,237].
Functionalized TWMs allow for smart applications, for instance, as photochromic, electrochromic, luminescent, and photothermal
conversion windows [63,126,132,140,146,151,224,237,241]. A photochromic smart window was prepared by infiltrating a BWS with
a photochromic material/MMA blend, showing photoswitchability between purple and transparency with illumination (Fig. 9a). Light
transmittance was controllable by the photochromic material-doped TW due to the reversible chemical structure of spiropyran/
merocyanie [132]. Inspired by this concept, Samanta et al. reported a double-stimulus-responsive chromic TW where the optical
performance was reversibly tailored by changing the temperature and UV radiation [206]. A smart electrochromic TW device was
assembled, and the transparent conducting electrode was prepared by coating poly(3,4-ethylenedioxythiophene):poly(styrene sul­
fonate) on TW. The color could be switched when 0.8 or − 0.5 V was applied to the electrochromic TW device (Fig. 9b) [63]. In a
luminescent lighting application, CdSe/ZnS quantum dots were incorporated into a TWM, resulting in a luminescent and light scat­
tering material that could be applied in lighting panels as a replacement for glass (Fig. 9c-d) [126,139,151]. The combination of
TEMPO-oxidized wood films and lignin enables excellent water resistance and UV-shielding performance (Fig. 9e) [224]. Due to the
high absorption in the near-infrared spectral region and low reflectance in the solar spectrum, this TWM displayed great potential in
photothermal conversion devices (Fig. 9f). Based on the above descriptions, these functional TWMs may have potential in applications
where glass is commonly used, such as windows, lighting covers/panels and energy-efficient buildings.

3.3.2. Optoelectronic devices: TWM as an alternative to plastics


There is a growing demand for optoelectronic devices, such as laptops, smart phones, televisions, portable electronics and displays
[242–249]. However, the substrates of these devices are typically made of nonbiodegradable and/or nonrenewable glass or plastics. In
particular, most plastics used for optoelectronic devices are fossil-based. They are frequently upgraded and discarded due to the rapid
development of advanced technologies, which have led to serious environmental contamination, for example, land, seabed and air
pollution (Fig. 10a) [15,250]. Moreover, these electronic wastes usually contain hazardous inks or semiconductors (e.g., gallium
arsenide or heavy metals) that can release toxic substances, causing major environmental concerns and human health problems [96].
This has motivated a search for excellent alternative substrates with high transparency, flexibility, printability and more environ­
mentally friendly features (Fig. 10b) [223,251]. The recent development of TWMs drives the production of fully bio-based and eco-
friendly materials, which can fulfill these requirements and have demonstrated great potential to gradually replace all-plastic-
based substrates in optoelectronic devices [89,125,223].
Due to their high transparency and low thermal expansion coefficient, TWMs have recently been explored as substrates for
replacing fossil-based plastics, enabling their application in flexible electronics [252–253]. A flexible TW substrate with high trans­
parency (80 % transmittance) was demonstrated by coating silver nanowire (AgNW) ink on a TW substrate using the Meyer method
(Fig. 11a). The as-prepared conductive substrate with a sheet resistance of 11 Ω sq-1 could illuminate a light-emitting diode (LED), even
when being bent [131]. Another study reported spray coating AgNWs on TW, achieving a much lower sheet resistance (3 Ω sq-1) and a
stable conductivity after 500 bending test cycles (Fig. 11b) [133]. By sandwiching a luminescent layer between two conductive
electrodes, a three-layer alternating current electroluminescent (ACEL) device was successfully fabricated with a luminance value of
18.36 cd m− 2 (voltage = 220 V, frequency = 1 kHz). More interestingly, the BWS can effectively limit the volume expansion of the
ACEL devices. Due to the addition of the BWS with a low coefficient of thermal expansion (0.1 × 10-6 K− 1), TW showed a lower
coefficient of thermal expansion (5 × 10-6 K− 1) than polyethylene terephthalate (PET) [(86.37 ± 3.62) × 10-6 K− 1] and epoxy [(76.89
± 3.53) × 10-6 K− 1], providing favorable dimensional stability of the TW substrates in ACEL devices. The flexible ACEL device also
demonstrated good thermal stability and waterproofing properties and could retain good luminescence properties at 100 ◦ C for 30 min,

17
S. Zhu et al. Progress in Materials Science 132 (2023) 101025

in water for 30 min, or even at 90 ◦ C and 90 % relative humidity for 30 min (Fig. 11c). The values were comparable to those of common
plastics, such as PET, nylon, polyethylene, polyvinyl chloride and epoxy [204]. Interestingly, various patterns can be fabricated,
indicating versatile material design concepts in high temperature-resistant and programmable ACEL devices (Fig. 11c) [204]. Based on
environmental concerns, a full cellulose-based TWF was demonstrated to be a biodegradable transistor. After placing the device in soil
and exposing it to nature for 6 days, the TWF electronic device disappeared from the soil, suggesting fast degradation of the device
[223]. Very recently, an environmentally friendly all-biobased TWM was developed by infiltrating a renewable and biodegradable CNF
or chitosan aqueous suspensions into a BWS. Due to the hydrogen bond interaction between CNF/chitosan and the BWS, a TWM with
high transmittance (80 %), adjustable haze (30 ~ 60 %), good thermal stability (315 ℃) and excellent UV shielding properties was
produced, holding great potential as a substrate for flexible electrodes [240].
TWMs have demonstrated applications in sensors due to their anisotropic structure and excellent flexibility [138,140,203]. For a
strain sensor, Fu et al. proposed the highly efficient utilization of all wood components combining a TWF and lignin-derived
carbonized fiber ink [138]. Through directly printing lignin-derived conducting ink on the TWF substrate, a fully wood-based flex­
ible film could be produced as a flexible strain sensor. The TWF electronic circuit showed a more stable electromechanical performance
than PET plastic, as the TWF circuit had less deformation/detachment from the substrate when the circuit was loaded. Moreover, these
circuits could activate LEDs when they were bent and folded, indicating favorable flexibility and conductivity (Fig. 11d left and
middle). The anisotropic structure of TWFs provides further opportunities in this regard. As a proof of concept, a hand-held TWF strain
sensor along the transverse direction (vertical with wood fiber orientation) showed variation in relative resistance with finger
movement (Fig. 11d right). Wang et al. reported novel stretchable and conductive TW by backfilling nontoxic polymerizable deep
eutectic solvents (PDES and acrylic-acid/choline chloride) into the pores of a BWS via in situ photopolymerization. Due to the hydrogen
bonding of PDES and the BWS, the obtained TW exhibited high transparency, stretchability, electrical conductivity and sensing ability
to external stimuli (e.g., strain and touch), which could be applied to monitor human activities and subtle pressures [203]. Based on
this work, Yang et al. further reported flexible TW using the same PDES, and they discovered the temperature sensing properties of the
TW. When the temperature increased or decreased, the conductivity of the flexible TW increased or decreased monotonously. This
favorable temperature-sensing feature makes TW a promising candidate as a temperature sensor [254]. Recently, superflexible TW
fabricated by impregnating biocompatible and flexible poly(ethylene glycol) diacrylate into a BWS was reported, serving as the TW-
based electrode for a pressure sensor after coating AgNWs. By sandwiching two TW-based electrodes and a stimuli-responsive layer
with a pyramidal microstructure, a flexible wood-based e-skin with capacitive and optical dual-response sensitivity was developed,
showing both naked eye- and instrument-readable features. This wood-based e-skin could make an electroluminescent response within
the range of pressures (0 ~ 150 kPa) that humans experience pain and maintain a high sensitivity (>0.15 kPa− 1) [144]. For the
application of TWMs in gas sensors, Liu et al. developed a smart construction material with timely and visible self-detection of
formaldehyde (FA) gas. It was fabricated by encapsulating a gas sensor consisting of a blend of multicolor carbon dots (CDs) and PVA.
The CDs/PVA TW demonstrated the ability to self-detect indoor FA by visual color-responsive ratiometric fluorescence and delayed
room temperature phosphorescence [140].
TWMs have the potential to replace plastic/glass substrates in solar cells or WLEDs because of their high light transmission and
scattering performance [125,135,140,169,255–256]. A light scattering TWM was used as a light management material for a GaAs solar
cell device with an enhanced total conversion efficiency of 14 ~ 18 % (Fig. 11e) [125,255]. This could be ascribed to two reasons: (1)
Light reflection was reduced using the light scattering TWM, resulting in more light propagating through the coating layer to the GaAs
solar cell. (2) The incident light was diffused due to the high haze (~80 %) of the light scattered TWM, extending the photon-traveling
path and improving light trapping performance. Similarly, an assembled TW perovskite solar cell showed a high power conversion
efficiency of 16.8 % and favorable long-term stability [135]. Beyond solar cells, the large light scattering effect also allows TWMs to be
candidates for the encapsulation material of WLEDs [140,169,198,201]. A green WLED TW was prepared by mixing citric acid-derived
multicolor CDs with acrylic acid, followed by incorporation into a BWS [169]. The TW-based encapsulation material, together with a
blue UV-chip, was assembled into a promising WLED device, which displayed excellent color characteristics with a Commission on
Illumination value of (0.33, 0.32) and good luminous stability for 7 days (Fig. 11f). Apart from citric acid-based CDs, lignin-derived
CDs doped with various nitrogen-containing compounds could also be brought into wood cells as trichromatic systems for TW-based
encapsulating materials [140,201]. By infiltrating the multicolor-emission lignin-based CDs and PVA or UV curing resin into a
delignified wood frame, outstanding encapsulation films of WLEDs were achieved and showed uniform white light under a UV lamp
[140,201], as well as resistance to water, acid, alkali and salt, with a stable light-emitting performance [198]. Thus, TWMs demon­
strate great potential to replace plastic/glass substrates for clean energy conversion/storage devices.
Based on the discussion in Section 3.2, TWMs benefit from their anisotropic structure, showing tunable optical, mechanical and
thermal properties. As a result, they are regarded as alternatives to glass and plastics in a wide range of applications, including energy-
efficient buildings, smart windows, flexible optoelectronics, sensors, energy conversion and storage devices. Structural design concepts
allow for the manipulation of the microstructure and chemical compositions of BWSs, as well as for laminated and densified structures.
The functional design concepts enable the achievement of additional functionalities, such as conductivity, thermal insulation, thermal
storage, luminescence, photon response and electroresponse. Combining the functional and structural performances of TWMs endows
them with great potential as an engineered replacement for glass and plastics and in other advanced applications (e.g., automobile and
aircraft materials).

4. Process–structure–property–application relationships

TWMs integrate load-bearing properties with photonics functions and light weight. The cellulose microfibril skeleton in the cell

18
S. Zhu et al. Progress in Materials Science 132 (2023) 101025

wall is a sophisticated reinforcement structure, and chemical/physical treatments can be used to tailor the nanostructured BWS. In
addition, functionalization strategies enable the addition of new functions/properties to TWMs for specific applications. To this end,
fundamentally understanding the process–structure–property–application relationships of TWMs is essential for engineering purposes.
(1) Process–structure relationship: Currently, wood nanotechnologies including various chemical and physical treatments are
applied to tailor the composition and microstructure, resulting in a fully/partly delignified BWS with a wide range of Bru­
nauer–Emmett–Teller specific surface areas (1 ~ 260 m2/g, Table 1). The influence of these treatments on the chemical reactions and
components of different wood cell walls (e.g., balsa, pine, poplar, ash, beech and basswood) is not fully understood. For example, the
extraction of the lignin component is complicated in bulk wood. Moreover, uniform delignification of a global wood structure,
particularly in the thickness direction, is always a challenge because of the limitation of chemical penetration into the cell wall. Thus,
efficient and controllable delignification processes are being developed. Furthermore, greener chemical treatment of wood with a high
porosity should be considered for future investigation, as a highly porous material provides many interactive sites for the function­
alization step. Low carbon emission processes could be adapted to a continuous roll-to-roll production system with aiming at large-
scale production [138]. Extending these wood nanotechnologies to other natural plants for designing new materials, such as
bamboo, reed and loofah templates, is also of great interest [257–266].
(2) Structure–property relationship: A basic understanding of the impact of the structure–property relationship is crucial for the
material design concept and adding new functionalities to TWMs. There is a lack of knowledge of the underlying structure–property
relationships of nanostructured wood and how to manipulate these specific nanoeffect properties. Wood nanoscience has developed
from multidisciplinary theories, such as the nanoconfinement effect, supermolecular topology, permeability theory and optic physics,
which are applied to answer the scientific questions of TWMs. For example, how do photons, electrons and ions travel/transport within
the nanostructure of TWMs? From a technological point of view, the high performance of TWMs (optical, thermal and mechanical
properties) can be achieved by adapting the volume fraction, thickness, lamination and chemical modification/functionalization of
BWSs. Specifically, by tailoring the wood thickness and volume fraction and applying lamination and interfacial design strategies,
TWMs can achieve suitable optical properties. To achieve the expected mechanical properties of TWMs, structural manipulation by
densification and lamination can be applied, as well as chemical modification. For thermal properties, changing the volume fraction of
the BWS or anisotropic microstructure can be adopted. Moreover, advanced characterization techniques (e.g., computational simu­
lation, X-ray synchrotron and micro computed tomography, and ultrahigh transmission electron microscopy), are required for the
development of wood nanoscience and nanotechnologies, and to obtain a fundamental understanding of the physics of TWMs. Further
investigations of the chemical interactions will help us to control their microstructure for tunable properties.
(3) Property–application relationship: Our review emphasizes the mechanical, optical and thermal properties, as well as relative
emerging applications (electronics, photonics, sensors and energy storage devices). Although many chemical modification and
functionalization strategies have been proposed to achieve multifunctional TWMs and have been demonstrated in various applications,
the development of TWMs is in an early stage. TW replies on the impregnation of refractive-index-matched polymers with additional
functions. An appealing strategy is to functionalize the BWS prior to conversion into TWMs, which may realize new opportunities for
making multiple functional TWMs possible. In addition, most applications of TWMs are at the proof-of-concept stage, with a
considerable amount of work required to achieve commercialization and large-scale production. The impact on our society may not be
realized immediately; thus, future research should pay attention to realistic demands and the connections between functionalities and
real applications in our lives.

5. Summary and future outlook

TWMs are an exciting option for next-generation wood-based materials. The top-down method is a powerful strategy to convert
bulk wood into functional and sustainable materials, making a large-scale production possible. Wood nanotechnology has been
developed for engineering the production of TWMs with combined structural and functional performances. This review has over­
viewed the state-of-the-art research covering the transparency mechanism, processing methods, structure, property, application and
their relationships.
One may ask the justification behind converting wood into a transparent material by the top-down method, specifically, instead of
the already-matured nanocellulose-based transparent materials prepared via the bottom-up approach. Although the formation of
nanocellulose-based materials provides greater control of the material structure, it requires multiple energy-intensive and time-
consuming steps, starting with nanocellulose deconstruction from the plant cell wall and ending with the reconstruction of new
material structures via bottom-up assembly. Note that the nanocellulose deconstruction process may consume ~ 500 kWh ton-1 to 30
000 kWh ton-1 energy depending on the native or modified state of the cell walls (e.g., fibers), leading to an estimated cost of over $50
~ 100 kg− 1 [267]. Furthermore, the inherent anisotropic hierarchical structure of wood is damaged, which requires the reconstruction
of small building blocks to achieve transparent materials.
The justification for making wood transparent lies in the fact that TWMs incorporate all the advantages of nanocelluloses, such as
their light weight, excellent mechanical and light transmission properties, and high thermal stability, but energy-intensive and time-
consuming deconstruction and reconstruction processes are not needed. The excellent inherent anisotropic characteristics of wood
have led to unique anisotropic optical, mechanical and thermal functionalities. The ease of functionalization of the BWS via chemical/
physical modification and nanoparticle impregnation has led to the fabrication of TWMs with additional features, such as high optical
transmittance, adjustable haze, mechanical flexibility, luminescence, photo- and electrochromic performance, heat and UV resistance,
flame retardancy, and gas sensitivity.
However, there are still some challenges to tackle in TWM research. One of the major challenges, for example, is to produce a

19
S. Zhu et al. Progress in Materials Science 132 (2023) 101025

Fig. 11. Diagram of TWM applied to flexible optoelectronic devices. (a) Schematic to demonstrate a LED lighted up by a flexible and conductive TW
before and after bending, respectively. Reproduced with permission [131]. Copyright 2018, Royal Society of Chemistry. (b) A flexible and
conductive TW with an excellent conductivity stability even after 500 bending tests. Reproduced with permission [133]. Copyright 2019, American
Chemical Society. (c) Schematic to diagram the ACEL device and its excellent luminescence stability and different illumination modes. Reproduced
with permission [204]. Copyright 2019, American Chemical Society. (d) Photographs of the conductive TWF and its application for strain sensors.
Reproduced with permission [138]. Copyright 2020, American Chemical Society. (e) Application of TW/GaAs solar cell. Reproduced with
permission [125]. Copyright 2016, Elsevier. (f) Luminous stability of the TW/CDs WLEDs in 7 days. Inset is a photograph of a lighting WLED.
Reproduced with permission [169]. Copyright 2018, American Chemical Society.

sufficiently large and thick TWM for commercial implementation. This is partly because the integrity of the BWS with uniform cell wall
penetration is difficult to maintain after delignification treatment. Lignin is known as the gluing material among the fibers. The
removal of lignin essentially makes the BWS delicate to handle for further processing. Although lignin-retaining/modified methods
have already been developed, more investigation is needed to optimize lignin modification or to develop new efficient delignification
methods that would not sacrifice the optical properties and thickness of the TWM.
Another challenge is to obtain optically clear wood with a haze below 10 %. Most of the currently reported TWMs have a haze value
of 40 ~ 80 % depending on their thicknesses. Recently, optically clear wood has been reported to have a high transmittance of 90 %

20
S. Zhu et al. Progress in Materials Science 132 (2023) 101025

and a low haze of 10 %, but the cellulose fraction was only 2.54 vol%. This led to a negligible improvement in the strength properties
compared to those of the constituents. On the other hand, the haze value of a nanocellulose-based transparent material, despite having
similar constituents with a nanocellulose content as high as ~ 70 %, can easily reach below 10 ~ 20 %. The elucidation of the un­
derlying reasons for this phenomenon may eventually lead to a strong, highly optically clear wood for next-generation applications.
To date, little attention has been focused on the trade-off of the stability and biodegradability of TWMs through their life cycle, as
well as the production cost. The optical and mechanical properties of TWMs may be slightly impaired when they are exposed to water,
fire or UV irradiation, owing to the biodegradable and hydrophilic characteristics of the wood components. Although there have been
reports on functionalization, such as UV absorption and hydrophobic modification, how to achieve a stable yet biodegradable TWM
with a certain stimulus response is a promising direction for future research. Future research should aim at lowering the production
cost of TWMs with bio-based polymers and environmentally friendly processes, as well as developing new processes suitable for
manipulating TWM structures with controlled micro- and nanostructures.
Overall, wood nanoscience and nanotechnology are new and developing topics in the material science field. A better understanding
of the science (the processes–structures–properties–applications relationships) of TWMs is essential for both academic and industrial
communities to drive their progress. A combination of emerging technologies (e.g., machine learning, artificial intelligence, 3D
printing and big data) will enable commercialization and support our sustainable modern society in the future. This will benefit to the
sustainable development of structural and functional TWMs in the extension of carbon storage and replacement of nondegradable
plastics. It is also crucial to address the fundamental science and physics of TWMs that have inspired and been integrated with other
disciplines, such as wood science, material chemistry and physics. All future challenges aside, derived from the abundant advantages
reported thus far, the global overview of the TWMs in the current review will serve as an indicator and establish new insights for
developing the topic. We foresee that TWMs have an enormous potential to replace glass and fossil-based plastics, expanding their
future application in sustainable smart buildings, optoelectronics, clean energy conversions, sensors, e-skins, soft robotics and other
emerging fields.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Data availability

Data will be made available on request.

Acknowledgments

S. L. Zhu and S. K. Biswas contributed equally to this work. We acknowledge the National Natural Science Foundation of China
(31901274), Canada Research Chairs Program (231928), 13th China Special Postdoctoral Science Foundation (2020T130303), China
Postdoctoral Science Foundation (2019M661854), Postdoctoral Science Foundation of Jiangsu Province (2019K142), Qing Lan Project
of Jiangsu Province (2019), 333 Project Foundation of Jiangsu Province (BRA2018337), the Royal Society of New Zealand Catalyst
Seeding grant (CSG-FRI1901), the funding from the New Zealand Ministry of Business, Innovation and Employment (MBIE) in the
framework of the Strategic Science Investment Fund (C04X1703, Scion Platforms Plan) for part of financial support. We are grateful to
Dr. A. Dickson (Scion, New Zealand) and Dr. T. Singh (Scion, New Zealand) for their comments and suggestions.

References

[1] Yu KL, Fan TX, Lou S, Zhang D. Biomimetic optical materials: integration of nature’s design for manipulation of light. Prog Mater Sci 2013;58(6):825–73.
https://doi.org/10.1016/j.pmatsci.2013.03.003.
[2] Mao Y, Hu L, Ren ZJ. Engineered wood for a sustainable future. Matter 2022;5(5):1326–9. https://doi.org/10.1016/j.matt.2022.04.013.
[3] Someya T, Bao Z, Malliaras GG. The rise of plastic bioelectronics. Nature 2016;540(7633):379–85. https://doi.org/10.1038/nature21004.
[4] Burgert I, Cabane E, Zollfrank C, Berglund L. Bio-inspired functional wood-based materials-hybrids and replicates. Int Mater Rev 2016;60(8):431–50. https://
doi.org/10.1179/1743280415y.0000000009.
[5] Toumpanaki E, Shah DU, Eichhorn SJ. Beyond what meets the eye: imaging and imagining wood mechanical-structural properties. Adv Mater 2021;33(28):
2001613. https://doi.org/10.1002/adma.202001613.
[6] Hill CA. Wood modification: chemical, thermal and other processes. West sussex, England: John Wiley & Sons; 2007.
[7] Rowell RM. Handbook of wood chemistry and wood composites. Boca Raton, Florida: CRC Press; 2005. 10.1201/9780203492437.
[8] Liu C, Luan PC, Li Q, Cheng Z, Xiang PY, Liu DT, et al. Biopolymers derived from trees as sustainable multifunctional materials: a review. Adv Mater 2020;
2001654. https://doi.org/10.1002/adma.202001654.
[9] Ajdary R, Tardy BL, Mattos BD, Bai L, Rojas OJ. Plant nanomaterials and inspiration from nature: water interactions and hierarchically structured hydrogels.
Adv Mater 2020:2001085. https://doi.org/10.1002/adma.202001085.
[10] Huang JL, Zhao B, Liu T, Mou J, Jiang ZJ, Liu J, et al. Wood-derived materials for advanced electrochemical energy storage devices. Adv Funct Mater 2019;29
(31):1902255. https://doi.org/10.1002/adfm.201902255.
[11] Cabane E, Keplinger T, Merk V, Hass P, Burgert I. Renewable and functional wood materials by grafting polymerization within cell walls. ChemSusChem 2014;
7(4):1020–5. https://doi.org/10.1002/cssc.201301107.
[12] Berglund LA, Burgert I. Bioinspired wood nanotechnology for functional materials. Adv Mater 2018;30(19):1704285. https://doi.org/10.1002/
adma.201704285.
[13] Mredha MTI, Jeon I. Biomimetic anisotropic hydrogels: advanced fabrication strategies, extraordinary functionalities, and broad applications. Prog Mater Sci
2022;124:100870. https://doi.org/10.1016/j.pmatsci.2021.100870.

21
S. Zhu et al. Progress in Materials Science 132 (2023) 101025

[14] Su Z, Yang Y, Huang Q, Chen R, Ge W, Fang Z, et al. Designed biomass materials for “green” electronics: a review of materials, fabrications, devices, and
perspectives. Prog Mater Sci 2022;125:100917. https://doi.org/10.1016/j.pmatsci.2021.100917.
[15] Jiang F, Li T, Li YJ, Zhang Y, Gong A, Dai JQ, et al. Wood-based nanotechnologies toward sustainability. Adv Mater 2018;30(1):e1703453. https://doi.org/
10.1002/adma.201703453.
[16] Zhu ZD, Xiao GF, Chen JQ, Fu SY. Wood nanotechnology: a more promising solution toward energy issues: a mini-review. Cellulose 2020;27(15):8513–26.
https://doi.org/10.1007/s10570-020-03404-2.
[17] Montanari C, Olsén P, Berglund LA. Sustainable wood nanotechnologies for wood composites processed by in-situ polymerization. Front Chem 2021;9(483):
682883. https://doi.org/10.3389/fchem.2021.682883.
[18] Chen FJ, Gong AS, Zhu MW, Chen G, Lacey SD, Jiang F, et al. Mesoporous, three-dimensional wood membrane decorated with nanoparticles for highly
efficient water treatment. ACS Nano 2017;11(4):4275–82. https://doi.org/10.1021/acsnano.7b01350.
[19] Fu QL, Ansari F, Zhou Q, Berglund LA. Wood nanotechnology for strong, mesoporous, and hydrophobic biocomposites for selective separation of oil/water
mixtures. ACS Nano 2018;12(3):2222–30. https://doi.org/10.1021/acsnano.8b00005.
[20] Guan H, Cheng ZY, Wang XQ. Highly compressible wood sponges with a spring-like lamellar structure as effective and reusable oil absorbents. ACS Nano 2018;
12(10):10365–73. https://doi.org/10.1021/acsnano.8b05763.
[21] Kim S, Kim K, Jun G, Hwang W. Wood-nanotechnology-based membrane for the efficient purification of oil-in-water emulsions. ACS Nano 2020;14(12):
17233–40. https://doi.org/10.1021/acsnano.0c07206.
[22] Cai CY, Wei ZC, Huang YZ, Fu Y. Wood-inspired superelastic mxene aerogels with superior photothermal conversion and durable superhydrophobicity for
clean-up of super-viscous crude oil. Chem Eng J 2021;421:127772. https://doi.org/10.1016/j.cej.2020.127772.
[23] Chen X, Zhu XB, He SM, Hu LB, Ren ZJ. Advanced nanowood materials for the water–energy nexus. Adv Mater 2021;33(28):2001240. https://doi.org/
10.1002/adma.202001240.
[24] Chen CJ, Li YJ, Song JW, Yang Z, Kuang YD, Hitz E, et al. Highly flexible and efficient solar steam generation device. Adv Mater 2017;29(30):1701756.
https://doi.org/10.1002/adma.201701756.
[25] Li T, Liu H, Zhao XP, Chen G, Dai JQ, Pastel G, et al. Scalable and highly efficient mesoporous wood-based solar steam generation device: localized heat, rapid
water transport. Adv Funct Mater 2018;28(16):1707134. https://doi.org/10.1002/adfm.201707134.
[26] Liu H, Chen CJ, Chen G, Kuang YD, Zhao XP, Song JW, et al. High-performance solar steam device with layered channels: artificial tree with a reversed design.
Adv Energy Mater 2018;8(8):1701616. https://doi.org/10.1002/aenm.201701616.
[27] Sun JG, Guo HY, Ribera J, Wu CS, Tu KK, Binelli M, et al. Sustainable and biodegradable wood sponge piezoelectric nanogenerator for sensing and energy
harvesting applications. ACS Nano 2020;14(11):14665–74. https://doi.org/10.1021/acsnano.0c05493.
[28] Sun JG, Guo HZ, Schadli GN, Tu KK, Schar S, Schwarze F, et al. Enhanced mechanical energy conversion with selectively decayed wood. Sci Adv 2021;7(11):
eabd9138. https://doi.org/10.1126/sciadv.abd9138.
[29] Chen CJ, Zhang Y, Li YJ, Dai JQ, Song JW, Yao YG, et al. All-wood, low tortuosity, aqueous, biodegradable supercapacitors with ultra-high capacitance. Energ
Environ Sci 2017;10(2):538–45. https://doi.org/10.1039/C6EE03716J.
[30] Xu WW, Liu CZ, Ren SX, Lee D, Gwon J, Flake JC, et al. A cellulose nanofiber-polyacrylamide hydrogel based on a co-electrolyte system for solid-state zinc ion
batteries to operate at extremely cold temperatures. J Mater Chem A 2021;9(45):25651–62. https://doi.org/10.1039/D1TA08023G.
[31] Wang J, Xu Z, Eloi J-C, Titirici M-M, Eichhorn SJ. Ice-templated, sustainable carbon aerogels with hierarchically tailored channels for sodium- and potassium-
ion batteries. Adv Funct Mater 2022;32(16):2110862. https://doi.org/10.1002/adfm.202110862.
[32] Zhang Y, Luo W, Wang CW, Li YJ, Chen CJ, Song JW, et al. High-capacity, low-tortuosity, and channel-guided lithium metal anode. P Natl A Sci 2017;114(14):
3584–9. https://doi.org/10.1073/pnas.1618871114.
[33] Song HY, Xu SM, Li YJ, Dai JQ, Gong A, Zhu MW, et al. Hierarchically porous, ultrathick, “breathable” wood-derived cathode for lithium-oxygen batteries. Adv
Energy Mater 2018;8(4):1701203. https://doi.org/10.1002/aenm.201701203.
[34] Chen CJ, Xu SM, Kuang YD, Gan WT, Song JW, Chen GG, et al. Nature-inspired tri-pathway design enabling high-performance flexible Li-O2 batteries. Adv
Energy Mater 2019;9(9):1802964. https://doi.org/10.1002/aenm.201802964.
[35] Wang F, Cheong JY, He Q, Duan G, He SJ, Zhang L, et al. Phosphorus-doped thick carbon electrode for high-energy density and long-life supercapacitors. Chem
Eng J 2021;414:128767. https://doi.org/10.1016/j.cej.2021.128767.
[36] Wang F, Cheong JY, Lee J, Ahn J, Duan G, Chen HL, et al. Pyrolysis of enzymolysis-treated wood: hierarchically assembled porous carbon electrode for
advanced energy storage devices. Adv Funct Mater 2021;31(31):2101077. https://doi.org/10.1002/adfm.202101077.
[37] Wang F, Liu XL, Duan GG, Yang H, Cheong JY, Lee J, et al. Wood-derived, conductivity and hierarchical pore integrated thick electrode enabling high areal/
volumetric energy density for hybrid capacitors. Small 2021;17(35):2102532. https://doi.org/10.1002/smll.202102532.
[38] Wang F, Zhang L, Zhang Q, Yang JJ, Duan GG, Xu WH, et al. Electrode thickness design toward bulk energy storage devices with high areal/volumetric energy
density. Appl Energ 2021;289:116734. https://doi.org/10.1016/j.apenergy.2021.116734.
[39] Huang J, Zhu HL, Chen YC, Preston C, Rohrbach K, Cumings J, et al. Highly transparent and flexible nanopaper transistors. ACS Nano 2013;7(3):2106–13.
https://doi.org/10.1021/nn304407r.
[40] Guan H, Meng JW, Cheng ZY, Wang XQ. Processing natural wood into a high-performance flexible pressure sensor. ACS Appl Mater Inter 2020;12(41):
46357–65. https://doi.org/10.1021/acsami.0c12561.
[41] Guan H, Dai XJ, Ni L, Hu JH, Wang XQ. Highly elastic and fatigue-resistant graphene-wrapped lamellar wood sponges for high-performance piezoresistive
sensors. ACS Sustain Chem Eng 2021;9(45):15267–77. https://doi.org/10.1021/acssuschemeng.1c05401.
[42] Wang ZX, Han XS, Zhou ZJ, Meng WY, Han XW, Wang SJ, et al. Lightweight and elastic wood-derived composites for pressure sensing and electromagnetic
interference shielding. Compos Sci Technol 2021;213:108931. https://doi.org/10.1016/j.compscitech.2021.108931.
[43] Huang W, Li HQ, Zheng LZ, Lai XJ, Guan H, Wei Y, et al. Superhydrophobic and high-performance wood-based piezoresistive pressure sensors for detecting
human motions. Chem Eng J 2021;426:130837. https://doi.org/10.1016/j.cej.2021.130837.
[44] Chen CJ, Song JW, Zhu SZ, Li YJ, Kuang YD, Wan JY, et al. Scalable and sustainable approach toward highly compressible, anisotropic, lamellar carbon
sponge. Chem 2018;4(3):544–54. https://doi.org/10.1016/j.chempr.2017.12.028.
[45] Nie KC, Wang ZS, Tang RX, Zheng L, Li CC, Shen XP, et al. Anisotropic, flexible wood hydrogels and wrinkled, electrodeposited film electrodes for highly
sensitive, wide-range pressure sensing. ACS Appl Mater Inter 2020;12(38):43024–31. https://doi.org/10.1021/acsami.0c13962.
[46] Fujisaki Y, Koga H, Nakajima Y, Nakata M, Tsuji H, Yamamoto T, et al. Transparent nanopaper-based flexible organic thin-film transistor array. Adv Funct
Mater 2014;24(12):1657–63. https://doi.org/10.1002/adfm.201303024.
[47] Koga H, Nogi M, Komoda N, Nge TT, Sugahara T, Suganuma K. Uniformly connected conductive networks on cellulose nanofiber paper for transparent paper
electronics. NPG Asia Mater 2014;6(3):e93. https://doi.org/10.1038/am.2014.9.
[48] Zhu HL, Fang ZQ, Wang Z, Dai JQ, Yao YG, Shen F, et al. Extreme light management in mesoporous wood cellulose paper for optoelectronics. ACS Nano 2016;
10(1):1369–77. https://doi.org/10.1021/acsnano.5b06781.
[49] Kong WQ, Wang CW, Jia C, Kuang YD, Pastel G, Chen CJ, et al. Muscle-inspired highly anisotropic, strong, ion-conductive hydrogels. Adv Mater 2018;30(39):
1801934. https://doi.org/10.1002/adma.201801934.
[50] Li T, Li SX, Kong WQ, Chen CJ, Hitz E, Jia C, et al. A nanofluidic ion regulation membrane with aligned cellulose nanofibers. Sci Adv 2019;5(2):eaau4238.
https://doi.org/10.1126/sciadv.aau4238.
[51] Li T, Zhang X, Lacey SD, Mi R, Zhao X, Jiang F, et al. Cellulose ionic conductors with high differential thermal voltage for low-grade heat harvesting. Nat Mater
2019;18(6):608–13. https://doi.org/10.1038/s41563-019-0315-6.
[52] Chen CC, Wang YR, Wu QQ, Wan ZM, Li DG, Jin YC. Highly strong and flexible composite hydrogel reinforced by aligned wood cellulose skeleton via alkali
treatment for muscle-like sensors. Chem Eng J 2020;400:125876. https://doi.org/10.1016/j.cej.2020.125876.

22
S. Zhu et al. Progress in Materials Science 132 (2023) 101025

[53] Chen CC, Wang YR, Zhou T, Wan ZM, Yang QL, Xu ZY, et al. Toward strong and tough wood-based hydrogels for sensors. Biomacromolecules 2021;22(12):
5204–13. https://doi.org/10.1021/acs.biomac.1c01141.
[54] Chen GG, Li T, Chen CJ, Kong WQ, Jiao ML, Jiang B, et al. Scalable wood hydrogel membrane with nanoscale channels. ACS Nano 2021;15(7):11244–52.
https://doi.org/10.1021/acsnano.0c10117.
[55] Fu QL, Medina L, Li YY, Carosio F, Hajian A, Berglund LA. Nanostructured wood hybrids for fire-retardancy prepared by clay impregnation into the cell wall.
ACS Appl Mater Inter 2017;9(41):36154–63. https://doi.org/10.1021/acsami.7b10008.
[56] Zhu HL, Fang ZQ, Preston C, Li YY, Hu LB. Transparent paper: fabrications, properties, and device applications. Energ Environ Sci 2014;7(1):269–87. https://
doi.org/10.1039/C3EE43024C.
[57] Li YY, Fu QL, Yang X, Berglund L. Transparent wood for functional and structural applications. Philos T R Soc A 2018;376(2112):20170182. https://doi.org/
10.1098/rsta.2017.0182.
[58] Li Y, Vasileva E, Sychugov I, Popov S, Berglund L. Optically transparent wood: recent progress, opportunities, and challenges. Adv Optical Mater 2018;6(14):
1800059. https://doi.org/10.1002/adom.201800059.
[59] Fang ZQ, Hou GY, Chen CJ, Hu LB. Nanocellulose-based films and their emerging applications. Curr Opin Solid St M 2019;23(4):100764. https://doi.org/
10.1016/j.cossms.2019.07.003.
[60] Okahisa Y, Yoshida A, Miyaguchi S, Yano H. Optically transparent wood–cellulose nanocomposite as a base substrate for flexible organic light-emitting diode
displays. Compos Sci Technol 2009;69(11–12):1958–61. https://doi.org/10.1016/j.compscitech.2009.04.017.
[61] Li T, Zhu MW, Yang Z, Song JW, Dai JQ, Yao YG, et al. Wood composite as an energy efficient building material: guided sunlight transmittance and effective
thermal insulation. Adv Energy Mater 2016;6(22):1601122. https://doi.org/10.1002/aenm.201601122.
[62] Biswas SK, Sano H, Shams MI, Yano H. Three-dimensional-moldable nanofiber-reinforced transparent composites with a hierarchically self-assembled
“reverse” nacre-like architecture. ACS Appl Mater Interfaces 2017;9(35):30177–84. https://doi.org/10.1021/acsami.7b09390.
[63] Lang AW, Li YY, De Keersmaecker M, Shen DE, Osterholm AM, Berglund L, et al. Transparent wood smart windows: polymer electrochromic devices based on
poly(3,4-ethylenedioxythiophene):poly(styrene sulfonate) electrodes. ChemSusChem 2018;11(5):854–63. https://doi.org/10.1002/cssc.201702026.
[64] Biswas SK, Sano H, Yang X, Tanpichai S, Shams MI, Yano H. Highly thermal-resilient AgNW transparent electrode and optical device on thermomechanically
superstable cellulose nanorod-reinforced nanocomposites. Adv Optical Mater 2019;7(15):1900532. https://doi.org/10.1002/adom.201900532.
[65] Biswas SK, Tanpichai S, Witayakran S, Yang X, Shams MI, Yano H. Thermally superstable cellulosic-nanorod-reinforced transparent substrates featuring
microscale surface patterns. ACS Nano 2019;13(2):2015–23. https://doi.org/10.1021/acsnano.8b08477.
[66] Jia C, Chen CJ, Mi RY, Li T, Dai JQ, Yang Z, et al. Clear wood toward high-performance building materials. ACS Nano 2019;13(9):9993–10001. https://doi.
org/10.1021/acsnano.9b00089.
[67] Mi RY, Li T, Dalgo D, Chen CJ, Kuang YD, He SM, et al. A clear, strong, and thermally insulated transparent wood for energy efficient windows. Adv Funct
Mater 2019;30(1):1907511. https://doi.org/10.1002/adfm.201907511.
[68] Kumar A, Jyske T, Petrič M. Delignified wood from understanding the hierarchically aligned cellulosic structures to creating novel functional materials: a
review. Adv Sustainable Syst 2021;5(5):2000251. https://doi.org/10.1002/adsu.202000251.
[69] Eichhorn SJ, Baillie CA, Zafeiropoulos N, Mwaikambo LY, Ansell MP, Dufresne A, et al. Review: current international research into cellulosic fibres and
composites. J Mater Sci 2001;36(9):2107–31. https://doi.org/10.1023/A:1017512029696.
[70] Moon RJ, Martini A, Nairn J, Simonsen J, Youngblood J. Cellulose nanomaterials review: structure, properties and nanocomposites. Chem Soc Rev 2011;40(7):
3941–94. https://doi.org/10.1039/C0CS00108B.
[71] De France K, Zeng ZH, Wu TT, Nyström G. Nanocellulose: functional materials from nanocellulose: utilizing structure-property relationships in bottom-up
fabrication. Adv Mater 2021;33(28):2170216. https://doi.org/10.1002/adma.202170216.
[72] Li T, Chen CJ, Brozena AH, Zhu JY, Xu L, Driemeier C, et al. Developing fibrillated cellulose as a sustainable technological material. Nature 2021;590(7844):
47–56. https://doi.org/10.1038/s41586-020-03167-7.
[73] Eichhorn SJ. Cellulose nanowhiskers: promising materials for advanced applications. Soft Matter 2011;7(2):303–15. https://doi.org/10.1039/C0SM00142B.
[74] Li M-C, Wu Q, Moon RJ, Hubbe MA, Bortner MJ. Rheological aspects of cellulose nanomaterials: governing factors and emerging applications. Adv Mater
2021;33(21):2006052. https://doi.org/10.1002/adma.202006052.
[75] Liu CZ, Li ZL, Li M-C, Chen WM, Xu WW, Hong S, et al. Lignin-containing cellulose nanofibers made with microwave-aid green solvent treatment for magnetic
fluid stabilization. Carbohyd Polym 2022;291:119573. https://doi.org/10.1016/j.carbpol.2022.119573.
[76] Ding QQ, Xu XW, Yue YY, Mei CT, Huang CB, Jiang SH, et al. Nanocellulose-mediated electroconductive self-healing hydrogels with high strength, plasticity,
viscoelasticity, stretchability, and biocompatibility toward multifunctional applications. ACS Appl Mater Inter 2018;10(33):27987–8002. https://doi.org/
10.1021/acsami.8b09656.
[77] Han JQ, Wang HX, Yue YY, Mei CT, Chen JZ, Huang CB, et al. A self-healable and highly flexible supercapacitor integrated by dynamically cross-linked electro-
conductive hydrogels based on nanocellulose-templated carbon nanotubes embedded in a viscoelastic polymer network. Carbon 2019;149:1–18. https://doi.
org/10.1016/j.carbon.2019.04.029.
[78] Han JQ, Wang SW, Zhu SL, Huang CB, Yue YY, Mei CT, et al. Electrospun Core-Shell Nanofibrous Membranes with Nanocellulose-Stabilized Carbon Nanotubes
for Use as High-Performance Flexible Supercapacitor Electrodes with Enhanced Water Resistance, Thermal Stability, and Mechanical Toughness. ACS Appl
Mater Interfaces 2019;11(47):44624–35. https://doi.org/10.1021/acsami.9b16458.
[79] Zheng CX, Lu KY, Lu Y, Zhu SL, Yue YY, Xu XW, et al. A stretchable, self-healing conductive hydrogels based on nanocellulose supported graphene towards
wearable monitoring of human motion. Carbohyd Polym 2020;250:116905. https://doi.org/10.1016/j.carbpol.2020.116905.
[80] Fang Y, Jing CY, Li GL, Ling SJ, Wang ZH, Lu P, et al. Wood-derived systems for sustainable oil/water separation. Adv Sustain Syst 2021;5(7):2100039.
https://doi.org/10.1002/adsu.202100039.
[81] Lu Y, Yue YY, Ding QQ, Mei CT, Xu XW, Wu QL, et al. Self-recovery, fatigue-resistant, and multifunctional sensor assembled by a nanocellulose/carbon
nanotube nanocomplex-mediated hydrogel. ACS Appl Mater Inter 2021;13(42):50281–97. https://doi.org/10.1021/acsami.1c16828.
[82] Niu ZX, Cheng WL, Cao ML, Wang D, Wang QX, Han JQ, et al. Recent advances in cellulose-based flexible triboelectric nanogenerators. Nano Energy 2021;87:
106175. https://doi.org/10.1016/j.nanoen.2021.106175.
[83] Yang XP, Biswas SK, Han JQ, Tanpichai S, Li MC, Chen CC, et al. Surface and interface engineering for nanocellulosic advanced materials. Adv Mater 2021;33
(28):2002264. https://doi.org/10.1002/adma.202002264.
[84] Zhao DW, Zhu Y, Cheng WK, Chen WS, Wu YQ, Yu HP. Cellulose-based flexible functional materials for emerging intelligent electronics. Adv Mater 2021;33
(28):2000619. https://doi.org/10.1002/adma.202000619.
[85] Zhu SL, Sun HY, Lu Y, Wang SL, Yue YY, Xu XW, et al. Inherently conductive poly (dimethylsiloxane) elastomers synergistically mediated by nanocellulose/
carbon nanotube nanohybrids toward highly sensitive, stretchable, and durable strain sensors. ACS Appl Mater Inter 2021;13(49):59142–53. https://doi.org/
10.1021/acsami.1c19482.
[86] Iwamoto S, Nakagaito AN, Yano H, Nogi M. Optically transparent composites reinforced with plant fiber-based nanofibers. Appl Phys A 2005;81(6):1109–12.
https://doi.org/10.1007/s00339-005-3316-z.
[87] Iftekhar Shams M, Nogi M, Berglund LA, Yano H. The transparent crab: preparation and nanostructural implications for bioinspired optically transparent
nanocomposites. Soft Matter 2012;8(5):1369–73. https://doi.org/10.1039/c1sm06785k.
[88] Li YY, Fu QL, Yu S, Yan M, Berglund L. Optically transparent wood from a nanoporous cellulosic template: combining functional and structural performance.
Biomacromolecules 2016;17(4):1358–64. https://doi.org/10.1021/acs.biomac.6b00145.
[89] Zhu MW, Song JW, Li T, Gong A, Wang YB, Dai JQ, et al. Highly anisotropic, highly transparent wood composites. Adv Mater 2016;28(26):5181–7. https://doi.
org/10.1002/adma.201600427.
[90] Fang ZQ, Zhu HL, Bao WZ, Preston C, Liu Z, Dai JQ, et al. Highly transparent paper with tunable haze for green electronics. Energ Environ Sci 2014;7(10):
3313–9. https://doi.org/10.1039/c4ee02236j.

23
S. Zhu et al. Progress in Materials Science 132 (2023) 101025

[91] Henriksson M, Berglund LA, Isaksson P, Lindström T, Nishino T. Cellulose nanopaper structures of high toughness. Biomacromolecules 2008;9(6):1579–85.
https://doi.org/10.1021/bm800038n.
[92] Nogi M, Iwamoto S, Nakagaito AN, Yano H. Optically transparent nanofiber paper. Adv Mater 2009;21(16):1595–8. https://doi.org/10.1002/
adma.200803174.
[93] Yano H, Sugiyama J, Nakagaito AN, Nogi M, Matsuura T, Hikita M, et al. Optically transparent composites reinforced with networks of bacterial nanofibers.
Adv Mater 2005;17(2):153–5. https://doi.org/10.1002/adma.200400597.
[94] Saito T, Kimura S, Nishiyama Y, Isogai A. Cellulose nanofibers prepared by TEMPO-mediated oxidation of native cellulose. Biomacromolecules 2007;8(8):
2485–91. https://doi.org/10.1021/bm0703970.
[95] Yano H, Sasaki S, Shams MI, Abe K, Date T. Wood pulp-based optically transparent film: a paradigm from nanofibers to nanostructured fibers. Adv Optical
Mater 2014;2(3):231–4. https://doi.org/10.1002/adom.201300444.
[96] Jung YH, Chang TH, Zhang HL, Yao CH, Zheng QF, Yang VW, et al. High-performance green flexible electronics based on biodegradable cellulose nanofibril
paper. Nat Commun 2015;6(1):7170. https://doi.org/10.1038/ncomms8170.
[97] Gui Z, Zhu HL, Gillette E, Han XG, Rubloff GW, Hu LB, et al. Natural cellulose fiber as substrate for supercapacitor. ACS Nano 2013;7(7):6037–46. https://doi.
org/10.1021/nn401818t.
[98] Fang ZQ, Zhu HL, Yuan YB, Ha D, Zhu SZ, Preston C, et al. Novel nanostructured paper with ultrahigh transparency and ultrahigh haze for solar cells. Nano Lett
2014;14(2):765–73. https://doi.org/10.1021/nl404101p.
[99] Rajala S, Siponkoski T, Sarlin E, Mettänen M, Vuoriluoto M, Pammo A, et al. Cellulose nanofibril film as a piezoelectric sensor material. ACS Appl Mater
Interfaces 2016;8(24):15607–14. https://doi.org/10.1021/acsami.6b03597.
[100] Lei T, Guan M, Liu J, Lin H-C, Pfattner R, Shaw L, et al. Biocompatible and totally disintegrable semiconducting polymer for ultrathin and ultralightweight
transient electronics. P Natl Acad Sci USA 2017;114(20):5107–12. https://doi.org/10.1073/pnas.1701478114.
[101] Zhou SY, Kong XY, Zheng B, Huo FW, Strømme M, Xu C. Cellulose nanofiber@conductive metal-organic frameworks for high-performance flexible
supercapacitors. ACS Nano 2019;13(8):9578–86. https://doi.org/10.1021/acsnano.9b04670.
[102] Zhao DW, Chen CJ, Zhang Q, Chen WS, Liu SX, Wang QW, et al. High performance, flexible, solid-state supercapacitors based on a renewable and
biodegradable mesoporous cellulose membrane. Adv Energy Mater 2017;7(18):1700739. https://doi.org/10.1002/aenm.201700739.
[103] Zhao DW, Zhang Q, Chen WS, Yi X, Liu SX, Wang QW, et al. Highly flexible and conductive cellulose-mediated PEDOT:PSS/MWCNT composite films for
supercapacitor electrodes. ACS Appl Mater Interfaces 2017;9(15):13213–22. https://doi.org/10.1021/acsami.7b01852.
[104] Wang ZH, Tammela P, Strømme M, Nyholm L. Cellulose-based supercapacitors: material and performance considerations. Adv Energy Mater 2017;7(18):
1700130. https://doi.org/10.1002/aenm.201700130.
[105] Liu A, Walther A, Ikkala O, Belova L, Berglund LA. Clay nanopaper with tough cellulose nanofiber matrix for fire retardancy and gas barrier functions.
Biomacromolecules 2011;12(3):633–41. https://doi.org/10.1021/bm101296z.
[106] Koga H, Saito T, Kitaoka T, Nogi M, Suganuma K, Isogai A. Transparent, conductive, and printable composites consisting of TEMPO-oxidized nanocellulose and
carbon nanotube. Biomacromolecules 2013;14(4):1160–5. https://doi.org/10.1021/bm400075f.
[107] Li YY, Zhu HL, Gu HB, Dai HQ, Fang ZQ, Weadock NJ, et al. Strong transparent magnetic nanopaper prepared by immobilization of Fe3O4 nanoparticles in a
nanofibrillated cellulose network. J Mater Chem A 2013;1(48):15278–83. https://doi.org/10.1039/C3TA12591B.
[108] Hamedi MM, Hajian A, Fall AB, Håkansson K, Salajkova M, Lundell F, et al. Highly conducting, strong nanocomposites based on nanocellulose-assisted aqueous
dispersions of single-wall carbon nanotubes. ACS Nano 2014;8(3):2467–76. https://doi.org/10.1021/nn4060368.
[109] Sehaqui H, Zimmermann T, Tingaut P. Hydrophobic cellulose nanopaper through a mild esterification procedure. Cellulose 2014;21(1):367–82. https://doi.
org/10.1007/s10570-013-0110-5.
[110] Zhou D, Zou HY, Liu M, Zhang K, Sheng Y, Cui JL, et al. Surface ligand dynamics-guided preparation of quantum dots-cellulose composites for light-emitting
diodes. ACS Appl Mater Interfaces 2015;7(29):15830–9. https://doi.org/10.1021/acsami.5b03004.
[111] Sehaqui H, Ezekiel Mushi N, Morimune S, Salajkova M, Nishino T, Berglund LA. Cellulose nanofiber orientation in nanopaper and nanocomposites by cold
drawing. ACS Appl Mater Inter 2012;4(2):1043–9. https://doi.org/10.1021/am2016766.
[112] Torres-Rendon JG, Schacher FH, Ifuku S, Walther A. Mechanical performance of macrofibers of cellulose and chitin nanofibrils aligned by wet-stretching: a
critical comparison. Biomacromolecules 2014;15(7):2709–17. https://doi.org/10.1021/bm500566m.
[113] Tang H, Butchosa N, Zhou Q. A transparent, hazy, and strong macroscopic ribbon of oriented cellulose nanofibrils bearing poly(ethylene glycol). Adv Mater
2015;27(12):2070–6. https://doi.org/10.1002/adma.201404565.
[114] Wang BC, Torres-Rendon JG, Yu JC, Zhang YM, Walther A. Aligned bioinspired cellulose nanocrystal-based nanocomposites with synergetic mechanical
properties and improved hygromechanical performance. ACS Appl Mater Interfaces 2015;7(8):4595–607. https://doi.org/10.1021/am507726t.
[115] Zhu MW, Wang YL, Zhu SZ, Xu LS, Jia C, Dai JQ, et al. Anisotropic, transparent films with aligned cellulose nanofibers. Adv Mater 2017;29(21):1606284.
https://doi.org/10.1002/adma.201606284.
[116] Wang S, Li T, Chen CJ, Kong WQ, Zhu SZ, Dai JQ, et al. Transparent, anisotropic biofilm with aligned bacterial cellulose nanofibers. Adv Funct Mater 2018;28
(24):1707491. https://doi.org/10.1002/adfm.201707491.
[117] Ye DD, Lei XJ, Li T, Cheng QY, Chang CY, Hu LB, et al. Ultrahigh tough, super clear, and highly anisotropic nanofiber-structured regenerated cellulose films.
ACS Nano 2019;13(4):4843–53. https://doi.org/10.1021/acsnano.9b02081.
[118] Qin CR, Yang W, Wang Y, Zhang LN, Lu A. Robust, magnetic cellulose/Fe3O4 film with anisotropic sensory property. Cellulose 2021;28(4):2353–64. https://
doi.org/10.1007/s10570-020-03634-4.
[119] Chen YW, Liu YH, Xia YM, Liu XQ, Qiang Z, Yang JY, et al. Electric field-induced assembly and alignment of silver-coated cellulose for polymer composite films
with enhanced dielectric permittivity and anisotropic light transmission. ACS Appl Mater Interfaces 2020;12(21):24242–9. https://doi.org/10.1021/
acsami.0c03086.
[120] Ray U, Zhu SZ, Pang ZQ, Li T. Mechanics design in cellulose-enabled high-performance functional materials. Adv Mater 2021;33(28):2002504. https://doi.
org/10.1002/adma.202002504.
[121] Rosén T, Hsiao BS, Söderberg LD. Elucidating the opportunities and challenges for nanocellulose spinning. Adv Mater 2021;33(28):2001238. https://doi.org/
10.1002/adma.202001238.
[122] Chen CJ, Kuang YD, Zhu SZ, Burgert I, Keplinger T, Gong A, et al. Structure-property-function relationships of natural and engineered wood. Nat Rev Mater
2020;5(9):642–66. https://doi.org/10.1038/s41578-020-0195-z.
[123] Rai R, Ranjan R, Dhar P. Life cycle assessment of transparent wood production using emerging technologies and strategic scale-up framework. Sci Total
Environ 2022;846:157301. https://doi.org/10.1016/j.scitotenv.2022.157301.
[124] Xia QQ, Chen CJ, Li T, He SM, Gao JL, Wang XZ, et al. Solar-assisted fabrication of large-scale, patternable transparent wood. Sci Adv 2021;7(5):7342–9.
https://doi.org/10.1126/sciadv.abd7342.
[125] Zhu MW, Li T, Davis CS, Yao YG, Dai JQ, Wang YB, et al. Transparent and haze wood composites for highly efficient broadband light management in solar
cells. Nano Energy 2016;26:332–9. https://doi.org/10.1016/j.nanoen.2016.05.020.
[126] Li YY, Yu S, Veinot JGC, Linnros J, Berglund L, Sychugov I. Luminescent transparent wood. Adv Optical Mater 2017;5(1):1600834. https://doi.org/10.1002/
adom.201600834.
[127] Gan WT, Xiao SL, Gao LL, Gao RN, Li J, Zhan XX. Luminescent and transparent wood composites fabricated by poly(methyl methacrylate) and γ-Fe2O3@YVO4:
Eu3+ nanoparticle impregnation. ACS Sustainable Chem Eng 2017;5(5):3855–62. https://doi.org/10.1021/acssuschemeng.6b02985.
[128] Yu ZY, Yao YJ, Yao JN, Zhang LM, Chen Z, Gao YF, et al. Transparent wood containing CsxWO3 nanoparticles for heat-shielding window applications. J Mater
Chem A 2017;5(13):6019–24. https://doi.org/10.1039/c7ta00261k.
[129] Fu QL, Yan M, Jungstedt E, Yang X, Li YY, Berglund LA. Transparent plywood as a load-bearing and luminescent biocomposite. Compos Sci Technol 2018;164:
296–303. https://doi.org/10.1016/j.compscitech.2018.06.001.

24
S. Zhu et al. Progress in Materials Science 132 (2023) 101025

[130] Li YY, Yang X, Fu QL, Rojas R, Yan M, Berglund L. Towards centimeter thick transparent wood through interface manipulation. J Mater Chem A 2018;6(3):
1094–101. https://doi.org/10.1039/c7ta09973h.
[131] Tang QH, Fang L, Wang YF, Zou M, Guo WJ. Anisotropic flexible transparent films from remaining wood microstructures for screen protection and AgNW
conductive substrate. Nanoscale 2018;10(9):4344–53. https://doi.org/10.1039/c7nr08367j.
[132] Wang LH, Liu YJ, Zhan XY, Luo D, Sun XW. Photochromic transparent wood for photo-switchable smart window applications. J Mater Chem C 2019;7(28):
8649–54. https://doi.org/10.1039/c9tc02076d.
[133] Zhang T, Yang P, Li YZ, Cao Y, Zhou Y, Chen M, et al. Flexible transparent sliced veneer for alternating current electroluminescent devices. ACS Sustainable
Chem Eng 2019;7(13):11464–73. https://doi.org/10.1021/acssuschemeng.9b01129.
[134] Montanari C, Li YY, Chen H, Yan M, Berglund LA. Transparent wood for thermal energy storage and reversible optical transmittance. ACS Appl Mater
Interfaces 2019;11(22):20465–72. https://doi.org/10.1021/acsami.9b05525.
[135] Li YY, Cheng M, Jungstedt E, Xu B, Sun LC, Berglund L. Optically transparent wood substrate for perovskite solar cells. ACS Sustainable Chem Eng 2019;7(6):
6061–7. https://doi.org/10.1021/acssuschemeng.8b06248.
[136] Mi RY, Chen CJ, Keplinger T, Pei Y, He SM, Liu DP, et al. Scalable aesthetic transparent wood for energy efficient buildings. Nat Commun 2020;11(1):3836.
https://doi.org/10.1038/s41467-020-17513-w.
[137] Chen L, Xu ZW, Wang F, Duan GG, Xu WH, Zhang GY, et al. A flame-retardant and transparent wood/polyimide composite with excellent mechanical strength.
Compos Commun 2020;20:100355. https://doi.org/10.1016/j.coco.2020.05.001.
[138] Fu QL, Chen Y, Sorieul M. Wood-based flexible electronics. ACS Nano 2020;14(3):3528–38. https://doi.org/10.1021/acsnano.9b09817.
[139] Fu QL, Tu KK, Goldhahn C, Keplinger T, Adobes-Vidal M, Sorieul M, et al. Luminescent and hydrophobic wood films as optical lighting materials. ACS Nano
2020;14(10):13775–83. https://doi.org/10.1021/acsnano.0c06110.
[140] Liu YS, Yang HY, Ma CH, Luo S, Xu MC, Wu ZW, et al. Luminescent transparent wood based on lignin-derived carbon dots as a building material for dual-
channel, real-time, and visual detection of formaldehyde gas. ACS Appl Mater Interfaces 2020;12(32):36628–38. https://doi.org/10.1021/acsami.0c10240.
[141] Höglund M, Garemark J, Nero M, Willhammar T, Popov S, Berglund LA. Facile processing of transparent wood nanocomposites with structural color from
plasmonic nanoparticles. Chem Mater 2021;33(10):3736–45. https://doi.org/10.1021/acs.chemmater.1c00806.
[142] Wang KL, Dong YM, Ling Z, Liu XR, Shi SQ, Li JZ. Transparent wood developed by introducing epoxy vitrimers into a delignified wood template. Compos Sci
Technol 2021;207:108690. https://doi.org/10.1016/j.compscitech.2021.108690.
[143] Montanari C, Ogawa Y, Olsén P, Berglund LA. High performance, fully bio-based, and optically transparent wood biocomposites. Adv Sci 2021;8(12):2100559.
https://doi.org/10.1002/advs.202100559.
[144] Tang QH, Zou M, Chang L, Guo WJ. A super-flexible and transparent wood film/silver nanowire electrode for optical and capacitive dual-mode sensing wood-
based electronic skin. Chem Eng J 2021;132152. https://doi.org/10.1016/j.cej.2021.132152.
[145] Chen H, Montanari C, Shanker R, Marcinkevicius S, Berglund LA, Sychugov I. Photon walk in transparent wood: scattering and absorption in hierarchically
structured materials. Adv Opt Mater 2022;10(8):2102732. https://doi.org/10.1002/adom.202102732.
[146] Liu L, Zhu G, Chen Y, Liu Z, Donaldson L, Zhan X, et al. Switchable photochromic transparent wood as smart packaging materials. Ind Crop Prod 2022;184:
115050. https://doi.org/10.1016/j.indcrop.2022.115050.
[147] Panzarasa G, Burgert I. Designing functional wood materials for novel engineering applications. Holzforschung 2022;76(2):211–22. https://doi.org/10.1515/
hf-2021-0125.
[148] Wan C, Liu X, Huang Q, Cheng W, Su J, Wu Y. A brief review of transparent wood: synthetic strategy, functionalization, and applications. Curr Org Synth 2021;
18:615–23. https://doi.org/10.2174/1570179418666210614141032.
[149] Wang J, Zhu JG. Prospects and applications of biomass-based transparent wood: an architectural glass perspective. Front Chem 2021;9:747385. https://doi.
org/10.3389/fchem.2021.747385.
[150] Qiu Z, Xiao ZF, Gao LK, Li J, Wang HG, Wang YG, et al. Transparent wood bearing a shielding effect to infrared heat and ultraviolet via incorporation of
modified antimony-doped tin oxide nanoparticles. Compos Sci Technol 2019;172:43–8. https://doi.org/10.1016/j.compscitech.2019.01.005.
[151] Li YY, Fu QL, Rojas R, Yan M, Berglund L. Lignin-retaining transparent wood. ChemSusChem 2017;10(17):3445–51. https://doi.org/10.1002/cssc.201701089.
[152] Müller U, Rätzsch M, Schwanninger M, Steiner M, Zöbl H. Yellowing and IR-changes of spruce wood as result of UV-irradiation. J Photoch Photobio B 2003;69
(2):97–105. https://doi.org/10.1016/s1011-1344(02)00412-8.
[153] Subba Rao AN, Nagarajappa GB, Nair S, Chathoth AM, Pandey KK. Flexible transparent wood prepared from poplar veneer and polyvinyl alcohol. Compos Sci
Technol 2019;182:107719. https://doi.org/10.1016/j.compscitech.2019.107719.
[154] Fink S. Transparent wood-a new approach in the functional study of wood structure. Holzforschung 1992;46:403–8. https://doi.org/10.1515/
hfsg.1992.46.5.403.
[155] Zhu HL, Luo W, Ciesielski PN, Fang ZQ, Zhu JY, Henriksson G, et al. Wood-derived materials for green electronics, biological devices, and energy applications.
Chem Rev 2016;116(16):9305–74. https://doi.org/10.1021/acs.chemrev.6b00225.
[156] Keplinger T, Wittel FK, Rüggeberg M, Burgert I. Wood derived cellulose scaffolds-processing and mechanics. Adv Mater 2020;33(28):2001375. https://doi.
org/10.1002/adma.202001375.
[157] Chen X, Ge-Zhang S, Han Y, Yang H, Ou-Yang W, Zhu H, et al. Ultraviolet-assisted modified delignified wood with high transparency. Appl Sci 2022;12(15):
7406. https://doi.org/10.3390/app12157406.
[158] Jiang Y, Zhang M, Weng M, Liu X, Rong X, Huang Q, et al. Hemicellulose-rich transparent wood: Microstructure and macroscopic properties. Carbohyd Polym
2022;296:119925. https://doi.org/10.1016/j.carbpol.2022.119925.
[159] Wu JM, Wu Y, Yang F, Tang CY, Huang QT, Zhang JL. Impact of delignification on morphological, optical and mechanical properties of transparent wood.
Compos Part A-Appl S 2019;117:324–31. https://doi.org/10.1016/j.compositesa.2018.12.004.
[160] Li JG, Chen CJ, Zhu JY, Ragauskas AJ, Hu LB. In situ wood delignification toward sustainable applications. Acc Mater Res 2021;2(8):606–20. https://doi.org/
10.1021/accountsmr.1c00075.
[161] Gierer J. Chemistry of delignification. Wood Sci Technol 1986;20(1):1–33. https://doi.org/10.1007/BF00350692.
[162] Ma RS, Xu Y, Zhang X. Catalytic oxidation of biorefinery lignin to value-added chemicals to support sustainable biofuel production. ChemSusChem 2015;8(1):
24–51. https://doi.org/10.1002/cssc.201402503.
[163] Gierer J. Chemistry of delignification. Wood Sci Technol 1985;19(4):289–312. https://doi.org/10.1007/BF00350807.
[164] Li HY, Guo XL, He YM, Zheng RB. A green steam-modified delignification method to prepare low-lignin delignified wood for thick, large highly transparent
wood composites. J Mater Res 2019;34(6):932–40. https://doi.org/10.1557/jmr.2018.466.
[165] Frey M, Widner D, Segmehl JS, Casdorff K, Keplinger T, Burgert I. Delignified and densified cellulose bulk materials with excellent tensile properties for
sustainable engineering. ACS Appl Mater Interfaces 2018;10(5):5030–7. https://doi.org/10.1021/acsami.7b18646.
[166] Martínez ÁT, Ruiz-Dueñas FJ, Martínez MJ, del Río JC, Gutiérrez A. Enzymatic delignification of plant cell wall: from nature to mill. Curr Opin Biotech 2009;
20(3):348–57. https://doi.org/10.1016/j.copbio.2009.05.002.
[167] Chen ZJ, Dang B, Luo XF, Li W, Li J, Yu HP, et al. Deep eutectic solvent-assisted in situ wood delignification: a promising strategy to enhance the efficiency of
wood-based solar steam generation devices. ACS Appl Mater Inter 2019;11(29):26032–7. https://doi.org/10.1021/acsami.9b08244.
[168] He YM, Li HY, Guo XL, Zheng RB. Delignified wood-based highly efficient solar steam generation device via promoting both water transportation and
evaporation. BioResources 2019;14(2):3758–67. https://doi.org/10.1016/j.solmat.2020.110910.
[169] Bi ZH, Li TW, Su H, Ni Y, Yan LF. Transparent wood film incorporating carbon dots as encapsulating material for white light-emitting diodes. ACS Sustainable
Chem Eng 2018;6(7):9314–23. https://doi.org/10.1021/acssuschemeng.8b01618.
[170] Khakalo A, Tanaka A, Korpela A, Hauru LKJ, Orelma H. All-wood composite material by partial fiber surface dissolution with an ionic liquid. ACS Sustain
Chem Eng 2019;7(3):3195–202. https://doi.org/10.1021/acssuschemeng.8b05059.

25
S. Zhu et al. Progress in Materials Science 132 (2023) 101025

[171] Khakalo A, Tanaka A, Korpela A, Orelma H. Delignification and ionic liquid treatment of wood toward multifunctional high-performance structural materials.
ACS Appl Mater Interfaces 2020;12(20):23532–42. https://doi.org/10.1021/acsami.0c02221.
[172] Wu Y, Yang LC, Zhou JC, Yang F, Huang QT, Cai YJ. Softened wood treated by deep eutectic solvents. ACS Omega 2020;5(35):22163–70. https://doi.org/
10.1021/acsomega.0c02223.
[173] Liu YZ, Chen WS, Xia QQ, Guo BT, Wang QW, Liu SX, et al. Efficient cleavage of lignin-carbohydrate complexes and ultrafast extraction of lignin oligomers
from wood biomass by microwave-assisted treatment with deep eutectic solvent. ChemSusChem 2017;10(8):1692–700. https://doi.org/10.1002/
cssc.201601795.
[174] Ren K, Xia QQ, Liu YZ, Cheng WK, Zhu Y, Liu Y, et al. Wood/polyimide composite via a rapid substitution compositing method for extreme temperature
conditions. Compos Sci Technol 2021;207:108698. https://doi.org/10.1016/j.compscitech.2021.108698.
[175] Jiang J, Zhu Y, Jiang F. Sustainable isolation of nanocellulose from cellulose and lignocellulosic feedstocks: Recent progress and perspectives. Carbohydr
Polym 2021;267:118188. https://doi.org/10.1016/j.carbpol.2021.118188.
[176] Liu Y, Deak N, Wang Z, Yu H, Hameleers L, Jurak E, et al. Tunable and functional deep eutectic solvents for lignocellulose valorization. Nat Commun 2021;12
(1):5424. https://doi.org/10.1038/s41467-021-25117-1.
[177] Jungstedt E, Östlund S, Berglund L. Transverse fracture toughness of transparent wood biocomposites by FEM updating with cohesive zone fracture modeling.
Compos Sci Technol 2022;109492. https://doi.org/10.1016/j.compscitech.2022.109492.
[178] Pokhrel D, Viraraghavan T. Treatment of pulp and paper mill waste water-a review. Sci Total Environ 2004;333(1–3):37–58. https://doi.org/10.1016/j.
scitotenv.2004.05.017.
[179] Xia QQ, Chen CJ, Yao YG, He SM, Wang XZ, Li JG, et al. In situ lignin modification toward photonic wood. Adv Mater 2021;33(8):2001588. https://doi.org/
10.1002/adma.202001588.
[180] Frey M, Schneider L, Masania K, Keplinger T, Burgert I. Delignified wood–polymer interpenetrating composites exceeding the rule of mixtures. ACS Appl Mater
Inter 2019;11(38):35305–11. https://doi.org/10.1021/acsami.9b11105.
[181] Cai HC, Wang ZQ, Xie D, Zhao PP, Sun JP, Qin DY, et al. Flexible transparent wood enabled by epoxy resin and ethylene glycol diglycidyl ether. J For Res 2020;
32(4):1779–87. https://doi.org/10.1007/s11676-020-01201-y.
[182] Montanari C, Olsén P, Berglund LA. Interface tailoring by a versatile functionalization platform for nanostructured wood biocomposites. Green Chem 2020;22
(22):8012–23. https://doi.org/10.1039/d0gc02768e.
[183] Yue DR, Fu GH, Jin ZY. Transparent wood prepared by polymer impregnation of rubber wood (hevea brasiliensis muell. arg). Bioresources 2021;16(2):
2491–502. https://doi.org/10.15376/biores.16.2.2491-2502.
[184] Hoglund M, Johansson M, Sychugov I, Berglund LA. Transparent wood biocomposites by fast UV-curing for reduced light-scattering through wood/thiol-ene
interface design. ACS Appl Mater Interfaces 2020;12(41):46914–22. https://doi.org/10.1021/acsami.0c12505.
[185] Tan Y, Wang K, Dong Y, Gong S, Shi SQ, Li J. High performance, shape manipulatable transparent wood based on delignified wood framework and
exchangeable dynamic covalent vitrimers. Chem Eng J 2022;448:137487. https://doi.org/10.1016/j.cej.2022.137487.
[186] Samanta P, Samanta A, Montanari C, Li Y, Maddalena L, Carosio F, et al. Fire-retardant and transparent wood biocomposite based on commercial thermoset.
Compos Part A-Appl S 2022;156:106863. https://doi.org/10.1016/j.compositesa.2022.106863.
[187] Foster KEO, Jones R, Miyake GM, Srubar WV. Mechanics, optics, and thermodynamics of water transport in chemically modified transparent wood composites.
Compos Sci Technol 2021;208:108737. https://doi.org/10.1016/j.compscitech.2021.108737.
[188] Jakob M, Mahendran AR, Gindl-Altmutter W, Bliem P, Konnerth J, Müller U, et al. The strength and stiffness of oriented wood and cellulose-fibre materials: A
review. Prog Mater Sci 2022;125:100916. https://doi.org/10.1016/j.pmatsci.2021.100916.
[189] Sun H, Ji T, Zhou X, Bi H, Xu M, Huang Z, et al. Mechanically strong, transparent, and biodegradable wood-derived film. Mater Chem Front 2021;5(21):
7903–9. https://doi.org/10.1039/d1qm00973g.
[190] Han XS, Ye YH, Lam F, Pu JW, Jiang F. Hydrogen-bonding-induced assembly of aligned cellulose nanofibers into ultrastrong and tough bulk materials. J Mater
Chem A 2019;7(47):27023–31. https://doi.org/10.1039/c9ta11118b.
[191] Li K, Wang SN, Chen H, Yang X, Berglund LA, Zhou Q. Self-densification of highly mesoporous wood structure into a strong and transparent film. Adv Mater
2020;32(42):2003653. https://doi.org/10.1002/adma.202003653.
[192] Wang K, Liu X, Dong Y, Ling Z, Cai Y, Tian D, et al. Editable shape-memory transparent wood based on epoxy-based dynamic covalent polymer with excellent
optical and thermal management for smart building materials. Cellulose 2022;29(14):7955–72. https://doi.org/10.1007/s10570-022-04754-9.
[193] Hu X, Zhang Y, Zhang J, Yang H, Wang F, Bin F, et al. Sonochemically-coated transparent wood with ZnO: Passive radiative cooling materials for energy saving
applications. Renew Energ 2022;193:398–406. https://doi.org/10.1016/j.renene.2022.05.008.
[194] Vasileva E, Li YY, Sychugov I, Mensi M, Berglund L, Popov S. Lasing from organic dye molecules embedded in transparent wood. Adv Opt Mater 2017;5(10):
1700057. https://doi.org/10.1002/adom.201700057.
[195] Gan WT, Gao LK, Xiao SL, Zhang WB, Zhan XX, Li J. Transparent magnetic wood composites based on immobilizing Fe3O4 nanoparticles into a delignified
wood template. J Mater Sci 2017;52(6):3321–9. https://doi.org/10.1007/s10853-016-0619-8.
[196] Koivurova M, Vasileva E, Li YY, Berglund L, Popov S. Complete spatial coherence characterization of quasi-random laser emission from dye doped transparent
wood. Opt Express 2018;26(10):13474–82. https://doi.org/10.1364/oe.26.013474.
[197] Liu YX, Lu CJ, Bian S, Hu K, Zheng KW, Sun QF. Reversible photo-responsive smart wood with resistant to extreme weather. J Mater Sci 2022:1–11. https://
doi.org/10.1007/s10853-021-06756-7.
[198] Zou M, Chen Y, Chang L, Cheng X, Gao L, Guo W, et al. Toward 90 μm superthin transparent wood film impregnated with quantum dots for color-converting
materials. ACS Sustain Chem Eng 2022;10(6):2097–106. https://doi.org/10.1021/acssuschemeng.1c07013.
[199] Zhang L, Jiang Y, Zhou L, Jiang Z, Li L, Che W, et al. Mechanical, thermal stability, and flame retardancy performance of transparent wood composite improved
with delaminated Ti3C2Tx (MXene) nanosheets. J Mater Sci 2022:1–12. https://doi.org/10.1007/s10853-021-06776-3.
[200] Chu T, Gao Y, Yi L, Fan C, Yan L, Ding C, et al. Highly fire-retardant optical wood enabled by transparent fireproof coatings. Adv Compos Hybrid Ma 2022:1–9.
https://doi.org/10.1007/s42114-022-00440-3.
[201] Gan J, Wu Y, Yang F, Zhang H, Wu X, Wang Y, et al. Wood-cellulose photoluminescence material based on carbon quantum dot for light conversion. Carbohyd
Polym 2022:119429. https://doi.org/10.1016/j.carbpol.2022.119429.
[202] Samanta A, Höglund M, Samanta P, Popov S, Sychugov I, Maddalena L, et al. Charge regulated diffusion of silica nanoparticles into wood for flame retardant
transparent wood. Adv Sustain Syst 2022;2100354. https://doi.org/10.1002/adsu.202100354.
[203] Wang M, Li RA, Chen GX, Zhou SH, Feng X, Chen Y, et al. Highly stretchable, transparent, and conductive wood fabricated by in situ photopolymerization with
polymerizable deep eutectic solvents. ACS Appl Mater Interfaces 2019;11(15):14313–21. https://doi.org/10.1021/acsami.9b00728.
[204] Zhang T, Yang P, Chen MZ, Yang K, Cao YZ, Li XH, et al. Constructing a novel electroluminescent device with high-temperature and high-humidity resistance
based on a flexible transparent wood film. ACS Appl Mater Interfaces 2019;11(39):36010–9. https://doi.org/10.1021/acsami.9b09331.
[205] Zhang LM, Wang A, Zhu TL, Chen Z, Wu YP, Gao YF. Transparent wood composites fabricated by impregnation of epoxy resin and W-doped VO2 nanoparticles
for application in energy-saving windows. ACS Appl Mater Interfaces 2020;12(31):34777–83. https://doi.org/10.1021/acsami.0c06494.
[206] Samanta A, Chen H, Samanta P, Popov S, Sychugov I, Berglund LA. Reversible dual-stimuli-responsive chromic transparent wood biocomposites for smart
window applications. ACS Appl Mater Interfaces 2021;13(2):3270–7. https://doi.org/10.1021/acsami.0c21369.
[207] Bisht P, Pandey KK, Barshilia HC. Photostable transparent wood composite functionalized with an UV-absorber. Polym Degrad Stabil 2021;189:109600.
https://doi.org/10.1016/j.polymdegradstab.2021.109600.
[208] Duan JF, Yang N, Li T, Saiyu Y, Zong SY, Wen HK, et al. Preparation and properties of a kind of high conductivity flexible wood with light switch
characteristics. J Appl Polym Sci 2021:e51843. https://doi.org/10.1002/app.51843.
[209] Aldalbahi A, El-Naggar ME, Khattab TA, Hossain M. Preparation of flame-retardant, hydrophobic, ultraviolet protective, and luminescent transparent wood.
Luminescence 2021;36(8):1922–32. https://doi.org/10.1002/bio.4126.

26
S. Zhu et al. Progress in Materials Science 132 (2023) 101025

[210] Wang D, Zhang Y, Zhang M, Wang Y, Li T, Liu T, et al. Wood-derived composites with high performance for thermal management applications.
Biomacromolecules 2021;22(10):4228–36. https://doi.org/10.1021/acs.biomac.1c00786.
[211] Wang K, Zhang T, Li C, Xiao X, Tang Y, Fang X, et al. Shape-reconfigurable transparent wood based on solid-state plasticity of polythiourethane for smart
building materials with tunable light guiding, energy saving, and fire alarm actuating functions. Compos Part B-Eng 2022::110260. https://doi.org/10.1016/j.
compositesb.2022.110260.
[212] Walker JC. Primary wood processing: principles and practice. Dordrecht: Springer Science & Business Media; 2006.
[213] Sorieul M, Dickson A, Hill SJ, Pearson H. Plant fibre: molecular structure and biomechanical properties, of a complex living material, influencing its
deconstruction towards a biobased composite. Materials 2016;9(8):618.
́
[214] J. Fahl en, en
́ LS. Cross-sectional structure of the secondary wall of wood fibers as affected by processing. J Mater Sci 2003; 38 (1): 119– 126. http://doi.org/
10.1023/A:1021174118468.
[215] Chen CJ, Hu LB. Nanoscale ion regulation in wood-based structures and their device applications. Adv Mater 2021;33(28):2002890. https://doi.org/10.1002/
adma.202002890.
[216] Chen P, Li Y, Nishiyama Y, Pingali SV, O’Neill HM, Zhang Q, et al. Small angle neutron scattering shows nanoscale PMMA distribution in transparent wood
biocomposites. Nano Lett 2021;21(7):2883–90. https://doi.org/10.1021/acs.nanolett.0c05038.
[217] Pang J, Baitenov A, Montanari C, Samanta A, Berglund L, Popov S, et al. Light propagation in transparent wood: efficient ray-tracing simulation and retrieving
an effective refractive index of wood scaffold. Adv Photonics Res 2021;2(11):2100135. https://doi.org/10.1002/adpr.202100135.
[218] Yaddanapudi HS, Hickerson N, Saini S, Tiwari A. Fabrication and characterization of transparent wood for next generation smart building applications.
Vacuum 2017;146:649–54. https://doi.org/10.1016/j.vacuum.2017.01.016.
[219] Jungstedt E, Montanari C, Östlund S, Berglund L. Mechanical properties of transparent high strength biocomposites from delignified wood veneer. Compos Part
A-Appl S 2020;133:105853. https://doi.org/10.1016/j.compositesa.2020.105853.
[220] Foster KEO, Hess KM, Miyake GM, Srubar 3rd WV. Optical properties and mechanical modeling of acetylated transparent wood composite laminates. Materials
(Basel) 2019;12(14):2256. https://doi.org/10.3390/ma12142256.
[221] Qin JK, Li XW, Shao YL, Shi KX, Zhao X, Feng TS, et al. Optimization of delignification process for efficient preparation of transparent wood with high strength
and high transmittance. Vacuum 2018;158:158–65. https://doi.org/10.1016/j.vacuum.2018.09.058.
[222] Wang Y, Wu Y, Yang F, Yang L, Wang J, Zhou J, et al. A highly transparent compressed wood prepared by cell wall densification. Wood Sci Technol 2022:1–18.
https://doi.org/10.1007/s00226-022-01372-3.
[223] Zhu MW, Jia C, Wang YL, Fang ZQ, Dai JQ, Xu LS, et al. Isotropic paper directly from anisotropic wood: top-down green transparent substrate toward
biodegradable electronics. ACS Appl Mater Interfaces 2018;10(34):28566–71. https://doi.org/10.1021/acsami.8b08055.
[224] Sun H, Bi H, Ren Z, Zhou X, Ji T, Xu M, et al. Hydrostable reconstructed wood with transparency, Excellent ultraviolet-blocking performance, and
photothermal conversion ability. Compos Part B-Eng 2022;109615. https://doi.org/10.1016/j.compositesb.2022.109615.
[225] Chen H, Baitenov A, Li Y, Vasileva E, Popov S, Sychugov I, et al. Thickness dependence of optical transmittance of transparent wood: chemical modification
effects. ACS Appl Mater Interfaces 2019;11(38):35451–7. https://doi.org/10.1021/acsami.9b11816.
[226] Vasileva E, Chen H, Li YY, Sychugov I, Yan M, Berglund L, et al. Light scattering by structurally anisotropic media: a benchmark with transparent wood. Adv
Opt Mater 2018;6(23):1800999. https://doi.org/10.1002/adom.201800999.
[227] Vasileva E, Baitenov A, Chen H, Li YY, Sychugov I, Yan M, et al. Effect of transparent wood on the polarization degree of light. Opt Lett 2019;44(12):2962–5.
https://doi.org/10.1364/ol.44.002962.
[228] Wu Y, Zhou JC, Yang F, Wang YJ, Wang J, Zhang J. A strong multilayered transparent wood with natural wood color and texture. J Mater Sci 2021;56(13):
8000–13. https://doi.org/10.1007/s10853-021-05833-1.
[229] Wu Y, Wang YJ, Yang F. Comparison of multilayer transparent wood and single layer transparent wood with the same thickness. Front Mater 2021;8(41):
633345. https://doi.org/10.3389/fmats.2021.633345.
[230] Zhou JC, Xu W. Toward interface optimization of transparent wood with wood color and texture by silane coupling agent. J Mater Sci 2022;57(10):5825–38.
https://doi.org/10.1007/s10853-022-06974-7.
[231] Building technologies office’s multi-year program plan for fiscal years 2016 through 2020. February 2016.
[232] Li HY, Guo XL, He YM, Zheng RB. House model with 2–5 cm thick translucent wood walls and its indoor light performance. Eur J Wood Wood Prod 2019;77(5):
843–51. https://doi.org/10.1007/s00107-019-01431-w.
[233] Xia RQ, Zhang WY, Yang YN, Zhao JQ, Liu Y, Guo HW. Transparent wood with phase change heat storage as novel green energy storage composites for
building energy conservation. J Clean Prod 2021;296. https://doi.org/10.1016/j.jclepro.2021.126598.
[234] Pielichowski K, Flejtuch K. Differential scanning calorimetry studies on poly(ethylene glycol) with different molecular weights for thermal energy storage
materials. Polym Advan Technol 2002;13(10–12):690–6. https://doi.org/10.1002/pat.276.
[235] Soares N, Costa JJ, Gaspar AR, Santos P. Review of passive PCM latent heat thermal energy storage systems towards buildings’ energy efficiency. Energ
Buildings 2013;59:82–103. https://doi.org/10.1016/j.enbuild.2012.12.042.
[236] Fan C, Gao Y, Li Y, Yan L, Zhuang Y, Zhang Y, et al. A flame-retardant and optically transparent wood composite. J Appl Polym Sci 2022:e52945. https://doi.
org/10.1002/app.52945.
[237] Al-Qahtani S, Aljuhani E, Felaly R, Alkhamis K, Alkabli J, Munshi A, et al. Development of photoluminescent translucent wood toward photochromic smart
window applications. Ind Eng Chem Res 2021;60(23):8340–50. https://doi.org/10.1021/acs.iecr.1c01603.
[238] Zhou J, Liu Y, Yang Z, Wang X, Li Y, Liu F, et al. Turing Pattern-Inspired Highly Transparent Wood for Multifunctional Smart Glass with Superior Thermal
Management and UV-Blocking Ability. Adv Sustain Syst 2022;6(8):2200132. https://doi.org/10.1002/adsu.202200132.
[239] Li YY, Gu X, Gao H, Li J. Photoresponsive wood composite for photoluminescence and ultraviolet absorption. Constr Build Mater 2020;261:119984. https://
doi.org/10.1016/j.conbuildmat.2020.119984.
[240] Van Hai L, Muthoka RM, Panicker PS, Agumba DO, Pham HD, Kim J. All-biobased transparent-wood: a new approach and its environmental-friendly packaging
application. Carbohyd Polym 2021;264:118012. https://doi.org/10.1016/j.carbpol.2021.118012.
[241] Liu S, Tso CY, Lee HH, Du YW, Yu KM, Feng S-P, et al. Self-densified optically transparent VO2 thermochromic wood film for smart windows. ACS Appl Mater
Interfaces 2021;13(19):22495–504. https://doi.org/10.1021/acsami.1c03803.
[242] Wu H, Kong DS, Ruan ZC, Hsu P-C, Wang S, Yu ZF, et al. A transparent electrode based on a metal nanotrough network. Nat Nanotechnol 2013;8(6):421–5.
https://doi.org/10.1038/nnano.2013.84.
[243] Green MA, Ho-Baillie A, Snaith HJ. The emergence of perovskite solar cells. Nat Photonics 2014;8(7):506–14. https://doi.org/10.1038/nphoton.2014.134.
[244] Liu Z, Xu J, Chen D, Shen GZ. Flexible electronics based on inorganic nanowires. Chem Soc Rev 2015;44(1):161–92. https://doi.org/10.1039/C4CS00116H.
[245] Fan FR, Tang W, Wang ZL. Flexible nanogenerators for energy harvesting and self-powered electronics. Adv Mater 2016;28(22):4283–305. https://doi.org/
10.1002/adma.201504299.
[246] Lou Z, Shen GZ. Flexible photodetectors based on 1D inorganic nanostructures. Adv Sci 2016;3(6):1500287. https://doi.org/10.1002/advs.201500287.
[247] Petrus ML, Schlipf J, Li C, Gujar TP, Giesbrecht N, Müller-Buschbaum P, et al. Capturing the sun: a review of the challenges and perspectives of perovskite solar
cells. Adv Energy Mater 2017;7(16):1700264. https://doi.org/10.1002/aenm.201700264.
[248] Han JQ, Lu KY, Yue YY, Mei CT, Huang CB, Wu QL, et al. Nanocellulose-templated assembly of polyaniline in natural rubber-based hybrid elastomers toward
flexible electronic conductors. Ind Crop Prod 2019;128:94–107. https://doi.org/10.1016/j.indcrop.2018.11.004.
[249] He J, Zhou R, Zhang Y, Gao W, Chen T, Mai W, et al. Strain-insensitive self-powered tactile sensor arrays based on intrinsically stretchable and patternable
ultrathin conformal wrinkled graphene-elastomer composite. Adv Funct Mater 2021;2107281. https://doi.org/10.1002/adfm.202107281.
[250] Dai SL, Chu YL, Liu DP, Cao F, Wu XH, Zhou JC, et al. Intrinsically ionic conductive cellulose nanopapers applied as all solid dielectrics for low voltage organic
transistors. Nat commun 2018;9(1):1–10. https://doi.org/10.1038/s41467-018-05155-y.

27
S. Zhu et al. Progress in Materials Science 132 (2023) 101025

[251] Zhou TL, Wang JW, Huang M, An R, Tan HP, Wei H, et al. Breathable nanowood biofilms as guiding layer for green on-skin electronics. Small 2019;15(31):
1901079. https://doi.org/10.1002/smll.201901079.
[252] Zheng CX, Yue YY, Gan L, Xu XW, Mei CT, Han JQ. Highly stretchable and self-healing strain sensors based on nanocellulose-supported graphene dispersed in
electro-conductive hydrogels. Nanomaterials (Basel) 2019;9(7):9070937. https://doi.org/10.3390/nano9070937.
[253] Zou WH, Sun DL, Wang ZH, Li RY, Yu WX, Zhang PF. Eco-friendly transparent poplar-based composites that are stable and flexible at high temperature. RSC
Adv 2019;9(37):21566–71. https://doi.org/10.1039/c9ra03550h.
[254] Yang L, Wu Y, Yang F, Wang W. Study on the preparation process and performance of a conductive, flexible, and transparent wood. J Mater Res Technol 2021;
15:5396–404. https://doi.org/10.1016/j.jmrt.2021.11.021.
[255] Jia C, Li T, Chen CJ, Dai JQ, Kierzewski IM, Song JW, et al. Scalable, anisotropic transparent paper directly from wood for light management in solar cells.
Nano Energy 2017;36:366–73. https://doi.org/10.1016/j.nanoen.2017.04.059.
[256] Zhang C, Lin T, Yin X, Wu X, Wei X. Preparation of transparent wood containing carbon dots for application in the field of white-LED. J Wood Chem Technol
2022:1–11. https://doi.org/10.1080/02773813.2022.2085749.
[257] Zhu KP, Lu Z, Cong S, Cheng G, Ma P, Lou Y, et al. Ultraflexible and lightweight bamboo-derived transparent electrodes for perovskite solar cells. Small 2019;
15(33):1902878. https://doi.org/10.1002/smll.201902878.
[258] Zhou C, Julianri I, Wang S, Chan SH, Li M, Long Y. Transparent bamboo with high radiative cooling targeting energy savings. ACS Mater Lett 2021;3(6):883–8.
https://doi.org/10.1021/acsmaterialslett.1c00272.
[259] Wang Y-Y, Guo F-L, Li Y-Q, Zhu W-B, Li Y, Huang P, et al. High overall performance transparent bamboo composite via a lignin-modification strategy. Compos
Part B-Eng 2022;235:109798. https://doi.org/10.1016/j.compositesb.2022.109798.
[260] Peng C, Zhong J, Ma X, Huang A, Chen G, Luo W, et al. Transparent, hard-wearing and bio-based organic/silica hybrid coating for bamboo with enhanced
flame retardant and antifungal properties. Prog Org Coat 2022;167:106830. https://doi.org/10.1016/j.porgcoat.2022.106830.
[261] Chen CJ, Li ZH, Mi RY, Dai JQ, Xie H, Pei Y, et al. Rapid processing of whole bamboo with exposed, aligned nanofibrils toward a high-performance structural
material. ACS Nano 2020;14(5):5194–202. https://doi.org/10.1021/acsnano.9b08747.
[262] Li ZH, Chen CJ, Mi RY, Gan WT, Dai JQ, Jiao ML, et al. A strong, tough, and scalable structural material from fast-growing bamboo. Adv Mater 2020;32(10):
1906308. https://doi.org/10.1002/adma.201906308.
[263] Liu C, Hong K, Sun X, Natan A, Luan PC, Yang Y, et al. An ‘antifouling’ porous loofah sponge with internal microchannels as solar absorbers and water pumpers
for thermal desalination. J Mater Chem A 2020;8(25):12323–33. https://doi.org/10.1039/D0TA03872E.
[264] Wu Y, Wang YJ, Yang F, Wang J, Wang XH. Study on the properties of transparent bamboo prepared by epoxy resin impregnation. Polymers 2020;12(4):863.
https://doi.org/10.3390/polym12040863.
[265] Dong XY, Si Y, Chen CJ, Ding B, Deng HB. Reed leaves inspired silica nanofibrous aerogels with parallel-arranged vessels for salt-resistant solar desalination.
ACS Nano 2021;15(7):12256–66. https://doi.org/10.1021/acsnano.1c04035.
[266] Wang X, Shan SY, Shi SQ, Zhang YL, Cai LP, Smith LM. Optically transparent bamboo with high strength and low thermal conductivity. ACS Appl Mater
Interfaces 2021;13(1):1662–9. https://doi.org/10.1021/acsami.0c21245.
[267] Bian H, Chen L, Gleisner R, Dai H, Zhu JY. Producing wood-based nanomaterials by rapid fractionation of wood at 80◦ C using a recyclable acid hydrotrope.
Green Chem 2017;19(14):3370–9. https://doi.org/10.1039/C7GC00669A.

Sailing Zhu is currently working toward her Ph.D. under the supervision of Prof. Jingquan Han in College of Materials Science and Engineering at
Nanjing Forestry University (China). Her research interests revolve around the wood-based composite materials, including their processing, char­
acterization, along with integration in flexible and wearable devices.

Subir Kumar Biswas received his Ph.D. (2019) in agricultural science (functional bio-based materials) from Kyoto University (Japan). Currently, he is
a postdoctoral researcher in Prof. Hiroyuki Yano’s group at Kyoto University. His interests cover wood nanotechnology including nanocelluloses and
the development of structurally hierarchical (nano)composites via hydrophilic-hydrophobic interaction of the components.

Zhe Qiu is currently pursuing his Ph.D. under the supervision of Prof. Yanjun Xie in College of Materials Science and Engineering at Northeast Forestry
University (China). Now, he is studying as a visiting student in Feng Jiang’s group in Department of Wood Science at the University of British
Columbia. His research interests focus on the wood-based functional materials, including wood modification, transparent wood, utilization of wood
structure and wood-based bionic materials.

28
S. Zhu et al. Progress in Materials Science 132 (2023) 101025

Yiying Yue is currently an associate professor and a master supervisor in College of Biology and the Environment at Nanjing Forestry University
(China). She received her Ph.D. in Renewable Natural Resources at Louisiana State University (USA) in 2015. Her research interests cover environ­
mental functional bio-based materials including photo-catalytic materials, filtering materials, oil/water separation materials, desalination materials
and intelligent detection materials etc.

Qiliang Fu received his PhD degree (2018) from the department of Fibre and Polymer Technology at KTH Royal Institute of Technology. He was
employed as a scientist at Scion (New Zealand Forest Research Institute Limited) in 2018-2022. Currently, he is a full professor in College of Materials
Science and Engineering at Nanjing Forestry University. Dr Fu carries out his independent research and focuses on wood nanotechnologies and
nanoscience, as well as wood-based nanocomposites and cellulosic functional nanomaterials.

Feng Jiang received his PhD degree in Macromolecular Science and Engineering from Virginia Tech in 2011. He is currently Assistant Professor in the
Department of Wood Science at the University of British Columbia, and Tier II Canada Research Chair in Sustainable Functional Biomaterials. He
serves as Editor of Carbohydrate Polymers and is the 2021 recipient of ACS CELL KINGFA Young Investigator Award. His research areas focus on
sustainable isolation and modification of lignocellulosic nanomaterials, as well as advanced manufacturing of bio-based products developments for
emerging thermal, environmental, electrical, and energy applications.

Jingquan Han received his Ph.D. in renewable natural resources from Louisiana State University (USA) in 2014. He is currently a Professor and Ph.D.
supervisor in College of Materials Science and Engineering at Nanjing Forestry University (China). His current research interests focus on the biomass-
based soft nanocomposite materials with sustainability, renewability and biocompatibility, including gels, elastomers, membranes, macro/nanofibers,
and their advanced applications in multifunctional wearable electronic devices.

29

You might also like