Download as pdf or txt
Download as pdf or txt
You are on page 1of 28

Methacrylated Kraft lignin as a double-edged sword:

a quantitative analysis of Kraft lignin-containing

objects printed by stereolithography

Stephen C. Chmely,1* Mostafa Rohi Gal,1 Aditi Arya,1 Wei-Shu Lin,1 Gamini P. Mendis2

1
Department of Agricultural and Biological Engineering, The Pennsylvania State University,

University Park, PA 16802 USA

2
School of Engineering, Penn State Behrend, Erie, PA 16563 USA

KEYWORDS Kraft, lignin, 3D printing, stereolithography, additive manufacturing

ABSTRACT Finding an alternative to fossil-based resins which is renewable, sustainable, and

affordable is an important step to enable large-scale industrial manufacturing using

stereolithography. Lignin has shown some potential to fulfil these requirements. However,

examples that contain Kraft lignin, which is produced industrially at a rate of 70 million tons per

year, are relatively scarce compared to examples using organosolv lignin or lignin model

compounds. In this work, we fractionated and modified Kraft lignin and blended it with a

commercially available stereolithography resin produce 3D printed objects. We thoroughly

characterized the lignin, the resins, and the 3D printed objects, and conclude that while our

modification scheme allows for smooth incorporation of large amounts Kraft lignin into resins (up

https://doi.org/10.26434/chemrxiv-2023-klv3z ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
1
to 10% (w/w), or 10 times the amount previously demonstrated as possible), and the lignin itself

aids in the photopolymerization process (Ec = 662(2) mJ·cm-2 for the base case, Ec = 0.763(1)

mJ·cm-2 for resins containing 10% (w/w) lignin), the printed objects are weaker (σ = 20(1) MPa)

and more brittle (ε = 1.12(5)%) than the control that contains no Kraft lignin (σ = 46.3(5) MPa,

ε = 3.95(9)%). Together, these results demonstrate the importance of lignin modifications to create

miscible blends for stereolithography resins but highlight the need for further understanding how

these modifications can simultaneously enhance photo- and mechanical properties of printed

objects. These are an important first step towards creating 3D printable objects that contain

industrially relevant lignin.

https://doi.org/10.26434/chemrxiv-2023-klv3z ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
2
INTRODUCTION

Additive manufacturing by 3D printing is an effective and innovative way to fabricate materials

compared to conventional manufacturing processes. It allows users to efficiently produce materials

with complex structures.1 One of the major 3D printing techniques, stereolithography printing

(SLA) creates solid objects by free radical photopolymerization of liquid resins which are cured

layer by layer by UV light.2 However, a major hinderance for moving SLA printing to industrial

scale is the existence of inexpensive photoactive resins. Accordingly, a large body of research is

focused on developing new photoactive resins with interesting combinations of properties.3

The performance properties of liquid resins and printed objects are strongly correlated to the

components included in the resin formulation, which include reactive diluents, oligomers, UV

blockers, and photoinitiators. In addition, various carbon nanomaterials, including graphene oxide,

nanocellulose, and carbon nanotubes have been used as reinforcing additives to improve the

mechanical and thermal properties of the printed objects.4–6 Since many resulting combinations

are not environmentally or economically sustainable, a growing need exists to find sustainable

alternatives to mitigate the environmental concerns brought about by using fossil-based resources

for industrial 3D printing.7

Lignocellulosic materials are known as the most abundant and renewable materials in the world.

Lignin is the second most abundant natural polymer after cellulose,8 and it is produced industrially

as a byproduct in pulp and paper industries. Unfortunately, only 1-2% of 70 million tons of lignin

produced is used to generate valuable products,9 and the remainder is combusted to generate low-

value process heat.10 Rapidly emerging environmental issues related to non-renewable petroleum

products has brought lignin to the forefront as a precursor to high-value biorenewable products.11–
14

https://doi.org/10.26434/chemrxiv-2023-klv3z ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
3
Unsurprisingly, lignin has been widely investigated as a component in various 3D printing

precursors.5,15 Investigations focused on using lignin in multicomponent resins for 3D printing by

SLA or other vat-polymerization techniques are vastly outnumbered by those using other 3D

printing technologies.16 Some examples of well-characterized resins containing phenolic

monomers that could be isolated from lignin exist.17–19 Although these examples demonstrate the

potential utility of lignin for additive manufacturing, isolating discrete molecules (e.g. vanillin,

guaiacol, etc.) from heterogenous technical lignin (e.g. Kraft, organosolv, etc.) is exceptionally

challenging from both an engineering and economic standpoint, and itself is the focus of a

tremendous amount of research.20

Instead, incorporation of technical lignin, such as Kraft or organosolv, is emerging as a method

to valorize these industrially relevant lignin samples without the separation costs incurred through

isolating discrete monomers. Whole technical lignin can be modified directly by acylation with,

e.g., methacrylic anhydride, which allows it to be incorporated into photoactive resins for 3D

printing. Our group has recently demonstrated that both hardwood21 and softwood22 organosolv

lignin can be modified and incorporated into 3D printing resins, and these resins display enhanced

performance properties because of the lignin present.

Isolation of discrete small molecules perhaps represents the most challenging and expensive

route for lignin valorization via 3D printing. In contrast, direct incorporation of Kraft lignin (the

most widely available and industrially relevant technical lignin sample available today), without

chemical modification, represents the least technically challenging and (potentially) most

ecologically friendly option. However, results using unmodified technical lignin in 3D printing

resins are mixed.

https://doi.org/10.26434/chemrxiv-2023-klv3z ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
4
For example, cellulose nanomaterials coated with 3-6% (w/w) lignin have been shown to greatly

enhance the mechanical and thermal performance of 3D printed composites.23,24 The presence of

lignin in these composites likely increases the matrix-fiber interactions between hydrophilic

cellulose nanomaterials and hydrophobic SLA resin mixtures.25 On the other hand, attempts to

incorporate unmodified Kraft lignin into a commercial SLA resin led to only a modest increase in

mechanical properties, and the total amount of lignin contained in these resins did not exceed 1%

(w/w).26 Given our success with chemical modification of organosolv lignin for incorporation into

3D printer resins, coupled with the wide availability of industrial Kraft lignin, we set out to

investigate how chemical modifications of Kraft lignin would affect the performance properties of

objects 3D printed from resins that contain it.

In this contribution, we modified fractionated Kraft lignin using the methacrylation strategy we

employed for organosolv lignin. We then then blended this material with a commercial SLA resin

and measured performance properties of the resin and objects 3D printed from it. We have shown

that the modifications act as a kind of “double-edged sword,” whereby they allow for a much

greater amount of lignin to be incorporated into resins, and have potentially beneficial effects on

resin photoproperties, but the mechanical properties of objects printed from them do not enjoy the

same improvements as those printed using organosolv lignin. We speculate on the reasons for this

behavior and present some potential solutions below.

MATERIALS AND METHODS

Reagents

Indulin AT softwood Kraft lignin was provided by the WestRock Company, Atlanta, GA USA.

It was fully characterized as described below. SainSmart Rapid UV General Purpose

Photopolymer Curing Resin (405 nm, transparent) was purchased from Amazon.com, Inc. 2-

https://doi.org/10.26434/chemrxiv-2023-klv3z ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
5
chloro-4,4,5,5-tetramethyl-1,3,2-dioxaphospholane (TMDP), methacrylic anhydride

(MilliporeSigma, Burlington, MA USA), chloroform-d, and dimethyl sulfoxide-d6 (Cambridge

Isotope Laboratories, Inc., Tewksbury, MA USA) were stored in a desiccator over anhydrous

calcium sulfate prior to use. Ultrapure water (specific resistance, ρ = 18.18 MΩ⋅cm) was used for

all preparations. All other reagents were used as received from their respective vendor unless

otherwise noted.

Lignin fractionation and preparation

Acetone fractionation of Kraft lignin was conducted as described elsewhere.27 Briefly, the parent

technical lignin was separated into two fractions using a concentration of 10% (w/v) suspension

of lignin in acetone. This mixture was allowed to stir at ambient temperature for 24 h.

Subsequently, the dissolved (low molecular weight, MW) fraction was separated from the solid

(high MW) fraction by filtration through a cellulose filter paper. The filtrate was collected, and the

low-MW lignin isolated by evaporating the solvent in a rotatory evaporator followed by drying

under vacuum at ambient temperature for 24 h. The high-MW precipitate was rinsed with acetone

followed by drying under vacuum at ambient temperature for 24 h.

The as-fractionated low-MW lignin was suspended in 0.2 N H2SO4 and stirred at ambient

temperature for 1 h to remove extractives or other industrial impurities. The material was

subsequently filtered, suspended in water at 70 °C, shaken, and centrifuged three times to remove

any traces of sulfuric acid. The fractionated and washed low-MW lignin was used for all

subsequent experiments unless otherwise noted.

Lignin characterization

Gel permeation chromatography was performed using a 1260 Infinity High Temperature GPC

System (Agilent Technologies, Santa Clara, CA, USA) equipped with a refractive index (RI)

https://doi.org/10.26434/chemrxiv-2023-klv3z ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
6
detector. The instrument and reference cell flow rates were set to 1 mL/min and all analysis and

calibration were performed in tetrahydrofuran (THF, HPLC grade, Sigma Aldrich, St. Louis, MO,

USA). A PL gel Mixed-C column (5 µm particle size, 7.5 x 300 mm, Agilent Technologies) was

used to perform the separation. A 200-µL injection volume and a run time of 15 minutes were used

for all samples. The molecular weights were calibrated using polystyrene standards with a range

of molecular weights from 128 Da to 2 MDa (EasiVial PS Tri-Pack, Agilent Technologies, Santa

Clara, CA, USA) were used to calibrate the system. Lignin samples were dissolved in THF at 40

°C for 1 hour before performing analysis. All lignin samples were acetylated prior to analysis

following a procedure described elsewhere.28

Hydroxyl (OH) group content of the lignin samples was measured by phosphitylation using
31
TMDP followed by P NMR spectroscopy using a Bruker Avance 500 spectrometer. Samples

were prepared using ~30 mg lignin. Sample preparation, data collection, and processing were

performed as described elsewhere.29

FTIR spectra were obtained using a Bruker Vertix 70 spectrometer equipped with a liquid

nitrogen cooled MCT detector. All samples were run in attenuated total reflection at an incident

angle of 45 degrees. A total of 500 scans were averaged at 6 cm-1 resolution and absorbance was

calculated by referencing to the clean bare crystal.

Lignin modification

The protocol for lignin acylation was adjusted from our previous method applied for

modification of organosolv hybrid poplar lignin.21,22 Briefly, 5 g lignin (8.2 mmol hydroxyl groups

per g lignin, 41 mmol OH total) was suspended in 110 mL 1,4-dioxane. To this suspension was

added 8.7 mL methacrylic anhydride (59 mmol, 1.4 equiv) and 0.24 g dimethylamino pyridine

(DMAP, 1.96 mmol, catalytic). The resulting mixture was heated to 60 °C and stirred for 48 h,

https://doi.org/10.26434/chemrxiv-2023-klv3z ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
7
after which time it was allowed to cool to ambient temperature and quenched with saturated

aqueous sodium bicarbonate. The precipitated modified lignin material was isolated by

centrifugation and washed with water until the washing liquid had a neutral pH. The isolated

methacrylated lignin (M-lignin) was then dried at 40 °C under vacuum for 48 h and ground to a

friable powder using a mortar and pestle.

Resin formulation and characterization

M-lignin was weighed and added to a known amount of SainSmart Rapid UV general purpose

photopolymer curing resin to generate 30-g batches of composite resins containing 1, 2, 5, and

10% (w/w) M-lignin. The formulated resins were mixed in a Thinky ARM-310 planetary mixer

(THINKY USA, Laguna Hills, CA USA) at 2000 rpm for 30 min to obtain uniform mixtures. The

final formulated resins were transferred into firmly sealed opaque white polypropylene containers

to protect them from light prior to printing.

Rheological properties of the formulated resins were measured using a TA Instruments

Discovery HR-3 hybrid rheometer (TA Instruments, New Castle, DE USA) with a cone and plate

geometry. The rheometer was configured with a 40-mm diameter cone at a truncation gap of

48 µm. Tests were performed by varying the shear rate from 0 to 100 Hz at 20 °C.

3D printing and mechanical testing

All objects were printed using an Elegoo Mars 2 PRO Mono LCD MSLA Resin 3D Printer

(Elegoo, Inc., Shenzhen, Guangdong, China). The build platform was calibrated according to

instructions provided by the manufacturer before each print. Printed objects included working

curve windowpanes21,22 and ASTM D638 Type-IV tensile testing specimens.30 CHITUBOX

(CBD-Tech, Shenzhen, Guangdong China) was used to set up the printing parameters for each

object.

https://doi.org/10.26434/chemrxiv-2023-klv3z ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
8
All printed objects were soaked in an acetone bath at ambient temperature for 2 min to ensure

all uncured resin was removed. After drying with a stream of compressed air, objects were post-

cured with a 400-W metal halide UV lamp (Uvitron International, Inc., West Springfield, MA

USA) with an irradiance of 200 mW/cm2. Objects were irradiated for a total of 1 minute and

inverted halfway through the process to ensure complete curing of both sides.

Ultimate tensile strength, Young’s modulus, and strain at break of the printed and post-cured

objects were measured using an Instron 5567 dual column universal testing machine (Instron,

Norwood, MA USA). A 30-kN static load cell was used according to the ASTM standard D63830

(Type IV specimen dimensions) with a 1 mm/min extension rate for all tests. Each property was

measured and reported as an average of at least 4 replications.

Statistical analyses

All measurements are reported as the mean of at least 4 replications. However, given the small

number of samples collected, a non-parametric Kruskal-Wallis one-way analysis of variance was

employed to determine if any statistical difference exists between the groups followed by a post-

hoc Mann-Whitney U test (95% confidence level) of the null hypothesis.31

RESULTS AND DISCUSSION

Lignin preparation, modification, and characterization

Given that Indulin AT is a softwood Kraft lignin from an industrial pulping process, we wanted

to ensure that we were working with clean material that was free of residual sugars or other process

reagent residues. In addition, we were interested in isolating low-MW material so that it would be

maximally soluble in the 3D printer resin that we chose for these studies. Accordingly, we used a

previously reported molecular weight fractionation using acetone, followed by a washing step

https://doi.org/10.26434/chemrxiv-2023-klv3z ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
9
using dilute sulfuric acid. This fractionation process allowed us to isolate 20% (w/w) of the lignin

as low-MW material.

We used size exclusion chromatography to analyze the molecular weight distribution of the

lignin samples before and after fractionation, the results of which are summarized in Table 1.

Table 1. Molecular weight distributions of parent Kraft and fractionated lignins.

Sample Mn (g/mol) Mw (g/mol) Ð

Kraft lignin 2684 3257 1.213

High-MW lignin 2717 3313 1.219

Low-MW lignin 1959 2135 1.09

As shown in Table 1, fractionation of Kraft lignin using acetone produces a lignin fraction with

a lower weight- and number-average molecular weight and a tighter dispersity. This is an important

result, since lower molecular weight lignin typically has a higher solubility in conventional organic

solvents, due primarily to the lower degree of branching and chain entanglement occurring in low-

MW lignin.32 We hypothesized that this enhanced solubility would be beneficial regarding

dispersion of the suitably modified lignin material in photoactive hydrophobic 3D printer resins.

Accordingly, we modified this low-MW lignin using methacrylic anhydride as demonstrated in

Scheme 1 below.

Scheme 1. Modification of softwood Kraft (primarily guaiacyl-type) lignin with methacrylic

anhydride.

https://doi.org/10.26434/chemrxiv-2023-klv3z ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
10
The modification using methacrylic anhydride necessarily produces methacrylic acid as a

byproduct of this reaction (Scheme 1, red molecules). This material needs to be removed by

neutralization and thorough washing, although raising the pH of lignin suspensions could also

catalyze the dissolution of lignin. Moreover, the atom economy of such a modification, which

produces a mole of acidic byproduct for every mole of hydroxyl group substituted, is not

particularly attractive for an industrial process. Still, recycling of this material and reformation of

the anhydride from it could be relatively straightforward and would have a beneficial effect on the

carbon footprint of this process. These investigations are ongoing in our laboratory.

Because the modification reaction targets aliphatic and aromatic hydroxyl groups of lignin, 31P

NMR spectroscopy and FTIR spectroscopy would be ideally suited to characterize the resulting

modified Kraft lignin (M-lignin). Figure 1 summarizes the results of these spectroscopic

investigations.

https://doi.org/10.26434/chemrxiv-2023-klv3z ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
11
Figure 1. (a) Fourier-transform infrared (FTIR) spectra of Kraft lignin (top, brown), fractionated

and rinsed low-MW Kraft lignin (middle, blue), and methacrylate modified M-lignin (bottom,
31
green). (b) P NMR spectra of the same samples (color scheme and position identical) with

integration ranges.

The FTIR spectra of the various lignins (Figure 1a) reveals a reduction in intensity of the peak

associated with –OH groups in lignin (3404 cm-1). This is due to the substitution of the

methacrylate groups at free hydroxyl groups as depicted in Scheme 1 above. Moreover, the FTIR

spectra show increasing intensities of peaks associated with C=O bond stretches (ν = 1723 cm-1),

C–O–C asymmetric stretches (ν = 1131 cm-1), and in-plane =CH2 scissoring bends (δ = 943 cm-1)

in M-lignin, all associated with the addition of the methacrylate moiety onto lignin.21

In addition, 31P NMR spectroscopy is a method used to quantify the amount of hydroxyl groups

and can be used to determine differences among substitutions at the aliphatic and aromatic

hydroxyl groups in lignin. As expected for softwood lignin, the 31P NMR spectrum (Figure 1b)

shows an abundance of G-type aromatic hydroxyl groups, with a vanishingly small amount of

groups associated with S- and H-type lignin.33 We measured a total hydroxyl group content of 8.2

https://doi.org/10.26434/chemrxiv-2023-klv3z ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
12
mmol OH per g lignin, which we used to calculate the amount of methacrylic anhydride to achieve

complete substitution.

We achieved transformation of the hydroxyl groups using methacrylic anhydride in the presence

of DMAP in 1,4-dioxane. The NMR spectra demonstrate that the degree of substitution we

achieved was almost 98%. We also note a relatively large peak associated with –COOH groups

(Figure 1, asterisk) and attributable to methacrylic acid (Scheme 1, red molecules) that we were

unable to rinse with successive sodium bicarbonate washes, which we have noted previously for

other organosolv lignins.21,22

We also confirmed the installation of methacrylate groups onto low-MW lignin using 1H-13C

HSQC NMR spectroscopy. The geminal vinylic protons associated with the installed C=C

moieties (which are not native to technical lignin) are clearly shown in Figure 2 below. We have
1
tentatively assigned these peaks by simulating a H-13C HSQC NMR spectrum of

permethacrylated vanillyl alcohol (molecule shown in Figure 2 inset) using software provided at

nmrdb.org.34–37

https://doi.org/10.26434/chemrxiv-2023-klv3z ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
13
Figure 2. Experimental 1H-13C NMR spectrum of M-lignin showing signals associated with the

geminal protons of installed methacrylate moieties which are not native to technical lignin. Peak

assignments were made by simulating the spectrum of the permethacrylated vanillyl alcohol

molecule shown.

Stereolithography resin formulation and characterization

We blended M-lignin with a commercial clear SLA resin at different loading percentages. Given

the viscosity of the existing commercial resin, we employed a benchtop kinetic mixer to

accomplish this blending procedure. The powdery lignin sample smoothly blended with the

viscous resin after no more than 30 min stirring in the kinetic mixer. The resins appear as relatively

homogenous brown mixtures as shown in Figure 3 below.

https://doi.org/10.26434/chemrxiv-2023-klv3z ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
14
Figure 3. Photograph of a representative sample of a lignin-commercial resin blend after mixing.

The sample appears to be a homogenous brown mixture with no apparent undissolved particles.

However, although the resins appear homogenous, we were interested in some quantitative

measures of their homogeneity. Resin viscosity is a crucial parameter in stereolithography. Resins

with low viscosity will flow easily during the printing process, which depends on a moving build

platform through the liquid resin bath. In contrast, resins with too high viscosity can affect printing

speed and resolution. Figure 4 summarizes the rheological properties of the lignin-containing

resins.

https://doi.org/10.26434/chemrxiv-2023-klv3z ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
15
Figure 4. Flow curves (A) and viscosity curves (B) of the commercial base resin (gray squares)

and lignin-containing resins containing 1% (green), 2% (pink), 5% (orange), and 10% (blue) (w/w)

lignin. The insets (highlighted in green) depict the areas between 10-30 Hz, which is a shear range

common for many desktop SLA printers.38

At low shear rates, the base resin behaves as Newtonian fluid, given it is a mixture of relatively

small molecular weight materials. In addition, resins containing up to 5% (w/w) lignin also exhibit

Newtonian behavior, although their viscosity is higher overall due to the presence of lignin. Most

importantly, in the shear rate regime of 10-30 Hz (Figure 4 green shaded region), common for

desktop SLA printers,38 the Newtonian behavior of these resins combined with their viscosities

that do not exceed 5 Pa∙s (a common upper limit for SLA resin viscosity)25 leads us to conclude

these resins would be suitable for 3D printing by stereolithography.

In the case of the resin containing 10% (w/w) lignin, the resulting non-Newtonian colloidal

suspension behaves as a jammed “fragile matter” system.39 In a jammed system, at low shear rates,

an apparent yield stress must be overcome to initiate flow. Once this apparent yield stress is

https://doi.org/10.26434/chemrxiv-2023-klv3z ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
16
overcome, the viscosity drops rapidly (shear thinning). However, at high shear rates, the viscosity

rises and potentially plateaus (shear thickening).40 This behavior is ubiquitous in colloidal

suspensions containing a high concentration of nonaggregating solid particles.41 These data imply

that at high concentrations, our M-lignin is perhaps not as uniformly dispersed as is apparent upon

visual inspection of the resins.

Even so, in the 10-30 Hz regime, the shear thinning behavior and acceptable viscosity of the

10% (w/w) suspension suggest the material is still printable. However, a common method to

shorten print times is to shorten lift and retract speeds of the printer, which are the speeds at which

the build platform moves away from and towards the bottom of the resin bath between layers,

respectively. If that speed is too fast, the shear rates experienced by the resin may exceed 30 Hz.

In this case, the shear thickening behavior of the resin might cause the lignin particles to flocculate,

which could create discontinuities in printed objects and detrimentally affect their performance

properties.

Next, we characterized the curing parameters of the resins. Although the curing parameters are

inherent to the liquid resins, measurement of a working curve that indicates penetration depth and

critical curing energy is accomplished by printing a windowpane object. These curves were

constructed as we have described previously, and the results of these measurements are

summarized in Figure 5 below.

https://doi.org/10.26434/chemrxiv-2023-klv3z ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
17
Figure 5. Plots of penetration depth (A) and critical curing energy (B) versus lignin loading (color

scheme matches Figure 4). Each measurement was performed 3 times by printing 3 separate

windowpane specimens, and the means (markers and values) and standard errors of the mean (error

bars and parenthetical values) are shown. Values in parentheses are the standard error of the

preceding digit. In some cases, error bars are smaller than the size of the markers and are thus

obscured. Non-parametric statistical analysis (Kruskal-Wallis H test followed by Mann-Whitney

U test) was used to determine statistical differences among the samples (letters a-d); samples that

are not statistically different share letters. The dotted lines are not mathematically meaningful and

are provided to guide the eye.

The penetration depth of curing light in a photosensitive resin (such as those used for SLA 3D

printing) follows a Beer-Lambert relation and is defined as the depth at which the intensity of the

https://doi.org/10.26434/chemrxiv-2023-klv3z ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
18
penetrating light falls to e-1 the intensity of the light at the surface of the resin.42 At a basic level,

it stands to reason that the penetration depth of a given resin will be lower if the material is darker.

This phenomenon is depicted in Figure 5a, whereby the measured penetration depth decreases with

additional brown lignin. Although the exact cause for this phenomenon is still unknown, we posit

that the electronic response of the resin material (governed by the sum of the electronic structures

of all the components) changes substantially upon addition of lignin. Since lignin contains

delocalized π-electrons that give rise to its light absorptive properties (i.e., lignin is brown43),

changes in reflectance and absorption in the resin due to lignin’s electronic structure necessarily

reduce the penetration depth of curing light.44

Results related to the critical curing energy displayed in Figure 5b are somewhat surprising.

Critical curing energy corresponds to the minimum energy required to activate the photoinitiator

and initiate the polymerization reaction. One might expect an inverse relationship, that as

penetration depth decreases, critical curing energy should increase: fewer photons find

photoinitiator molecules deep within the resin, and therefore more overall are required to initiate

a polymerization reaction. However, as shown in Figure 5b, as the amount of lignin increases, the

critical curing energy decreases, in fact by 2-3 orders of magnitude depending on the amount of

lignin used.

This result surprised us, but we have developed a tentative explanation. Technical Kraft lignin

contains several aromatic moieties that are conjugated with aliphatic ketones, which is a result of

oxidizing conditions during the pulping process that convert aliphatic –OH groups to ketones (or

aldehydes or carboxylic acids). As such, lignin can act as a Norrish Type-II photoinitiator, which

is a class of molecules known to absorb in the near-UV-to-visible range (our 3D printer emits at

405 nm) and form energetic species that can abstract protons from neighboring molecules to form

https://doi.org/10.26434/chemrxiv-2023-klv3z ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
19
radicals and initiate photopolymerizations.45 While our acetylation method converts nearly all the

free –OH groups in lignin to methacrylate esters, there are some phenolic –OH groups that remain

even after conversion (see Figure 1b above). Under the right conditions, these could be converted

to quinones during photoinitiation, which are chemical species known to be Norrish Type-II

photoinitiators.46–48 In effect, the lignin is acting as additional photoinitiator, thereby aiding in the

polymerization reaction. This is an exciting result, as if it could be controlled, lignin could be used

to alter the light absorption properties of 3D printer resins by controlling its electronic structure

via chemical modifications. These investigations are ongoing in our laboratory.

Printed objects and mechanical testing

With lignin-containing resins in hand, we were interested in probing what effect the modified

lignin had on the mechanical properties of printed objects. We printed tensile testing specimens

using a desktop SLA printer and subjected the resulting specimens to static tensile testing to

measure tensile strength, elongation, and Young’s modulus. The results of these measurements are

summarized in Figure 6 below.

https://doi.org/10.26434/chemrxiv-2023-klv3z ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
20
Figure 6. Plots of ultimate tensile strength (A), strain at break (B), and Young’s modulus (C) of

printed objects versus lignin loading (color scheme matches Figures 4 and 5). Each measurement

was performed at least 4 times by printing separate tensile testing specimens. Types of information

displayed and statistical analyses are identical to Figure 5.

The tensile tests we performed indicate that the tensile strength of the commercial resin

decreases upon addition of lignin, with a statistically significant decrease of 37% upon addition of

5% (w/w) lignin, and a decrease of nearly 57% upon addition of 10% (w/w) lignin (Figure 6a).

We attribute this to the relatively low molecular weight of the lignin added. As the percentage of

low molecular weight material increases as a function of the total mixture, the intermolecular

https://doi.org/10.26434/chemrxiv-2023-klv3z ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
21
interactions between chains become less pronounced, thereby leading to an overall decrease in

strength.

Furthermore, the strain at break of the commercial resin decreases upon addition of lignin, with

a statistically significant 74% decrease in strain at break for the sample containing 10% (w/w)

lignin (Figure 6b). In general, highly crosslinked photopolymers are relatively brittle. Since the

lignin oligomers we add could each be multifunctional (each having multiple –OH groups

acetylated with methacrylic esters), addition of lignin would substantially increase the crosslink

density of the resulting printed objects. In that case, the objects would become much more brittle

as depicted in Figure 6.

Moreover, we noted an overall decrease in Young’s modulus upon addition of lignin, with a

statistically significant decrease of 13% for the sample containing 10% (w/w) lignin (Figure 6c).

We attribute this decrease in flexibility to the crosslinking phenomenon described above. As

crosslink density increases with lignin content, polymer chains are less able to slide past one

another in the printed object, and the overall flexibility of the object is diminished.

CONCLUSIONS

We have reported on a method to create Kraft-lignin-containing stereolithography resins that

can be 3D printed using a desktop 3D printer. Fractionation of Kraft lignin using acetone afforded

a low-MW fraction that we were able to acetylate with methacrylate ester groups nearly

quantitatively. This low-MW, highly substituted material allowed us to create new resins from

commercial SLA resins with ten times the amount of Kraft lignin previously reported. The resins

were relatively homogenous, except at the highest lignin loading, and we were able to successfully

print tensile testing specimens. Working curve measurements suggest that the lignin participates

in the photopolymerization process, perhaps by absorbing light and generating active species that

https://doi.org/10.26434/chemrxiv-2023-klv3z ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
22
can then react to form radicals for polymerization. However, although we chose to incorporate

fractionated low-MW lignin to enhance its overall solubility in the resin materials, these low-MW

oligomers likely contribute to a decrease in strength to the printed objects. We were able to include

a much larger overall percentage of Kraft lignin in our samples than previously reported, but at the

expense of the tensile strength of the printed objects. In addition, our acetylation method to add

methacrylate esters to lignin probably enhanced its interactions with the resulting photopolymer

through covalent interactions. However, the increase in crosslink density imparted by addition of

multifunctional lignin oligomers increased the brittleness of the objects and decreased their

flexibility. These results are simultaneously promising and frustrating and they highlight the need

to further our fundamental understanding of how lignin behaves in photopolymer resins. These

investigations are ongoing in our laboratory.

https://doi.org/10.26434/chemrxiv-2023-klv3z ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
23
Corresponding Author

* E-mail: sc411@psu.edu

ORCID

Stephen C. Chmely: 0000-0002-2637-9974

Author Contributions

The manuscript was written through contributions of all authors. SCC conceived and designed

the experiments, obtained funding, interpreted data, and wrote the manuscript. WSL collected

rheology data for the resins. GPM collected GPC data for the lignin samples. MRG and AA

collected data related to 3D printing and all performed the remaining characterizations. All

authors have edited and given approval to the final version of the manuscript.

Funding Sources

This work was supported by funding from the United States Department of Agriculture USDA-

NIFA Project PEN04671 and Accession number 1017582. SCC and MRG also acknowledge

support from USDA-NIFA under grant number: 2020-68012-31881. The findings and conclusions

in publication have not been formally disseminated by the United States Department of Agriculture

and should not be construed to represent any agency determination or policy.

Notes

The authors declare no competing financial interest.

ACKNOWLEDGMENT

The authors wish to acknowledge the Penn State Materials Characterization Lab for use of the

Bruker Vertix 70 FTIR spectrometer. We also wish to acknowledge the support of Dr. Christy

https://doi.org/10.26434/chemrxiv-2023-klv3z ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
24
George and Dr. Tapas Mal for their help with collecting NMR data, and the support of Jerry

Magraw for his help with collecting GPC data.

REFERENCES

(1) Gebhardt, A. Basics, Definitions, and Application Levels. In Understanding Additive


Manufacturing; Gebhardt, A., Ed.; Hanser, 2011; pp 1–29.
https://doi.org/10.3139/9783446431621.001.
(2) Chaudhary, R.; Fabbri, P.; Leoni, E.; Mazzanti, F.; Akbari, R.; Antonini, C. Additive
Manufacturing by Digital Light Processing: A Review. Prog. Addit. Manuf. 2023, 8 (2), 331–
351. https://doi.org/10.1007/s40964-022-00336-0.
(3) Voet, V. S. D.; Guit, J.; Loos, K. Sustainable Photopolymers in 3D Printing: A Review on
Biobased, Biodegradable, and Recyclable Alternatives. Macromol. Rapid Commun. 2021, 42
(3), 2000475. https://doi.org/10.1002/marc.202000475.
(4) Farahani, R. D.; Dubé, M.; Therriault, D. Three-Dimensional Printing of Multifunctional
Nanocomposites: Manufacturing Techniques and Applications. Adv. Mater. 2016, 28 (28),
5794–5821. https://doi.org/10.1002/adma.201506215.
(5) Yang, J.; An, X.; Liu, L.; Tang, S.; Cao, H.; Xu, Q.; Liu, H. Cellulose, Hemicellulose, Lignin,
and Their Derivatives as Multi-Components of Bio-Based Feedstocks for 3D Printing.
Carbohydr. Polym. 2020, 250, 116881. https://doi.org/10.1016/j.carbpol.2020.116881.
(6) Wang, Q.; Sun, J.; Yao, Q.; Ji, C.; Liu, J.; Zhu, Q. 3D Printing with Cellulose Materials.
Cellulose 2018, 25 (8), 4275–4301. https://doi.org/10.1007/s10570-018-1888-y.
(7) Maines, E. M.; Porwal, M. K.; Ellison, C. J.; Reineke, T. M. Sustainable Advances in
SLA/DLP 3D Printing Materials and Processes. Green Chem. 2021, 23 (18), 6863–6897.
https://doi.org/10.1039/D1GC01489G.
(8) Glasser, W. G. About Making Lignin Great Again—Some Lessons From the Past. Front.
Chem. 2019, 7.
(9) Fernando, S.; Adhikari, S.; Chandrapal, C.; Murali, N. Biorefineries: Current Status,
Challenges, and Future Direction. Energy Fuels 2006, 20 (4), 1727–1737.
https://doi.org/10.1021/ef060097w.
(10) Kai, D.; Tan, M. J.; Chee, P. L.; Chua, Y. K.; Yap, Y. L.; Loh, X. J. Towards Lignin-Based
Functional Materials in a Sustainable World. Green Chem. 2016, 18 (5), 1175–1200.
https://doi.org/10.1039/C5GC02616D.
(11) Upton, B. M.; Kasko, A. M. Strategies for the Conversion of Lignin to High-Value
Polymeric Materials: Review and Perspective. Chem. Rev. 2016, 116 (4), 2275–2306.
https://doi.org/10.1021/acs.chemrev.5b00345.
(12) Llevot, A.; Grau, E.; Carlotti, S.; Grelier, S.; Cramail, H. From Lignin-Derived Aromatic
Compounds to Novel Biobased Polymers. Macromol. Rapid Commun. 2016, 37 (1), 9–28.
https://doi.org/10.1002/marc.201500474.
(13) Isikgor, F. H.; Becer, C. R. Lignocellulosic Biomass: A Sustainable Platform for the
Production of Bio-Based Chemicals and Polymers. Polym. Chem. 2015, 6 (25), 4497–4559.
https://doi.org/10.1039/C5PY00263J.

https://doi.org/10.26434/chemrxiv-2023-klv3z ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
25
(14) Ten, E.; Vermerris, W. Functionalized Polymers from Lignocellulosic Biomass: State of
the Art. Polymers 2013, 5 (2), 600–642. https://doi.org/10.3390/polym5020600.
(15) Sabbatini, B.; Cambriani, A.; Cespi, M.; Palmieri, G. F.; Perinelli, D. R.; Bonacucina, G.
An Overview of Natural Polymers as Reinforcing Agents for 3D Printing. ChemEngineering
2021, 5 (4), 78. https://doi.org/10.3390/chemengineering5040078.
(16) Ebers, L.-S.; Arya, A.; Bowland, C. C.; Glasser, W. G.; Chmely, S. C.; Naskar, A. K.;
Laborie, M.-P. 3D Printing of Lignin: Challenges, Opportunities and Roads Onward.
Biopolymers 2021, 112 (6), e23431. https://doi.org/10.1002/bip.23431.
(17) Ding, R.; Du, Y.; Goncalves, R. B.; Francis, L. F.; Reineke, T. M. Sustainable near UV-
Curable Acrylates Based on Natural Phenolics for Stereolithography 3D Printing. Polym.
Chem. 2019, 10 (9), 1067–1077. https://doi.org/10.1039/C8PY01652F.
(18) Chin, K. C. H.; Cui, J.; O’Dea, R. M.; Epps, T. H. I.; Boydston, A. J. Vat 3D Printing of
Bioderivable Photoresins – Toward Sustainable and Robust Thermoplastic Parts. ACS Sustain.
Chem. Eng. 2023, 11 (5), 1867–1874. https://doi.org/10.1021/acssuschemeng.2c06313.
(19) Bassett, A. W.; Honnig, A. E.; Breyta, C. M.; Dunn, I. C.; La Scala, J. J.; Stanzione, J. F.
I. Vanillin-Based Resin for Additive Manufacturing. ACS Sustain. Chem. Eng. 2020, 8 (14),
5626–5635. https://doi.org/10.1021/acssuschemeng.0c00159.
(20) Abu-Omar, M. M.; Barta, K.; Beckham, G. T.; Luterbacher, J. S.; Ralph, J.; Rinaldi, R.;
Román-Leshkov, Y.; Samec, J. S. M.; Sels, B. F.; Wang, F. Guidelines for Performing Lignin-
First Biorefining. Energy Environ. Sci. 2021, 14 (1), 262–292.
https://doi.org/10.1039/D0EE02870C.
(21) Sutton, J. T.; Rajan, K.; Harper, D. P.; Chmely, S. C. Lignin-Containing Photoactive Resins
for 3D Printing by Stereolithography. ACS Appl. Mater. Interfaces 2018, 10 (42), 36456–
36463. https://doi.org/10.1021/acsami.8b13031.
(22) Sutton, J. T.; Rajan, K.; Harper, D. P.; Chmely, S. C. Improving UV Curing in Organosolv
Lignin-Containing Photopolymers for Stereolithography by Reduction and Acylation.
Polymers 2021, 13 (20), 3473. https://doi.org/10.3390/polym13203473.
(23) Feng, X.; Yang, Z.; Chmely, S.; Wang, Q.; Wang, S.; Xie, Y. Lignin-Coated Cellulose
Nanocrystal Filled Methacrylate Composites Prepared via 3D Stereolithography Printing:
Mechanical Reinforcement and Thermal Stabilization. Carbohydr. Polym. 2017, 169, 272–
281. https://doi.org/10.1016/j.carbpol.2017.04.001.
(24) Feng, X.; Yang, Z.; Wang, S.; Wu, Z. The Reinforcing Effect of Lignin-Containing
Cellulose Nanofibrils in the Methacrylate Composites Produced by Stereolithography. Polym.
Eng. Sci. 2022, 62 (9), 2968–2976. https://doi.org/10.1002/pen.26077.
(25) Battisto, E. W.; Sarsfield, S. R.; Lele, S. R.; Williams, T.; Catchmark, J. M.; Chmely, S.
C. Enhancing the Matrix–Fiber Interface with a Surfactant Leads to Improved Performance
Properties of 3D Printed Composite Materials Containing Cellulose Nanofibrils. ACS Appl.
Mater. Interfaces 2022, 14 (39), 44841–44848. https://doi.org/10.1021/acsami.2c12363.
(26) Zhang, S.; Li, M.; Hao, N.; Ragauskas, A. J. Stereolithography 3D Printing of Lignin-
Reinforced Composites with Enhanced Mechanical Properties. ACS Omega 2019, 4 (23),
20197–20204. https://doi.org/10.1021/acsomega.9b02455.
(27) Karaaslan, M. A.; Cho, M.; Liu, L.-Y.; Wang, H.; Renneckar, S. Refining the Properties
of Softwood Kraft Lignin with Acetone: Effect of Solvent Fractionation on the
Thermomechanical Behavior of Electrospun Fibers. ACS Sustain. Chem. Eng. 2021, 9 (1),
458–470. https://doi.org/10.1021/acssuschemeng.0c07634.

https://doi.org/10.26434/chemrxiv-2023-klv3z ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
26
(28) Guerra, A.; Lucia, L. A.; Argyropoulos, D. S. Isolation and Characterization of Lignins
from Eucalyptus Grandis Hill Ex Maiden and Eucalyptus Globulus Labill. by Enzymatic Mild
Acidolysis (EMAL). 2008, 62 (1), 24–30. https://doi.org/10.1515/HF.2008.004.
(29) Bozell, J. J.; O’Lenick, C. J.; Warwick, S. Biomass Fractionation for the Biorefinery:
Heteronuclear Multiple Quantum Coherence–Nuclear Magnetic Resonance Investigation of
Lignin Isolated from Solvent Fractionation of Switchgrass. J. Agric. Food Chem. 2011, 59
(17), 9232–9242. https://doi.org/10.1021/jf201850b.
(30) D20 Committee. Test Method for Tensile Properties of Plastics; ASTM International.
https://doi.org/10.1520/D0638-14.
(31) Smalheiser, N. R. Chapter 12 - Nonparametric Tests. In Data Literacy; Smalheiser, N. R.,
Ed.; Academic Press, 2017; pp 157–167. https://doi.org/10.1016/B978-0-12-811306-6.00012-
9.
(32) Sameni, J.; Krigstin, S.; Sain, M. Solubility of Lignin and Acetylated Lignin in Organic
Solvents. BioResources 2017, 12 (1), 1548–1565. https://doi.org/10.15376/biores.12.1.1548-
1565.
(33) Campbell, M. M.; Sederoff, R. R. Variation in Lignin Content and Composition
(Mechanisms of Control and Implications for the Genetic Improvement of Plants). Plant
Physiol. 1996, 110 (1), 3–13. https://doi.org/10.1104/pp.110.1.3.
(34) Banfi, D.; Patiny, L. Www.Nmrdb.Org: Resurrecting and Processing NMR Spectra On-
Line. CHIMIA 2008, 62 (4), 280–280. https://doi.org/10.2533/chimia.2008.280.
(35) Castillo, A. M.; Patiny, L.; Wist, J. Fast and Accurate Algorithm for the Simulation of
NMR Spectra of Large Spin Systems. J. Magn. Reson. 2011, 209 (2), 123–130.
https://doi.org/10.1016/j.jmr.2010.12.008.
(36) Aires-de-Sousa, J.; Hemmer, M. C.; Gasteiger, J. Prediction of 1H NMR Chemical Shifts
Using Neural Networks. Anal. Chem. 2002, 74 (1), 80–90.
https://doi.org/10.1021/ac010737m.
(37) Steinbeck, C.; Krause, S.; Kuhn, S. NMRShiftDBConstructing a Free Chemical
Information System with Open-Source Components. J. Chem. Inf. Comput. Sci. 2003, 43 (6),
1733–1739. https://doi.org/10.1021/ci0341363.
(38) Elbadawi, M. Polymeric Additive Manufacturing: The Necessity and Utility of Rheology;
IntechOpen, 2018. https://doi.org/10.5772/intechopen.77074.
(39) Cates, M. E.; Wittmer, J. P.; Bouchaud, J.-P.; Claudin, P. Jamming, Force Chains, and
Fragile Matter. Phys. Rev. Lett. 1998, 81 (9), 1841–1844.
https://doi.org/10.1103/PhysRevLett.81.1841.
(40) Wagner, N. J.; Brady, J. F. Shear Thickening in Colloidal Dispersions. Phys. Today 2009,
62 (10), 27–32. https://doi.org/10.1063/1.3248476.
(41) Barnes, H. A. Shear‐Thickening (“Dilatancy”) in Suspensions of Nonaggregating Solid
Particles Dispersed in Newtonian Liquids. J. Rheol. 1989, 33 (2), 329–366.
https://doi.org/10.1122/1.550017.
(42) Rapid Prototyping & Manufacturing— Fundamentals of Stereolithography. J. Manuf. Syst.
1993, 12 (5), 430–433. https://doi.org/10.1016/0278-6125(93)90311-G.
(43) Falkehag, S. I.; Marton, J.; Adler, E. Chromophores in Kraft Lignin. In Lignin Structure
and Reactions; Marton, J., Ed.; AMERICAN CHEMICAL SOCIETY: WASHINGTON, D.C.,
1966; Vol. 59, pp 75–89. https://doi.org/10.1021/ba-1966-0059.ch007.
(44) Fox, M. Optical Properties of Solids; Oxford University Press, 2010.

https://doi.org/10.26434/chemrxiv-2023-klv3z ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
27
(45) He, X.; Gao, Y.; Nie, J.; Sun, F. Methyl Benzoylformate Derivative Norrish Type I
Photoinitiators for Deep-Layer Photocuring under Near-UV or Visible LED. Macromolecules
2021, 54 (8), 3854–3864. https://doi.org/10.1021/acs.macromol.0c02868.
(46) Zhang, J.; Launay, K.; Hill, N. S.; Zhu, D.; Cox, N.; Langley, J.; Lalevée, J.; Stenzel, M.
H.; Coote, M. L.; Xiao, P. Disubstituted Aminoanthraquinone-Based Photoinitiators for Free
Radical Polymerization and Fast 3D Printing under Visible Light. Macromolecules 2018, 51
(24), 10104–10112. https://doi.org/10.1021/acs.macromol.8b02145.
(47) Sautrot-Ba, P.; Jockusch, S.; Malval, J.-P.; Brezová, V.; Rivard, M.; Abbad-Andaloussi,
S.; Blacha-Grzechnik, A.; Versace, D.-L. Quinizarin Derivatives as Photoinitiators for Free-
Radical and Cationic Photopolymerizations in the Visible Spectral Range. Macromolecules
2020, 53 (4), 1129–1141. https://doi.org/10.1021/acs.macromol.9b02448.
(48) Zhang, J.; Lalevée, J.; Hill, N. S.; Launay, K.; Morlet-Savary, F.; Graff, B.; Stenzel, M. H.;
Coote, M. L.; Xiao, P. Disubstituted Aminoanthraquinone-Based Multicolor Photoinitiators:
Photoinitiation Mechanism and Ability of Cationic Polymerization under Blue, Green, Yellow,
and Red LEDs. Macromolecules 2018, 51 (20), 8165–8173.
https://doi.org/10.1021/acs.macromol.8b01763.

TOC Graphic

https://doi.org/10.26434/chemrxiv-2023-klv3z ORCID: https://orcid.org/0000-0002-2637-9974 Content not peer-reviewed by ChemRxiv. License: CC BY-NC-ND 4.0
28

You might also like