Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text...

106
science.sciencemag.org/cgi/content/full/science.aau1567/DC1 Supplementary Materials for A sustainable wood biorefinery for lowcarbon footprint chemicals production Yuhe Liao*, Steven-Friso Koelewijn, Gil Van den Bossche, Joost Van Aelst, Sander Van den Bosch, Tom Renders, Kranti Navare, Thomas Nicolaï, Korneel Van Aelst, Maarten Maesen, Hironori Matsushima, Johan M. Thevelein, Karel Van Acker, Bert Lagrain, Danny Verboekend, Bert F. Sels* *Corresponding author. Email: [email protected] (B.F.S); [email protected] or [email protected] (Y.L.) Published 13 February 2020 on Science First Release DOI: 10.1126/science.aau1567 This PDF file includes: Materials and Methods Supplementary Text Figs. S1 to S53 Tables S1 to S14 References

Transcript of Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text...

Page 1: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

science.sciencemag.org/cgi/content/full/science.aau1567/DC1

Supplementary Materials for

A sustainable wood biorefinery for low–carbon footprint chemicals

production Yuhe Liao*, Steven-Friso Koelewijn, Gil Van den Bossche, Joost Van Aelst, Sander Van

den Bosch, Tom Renders, Kranti Navare, Thomas Nicolaï, Korneel Van Aelst, Maarten

Maesen, Hironori Matsushima, Johan M. Thevelein, Karel Van Acker, Bert Lagrain,

Danny Verboekend, Bert F. Sels*

*Corresponding author. Email: [email protected] (B.F.S); [email protected] or

[email protected] (Y.L.)

Published 13 February 2020 on Science First Release

DOI: 10.1126/science.aau1567

This PDF file includes:

Materials and Methods

Supplementary Text

Figs. S1 to S53

Tables S1 to S14

References

Page 2: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

2

Table of Contents 1. Materials ....................................................................................................................................6

1.1 Commercial chemical and materials .....................................................................................6 1.2 Synthesis of 4-n-propylsyringol (2,6-dimethoxy-4-n-propylphenol) ....................................6

2. Methods ...................................................................................................................................... 7 2.1 Lignocellulose compositional analysis ..................................................................................7 2.2 Catalyst preparation ...............................................................................................................8 2.3 Catalyst characterization .......................................................................................................8 2.4 Wood processing to lignin oil, lignin oil purification, product separation, and characterization ...........................................................................................................................9 2.5 Gas-phase reactions .............................................................................................................11 2.6 Preparation of condensate from hydroprocessing and dealkylation products .....................12 2.7 Zeolite sorption experiments: the uptake of 4-n-propylphenol and 4-isopropyl-3-methylphenol on zeolites ...........................................................................................................13 2.8 Saccharification-fermentation of RCF birch wood carbohydrate pulp toward bio-ethanol 13 2.9 Resin and varnish synthesis .................................................................................................14 2.10 Techno-economic analysis ................................................................................................15 2.11 Life-cycle assessment ........................................................................................................19

Supplementary text ..................................................................................................................... 22 ST1 The transition of fossil-based phenol to bio-phenol ..........................................................22 ST2 Catalytic funneling approaches .........................................................................................23 ST3 RCF birch wood phenolic oligomer characterization ........................................................23 ST4 Monomers extraction .........................................................................................................23 ST5 Hydroprocessing of phenolic monomers: the impact of metal type, loading, and support on selectivity .............................................................................................................................23 ST6 Hydroprocessing of phenolic monomers: process conditions ...........................................24 ST7 Hydroprocessing of phenolic monomers: mechanistic considerations ..............................24 ST8 Dealkylation: results and discussion ..................................................................................27 ST9 Semi-simultaneous saccharification-fermentation of RCF birch wood carbohydrate pulp ....................................................................................................................................................28 ST10 Synthesis of ink varnishes from RCF birch wood lignin oligomers ................................28 ST11 Loss of methanol in birch wood conversion .....................................................................30 ST12 Techno-economic analysis for this integrated biorefinery ..............................................30 ST13 Life-cycle assessment ......................................................................................................30 ST14 Carbon flow .....................................................................................................................32

Figures ..........................................................................................................................................33 Fig. S1 The global demand pattern of phenol and propylene and the suggested application of bio-phenol, bio-phenolic oligomers, and bio-propylene ...........................................................33 Fig. S2 Routes to phenol ...........................................................................................................34 Fig. S3 Reproducibility and impact of birch type on the outcome of the reductive catalytic fractionation of wood ................................................................................................................35 Fig. S4 Integrated production of phenol from lignocellulose via reductive catalytic fractionation and chemo-catalysis (catalytic funneling) ...........................................................36 Fig. S5 Aliphatic regions of 2D HSQC NMR spectra from the RCF birch wood lignin phenolic oligomers ...................................................................................................................................37

Page 3: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

3

Fig. S6 31P-NMR spectrum of phosphitylated RCF birch wood lignin phenolic oligomers derivatized with TMDP and with cholesterol as internal standard ...........................................38 Fig. S7 Systematic extraction study to isolate monomers from birch wood lignin oil using different n-hexane/lignin oil mass ratios ...................................................................................39 Fig. S8 Hydroprocessing of 4-n-propylguaiacol to n-propylphenols and ethylphenols: metal, loading, and support effects ......................................................................................................40 Fig. S9 Gas chromatograms of conversion of 4-n-propylguaiacol (PG) over different catalysts ....................................................................................................................................................41 Fig. S10 The influence of WSHV in the conversion of propylguaiacol ...................................42 Fig. S11 The influence of H2 partial pressure in the conversion of propylguaiacol over 64 wt.% Ni/SiO2 ......................................................................................................................................43 Fig. S12 The influence of reaction temperature in the conversion of propylguaiacol over 64 wt% Ni/SiO2 ..............................................................................................................................44 Fig. S13 Stability test and catalyst regeneration for the conversion of propylguaiacol over 64 wt.% Ni/SiO2 .............................................................................................................................45 Fig. S14 Gas chromatograms of different substrates conversion over 64 wt.% Ni/SiO2 ..........46 Fig. S15 Hydroprocessing of 4-n-propylguaiacol over 64 wt.% Ni/SiO2 at different WHSVs 47 Fig. S16 Hydroprocessing of different model compounds (4-n-propylcatechol, 4-n-propylanisole, and 4-n-propylpheol) over 64 wt.% Ni/SiO2 to understand the main reaction pathway of hydroprocessing step ..............................................................................................48 Fig. S17 Hydroprocessing of 4-n-propylsyringol over 64 wt.% Ni/SiO2 at different WHSVs 49 Fig. S18 The reaction network of hydroprocessing of 4-n-propylguaiacol and 4-n-propylsyringol over 64 wt.% Ni/SiO2 .......................................................................................50 Fig. S19 Gas chromatograms of the crude unseparated mixture of pine (A) and birch (B) wood lignin monomers before and after hydroprocessing over 64 wt.% Ni/SiO2 ..............................51 Fig. S20 Gas chromatograms of light fraction (gas, incondensable at 0-5 oC) in the conversion of 4-n-propylguaiacol over 64 wt.% Ni/SiO2 ............................................................................52 Fig. S21 Gas chromatogram of the condensate from hydroprocessing products ......................53 Fig. S22 Schematic illustration of access and diffusion characteristics of small (compared to the micropores) and bulky molecules in the microporous (left) and hierarchical (right) ZSM-5 ....................................................................................................................................................54 Fig. S23 Uptake of 4-n-propylphenol and 4-isopropyl-3-methylphenol at 40 oC on ferrierite (FER), ZSM-5 (Z40-P, Z140-P and Z140-H) and USY (USY-40) zeolites .............................55 Fig. S24 The N2 isotherms , DFT pore size distribution, and NH3-TPD profiles of Z140-P and Z140-H ......................................................................................................................................56 Fig. S25 The Gibbs free energy change (ΔG) of dealkylation of n-propylphenols as a function of temperature ...........................................................................................................................57 Fig. S26 Dealkylation of the hydroprocessing products from the extracted monomers of pine wood lignin oil over Z140-H at different temperatures ............................................................58 Fig. S27 Dealkylation of the hydroprocessing products from the extracted monomers of birch wood lignin oil over Z140-H .................................................................................................... 59 Fig. S28 Gas chromatograms of pine-derived crude alkylphenols before (red) and after (black) dealkylation over Z140-H at 470 oC...........................................................................................60 Fig. S29 Gas chromatogram of the condensate from dealkylation of the hydroprocessing products of birch wood lignin crude monomers .. ..................................................................... 61 Fig. S30 Catalytic conversion of cresols over zeolites ..............................................................62

Page 4: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

4

Fig. S31 The reaction network of dealkylation of 4-n-propylphenol over acid zeolites ...........63 Fig. S32 Dealkylation of 4-isopropyl-3-methylphenol over Z140-H and Z140-P ....................64 Fig. S33 Dealkylation of 4-n-propylphenol over Z140-H .........................................................65 Fig. S34 Dealkylation of 4-ethylphenol over Z140-H ..............................................................66 Fig. S35 Dealkylation of n-propylbenzene over Z140-H ..........................................................67 Fig. S36 Gas chromatogram of light fraction (gas, incondensable fraction at 0-5 oC) in the Poraplot Q column for dealkylation of n-propylphenols over Z140-H in the presence of H2 and water ..........................................................................................................................................68 Fig. S37 Mass flows of lignin valorization . ............................................................................. 69 Fig. S38 General scheme of the step-wise synthesis of ink from phenolics .............................70 Fig. S39 Chemical structure of rosin-modified resin ................................................................71 Fig. S40 Gel permeation chromatograms of birch wood lignin oligomers (from RCF), methanosolv birch wood lignin, and commercial acetosolv spruce wood lignin ......................72 Fig. S41 Resins synthesized from para-nonylphenol (left), self-prepared methanosolv birch wood lignin (middle), and RCF birch wood lignin oligomers (right) .......................................73 Fig. S42 Gel permeation chromatograms of resins based on (i) para-nonylphenol, (ii) self-prepared methanosolv birch wood lignin, and (iii) RCF birch wood lignin oligomers ............74 Fig. S43 Assessment of emulsification capacity (EC) of varnishes based on different resins .75 Fig. S44 Varnishes prepared from different resins (commercial resin, self-prepared para-nonylphenol based resin, methanosolv birch wood based resin, RCF birch wood lignin oligomers based resin) ...............................................................................................................76 Fig. S45 Graphical representation of the process flow diagram (PFD) used to model the conversion process of birch wood toward phenol, propylene, lignin phenolic oligomers and carbohydrate pulp ......................................................................................................................77 Fig. S46 Sankey diagram providing a graphical representation of all major process streams ..78 Fig. S47 (A) Mass balance of this integrated biorefinery (assuming conversion of 1 ton dry and extracted birch wood); (B) Mass balance of monomers conversion (on the basis of 1 ton dry and extracted birch wood) .........................................................................................................79 Fig. S48 CAPEX (Capital expenditure) distribution for the different subunits of the biorefinery ....................................................................................................................................................80 Fig. S49 Tornado diagram illustrating the results from the sensitivity analysis on several design and cost assumptions .....................................................................................................81 Fig. S50 System boundary of LCA of the production process of phenol from crude fossil oil 82 Fig. S51 System boundary of LCA of the production process of phenol from lignocellulose (birch wood) ..............................................................................................................................83 Fig. S52 GWPs [in kg of CO2-equivalent per kg of product (kg CO2e per kg product)] of phenol, propylene, oligomers, and carbohydrate pulp production from lignocellulose (birch wood) with different hydrogen sources and without/with sustainable forest management strategies in this integrated biorefinery ......................................................................................84 Fig. S53 Sensitivity analysis for GWP of phenol ......................................................................85

Tables ...........................................................................................................................................86 Table. S1 Monomer yield and distribution obtained from RCF of birch and pine wood ..........86 Table. S2 Semi-quantitative integration results of the 2D HSQC NMR spectra shown in fig. S5, for the lignin sub-units, inter-unit linkages and end-groups from the RCF birch wood lignin phenolic oligomers ..........................................................................................................87

Page 5: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

5

Table. S3 Hydroxyl group content in RCF birch wood lignin phenolic oligomers as determined by 31P-NMR ...............................................................................................................................88 Table. S4 The condensate composition of hydroprocessing products ......................................89 Table. S5 Properties of Z140-P and Z140-H .............................................................................90 Table. S6 Overview of the evaluation of resins synthesized from different phenol sources ....91 Table. S7 Overview of the evaluation of ink varnishes synthesized from different resins .......92 Table. S8 Overview of the mass and energy flows in the integrated (birch wood-to-phenol, propylene, lignin oligomers, and carbohydrate pulp) biorefinery .............................................93 Table. S9 List of economic parameters .....................................................................................94 Table. S10 Manufacture cost .....................................................................................................95 Table. S11 Summary of economics for the integrated biorefinery design using birch wood ...96 Table. S12 Global warming potentials (GWPs) of phenol, propylene, and nonylphenol production from crude fossil oil through Hock process, steam cracking of naphtha, and alkylation of phenol with nonenes, respectively .......................................................................97 Table. S13 Global warming potential (GWP) of phenol production from birch wood in this integrated biorefinery ................................................................................................................98 Table. S14 Global warming potentials (GWPs) of propylene, oligomers, carbohydrate pulp production from birch wood in this integrated biorefinery .......................................................99

References ..................................................................................................................................100

Page 6: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

6

1. Materials 1.1 Commercial chemical and materials 5 wt.% Pd, Pt, Ru, or Rh on Al2O3, and Ru, Pd on carbon were purchased from Sigma Aldrich. HTC Ni400 (17 wt.% Ni/Al2O3), HTC Ni500 (21 wt.% Ni/Al2O3), HTC Co 2000 (25 wt.% Co/Al2O3), Pricat 50/8 (51 wt.% Cu/SiO2), Pricat Ni 52/35 (50 wt.% Ni/Kieseguhr-Cr2O3), Pricat Ni 55/5 (55 wt.% Ni/Kieseguhr), and Pricat Ni 62/15 (60 wt.% Ni/kieseguhr-Al2O3) were supplied by Johnson Matthey. Ni 5249P (64 wt.% Ni/SiO2) was provided by Strem Chemicals. 65 wt.% Ni/SiO2-Al2O3 was purchased from Sigma Aldrich. These catalysts were used as received. CBV 28014 (NH4-ZSM-5, Si/Al=140, code: Z140-P), CBV 8014 (NH4-ZSM-5, Si/Al=40, code: Z40-P), CBV 2314 (NH4-ZSM-5, Si/Al=12, code: Z12-P), and CBV 780 (H-USY, Si/Al=40, code: USY-40) were purchased from Zeolyst. These NH4-form zeolites were converted into the protonic form by calcination in static air before catalytic test (vide infra). HSZ-770NAA (Na-FER, Si/Al=46.5, code: FER) was purchased from TOSOH. Na-FER was transferred into the protonic form by ion-exchange (to the NH4-form), followed by calcination (vide infra). Cellic CTEC2 enzyme mixture was purchased from Sigma Aldrich.

4-n-Propylguaiacol (99%), 4-ethylguaiacol (98%), 4-n-propylphenol (97%), 4-n-propylanisole (99%), 2-isopropylphenol (98%), 4-ethylphenol (98%), 4-cresol (99%), 2-cresol (99%), 4-isopropyl-3-methylphenol (99%), syringol (98%), 4-methylsyringol (97%), isoeugenol (98%), methanol (MeOH, >99%), DMSO-d6 (99.8 %), chloroform-d (CDCl3, 99.8 atom% D), N-methyl-N-(trimethylsilyl)trifluoroacetamide (MSTFA, ≥98.5%), n-hexane (99%), anhydrous pyridine (99.8%), dichloromethane (DCM, 99%), sodium hydroxide (NaOH, >98%), 2-chloro-4,4,5,5-tetramethyl-1,3,2-dioxaphospholane (95%), Chromium (III) acetylacetonate (97%), cholesterol (≥99%), concentrated hydrochloric acid (HCl, 36.5-38 wt.%), m-xylene and ammonium nitrate (NH4NO3, >99.5%) were ordered from Sigma Aldrich. Ethanol (EtOH, >99%) was purchased from Fischer Chemical Ltd. 4-Allyl-2,6-dimethoxyphenol (98%), Ni(NO3)2·6H2O (98%), TiO2 (anatase, > 150 m2 g-1), and 3-n-propylphenol (98%) were purchased from Alfa Aesar. These chemicals were analytic reagents and were used as received without further treatment. Tetrahydrofuran (THF, > 99.9%, HPLC grade) was purchased from Sigma Aldrich. Mineral oil PKWF 6/9 AR Blend and Mineral oil PKWF 6/9 AFN were provided by Haltermann Carless. Distilled water was used in reaction system. Birch and pine were provided by a local sawmill (Ecobois, Ghent, Belgium).

1.2 Synthesis of 4-n-propylsyringol (2,6-dimethoxy-4-n-propylphenol) 4-n-Propylsyringol was synthesized by hydrogenation of 4-allyl-2,6-dimethoxyphenol in MeOH over 5 wt.% Pd/C. Typically, 10 g of 4-allyl-2,6-dimethoxyphenol, 0.5 g of 5 wt.% Pd/C and 50 mL of MeOH were placed into a 100 mL stainless steel batch reactor (Parr Instruments Co.). The reactor was flushed three times with N2 (8 bar) to expel residual oxygen, and subsequently pressurized with 30 bar H2 at room temperature. Then, the mixture was heated to 60 oC and stirred at 750 rpm, and kept at this temperature for 3 hours (h). After reaction, the reaction mixture was filtered to remove Pd/C catalyst and the MeOH was removed by rotary evaporation to quantitatively yield 4-n-propylsyringol as an almost colorless oil. Yield: 10.10 g (>99%). [1H]NMR (300 MHz, CDCl3, 25oC, TMS): δH = 0.94 (t, 3JH,H= 7.3 Hz, 3H, -CH2CH3), 1.62 (sex, 3JH,H= 7.3 Hz, 2H, -CH2CH2CH3), 2.51 (t, 3JH,H= 7.3 Hz, 2H, -ArCH2CH2-), 3.88 (s, 6H, -OCH3), 5.36 (s, 1H, -OH) and 6.40 ppm (s, 2H, -ArH); [13C]NMR (100 MHz, CDCl3, 25oC, TMS): δC =

Page 7: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

7

146.9, 133.9, 132.7, 105.1, 56.3, 38.4, 25.0 and 13.9 ppm; MS (70 eV, EI): m/z (%): 196 (36), 167 (100).

2. Methods 2.1 Lignocellulose compositional analysis The lignocellulose compositional analysis was carried out as described in our previous work (16, 26). Birch wood is composed of 19.5 wt.% Klason lignin, 39.3 wt.% C6 sugars, 20.7 wt.% C5 sugars, 2.5 wt.% extractives; the rest is considered ashes (~0.34 wt.%), water (~5 wt.%) and some unknowns such as acid soluble lignin. Pine wood has 27.6 wt.% Klason lignin, 39.4 wt.% C6 sugars, 19.7 wt.% C5 sugars, 2.4 wt.% extractives; the rest is considered ashes (~0.33 wt.%), water (~5 wt.%) and some unknowns.

The performed procedures are as follows:

Dry wood was grinded and screened to a size of 250-500 µm for the analysis. The extractives such as waxes, fats, resins and terpenoids/steroids were removed through a Soxtec extraction, as their presence may affect the determination of Klason lignin. Typically, 2-3 g of oven-dried (80 oC) wood was loaded into porous thimbles and immersed completely into a boiling solvent mixture of toluene and ethanol (V-to-V of 2) for 15 minutes (min). Subsequently, a 3 h standard extraction was conducted. Then, ethanol was used to wash the cooled samples, followed by drying in an oven (80 oC) overnight. These extracted samples were used to determine the biomass composition. The content of extractives was determined gravimetrically by comparing the mass of the extracted sample with the mass of the original non-extracted sample.

Klason lignin content was determined according to an experimental procedure from Lin & Dence (27). 1 g of extracted sample was loaded into a 50 mL beaker, which was left for 2 h with magnetic stirring at room temperature after adding 15 mL of 72 wt.% H2SO4 solution (density of 1.6338 g mL-1). Next, the content of the beaker (including solution and solid residue) and the amount of water (~50 mL) from the beaker washing were transferred into a round-bottom-flask containing already 300 to 400 mL of water. A 3 wt.% H2SO4 (density of 1.0184 g mL-1) solution was obtained by adding additional water (density of 1 g mL-1) to a total volume of 577.54 mL. Afterwards, the content of the flask was heated to boil the solution for 4 h under reflux. A lignin precipitate was obtained on the filter after filtration of the hot solution, which was washed extensively with hot distilled water and dried (80 oC) in an oven for one night. The content of the Klason lignin was determined by measuring the weight of the residue on the filter. Three parallel experiments were conducted to average the content. Standard error of 0.2 wt.% was usually measured.

Carbohydrate content was measured by a standard sugar determination method, modified for hydrolysis of cellulose-rich materials (28-30). Typically, triplicate samples of 10 mg were hydrolyzed for 2 h in a concentrated H2SO4 solution (13 M, 1 mL) at room temperature, followed by hydrolysis at 100 °C for another 2 h in a diluted H2SO4 solution (2 M, 6.5 mL). The obtained sugars were reduced to sugar alcohols. Briefly, 3 mL of hydrolysate, 1.5 mL of 25 wt.% aqua ammonia, 1 mL of a 1 mg mL-1 β-D-allose solution of water and (saturated) benzoic acid (V-to-V of 1) as internal standard, and droplets of 2-octanol to avoid excessive foaming, were mixed. NaBH4 (0.2 mL of a 200 mg mL-1 NaBH4 solution containing 2 M NH3) was used to reduce the sugars at 40 oC for 0.5 h. Acetic acid (0.4 mL) was added to terminate the reaction.

Page 8: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

8

The sugar alcohols were quantified by GC analysis after acetylation. Briefly, sugar alcohol acetates were obtained from the reaction of 0.5 mL of the reduced samples and 5 mL of acetic acid anhydride in the presence of 0.5 mL of 1-methylimidazole. EtOH (1 mL) was added after 10 min. After another 5 min, 10 mL of water was added to stop the reaction. 0.5 mL of a 0.4 g L-1 bromophenol water solution and 10 mL of a 7.5 M KOH solution were added into the reaction vessel (cooled on ice bath) to color the aqueous phase blue. The ethyl acetate phase, containing the acetylated sugar alcohols, was separated by a Pasteur pipette. Anhydrous Na2SO4 was used to dry the ethyl acetate phase. GC analysis was carried out at 225 °C with injection and detection temperatures at 270 °C on a Agilent 6890 chromatograph, equipped with a Supelco SP-2380 column, and a flame ionization detector (FID). The addition of water during hydrolysis was corrected by a factor to determine the content of carbohydrates. The standard errors are less than 1 wt.% for glucose and xylose.

2.2 Catalyst preparation Hierarchical ZSM-5 (i.e. tailor-made ZSM-5 in this work) was obtained by post-modification (here: a combination of alkaline and acid treatment) of a commercial microporous ZSM-5 (Si/Al=140, code: Z140-P). For the alkaline treatment, the parent sample (Z140-P, 3.3 g) was introduced into a stirred 0.2 M NaOH solution (100 mL, 65 oC) for 0.5 h. Then, the suspension was rapidly cooled with water. The resultant solid was filtered, washed extensively with distilled water, and dried at 65 °C overnight. The alkaline treated sample (2 g) was subsequently treated in 0.1 M HCl solution (200 mL) at 65 oC (solution temperature) for 6 h.

Prior to catalytic evaluation, all zeolite samples were transferred into the protonic form via three consecutive ion exchange steps in 0.1 M aqueous NH4NO3 solution (room temperature, 12 h, 1 g of zeolite per 100 mL of solution), followed by calcination in static air (5.5 h at 550 °C for all zeolites at a 5 °C min-1 ramping rate). The obtained hierarchical ZSM-5 is coded as Z140-H.

Ni/TiO2 was prepared by incipient wetness impregnation of the TiO2 (anatase) with an aqueous solution of Ni(NO3)2·6H2O. The sample was dried at 60 oC for 24 h. Afterwards, the sample was reduced at 460 °C (heating rate 5 °C min-1 ) for 1.5 h under a flow of H2 (120 mL min-1 g-1).

2.3 Catalyst characterization Nitrogen sorption (N2-sorption) measurements were carried out at -196 °C with a Micromeritics TriStar 3000 instrument. The samples were degassed under a flow of N2 at 300 oC for 6 h prior to the sorption. The t-plot method was used to distinguish between micro- and mesopores. Due to the presence of an artefact at the pressures corresponding to a thickness of 0.35-0.50 nm for Z140-P and Z140-H, the thickness range used to process the t-plot analysis for these samples was chosen from 0.50 till 0.70 nm (31). The external surface area (denoted Smeso) includes the mesopores and the external surface area of the crystals, which is determined from the t-plot. The DFT model was applied to the adsorption branch of the isotherm to determine the pore size distribution.

Ammonia temperature-programmed desorption (NH3-TPD) measurements were performed in a home-made flow instrument equipped with a Pfeiffer Omnistar quadrupole mass spectrometer for the desorbed gas (NH3, m/z=16). For this, the sample (100 mg) was first pretreated in a helium flow at 400 oC (ramping rate of 5 oC min-1) for 2 h. The adsorption of NH3 was performed at 200 oC for 0.5 h, followed by flushing with helium for 0.5 h at the same temperature. Afterwards, the sample was heated to 1000 oC at a ramping rate of 10 oC min-1 in a helium flow to obtain the NH3-TPD profiles.

Page 9: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

9

Pyridine probe FTIR (Py-FTIR) was used to determine the density of Brønsted and Lewis acid sites of zeolite in a Nicolet 6700 spectrometer. Briefly, the sample was pelletized into a self-supported wafer (around 10 mg cm-2), followed by pretreatment under 1 mbar at 400 oC for 1 h. Afterwards, the background was recorded and 25 mbar pyridine vapor was introduced into the IR cell and adsorbed at 50 oC until saturation. Before being measured at 150 oC with 256 scans, the sample was degassed at 150 oC, 250 oC, and 350 oC for 20 minutes, respectively. The band areas of adsorbed pyridine at 1450 and 1550 cm-1 related to the total concentration of Lewis (L) and Brønsted (B) acid sites, respectively. The integrated molar extinction coefficients (ε(B)= 1.67 cm µmol-1 and ε(L)= 2.94 cm µmol-1) were determined by a reported method of Emeis (32).

Carbon monoxide chemisorption (CO-chemisorption) was conducted to determine the metal dispersion of nickel catalysts (i.e. 17 wt.% Ni/Al2O3, 21 wt.% Ni/Al2O3, 32 wt.% Ni/Al2O3, 16 wt.% Ni/SiO2, 55 wt.% Ni/Kieselguhr, and 64 wt.% Ni/SiO2) using the same equipment as for the NH3-TPD. Typically, 15 mg of sample was pretreated with H2 for 0.5 h at 400 oC (using a ramping rate of 5 oC min-1). Afterwards, the sample was flushed with helium and cooled to room temperature. CO-pulses were fed to a helium flow at a regular time interval (5 min) through a calibrated loop (20 µL). The CO concentration in the effluent was quantified by the mass spectrometer (CO, m/z = 28). For the calculation of metal dispersion, a CO to Ni stoichiometry of unity was assumed here.

Hydrogen temperature programmed reduction (H2-TPR) was conducted in the same setup for the NH3-TPD. The spent catalyst was loaded into the reactor tube without any pretreatment. Then, the sample was heated to 800 oC at a ramping rate of 10 oC min-1 in a reducing mixture (5% H2 in N2) flow to obtain the H2-TPR profile.

2.4 Wood processing to lignin oil, lignin oil purification, product separation, and characterization 2.4.1 Reductive catalytic fractionation (RCF) of wood to produce lignin oil and carbohydrate pulp For the production of the lignin-derived phenolic monomers, oligomers, and carbohydrate pulp, a 2 L stirred batch reactor (Parr Instruments Co.) was loaded with 150 g of wood chips (particle size: <10 mm), 800 mL of MeOH and 15 g of 5 wt.% Ru/C. The reaction vessel was closed and flushed three times with N2 (8 bar) in order to remove the residual oxygen. High pressure H2 (30 bar at room temperature) was applied on the reaction mixture before heating, and the reactor is stirred at 720 rpm. The reaction was performed at 235 °C. After 3 hours, the reaction was terminated by rapid cooling with compressed air flow and water. The reactor content was filtered in order to separate the solid fraction, containing the carbohydrate pulp and the catalyst, and the liquid fraction, containing the lignin oil and some soluble sugar products. To collect all liquid fraction, the solid residue was washed with EtOH. Afterwards, MeOH and EtOH were removed from the liquid phase by rotary evaporation to yield a crude brownish colored lignin oil, containing some soluble sugar products next to phenolic monomers and oligomers.

A threefold liquid-liquid extraction with water and dichloromethane (DCM) at a mass ratio of 1/3/3 (crude lignin oil/DCM/water) was performed to separate the soluble sugar products from the lignin-derived products, prior to gas chromatographic analysis and lignin monomers separation (vide infra). Note that >99 wt.% of the lignin derived monomers in the lignin oil is present in the DCM phase, while >99 wt.% of sugar products is presented in water phase. Evaporation of DCM yielded the sugar-free lignin oil, consisting of phenolic monomers and oligomers (see fig. S7). The

Page 10: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

10

weight of the sugar-free lignin oil was used to calculate the degree of delignification (on the basis of the Klason lignin weight) and to quantify the lignin products.

The phenolic monomers were quantified using a Gas Chromatograph (GC, Agilent 6890) equipped with a HP5 column and a FID. 2-Isopropylphenol was used as the internal standard. The following parameters were used in the GC analysis: injection and detection temperature of 300 °C, column temperature program: 50 °C (2 min), 15 °C min-1 to 150 °C, 10 °C min-1 to 220 °C and 20 °C min-

1 to 290 °C (12 min). Commercial standards (‘Materials’ section) were applied to determine the response factors of the monomeric products. For the non-commercially available monomers (such as 4-n-propanolsyringol and 4-ethylsyringol), response factors were inferred based on (i) the response factors of analogues, and (ii) application of the basic principles of the ‘effective carbon number method’ (33). The product yield (in wt.%) is defined as the ‘mass of products per mass of Klason lignin ×100%’. The mass of oligomers are determined gravimetrically as the mass difference between the lignin oil and phenolic monomers. The experimental errors of the method were determined by performing different batches of RCF (n=5; on the same wood) and different sources of wild birch wood (n=2) are shown in fig. S3. Clearly, the RCF experiments and of the analytical methods is highly reproducible, while the origin of the birch wood has no significant impact on the results.

Gel-permeation chromatography (GPC) was used to determine the molecular weight distribution of the lignin oil (see later for examples), distinguishing the phenolic monomers and the oligomers. Typically, a lignin oil sample was added into THF (around 10 mg mL-1) and subsequently filtrated with a 0.2 µm polytetrafluoroethylene (PTFE) membrane. GPC analysis was conducted at 40 oC on a Waters e2695 equipped with a Varian M-Gel column (3 µm, mixed), and a Waters 2988 Photodiode array detector (UV detection at 280 nm). THF was used as the elute (1 mL min-1). Polystyrene was used to calibrated the GPC.

2.4.2 Extraction of wood lignin-derived monomers To isolate the lignin-derived phenolic monomers from the sugar-free lignin oil, liquid-liquid extraction was applied.

After removal of the soluble sugars (vide supra), the purified lignin oil was subjected to a three or fourfold reflux extraction with n-hexane (at 80 oC of oil bath for 3 h), and the extract was distilled in vacuo to obtain a transparent yellowish oil. This oil presents the concentrated fraction of the phenolic monomers (see fig. S7 for the experimental results).

A systematic extraction study was conducted to find the optimum trade-off between (i) a high extraction efficiency, (ii) a low solvent usage, and (iii) a low oligomers co-extraction in the hexane phase. Typically, 2 g of lignin oil was therefore extracted with 1, 3 or 5 weight equivalents of n-hexane, and repeated three (or four) times, with intermediate extract analysis by GC and GPC. These extracts were compared to the initial lignin oil and the extracted residue. The results are shown in fig. S7 and the supplementary text ST4.

2.4.3 Characterization of birch wood lignin phenolic oligomers via NMR The RCF birch wood lignin phenolic oligomers (70 mg) were dissolved in 0.6 ml DMSO-d6 and loaded into a NMR tube. The two-dimensional 1H-13C HSQC NMR experiment was conducted at 298 K using a Bruker Avance III HD 400 MHz console with a Bruker AscendTM 400 Magnet, equipped with a 5 mm PABBO probe. A Bruker standard pulse sequence (‘hsqcedetgpsp.3’) was used with the following parameters: spectral width in F2 dimension (1H) of 13 ppm using 2048

Page 11: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

11

data points, a spectral width in F1 dimension (13C) of 165 ppm, using 256 data points, a total of 16 scans were recorded with a 2s interscan delay (D1). Bruker’s Topspin 4.0.2 software was used for data processing and volume integration. The solvent peak of DMSO was used as the internal reference (δc/δH: 39.5 ppm/2.49 ppm) following by manually phasing and automatic baseline correction. 31P-NMR measurements were performed in triplicate using a standard phosphitylation procedure (34, 35). A solvent solution (1.6 pyridine : 1 CDCl3 (V/V)) was used to make stock solutions of the internal standard (cholesterol, 20 mg mL1) and relaxation agent (chromium acetylacetonate, 10 mg mL-1). An amount of RCF birch wood lignin phenolic oligomers (ca. 20 mg) was accurately weighed, then 100 µl of the internal standard solution and 50 µl of the relaxation agent solution was added, next to 400 µl of solvent solution. Subsequently, 75 µl of 2-chloro-4,4,5,5-tetramethyl-,1,3,2-dioxaphospholane (TMDP) was added and the sample was thoroughly mixed before transferring them to the NMR tube. 31P-NMR spectra were obtained on a Bruker Avance III 400 MHz NMR using a standard phosphorous pulse program (256 scans, 5s interscan delay, O1P 140 ppm). The chemical shifts were calibrated by assigning the sharp peak of residual water + TMDP at 132.2 ppm and automatic baseline correction was applied.

2.5 Gas-phase reactions The above isolated mixture of phenolics monomers (and also some model compounds) will be subjected to two catalytic reactions in sequence to produce phenol and propylene. Therefore, two different gas-phase catalytic tests, hydroprocessing and dealkylation, were carried out (separately) in a home-made fixed-bed reactor, equipped with four parallel quartz reactor tubes (inner diameter of 3 mm) under ambient pressure.

2.5.1 Hydroprocessing reaction In a typical hydroprocessing experiment, 60 mg of catalyst, pelletized to a 0.125-0.25 mm fraction, was loaded into the four quartz reactor tubes (15 mg catalyst per tube) and held by two layers of quartz wool. The catalyst was diluted with quartz powder (0.125-0.25 mm) to reduce the local hot spots and to improve the temperature distribution, yielding a catalyst bed with a height of ca. 15 mm. Reactor temperature in axial direction of the oven at height of the catalyst bed is homogeneous. The substrate was brought into the reactor by using a nitrogen (N2) flow, passing through a saturator (in the case of crude mixture of lignin monomers, it was fed into the reactor using a syringe pump). Hydrogen was introduced into the reactor directly using mass flow controller. The molar composition of the gas mixture in the reactor before reaction is 0.02/0.4/0.58 (for substrate/H2/N2). The effluent gases were analyzed using an online GC (HP4890D) equipped with two parallel columns (HP1 column and Porapolt Q column), both connected with a FID. Condensable products (boiling point >30 oC) were collected by a glass collector at 0-5 oC, and characterized offline by using a Agilent Technologies 7890A GC, equipped with a FID and a mass spectrometer, and separated by a HP1 column. GC analysis of the light fraction (that is, permanent gases and the incondensable part at 0-5 oC) was performed on an Interscience Trace GC equipped with Hayesep Q and RTX-1 columns, a FID, and a thermal conductivity detector (TCD). Since some isomers such as 4-n-propylphenol and 3-n-propylphenol cannot be distinguished by the online GC, silylation of the collected products was applied prior to the offline analysis (36).

The different WHSV experiments were conducted by changing the amount of catalyst loading, while keeping the catalyst bed height (with changing the amount of quartz powder) and total flow constant. Different H2 partial pressure experiments were conducted by replacing the N2 by H2.

Page 12: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

12

WARNING! Use caution when taking out the reactor tubes and handling the catalysts as they may be pyrophoric.

Calculations of the catalytic outcome were done according to the following definitions:

The selectivity to products (%)=yield of products/theoretical yield of products×100%; Conversion of monomers (%) = (Nmonomers in-Nmonomers out)/Nmonomers in×100%, where Nmonomers is the mole of monomers. The unit of weight hourly space velocity (WHSV) in this work is gsubstrate gcat

-1 h-1 (i.e. h-1).

2.5.2 Dealkylation reaction to biophenol and biopropylene

The above hydroprocessed mixture of phenolic monomers was collected and used here to undergo catalytic dealkylation to form biophenol and biopropylene. In a typical dealkylation experiment, 120 mg of zeolite catalyst, pelletized to a 0.125-0.25 mm fraction, was loaded into the four quartz reactor tubes (30 mg catalyst per tube) and held by two layers of quartz wool, yielding a catalyst bed of ca. 13 mm. The alkylphenols (i.e. model compounds) and water were introduced into the reactor by using a N2 flow, passing through the related saturator. When condensates (of the hydroprocessing reaction) were tested, a syringe pump was used. The molar composition of the gas-phase before reaction is 0.02/0.12/0.86 (alkylphenols/water/N2). Dealkylation of 4-n-propylphenol in the presence of H2 was also conducted (as test reaction) by replacing N2 with H2.

The effluent gases were characterized by the above mentioned online GC equipped with two FIDs, a HP1 column and a Porapolt Q column. The condensable products (boiling point >30 oC) from the dealkylation reaction were also collected and analyzed by the above described method.

2.5.3 Dealkylation of n-propylbenzene For dealkylation of n-propylbenzene (test reaction), water was not fed into the reactor and the molar composition of the gas phase of 0.02/0.98 (n-propylbenzene/N2) was used. The total flow of n-propylbenzene dealkylation is same as that of the flow of the alkylphenols dealkylation.

Calculations of the catalytic outcome of the dealkylation experiments were done as follows. Conversion of dealkylation (%)=(Nsubstrate in-Nsubstrate out)/Nsubstrate in×100%, Nsubstrate is the mole of substrate in the reaction stream (including the isomers); Selectivity to phenol and olefin (%)=(Yphenol+Yolefin)/Conversion×100%; Selectivity to benzene and propylene (%)=(Ybenzene+Yolefin)/Conversion×100%, Y is carbon molar yield.

2.5.4 Disproportionation and transalkylation of cresols To improve the phenol yield, disproportionation and transalkylation of the formed cresols (side products) were attempted. In a typical disproportionation and transalkylation reaction of cresols, 120 mg of zeolite (a 0.125-0.25 mm fraction) was loaded into the four quartz reactor tubes (30 mg catalyst per tube) and held by two layers of quartz wool, yielding a catalyst bed of ca. 13 mm. The para-cresol and water were introduced into the reactor by using a N2 flow, passing through the related saturator. Note that para-cresol can undergo isomerization. The para-cresol transalkylates with the isomers (i.e. ortho- and meta-cresols).

Calculations of the catalytic outcome of the disproportionation and transalkylation experiments were done as follows: Conversion of cresols (%)=(Ncresols in-Ncresols out)/Ncresols in×100%, Ncresols is the mole of cresols (including all isomers) in the reaction stream. The theoretical yield of phenol from cresols disproportionation and transalkylation is 50% of the conversion. Selectivity to phenol (%)=yield of phenol/theoretical yield ×100%.

Page 13: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

13

2.6 Preparation of condensate from hydroprocessing and dealkylation products Pure 4-n-propylguaiacol, extracted pine wood lignin monomers, and birch wood lignin monomers (vide supra) were used as feedstock in the aforementioned hydroprocessing reactor unit without any purification steps. The hydroprocessing reaction was conducted at 285 oC over 64 wt.% Ni/SiO2 for the model compound 4-n-propylguaiacol and the pine lignin-derived phenolics monomers at 6.0 h-1 WHSV. For birch lignin-derived monomers, the reaction was conducted at 305 oC over 64 wt.% Ni/SiO2 at 5.3 h-1 WHSV.

The reaction was monitored in function of time (by online GC analysis), and the reaction was stopped when the conversion of the phenolic monomer(s) became lower than 93% (as a result of catalyst deactivation) in order to reduce the amount of unconverted monomers in the product stream. Afterwards, fresh catalyst was filled to carry out the same reaction as to obtain enough substrate (> 10 g with more than 10 experiments) for dealkylation experiments.

The condensates of hydroprocessing products were collected(vide supra) and combined (gas chromatogram see fig. S21), and they were fed to the dealkylation process without any purification.

During the dealkylation process, the mixture of dealkylation products, composing mainly of phenol, was also collected using the same methodology (vide supra). However, only small amount of dealkylation product was collected (gas chromatogram see fig. S29) due to safety regulation imposed by the university (8-hour shift and short-term (<15 minutes) permissible exposure limits are 5 ppm and 15.6 ppm, respectively). The biophenol can be purified with silica column chromatography using n-heptane/ethyl acetate = 8: 1 as eluent. The solvent was removed by rotary evaporation.

2.7 Zeolite sorption experiments: the uptake of 4-n-propylphenol and 4-isopropyl-3-methylphenol on zeolites The adsorption of 4-n-propylphenol and 4-isopropyl-3-methylphenol was conducted in 5 mL of toluene (10 mg mL-1) over 0.25 g of zorptioneolites at 40 oC for 12 h. After adsorption, the zeolites and solution were separated by centrifuging (4000 rpm, 5 min).

The uptake was determined by measuring the concentration difference before and after adsorption via GC. meta-Xylene was used as the internal standard. The experimental data are presented in fig. S23.

2.8 Saccharification-fermentation of RCF birch wood carbohydrate pulp toward bio-ethanol Utilization of the RCF birch wood pulp to bioethanol is demonstrated and the following procedures were followed. Semi-simultaneous saccharification-fermentation of the carbohydrate pulp (containing Ru/C catalyst), obtained after RCF (vide supra), was based on an NREL protocol (37), modified with a 24 h pre-incubation at 50 °C and 15 wt.% dry mass (stirring rate=300 rpm).

7.5 g of RCF pulp was suspended in 49 mL of a 100 mM citrate buffer (pH=5) containing the commercial CTEC2 enzyme mixture (Novozyme Corp.) at an enzyme load of 18 FPU g-1 DS (based on a reported enzyme activity of 119 FPU mL-1 (38)). This enzyme mixture comprises cellulases, hemicellulases and β-glucosidase to convert both xylan and glucan into their respective sugar monomers. After 24 h, the slurry was cooled down to ambient temperatures, after which 2 g L-1 urea was added and the pH adjusted to 5.2 with NaOH.

The engineered yeast MDS130 was grown overnight and added to the slurry to a final OD600=4. MDS130 is a descendant of the GSE16-T18 yeast strain which is improved for fermentation of

Page 14: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

14

xylose and increased inhibitor tolerance provided by GlobalYeast (39, 40). Samples were taken every 24 h for a total time of 264 h (24 h saccharification + 240 h saccharification-fermentation). The concentration of ethanol and both monomeric sugars (xylose and glucose) were determined with HPLC (Metacarb 87H) using H2SO4 as eluent (0.5 mM) at a flow rate of 0.7 mL min-1. The experiments were performed in duplicate. The data are collected in Fig. 3A.

2.9 Resin and varnish synthesis The mixture of phenolic oligomers from the RCF birch wood lignin oil was explored for the synthesis of rosin-based resins and ink varnishes.

For the synthesis of rosin-modified phenolic resins for printing ink, the reader is referred to (41). The phenolic part comprises 60 wt.% cardanol (a byproduct of cashew nut processing) and 40 wt.% others (here: the classic para-nonylphenol as benchmark, and two lignin sources: methanosolv lignin and the mixture of RCF birch wood phenolic oligomers), which were polymerized using paraformaldehyde. The preparation method of methanosolv lignin is mentioned in caption of table S6. Tall oil rosin was used as source of rosin acids (i.e. abietic acid), whereas maleic acid anhydride was added to form dicarboxylic adducts on the rosin. Glycerol and pentaerythritol were added for cross-linking via esterification. MgO was used as catalyst.

A schematic structure of the rosin-modified phenolic resin is presented in fig. S39. Subsequently, an ink varnish was prepared by mixing the above obtained phenolic resins (para-nonylphenol, organosolv and RCF birch wood phenolic oligomers based), rapeseed oil methyl ester and linseed oil. The relative ratio between these components was set to ensure a viscosity of around 100 Pa.s.

2.9.1 Gel permeation chromatography (for resins) For gel permeation chromatography of resins, 25 (± 5) mg of resin was solubilized in 10 mL of solvent. The solvent consisted of a mixture of THF (2.5 L), acetic acid (50.0 mL) and carbon disulphide (2.0 mL). The latter was used as a flow marker. Samples were left to sit overnight to make sure that the resins were completely solubilized.

Analysis was performed using an Agilent Plgel 10 µm Mixed-B (300 × 7.5 mm) column, protected by a Agilent PLgel 10 μm guard column (50 × 7.5 mm). Columns were placed in a column thermostat (Separations Analytical Instruments), which was connected to a Waters 717 plus auto sampler, a Waters 515 programmable HPLC pump, and a Waters 2410 differential refractive index detector (RID). A flow rate of 1 mL min-1 was installed, with the eluent being a mixture of 50.0 mL acetic acid and 2.5 L THF. The column temperature was set at 25 °C, the injector temperature was 25 °C, and the detector temperature was 30 °C. 100 µL of sample was injected. The run time equalled 45 min per sample. EasiCal polystyrene standards PS-1 (Agilent) were used for calibration.

2.9.2 Viscosity measurement (for resins and ink varnishes) The viscosity for resins was determined for a 40 wt.% concentration in mineral oil PKWF 6/9 AR Blend. Therefore, a weighed amount of resin (20.0 g) was added to a metal test tube (30 mm outer diameter, 205 mm length), together with mineral oil (30.0 g) and a magnetic stirring bar. A PT100 thermocouple was inserted and the test tube was placed in a Thermotronic (Novomatics) apparatus. A standard program was used. First, the contents of the test tube were heated to 230 °C (10 oC min-

1) under constant stirring (1200 rpm). The solution was kept at 230 oC for 2 min, followed by cooling to 90 °C (20 oC min-1). The test tube was emptied at 90 oC.

Page 15: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

15

Viscosity measurements were performed using a Physica MC300 rotational rheometer (Anton Paar), equipped with a conical plate (25 mm radius). The temperature was set at 23 °C. The shear rate was linearly increased from 0 to 50 Hz over a time span of 180 s. The viscosity measured at 25 Hz is used for comparison. The flow behaviour index (𝑛𝑛) was determined by applying the following Power Law model: ln(𝜏𝜏) = ln(𝐾𝐾) + 𝑛𝑛. ln(𝛾𝛾)

with 𝜏𝜏 being the shear stress (Pa.s), 𝛾𝛾 the shear rate (s-1), and 𝐾𝐾 the consistency coefficient (Pa.sn) (42).

The flow behaviour index (n) indicates the deviation from Newtonian flow behaviour (n=1). For shear-thickening fluids, n is larger than 1. For shear-thinning fluids, n is smaller than 1. Resins (solubilized in mineral oil) and varnishes typically display shear-thinning behaviour.

2.9.3 Filtration test (for resins) A weighed amount of resin (50 g) and xylene (250 g, technical grade isomeric mixture) were added to a 500 mL Erlenmeyer flask. The mixture was heated under reflux for a minimum of 2 h. Subsequently, the solution was cooled down to room temperature and poured over a stainless steel filter (500 mesh). The Erlenmeyer flask was rinsed with toluene, which was also poured over the filter. The residue on the filter was washed with xylene and was subsequently dried in an oven at 80 °C for 1 h. The mass of the residue was determined gravimetrically, and is expressed in ppm (= mgresidue ×1000 / 50 gresin).

2.9.4 Determination of acid value (for resins) A weighed amount of resin (1-2 g) was added to a 100 mL Erlenmeyer flask. Next, 40 mL of a 2:1 (V-to-V) xylene:isopropanol mixture containing phenolphthalein (indicator) was added. The resin was solubilized by heating under reflux, followed by cooling to room temperature. The solution was titrated with a 0.1 N KOH solution in methanol until the solution turned pink. The acid value is expressed as mgKOH / gresin. 2.9.5 Cloud point determination (for resins) The solid resin was crushed on a clean table using a spatula to obtain particles in the range of 1-5 mm. A weighed amount of crushed resin (2 g) was added to a glass test tube (22 mm outside diameter, 200 mm length). Subsequently, the resin was solubilized in mineral oil (PKWF 6/9) to obtain a 10 wt.% solution. A PT100 thermocouple was inserted and the test tube was placed in a Chemotronic apparatus (Novomatics). A standard program was used. First, the solution was heated to 230 °C at constant stirring of 1100 rpm. The temperature was kept constant for 2 min (800 rpm). Subsequently, the solution was cooled to 30 °C (1000 rpm). The resulting cloud point is expressed in °C and corresponds to the temperature below which the solution gets a turbid appearance and the resin is not fully soluble anymore.

2.9.6 Assessment of emulsification capacity (for ink varnishes) The emulsification capacity was investigated using a High Speed Lithotronic Emulsification Tester (Novomatics) apparatus. A weighed amount of varnish (25 g) was added to the metal sample holder, which was subsequently placed in the apparatus. The temperature of the sample was increased to 40 °C at a constant stirring rate (1200 rpm), resulting in a constant (stable) torque. Next, water was injected at a constant flow rate of 2 mL min-1, which resulted in an increase of the measured torque. The measurement was stopped when 25 mL of water was added.

2.10 Techno-economic analysis

Page 16: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

16

2.10.1 Process design details The presented unit is designed for the conversion of wood chips to produce phenol (purity: 99%), propylene (purity: 99.9%), lignin oligomers and raw carbohydrate pulp as the main products. This unit is a scale up from the lab installation described in this work.

The overall process is further divided into four subunits. In the first subunit, the lignocellulose feedstock is fractionated via in-situ extraction by methanol. The fractionated lignin intermediateds are depolymerized and stabilized by the use of a redox catalyst. After reaction, filtration of the solid carbohydrate pulp and evaporation of the solvent yield a liquid product mixture consisting mostly of lignin-derived phenolic monomers and oligomers, next to some soluble carbohydrates. In the second subunit, being the separation unit, water-soluble components like sugar-derived products are removed to obtain a purified lignin oil (i.e. refined lignin oil), after which the lignin monomers and oligomers are further separated by liquid-liquid extraction. The lignin monomers stream is then further processed in the third subunit by hydroprocessing, where a mixture of alkylphenols is formed from the lignin-derived phenolic monomers. In the final subunit, the obtained mixture of alkylphenolic products from hydroprocessing undergoes dealkylation and distillation to generate purified streams of phenol (99%) and propylene (99.9%).

A simplified process flow diagram (PFD) is provided in fig. S45. In addition, an overview of all major process streams is represented in the Sankey diagram in fig. S46, accompanied with table S8, listing the corresponding energy and mass balances.

The process simulation is executed using Chemcad 7.1. For creating new components like cellulose, hemicellulose, lignin monomers, and lignin oligomers the group contribution method Jobak was used. For the prediction of the properties and equilibrium data of lignin monomers the UNIFAC method was used. For modeling of the individual subunits a combination of the PSRK equation of state and the UNIFAC polymer model was used.

The unit design is based on the following core data: • Production capacity of 100 kilo-tons of bio-based phenol per year • Reaction conditions of RCF: 250 °C, 100 bar, and a reaction time of 1.5 h • Wood / Methanol / Hydrogen mass ratio in RCF: 1 / 4.4 / 0.01 • Crude lignin oil / Dichloromethane / Water mass ratio during separation: 1 / 2.6 / 2.6 • Refined (or purified) lignin oil / n-hexane mass ratio during separation: 1 / 3.5 • Reaction conditions of hydroprocessing: 300 °C and 3 bar (compensating for pressure drop

in the catalyst bed on a large scale) • Monomers / Hydrogen mass ratio in hydroprocessing:10 / 1 • Reaction conditions of dealkylation: 400 °C and 2 bar (compensating for pressure drop in

the catalyst bed on a large scale) • Alkylphenols/water mass ratio in dealkylation: 2 / 1 • Life time of 2 years is assumed for all catalysts Note: Wood is on dry and extracted basis. Pressures are provided in absolute pressure (not in bar gague, as relative overpressure). The results (i.e. the retention of carbohydrate pulp, monomers yield, oligomers yield) of RCF are almost the same for a reaction time of 1.5 h and 3 h as demonstrated on lab scale.

The reductive catalytic fractionation (RCF) process. According to the design (fig. S45), the hardwood chips are fed into the reactors via alternating dosing hoppers, which are evacuated and

Page 17: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

17

flushed with nitrogen to remove oxygen. Here hydrogen and fresh methanol are also injected into the suspension. Seventeen 250 m3 reactor systems, including catalyst beds filled with 5 wt.% Ru/C catalyst, are used in parallel to provide the minimum extraction contact time of 1.5 hours. Homogenization in the reactor occurs via external solvent circuits. The calculated heat of reaction is -7426 kJ mol-1

lignin, implying an exothermic reaction. After reaction, the solid fraction, i.e. the carbohydrate pulp, is isolated via a combination of a hydrocyclone and dead-end filters with backflush. The high solid stream is expanded using two steps to remove most of the solvent. Subsequently, the solids are dried in a fluidized bed.

As a result of the exothermic reaction, an isothermal regime can be reached with two independent circulation loops and addition of raw material. In the first circuit, after catalytic stabilization and separation of solids, a fraction of the liquid stream (e.g. 10%) containing the dissolved crude lignin oil in methanol is expanded from 100 bar to 3 bar in a rotating liquid turbine (with integrated power generator). The solvent (methanol) is then removed via distillation and introduced back into the reactor at a low temperature of 51 °C. In a partial condenser, most of the methane (obtained from hydrogenolysis of methanol in the RCF) and hydrogen is recuperated together with methyl acetate (from the acetyl group of hemicellulose) and a very small amount of methanol. The gaseous stream from the partial condenser is cooled down to separate hydrogen and methane, which is sent to the hydroprocessing subunit. Because of the presence of the methanol/methyl acetate azeotrope, a further separation is not economically acceptable. Therefore, this stream is sent to the incinerator in the trigeneration unit. In the second circuit, the other part (e.g. 90%) of the liquid stream is diverted to an external heat exchanger where the stream is cooled down and steam is generated. The use of 2 circuits is implemented in order to reduce the required cooling capacity and the equipment size for evaporation of the internal solvent system. One consequence of this design is the up-concentration of the stabilized lignin components in the reaction solvent.

Separation. A liquid-liquid extraction is performed on the cooled crude lignin oil. Water is added to dissolve methyl xylose and other soluble sugar-derived components, whereas the lignin-derived components are extracted by dichloromethane (DCM). The DCM stream is distillated to enable the recycling of DCM and the isolation of the refined (or purified) lignin oil. The water fraction is concentrated by distillation. Recovered water is introduced back to the process. The bottom fraction, containing solubilized sugar-derived components is treated as wastewater.

In the second extraction and distillation step, lignin monomers are extracted from the refined lignin oil by use of n-hexane. Here, the heavy fraction contains highly concentrated lignin oligomers, which can be used for the synthesis of resins and varnishes (vide supra). After n-hexane evaporation, the light fraction containing mostly lignin monomers is further processed in the hydroprocessing subunit.

Hydroprocessing. In the hydroprocessing subunit the gaseous lignin monomers are converted into a mixture of alkylphenols (containing a small amount of alkylbenzenes), methane and water. More than stoichiometrically required hydrogen is added in the form of a sub-cooled gaseous stream coming from the RCF subunit and additional fresh hydrogen. The reaction will take place in a plug flow reactor with a nickel-based catalyst (e.g., 64 wt.% Ni/SiO2) at isothermal conditions (ca. 300 °C). The calculated heat of the reaction is -343.5 kJ mol-1

lignin monomer. For cooling of the stream, high pressure condensate is used. The generated steam is reused in other subunits to heat the distillation reboilers. The outgoing intermediate product (without any purification) is processed in the dealkylation subunit.

Page 18: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

18

Dealkylation and distillation. The dealkylation reaction is taken place in a fire heated tube reactor by using a zeolite catalyst at 400 °C and 2 bar with added superheated steam. The calculated reaction heat is 174.4 kJ mol-1. The intermediate product from the hydroprocessing subunit is preheated by the outgoing stream from the reactor. Part of the collected off-gas is used as a fuel.

The simulation indicates a challenging separation of phenol due to the presence of both cresol and water in the product mixture. To reach a phenol purity of >99% it is necessary to first decrease the amount of water under 1 wt.% before further purification of phenol. For propylene purification the cryogenic distillation was designed with a column top temperature of -129 °C. Methane, ethylene and unreacted hydrogen are separated in a partial condenser and send to the incinerator to provide heating, cooling, and electriciity through a trigeneration system. For the cooling of the cryogenic part, a liquid nitrogen loop with gas recompression is proposed.

2.10.2 Energy and environmental impact optimization To optimize the energy consumption, standard industrial solutions have been considered such as preheating the distillation feed by the bottom product, usage of multiple independent steam pressure levels, and low potential energy recuperation. Moreover, in the presented design the amount of wastewater is decreased by the incorporation of an additional distillation unit to recycle water to the separation unit. A trigeneration unit was also designed for a better overall energy recuperation, resulting in an energy saving of 285481 kWh. In the first unit, the produced off-gas is incinerated in a gas turbine to produce electric power. In the second and third unit, the produced high temperature flue gas is used partially for steam generation and partially for absorption cooling to feed the low temperature cooling loop for the cryogenic distillation (for propylene purification). Afterwards the cooled flue gas is removed as exhaust. Based on the simulation, this unit can be fully independent on the field of electric energy, steam, and cooling (table S8). 2.10.3 Economic analysis Assumptions. To analyze the economic profitability of the proposed wood biorefinery (to pulp, phenol, propylene and phenolic oligomers), the detailed factors method was used. The method estimates the costs for physical facilities and operating costs as percentages of the total equipment costs and revenues. The selection of equipment is based on reference (43). Prices are updated with current indexes. The construction costs are calculated as a percentage of the total equipment cost. Note that due to the use of hydrogen in the process, the standard construction material is stainless steel 316 to avoid hydrogen cracking. For the steam system and fire heated exchanger, Inconel is chosen as a construction material. Profitability factors

• Project length: 20 years • Operation time: 8766 h per year • Salvage value: 5% • Depreciation method: Straight line depreciation (allows equipment to be depreciated at a

constant rate. Each year the depreciation factor is 1/n, where n is years for depreciation) • Depreciation period: 10 years • Fixed costs, raw material and product prices remain constant over time

Total plant cost (fixed capital)

Page 19: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

19

• Equipment cost • Other direct costs (Specified as % of Equipment cost)

• Installation: 15% • Piping: 45% • Instrumentation: 10% • Building & Structure: 10% • Auxiliaries: 25% • Outside lines: 15%

• Indirect costs (Specified as % of Equipment cost + Direct costs) • Engineering & Construction: 30% • Contingency: 20%

Working capital: 30% of total revenue

Start-up expense: 7.5% of total plant cost

Cost of selling goods: 5% of total revenue

SARE (Sales Administration and Research Expenses): 10% of total revenue

Tax: 50% on income

Manufacturing costs

• Labor expenses - Operatives expense: 2629800 € (4 shifts of each 10 employees) - Supervision expense: 394470 € - Laboratory expense: 262980 €

• Office overhead: 50% of Operatives + Supervision expenses • Supplies: 2% of Total plant cost • Property tax: 5% of Total plant cost • Maintenance: 10% of Total plant cost

2.11 Life-cycle assessment Life-cycle assessment (LCA) is a technique using category indicators such as global warming potential (GWP) and terrestrial toxicity to assess the sustainability impacts. The goal of this study is to compare the GWP associated with the production of 1 kg phenol and 1 kg propylene from crude fossil oil (via the Hock process and steam cracking of naphtha, respectively) to that of the production of 1 kg phenol and 1 kg propylene from lignocellulose (birch wood chips) using the proposed integrated biorefinery. Besides this, the GWP values associated with the production of 1 kg phenolic oligomers and 1 kg carbohydrate pulp are investigated for this integrated biorefinery.

The system boundary of the study is cradle-to-gate, i.e. the analysis accounted for all stages in life-cycle of production for products, starting from the extraction of raw materials, transport of raw materials and further on the production of products. For phenol production from crude fossil oil (fig. S50), the system being considered for the study includes the extraction of crude oil, processing of crude oil to benzene and propylene, alkylation of benzene and propylene to cumene and phenol production by cumene oxidation and decomposition (Hock Process (44)). Likewise, extraction of crude oil and processing of crude oil to propylene are considered in the system boundary of

Page 20: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

20

propylene production from crude fossil oil (fig. S50). For phenol production from lignocellulose (fig. S51), the system includes forest management and harvesting of birch, processing of wood (including debarking, chipping, etc.) to wood chips, transport of wood chips, and the conversion of wood chips to phenol. The conversion process consists of reductive catalytic fractionation (RCF), separation, hydroprocessing and dealkylation/distillation. Note that conversion of phenol and propylene to final product, their use phase and end-of-life of the final product are not included in the system boundary because these phases are identical for both fossil-based and lignocellulose based phenol and propylene.

GWP of the phenol production from the two feedstocks was studied using GaBi 8.7.0.18 software. Ecoinvent v3.3 (45) and ThinkStep Gabi professional database 2016 (46) were used to determine the environmental impact of existing (current) technologies and production processes. The complete inventory of the Hock process (including the upstream processes, i.e. production of crude fossil oil) was available in the ThinkStep Gabi professional database 2016, and has been used for this analysis (47), whereas for the production of phenol from lignocellulose, ThinkStep Gabi professional database 2016, Ecoinvent v3.3 along with existing scientific literature was referred to complete the inventory. The Ecoinvent v3.3 database was used for the inventory of existing production processes (e.g. forest management and harvesting, methanol production, wastewater treatment, etc.) and existing flows (e.g. fresh water, air etc.). All the processes that were available in the Ecoinvent database have been marked in blue in fig. S51. All the processes and flows for which data were referred from scientific literature are marked in green. This includes GWP for ruthenium (48). The data for the production process of phenol from lignocellulose (birch wood) is determined from laboratory experiments, which is simulated to industrial scale accompanied by its techno-economic analysis (vide supra). This process includes reductive catalytic fractionation, separation, hydroprocessing and dealkylation/distillation (marked in orange in fig. S51). In the trigeneration process, the off-gases collected from other processes are incinerated to generate steam, cooling, and electricity, which are integrated to provide energy to other parts of the process. The excess of electricity can be sold to the distribution network. This is accounted for in the model as an avoided impact of producing electricity as to the EU-28 average electricity generation mix.

The analysis also included the treatment of wastes and emissions. The process produces wastewater, which is treated before being released into the environment. The exhaust that is produced at the end of trigeneration does not need to be treated and is released directly to the atmosphere. It contains mainly CO2 and N2. This study is representative on a European scale. The background data were collected for EU-28. Wherever the dataset for EU-28 was not available, a dataset with the geographical average across the world was used.

The assessment method used was CML 2001 - Jan 2016. GWP in CML 2001 considers a 100-year time-scale, includes biogenic carbon, and takes into account the land use change and represents GWP in kilograms of CO2-equivalents.

The production of phenol from birch wood also yields propylene, phenolic oligomers and carbohydrate pulp (which can all be further valorized). The total emission of the production process is allocated between products based on their economic value. The price and mass considered for each of the co-product are listed in tables S8 and S9, respectively.

The oligomers and carbohydrate pulp are outputs of the reductive catalytic fractionation (RCF) and separation process (fig. S51). Hence the emissions associated with reductive catalytic fractionation, separation, trigeneration, and the upstream process (i.e. forest management and

Page 21: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

21

harvesting) are allocated between the products of reductive catalytic fractionation (i.e. oligomers, carbohydrate pulp and monomers). The economic value of these products and the amount in which they are produced result in the allocation of environmental burdens among these co-products as follows: oligomers (25.6%), carbohydrate pulp (59.7%), monomers (14.7%).

Propylene and phenol are produced at the end of the dealkylation/distillation (fig. S51). Hence the emissions associated with the hydroprocessing, dealkylation/distillation and trigeneration are allocated between phenol and propylene. The economic value of these products (phenol and propylene) and the amount in which they are produced result in the allocation of environmental burdens among these co-products as follow: propylene (22.6%) and phenol (77.4%). The GWPs of phenol, propylene, oligomers, and carbohydrate pulp are calculated based on the allocation method.

The aim of the sensitivity analysis is to gain an overview of the parameters that have strong impacts on the results. It gives useful information about the sensitivity of the model to parameter uncertainties. In this study, a local sensitivity analysis was carried out with one-at-a-time (OAT) approach. In this OAT approach one input parameter is varied at a time, keeping other parameters fixed to see how much that parameter influences the results. The parameters were varied by 10%. A sensitivity analysis was also performed for the allocation method by varying the price of products (here, scope was limited to sensitivity of phenol). The sensitivity ratio (SR) was calculated for each parameter in the analysis for 10% variation. SR is the ratio of change in the final results to the change in the parameter.

𝑆𝑆𝑆𝑆𝑛𝑛𝑛𝑛𝑛𝑛𝑛𝑛𝑛𝑛𝑛𝑛𝑛𝑛𝑛𝑛𝑛𝑛 𝑟𝑟𝑟𝑟𝑛𝑛𝑛𝑛𝑟𝑟 = Δ𝑟𝑟𝑆𝑆𝑛𝑛𝑟𝑟𝑟𝑟𝑛𝑛𝑛𝑛

𝐼𝐼𝑛𝑛𝑛𝑛𝑛𝑛𝑛𝑛𝑟𝑟𝑟𝑟 𝑟𝑟𝑆𝑆𝑛𝑛𝑟𝑟𝑟𝑟𝑛𝑛𝑛𝑛Δ𝑝𝑝𝑟𝑟𝑟𝑟𝑟𝑟𝑝𝑝𝑆𝑆𝑛𝑛𝑆𝑆𝑟𝑟

𝐼𝐼𝑛𝑛𝑛𝑛𝑛𝑛𝑛𝑛𝑟𝑟𝑟𝑟 𝑝𝑝𝑟𝑟𝑟𝑟𝑟𝑟𝑝𝑝𝑆𝑆𝑛𝑛𝑆𝑆𝑟𝑟

Page 22: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

22

Supplementary Text ST1. The transition of fossil-based phenol to bio-phenol Phenol is a bulk chemical in today’s industry with a global annual production projected at 13.5 million tonnes in 2026 (49). Its main downstream use comprises the production of bisphenol A (46%), phenolic resins (28%), caprolactam (13%), aniline (3%), and alkylphenols (3%) (fig. S1). Industrial phenol production currently proceeds through the Hock process, involving exothermic autoxidation of cumene, obtained from benzene alkylation with propylene, followed by acid-catalyzed decomposition of cumene hydroperoxide into equimolar amounts of phenol and acetone (fig. S2) (44, 50). Besides the fossil, non-renewable nature of the feedstock, and the use of dangerous intermediates/catalysts, such as hydroperoxide (explosive) and sulfuric acid (corrosive), the overall phenol yield (on benzene per single-pass basis) in the current process is only 5% (51). Moreover, the overproduction of acetone is a potential market burden (50).

Lignocellulose, as an abundant feedstock of renewable carbon, is a prime candidate for production of renewable fuels, chemicals and materials (3, 4). In contrast to relatively oxygen-free fossil oil, oxygen-containing functional groups are plentiful in lignocellulose, but the latter is often too functional to match petroleum-derived chemicals and materials. Selective defunctionalization strategies, which compete with the usually less selective and exothermic functionalization steps in petrochemistry, are therefore required (52). Lignin, a natural alkyl-phenolic bio-polymer and second largest constituent in lignocellulose (4, 8, 14, 53), could produce vast amounts of bio-phenol, providing selective and industrially feasible lignin conversion strategies are available. Literature reports many (catalytic) routes to convert lignin into chemicals and fuels (14), but the on-purpose lignin-to-phenol route is studied less intensively (vide infra) (19, 54, 55). The main challenge lies in finding a strategy that combines (i) a high degree of lignocellulose delignification, with (ii) selective cleavage of CAr-O and CAr-Cα bonds (iii) without destructing the phenolic entity, while (iv) keeping the carbohydrates unchanged.

The few available lignin-to-phenol reports include thermal and catalytic hydrotreatment, but none of them are in commercial use due to low phenol yields. For instance, direct catalytic hydrogenolysis of technical lignin (cf. Noguchi process) gives 3 wt.% phenol yield (on the basis of lignin weight) (19). Phenol from Kraft lignin is reported in the LignolTM process by combining catalytic hydrocracking and non-catalytic thermal hydrodealkylation (fig. S2) (56). Here, formation of large amounts of heavy oil, light distillate, benzene, and light alkanes, likely caused by high temperature (350 to 450 oC) and pressure (up to 170 bar), restricts the yield of phenol (19, 56).

Recently, more practical lignin-first lignocellulosic biorefinery concepts were elaborated, based on the active stabilization of reactive intermediates obtained from in planta lignin to avoid irreversible condensation of as described in the manuscript (12-16, 57, 58). Unlike previous attempts, which used recalcitrant technical lignin, this in planta lignin depolymerization strategy produces a select number of methoxylated and alkylated phenolic monomers in close-to-theoretical yields, viz. 20 and 50 wt.% for soft- and hardwoods, respectively. Since the monomeric fraction only contains few and structural alike methoxy(alkyl)phenols, their isolation and conversion into phenol are practical to handle. Therefore, a distinct catalytic strategy is herein proposed that funnels the crude mixture of methoxylated and alkylated phenolic monomers, obtained from RCF of wood, into phenol and propylene (fig. S2). This funnel strategy involves hydroprocessing over a metal catalyst, followed by catalytic dealkylation over acidic hierarchical zeolites. The concept of a catalytic funnel approach is described in detail in fig. S4 and supplementary text ST2.

Page 23: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

23

ST2. Catalytic funneling approaches The concept of biological funneling was introduced recently using micro-organism to transform complex mixture of monomeric aromatic compounds, as typically produced by lignin depolymerization (such as by alkaline treatment), into a specific metabolite (11, 59).

In reference to this, fig. S4 shows the selective conversion of the mixture of lignin monomers obtained from RCF of wood through hydroprocessing into alkylphenols, which can be further quantitatively dealkylated into phenol and propylene in this work. Therefore, the mixture of phenolic monomers is selectively transformed into one single bio-phenol product with small amount of others such as cresols and benzene. This catalytic defunctionalization strategy may be regarded similar to the concept of biological funneling (11, 59), which we therefore denominated and emphasized as a ‘catalytic funneling’ strategy.

ST3. RCF birch wood phenolic oligomer characterization After the RCF of birch wood, less than 6% of native lignin linkages remain in the oligomers, proving the efficient cleavage of the ether linkages. Interestingly, the selectivity to the γ-OH substituted propyl end-groups is relatively high (fig. S5, table S2). This results in an unexpected high amount of aliphatic OH in the birch wood lignin oligomers, since the selectivity to γ-OH substitution of lignin monomers is less than 10% (table S1). Ongoing research efforts are directed toward understanding the difference in this selectivity.

ST4. Monomers extraction The high monomer extraction efficiency of n-hexane originates from the relative hydrophobic nature of the propyl phenolic monomers, whereas the oligomers are too functional to dissolve in n-hexane (fig. S6. and table S3). Propanol monomers, if formed (e.g., with other catalysts than Ru/C, such as supported Pd and Ni catalysts), are mostly retained in the extracted residue due to their more hydrophilic nature (fig. S7).

ST5. Hydroprocessing of phenolic monomers: the impact of metal type, loading, and support on selectivity The major concern of hydroprocessing is the selectivity. In this work, the formation of n-propylphenols is the key, while aromatic hydrogenation and loss of phenolic hydroxyl should be avoided as much as possible. To investigate the impact of the metal catalyst (type and loading), a model compound 4-n-propylguaiacol was hydroprocessed in a fixed bed reactor under selected reaction circumstances.

fig. S8A shows the conversion and selectivity to n-propylphenols for different metals (supported on Al2O3 and SiO2 and tested at similar reaction conditions). 51 wt.% Cu/SiO2, 5 wt.% Ru/Al2O3, 25 wt.% Co/Al2O3 and 5 wt.% Pd/Al2O3 give low (<10%) PG conversion, and are not further considered. Higher conversion (50%) is reached for 5 wt.% Rh/Al2O3, but with low selectivity to PPs and EPs (<20%), caused by formation of undesirable alkylbenzenes and heavy products (fig. S9A). By substituting Rh with Pt, both higher PG conversion (76%) and selectivity to PPs and EPs (60%) are obtained (figs. S8A and S9B). An excellent 86% selectivity to PPs and EPs is achieved for 17 wt.% Ni/Al2O3 (at 60% conversion). In summary, for the series of Al2O3 supported catalysts, Ni and Pt are clearly preferred, whereas Co, Ru and Rh show lower catalytic activity and/or selectivity. Though alumina-supported (5 wt.%) Pt gave higher conversion than 17 wt.% Ni , the former is less selective by producing propylcatechol and more alkylbenzenes (figs. S8A, S9B, and S9C). The presence of propylcatechol corresponds to the reported results that catechol is the major

Page 24: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

24

product in hydrodeoxygenation of guaiacol over Pt/Al2O3 (60). Higher Ni content, up to 64 wt.% (on SiO2), eventually leads to full conversion with 84% selectivity to PPs and EPs (figs. S8A and S9D). At similar conversion, 64 wt.% Ni/SiO2 gives a slightly higher selectivity than 17 wt.% Ni/Al2O3 (89 vs. 85%) (fig. S8A). Change of contact time did not change the order of catalyst selectivity performance (fig. S10 and Fig. 2C). Besides, non-noble metal catalyst are preferred for cost reason. Therefore, this research selected Ni as the preferred metal, and more tuning of the Ni catalysis was carried out for the hydroprocessing reaction.

To show the effect of activity, different Ni loadings (on SiO2 and Al2O3) were tested (Fig. 2B), and the nickel dispersion was determined by CO chemisorption. The results of the catalytic activity are plotted against the amount of surface Ni atoms in Fig.2B. As can be seen, the activity proportionally relates to the amount of the surface Ni atoms of the catalysts: the higher the amount of surface Ni atoms (per catalyst weight), as measured by CO chemisorption, the higher the catalytic activity (per catalyst weight). This means that Ni catalysts with high Ni loading and Ni dispersion are most preferred. Interestingly, all the Ni catalysts show similarly high selectivity (>80%) toward alkylphenols (fig. S8B). The results of the tested samples thus show no indication that surface sensitivity effects play a key role in the selective hydroprocessing of alkylguaiacols.

To show the effect of the support, the catalysis of Ni on SiO2, (γ-)Al2O3, and TiO2 (anatase phase) are compared (fig. S8B and Fig. 2C). It is known that (γ-)Al2O3 contains more acidic sites compared to silica. At the same conversion levels, silica supported Ni is always more selective (fig. S8B and Fig. 2C). Detailed inspection of the side products reveals substantial formation of transalkylation products on the acidic alumina support, whereas these products are formed less in the presence of Ni on silica. A comparison of the chromatograms is illustrative for the selectivity difference (fig. S8C). It has been demonstrated that the reducible support facilitates hydrodeoxygenation of oxygenates to hydrocarbon. Anatase TiO2 was therefore selected as an example to investigate the effect of reducibility on the selectivity. The selectivity to n-propylbenzene is 7.7% for 20 wt.% Ni/TiO2, which is higher than the 2.6% of 65 wt.% Ni/SiO2 at similar conversion (ca. 85%). Besides, the n-propylcyclohexane was observed for 20% Ni/TiO2 rather than 65 wt.% Ni/SiO2. Therefore, the non-acidic and non-reducible support (in this study is SiO2) are most preferred.

ST6. Hydroprocessing of phenolic monomers: process conditions Process conditions were varied for hydroprocessing of PG into n-propylphenols over 64 wt.% Ni/SiO2. The conversion of PG decreases with increasing the WHSV. While the selectivity to PPs and EPs slightly increases with increasing WHSV (fig. S10). Although the conversion of PG slightly increases with increasing the H2 partial pressure from 0.4 bar to 1 bar, this has no influence on selectivity to PPs and EPs (fig. S11). Since hydrogenation of aromatic can occur at higher hydrogen pressure (e.g. 5 bar) over Ni/SiO2 catalysts (61), the reaction was conducted at low hydrogen pressure (e.g. 0.4 bar) in this work. The temperature has a significant influence on the catalytic performance. Highest yields are obtained between 280 and 330 °C. Significant coke formation (and therefore more rapid catalyst deactivation) appears above 365 °C (fig. S12).

ST7. Hydroprocessing of phenolic monomers: mechanistic considerations The reaction network of the conversion of guaiacol into saturated chemicals like cyclohexane, cyclohexanol has been well studied. It usually proceeds through two parallel pathways: (i) saturation of aromatic ring, followed by methoxyl substituent removal, and (ii) removal of methoxyl substituent, followed by ring saturation (62-65). However, for conversion of alkyl

Page 25: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

25

substituted monomers (particularly, alkylsyringols) into aromatics, especially into alkylphenols, the reaction pathway is less studied and discussed. Therefore, the reaction pathway of PG and PS to n-propylphenols was investigated into more detail here.

To gain preliminary insight into the reaction network of hydroprocessing of PG to n-propylphenols, the influence of contact time (i.e. WHSV) was investigated by changing the amount of catalyst (fig. S15; see the above experimental part). As expected, the conversion of propylguaiacol decreases with increasing WHSV (decreasing contact time). The selectivity to n-propylphenols only slightly decreases with increasing contact time (and conversion) (fig. S15). This is accompanied by an increase of the selectivity toward cresols and ethylphenols, which indicates that cresols and ethylphenols are secondary products of n-propylphenols (via hydrodealkylation (66)). Similarly, the selectivity to n-propylbenzene also slightly increases with the contact time, which suggests that n-propylbenzene is derived from n-propylphenols. The n-propylcresols, and small amount of 5-n-propylguaiacol and 4-n-propylveratrol are likely formed by transalkylation (17). The same type of reaction has also been observed during hydroprocessing of guaiacol (60, 61). It was reported that n-propylphenol can undergo isomerization (for instance from 4-n-propylphenol to 3-n-propylphenol) over an acidic catalyst (22). However, 5-n-propylguaiacol is unlikely formed by isomerization (i.e. propyl chain shift) as 5-n-propylguaiacol was not observed when the reaction was conducted under N2 atmosphere over 64 wt.% Ni/SiO2. The small amount of n-propylanisole can be formed by dehydroxylation of n-propylguaiacol (61). Besides, it can be obtained from transmethylation of n-propylphenols.

Although the investigation of contact time variation on product distribution provides some mechanistic insights, the pathway that leads to n-propylphenols is not revealed. Earlier work on conversion of guaiacol to aromatics has shown that the C-O bonds were cleaved by either demethoxylation or demethylation followed by dihydroxylation (67-69). These were usually determined by identification of catechol/methane (i.e. demethylation product) or phenol/methanol (i.e. demethoxylation product). In this work, although methane was obtained as the main side product, 4-n-propylcatechol was not observed in the product stream over 64 wt.% Ni/SiO2. Meanwhile, methanol, when used as prime substrate, was fully converted under the same hydroprocessing circumstances (285 oC, 0.9 h-1 WHSV). Therefore, feedstock with different functional groups such as 4-n-propylcatechol, 4-n-propylanisole, 4-n-propylphenol were used to further investigate the reaction network (fig. S16). The catalytic results of 4-n-propylcatechol hydroprocessing display that it can be selectively (> 90%) converted into 3-n- and 4-n-propylpenols over 64 wt.% Ni/SiO2 (fig. S16). This indicates that demethylation is a possible pathway in the conversion of 4-n-propylguaiacol to n-propylphenols. Since isomers were not observed by conversion of 4-n-propylphenol with 64 wt.% Ni/SiO2 under N2 atmosphere at the same conditions, the 3-n-propylphenol was obtained by CAr-O bond cleavage at the para-position to the propyl side chain of 4-n-propylcatechol. This corresponds to the reported results (17). In order to further investigate the reaction pathway, 4-n-propylanisole was used as the feedstock. In contrast to 4-n-propylguaiacol, 4-n-propylanisole shows lower reactivity, which is consistent with the reported results of anisole and guaiacol conversion (70). This suggests that n-propylanisole is not the main intermediate to form n-propylphenols, otherwise a high selectivity to n-propylanisole would be expected. This is also consistent with the higher bond energy of the CAr-O bond of the hydroxyl group (414 kJ mol-1) compared to that of the CAr-O bond (356 kJ mol-1) and aliphatic C-O bond (245 kJ mol-1) of the methoxyl group in the monomeric guaiacyl unit (71). The presence of 4-n-propylphenol in the products of 4-n-propylanisole hydroprocessing indicates the occurrence of demethylation. The low selectivity to 4-n-propylphenol from 4-n-propylanisole also excludes

Page 26: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

26

the main contribution of n-propylanisole in the conversion of PG into n-propylphenols as high selectivity to n-propylbenzene would be obtained through this route. The selectivity to 4-n-propylphenol increases by decreasing the contact time in the hydroprocessing of 4-n-propylanisole, suggesting that the obtained 4-n-propylphenol was further converted to n-propylbenzene. This was also demonstrated by using 4-n-propylphenol as the substrate, showing formation of n-propylbenzene. In the case of 4-n-propylguaiacol, the demethylation product - 4-n-propylcatechol - was not observed. This is likely caused by a strong adsorption on silica of the highly polar 4-n-propylcatechol, which is further converted at the catalyst surface into n-propylphenols via dehydroxylation. This was also observed in gas-phase hydrodeoxygenation of guaiacol to aromatic hydrocarbon over Fe/SiO2 (72). In addition, 4-n-propylcatechol shows higher reactivity compared to propylguaiacol. For instance, the conversion of 4-n-propylcatechol and propylguaiacol is 88.2% and 64.5% at same reaction conditions, respectively. The molar yield of n-propylbenzene from 4-n-propylanisole reaches 23% and 17% at 285 oC with 8.2 h-1 and 16.4 h-1 WHSV, respectively. While the conversion of 4-n-propylphenol is 27% and 17% with 48% and 55% selectivity to n-propylbenzene at 285 oC with 7.4 h-1 (same mole flow of 4-n-propylanisole with 8.2 h-1 WHSV) and 14.8 h-1, respectively. These results show that the yield of n-propylbenzene from 4-n-propylanisole is around 40-50% higher than the yield from 4-n-propylphenol at same conditions (temperature and contact time). This difference can be explained by the direct demethoxylation of 4-n-propylanisole to n-propylbenzene, though the aliphatic C-O bond energy is lower than that of CAr-O bond energy of the methoxyl functionality. This demethoxylation is also observed in other catalytic systems for conversion of guaiacol (65, 68). Therefore, both direct demethoxylation and tandem demethylation-dehydroxylation contribute to the n-propylphenols formation from PG. Compared to PG, PS contains one more methoxyl group, which is rarely studied in literature. Although there are a few studies on the conversion of (allyl)syringol to (propyl)cyclohexane (62), the reaction pathway of the C-O bond cleavage and aromatic production were seldom revealed. Particularly, selective conversion of PS into n-propylphenols, the reaction of interest for this work, has not been investigated yet. Although several reports focused on catalytic transformation of PG to n-propylphenols in liquid phase with batch reactor (73, 74). unsatisfied results were obtained for hydroprocessing of PS.

In order to gain understanding of the formation of n-propylphenols from 4-n-propylsyringol, the reaction was conducted at 305 oC with varying contact times (i.e. WHSV). The conversion of PS is decreased with increasing WHSV (fig. S17). The selectivity to n-propylphenols is increased with decreasing WHSV, which is accompanied by the decrease of the selectivity to PG and 3-methoxy-5-n-propylphenol. The highest selectivity reaches 75% at full conversion of 4-n-propylsyringol and one methoxyl-substituted intermediates. This illustrates that PG and 3-methoxy-5-n-propylphenol are the intermediates in the reaction pathway of PS to n-propylphenols. The presence of 3-methoxy-5-n-propylphenol with 17% selectivity at high WHSV implies the occurrence of demethylation of PS, which produces 3-methoxy-5-n-propylbenzene-1,2-diol. Though 3-methoxy-5-n-propylbenzene-1,2-diol was not observed (at least with selectivity above 1%), it is likely further converted into either 3-methoxy-5-n-propylphenol or PG as a result of dehydroxylation. PG can also be obtained via demethoxylation of PS. The slight decrease of the selectivity to others might be ascribed to the presence of some side products e.g. methylated propylguaiacol derived from PS, which can be further transformed. Due to the small amount and complexity of these products, this was not studied in detail here. The selectivity to n-

Page 27: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

27

propylbenzene, cresols and ethylphenols is slightly increased with increasing contact time, which is similar to the results of PG hydroprocessing (vide supra).

Overall, n-propylphenols were obtained by tandem demethylation-dehydroxylation and direct demethoxylation pathways (fig. S18). Demethoxylation of PS obtains PG, while demethylation forms 3-methoxy-5-n-propylbenzene-1,2-diol (not observed in the products), which is likely converted into 3-methoxy-5-n-propylphenol and PG via dehydroxylation (fig. S18). For hydroprocessing of PG, direct demethoxylation yields 4-n-propylphenol. Demethylation of PG gives 4-n-propylcatechol (not observed in the products), which is dehydroxylated into 3-n-propylphenol and 4-n-propylphenol (fig. 18). In the case of 3-methoxy-5-n-propylguaiacol hydroprocessing, 3-n-propylphenol can be obtained either by direct demethoxylation or tandem demethylation-dehydroxylation pathways (fig. S18).

ST8. Dealkylation: results and discussion ST8.1. Catalyst selection Since the crude alkylphenol stream obtained from hydroprocessing contains impurities such as alkylbenzene and bulkier molecules such as propylcresols (table S4), commercial microporous ZSM-5 might be inadequate (fig. S22). More sterically demanding alkylphenols (e.g. propylcresols and 3-n-propylphenol), herein exemplified by 4-isopropyl-3-methylphenol (4-iPMP, model compound of propylcresols), were indeed more difficult to convert due to pore restrictions (vide infra). Uptake experiments indicated that such bulky molecules were limited in entering zeolite micropores, and hence accessing the active sites (fig. S23). Besides, n-propylbenzene was shown to lead to catalyst deactivation due to coking (75). To circumvent the site-access restriction and coke formation, hierarchization of a microporous ZSM-5 (Z140-P) by post-modification (see details in Methods section) was conducted to obtain a zeolite (Z140-H) with a balanced network of micro- and mesopores (figs. S22 and S24, table S5) for dealkylation of the crude alkylphenol stream.

ST8.2. Conversion of cresols Another group of side products in the products streams (from both hydroprocessing and dealkylation) are cresols. Dealkylation of pure cresols with Z140-H was shown to be suboptimal (fig. S30), because monomolecular removal of methyl from the aromatic core is significantly more difficult than the removal of ethyl and propyl. Cresols are thus best separated after dealkylation. Hereto, existing technology is available, e.g. by batch distillation as applied in coal tar processing (44). If desirable, this separated stream of cresols can be converted to phenol and xylenols in an additional step (not included in Fig. 1) by using large micropore zeolites (e.g. USY) and bimolecular pathways such as disproportionation (76). For instance, a commercial USY zeolite (Si/Al=40) showed up to 90% of the theoretical selectivity with a thermodynamically limited conversion of about 57% (figs. S30E-F). Overall, integration of this cresol upgrading step can afford a phenol yield improvement of 5%, while xylenols can be isolated via distillation and used as antioxidants (77).

ST8.3. Dealkylation of alkylphenols In order to demonstrate the key role of hierarchization, dealkylation of model substrates (4-isopropyl-3methylphenol, PPs and EPs) was studied in detail. As illustrated for sterically hindered 4-isopropyl-3-methylphenol, Z140-H clearly outperformed the parent microporous ZSM-5 catalyst (Z140-P) in terms of conversion rate and stability without loss of selectivity (figs. S32). Likewise, (bulky) isomers of n-PP and EP in the crude stream can be selectively dealkylated to

Page 28: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

28

phenol and olefins over Z140-H at high temperature (figs. S33-S34). Besides a minor initial deactivation, likely due to coking on the strongest acid sites, stable conversion of PPs (92%) at excellent selectivity to phenol and propylene (>97%) was achieved with Z140-H (fig. S33D). Although large microporous zeolites such as USY are capable of transforming sterically demanding alkylphenols, they are lacking the (transition-state) pore confinement for shape-selective conversion (22). As a result, such zeolites show poor selectivity performance forming more side products (e.g. cresols) due to occurrence of disproportionation, transalkylation and C-C cracking (fig. S31). The confinement governed by the micropores within Z140-H is thus essential to achieve high selectivity.

ST8.4. Dealkylation of n-propylbenzene n-Propylbenzene is an impurity in the crude alkylphenol stream. The earlier reports mentioned that commercial (non-hierarchical) ZSM-5 catalysts are not stable due to severe coke formation in the dealkylation of n-propylbezene (75). The commercial ZSM-5 (Si/Al=12) is indeed not stable for dealkylation of n-propylbenzene (fig. S35), but n-propylbenzene is stably converted to benzene and propylene over Z140-H (fig. S35). These results underline the importance of hierarchical pore structure (facilitating the diffusion of coke precursor) and/or acidity modification(reducing the coke formation with low acidity) to achieve stable and robust catalysis (fig. S35).

ST8.5. Light-off dealkylation experiments (figs. S30, S32-S35) Catalytic stability tests show that Z140-H is slight deactivation at the initial stage and then stable during dealkylation of alkylphenolic model compounds (i.e., 4-isopropyl-3-methylphenol, 4-n-propylphenol, 4-ethylphenol, cresol), n-propylbenzene, and real alkylphenols (figs. S30, S32-S35). These results indicate the reliability of the light-off experiments (increasing the temperature from 200 oC to 475 oC with 1 oC min-1 to evaluate the catalytic performance).

ST9. Semi-simultaneous saccharification-fermentation of RCF birch wood carbohydrate pulp The carbohydrate pulp (containing Ru/C catalyst), obtained after RCF of birch, was subjected to semi-simultaneous saccharification fermentation (SSF) under high solid content conditions (15 wt.%). During the first 24 h, enzymatic saccharification was performed using CTEC2 as the enzyme mixture, resulting in 36.8 g L−1 of glucose and 22.3 g L−1 of xylose (Fig. 3A). Subsequently, MDS130 yeast suspension was added for the fermentation of the hydrolysate toward bio-ethanol. At this stage, further enzymatic saccharification occurs in parallel. After 264 hours, an ethanol titer of 40.2 g L−1 is reached. Though further optimization of the SSF set-up and protocol has to be performed in order to increase the process productivity, this demonstration shows the potential for fermentative conversion of RCF carbohydrate pulps (containing Ru/C) into bio-ethanol at high solid content. Ongoing research efforts are also directed toward the production of a catalyst-free RCF pulp and an improved understanding and control over inhibiting factors.

ST10. Synthesis of ink varnishes from RCF birch wood lignin oligomers Besides monomers (for conversion to bio-phenol and bio-propylene), the RCF birch wood lignin oil consists of ca. 30 wt.% oligomers. To demonstrate its valorization potential, the oligomers were used as (high-value) ingredient for the synthesis of ink varnishes for lythographic printing ink. Fig. S38 summarizes a step-wise scheme, which involves the synthesis of a rosin-modified resin, followed by an ink varnish formulation.

Page 29: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

29

Currently, fossil alkylphenol-formaldehyde resole is used for the preparation of rosin-modified phenolic resin (41). The resole part, prepared from para-nonylphenol, cardanol and (para)formaldehyde, is linked to rosin acids (e.g. abietic acid, levopimaric acid). Polyols such as glycerol or pentaerythritol are added to form a cross-linked network. Additionally, maleic acid adducts can be incorporated through the addition of maleic acid anhydride, which allows for additional cross-linking (fig. S39). For more information, the reader is referred to reference (41).

The para-nonylphenol in the resole part can possibly be substituted by RCF oligomers. para-Nonylphenol is currently obtained from fossil resources, and even more importantly, is a known endocrine disrupter (25). To evaluate the complete substitution of para-nonylphenol by RCF lignin oligomers, a resin and ink varnish were prepared, and their properties compared to resins/varnishes from para-nonylphenol (commercial reference), commercial acetosolv spruce wood lignin, and self-prepared methanosolv birch wood lignin (GPC results see fig. S40).

Resin synthesis was successful for para-nonylphenol, methanosolv birch wood lignin (the preparation method was mentioned in the caption of table S6), and RCF birch wood lignin oligomers. For the commercial acetosolv spruce lignin, however, the filtration step yielded a (too) large amount of particles so that the filter eventually blocked. Because of the high particle amount, the tested commercial organosolv lignin is not suitable for the envisioned application. Therefore, further testing was not representative.

The evaluation of the resins prepared from the other three phenol sources is summarized in table S6. During the resin synthesis, the viscosity was monitored in function of the reaction time. In case of the methanosolv birch wood lignin, the viscosity increase was slow, which suggests that the methanosolv birch wood lignin is not effectively incorporated in the resole chains. Therefore, vacuum was applied to increase the viscosity as to reach the appropriate specifications (table S6). The application of vacuum though is less preferable for industrial implementation and should be kept minimal. Moreover, the methanosolv birch wood lignin based resin yielded a relatively high amount of particles upon filtration (314 ppm), which is beyond commercial specifications (< 200 ppm).

For the RCF birch wood lignin oligomers, the required vacuum time and residue on the filter were within the acceptable ranges, indicating that the RCF birch wood lignin oligomers were effectively incorporated. The rheology of the RCF birch wood lignin oligomers based resin was comparable to that of the para-nonylphenol based resin (table S6); both the viscosity and flow-behaviour index were similar. In addition, the acid value was determined (41), which is preferably lower than 30 to ensure stability of the varnish (vide infra). This was the case for the RCF birch wood lignin oligomers based resin (as well as for the methanosolv birch wood lignin based resin). Fig. S41 shows the synthesized resin. Finally, GPC of the resins is presented in fig. S42. The molecular weight distribution of the different resins is comparable, except for the high Mw shoulder in the chromatogram of the para-nonylphenol based resin.

In addition, ink varnishes were prepared from the three self-synthesized resins, which were compared with an ink varnish based on commercial resins (table S7). Most importantly, the emulsification capacity was assessed, which demonstrated that all four varnishes could form a stable emulsion (fig. S43). The varnishes showed a similar water balance. In addition, the rheology was assessed. The methanosolv birch wood lignin and RCF birch wood lignin oligomers based varnishes showed a slightly lower elasticity compared to the commercial and para-nonylphenol based reference. Finally, it was noted that the colour of the lignin-based varnishes was somewhat

Page 30: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

30

darker than the commercial references (fig. S44). Adjusting this dark colour is set as an objective for future work.

Overall, it was shown that the RCF birch wood lignin oligomers fraction is a viable alternative for para-nonylphenol, and outperforms the commercial acetosolv lignin and self-prepared methanolsolv lignin.

ST11. Loss of methanol in birch wood conversion The total loss of methanol in this integrated biorefinery was estimated. This loss is caused by (i) hydrogenolysis over Ru/C during RCF, (ii) incorporation into products (solubilized sugars, methyl acetate, lignin oil), and (iii) recycling losses during distillation. The gas phase products of RCF were quantified by GC, and around 0.24% of methanol was converted into CH4. Conversion of acetyl groups consume 0.71% of methanol to methyl acetate. 0.35% of methanol was incorporated into the solubilized sugars and lignin oil. The simulation of methanol distillation showed that 0.06% of methanol is lost. Overall,1.4% of methanol is consumed in this integrated biorefinery. The mass flow of methanol loss can be found in fig. S47, which shows that 61 kg methanol is required for 1 tonne birch wood.

ST12. Techno-economic analysis for this integrated biorefinery ST12.1. CAPEX (Capital expenditure) Based on the developed model and the above described economic assumptions, CAPEX for this integrated biorefinery was calculated. CAPEX or capital expense is the cost to build and maintain new facilities. In general this is the cost for developing or providing non-consumable parts for an installation. For the designed biorefinery, the total CAPEX was calculated at 477 MM€. The distribution of the capital expenditure between the different subunits is provided in fig. S48. The largest contributors for the capital expenditure are the trigeneration unit and the RCF unit, constituting 40.6% and 49.0%, respectively. Investing in a trigeneration unit however is justified by its positive impact on the manufacture cost (internal production of steam, cooling power, and electricity). The relatively high CAPEX cost of RCF is related to high pressure reactor equipment and the presence of hydrogen throughout the process.

ST12.2. Sensitivity analysis on the design and cost assumptions A sensitivity analysis was performed to evaluate the influence of several design and cost assumptions on the overall profitability of the process. Fig. S49 shows the tornado diagram in which the deviation in the internal rate of return is plotted when varying a single parameter, compared to the basic case. The parameters for the basic case are defined in the caption of fig. S49.

The internal rate of return for the basic case is 23.33%, resulting in a payout time of 4.19 years. The biggest impact on the pay-out time results from the variation of the product price and the feedstock price. If the product price decreases by 50% or the feedstock cost increases with 200% the operation will be close to a break even, and thus non-profitable (negative internal rate of return). On the other hand if the product price increases by 200% or feedstock prices drop by 50%, the internal rate of return increases with 140% or 63%, respectively. With regard to the processing parameters, most gains can be made in the RCF subunit, either by reducing the methanol-to-wood ratio or by lowering the required reaction time.

ST13. Life-cycle assessment

Page 31: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

31

The global warming potentials (GWPs) of fossil-based phenol, propylene, and nonylphenol (oligomers were used to replace it in the printing-ink production in this study) are shown in table S12. The GWPs of phenol, propylene and nonylphenol from fossil oil are 1.73, 1.47 and >1.58 kg of CO2-equivalent per kg, respectively. The GWP of nonylphenol is roughly calculated, which is explained in the caption of table S12. Table S13 shows that the GWP of bio-phenol from lignin in this integrated approach is 0.736 kg of CO2-equivalent per kg of phenol, which is much lower than the fossil-derived phenol (1.73 kg of CO2-equivalent per kg of phenol). The reductive catalytic fractionation process shows a negative GWP for bio-phenol as the lignocellulose feedstock is a carbon stock. The main contributor of the GWP of bio-phenol is from conversion of monomers fraction to phenol (via hydroprocessing, dealkylation and distillation), which is attributed to the non-renewable H2 (GWP=8.20 kg of CO2-equivalent per kg of H2) used for hydroprocessing to obtain partial deoxygenated products-alkylphenols. Similarly, bio-propylene has a GWP of 0.469 kg of CO2-equivalent of per kg of propylene, which is also much lower than the fossil-derived propylene (table S14). The GWPs of oligomers and carbohydrate pulp obtained from reductive catalytic fractionation process are -0.949 and -0.217 kg of CO2-equivalent per kg, respectively. The oligomers show a great potential to replace para-nonylphenol in term of greenhouse gas emission.

This integrated biorefinery to make phenol has two major differences compared to Hock process, i.e. renewable carbon source (wood) vs. non-renewable carbon source (oil) and defunctionalization (e.g. hydroprocessing) vs. functionalization (e.g. oxidation). Beside, this study indicates that both lignocellulose and H2 have significant effect on the GWP of phenol. Therefore, different scenarios where H2 is from different source and the forest is more sustainable managed are investigated.

By using the hydrogen from steam reforming but with a higher GWP (11.89 kg of CO2-equivalent per kg of H2, non-renewable H2 I in fig. S52A-B, which is the highest case in the literature for H2 production (78)), the GWPs of bio-phenol and bio-propylene increase to 1.354 and 0.864 kg of CO2-equivalent per kg, respectively, which are still lower than the fossil-based phenol and propylene. In addition to steam reforming of natural gas to produce H2, electrolysis of water is another way to make H2. If the electricity from renewable energy such as wind energy can be used for electrolysis, the GWP of H2 (renewable H2 III in fig. S52A-B) is 0.97 kg of CO2-equivalent per kg of H2 (79). The GWPs of bio-phenol and bio-propylene decrease to -0.475 and -0.303 kg of CO2-equivalent per kg, respectively. The negative sign indicates that there is net carbon sequestration as the result of using renewable biomass. Since the oligomers and carbohydrate pulp are produced in the reductive catalytic fractionation step, the GWP is more negative but less influenced by different H2 sources. These results indicate that the contribution of H2 in the greenhouse gas emission of biorefinery should be taken into account to evaluate the sustainability.

To aim for sustainable development, the forest should be sustainably managed to keep the balance between ecological, economic and socio-cultural aspects. For the process of lignocellulose production (i.e. forest management, harvesting, wood chipping, transport), the process used from the Ecoinvent v3.3 database is an average over the European continent (without sustainable forest management in fig. S52C- D). However, in the scenario where the sustainable forest management strategies are applied, the benefits of this process could be much higher. This is exhibited in fig. S52C-D. The GWP of phenol from the reductive catalytic fractionation decreases from -0.772 to -4.64 kg of CO2-equivalent per kg of phenol, which leads to the total GWP of phenol decreased to -3.135 from 0.736 kg of CO2-equivalent per kg of phenol. Similarly, the GWP of bio-propylene decreases to -2.00 kg of CO2-equivalent per kg of propylene. As expected, the GWPs of oligomers

Page 32: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

32

and carbohydrate significantly decrease to -6.64 and -1.52 kg of CO2-equivalent per kg, respectively. Hence, the choice for using the average European number can be called conservative. The GWPs of bio-phenol, bio-propylene, oligomers and carbohydrate pulp can be improved to -4.35, -2.77, -6.95 and -1.59 kg of CO2-equivalent per kg, respectively, by using sustainable forest management strategies and renewable H2 simultaneously. Fig. S53 shows the sensitivity ratio (SR) for each parameter with -10% variation. The SR is less than 1 for all parameters except lignocellulose (i.e. the amount of wood) and the hydrogen consumed in hydroprocessing (assuming that the rest H2 after hydroprocessing and products from hydroprocessing used for energy are same). This indicates that change in the value of a parameter by 1%, changes the results by less than 1%. However, variations in the amount of lignocellulose and H2 amount in the hydroprocessing have significant impacts on the final results. The decrease in lignocellulose amount by 1% increases the GWP of the process by approximately 1.8%. This suggests that the value for the amount of lignocellulose needed in the process should be accurately known before concluding the GWP. The decrease in hydrogen amount consumed in hydroprocessing by 1% decrease the GWP of the process by approximately 1.5%.

The sensitivity analysis of the allocation method (via changing the economic value of co-products) is shown in fig. S53B. The price of propylene, oligomers, carbohydrate pulp or all of them was decreased by 10%. The SR is less than 1, indicating that a 1% decrease in the price of oligomers, carbohydrate pulp, or all of them decrease the GWP allocated to phenol by less than 1%. While a 1% decreases in the price of propylene increases the GWP allocated to phenol by less than 1%.

Overall, phenol, propylene and oligomers (to replace para-nonylphenol) production from birch wood have much lower global warming potential (including formation of the biogenic CO2) than phenol, propylene and para-nonylphenol production from crude fossil oil. The carbohydrate pulp has a negative GWP. The sensitivity analysis (only on phenol) highlights that the GWP results of phenol are most affected by the ‘amount of lignocellulose’. If the oligomers, carbohydrate pulp or all co-products are valued lower than in the current analysis, the GWP allocated to the phenol production decreases further.

ST14. Carbon flow For the carbon flow: birch wood components: cellulose C6H10O5, hemicellulose C5H8O4, lignin C10.75H13.50O3.75. The molecular formula of birch wood lignin is determined by the ratio of guaiacyl (G) unit and syringyl (S) unit in lignin (G/S=1:3). Carbon efficiency is the proportion of carbon in the products (carbohydrate pulp, propylene, phenol and oligomers) to the initial carbon in cellulose, hemicellulose, and lignin. The total carbon efficiency of this integrated biorefinery is 77.6%.

Page 33: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

33

Figures

Fig. S1. The global demand pattern of phenol and propylene and the suggested application of bio-phenol, bio-phenolic oligomers, and bio-propylene. For production of polypropylene, the minimum purity of propylene is 99.5% (polymer grade), while lower purity can be used for the production of chemicals.

Page 34: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

34

Fig. S2. Routes to phenol from (i) fossil oil by the current industrial Hock process, (ii) technical (isolated) lignin by the LignolTM process, and (iii) in planta lignin by the proposed process.

Page 35: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

35

Fig. S3. Reproducibility and impact of birch type on the outcome of the reductive catalytic fractionation of wood. Klason lignin content (wt.%) in the wood; Phenolic monomers and oligomers yield (wt.%) on the basis of Klason lignin weight.

Page 36: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

36

Fig. S4. Integrated production of phenol from lignocellulose via reductive catalytic fractionation and chemo-catalysis (catalytic funneling). Fractionation of lignocellulose can produce streams of lignin monomers (to yield bio-phenol and bio-propylene), lignin phenolic oligomers (to produce bioresins, varnish and ink) and a carbohydrate pulp (e.g., for bio-ethanol production).

Page 37: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

37

Fig. S5. Aliphatic regions of 2D HSQC NMR spectra from the RCF birch wood lignin phenolic oligomers.

Page 38: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

38

Fig. S6. 31P-NMR spectrum of phosphitylated RCF birch wood lignin phenolic oligomers derivatized with TMDP and with cholesterol as internal standard.

Page 39: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

39

Fig. S7. Systematic extraction study to isolate monomers from birch wood lignin oil using different n-hexane/lignin oil mass ratios. (A) Total amount of n-hexane extracted monomers against intial lignin oil composition, with indication of the upper-limit of extraction; (B) Extraction efficiency in function of the solvent usage and number of extraction steps; (C) GPC of the initial lignin oil (in black), and the extracted residue (in red) and extract (in blue) after extraction with 1:1 (w/w) n-hexane to lignin oil (four times). Note the propanol monomers situated around 290 g mol-1; (D) Pictures of extracts. Legend: n-hexane-unextractable monomers are 4-propanolguaiacol and 4-propanolsyringol as most of them remains in the residue, while n-hexane-extractable monomers are the other monomers. The extraction efficiency is defined as the amount of n-hexane extracted monomers divided by the total amount of n-hexane-extractable monomers, expressed as percentage.

Page 40: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

40

Fig. S8. Hydroprocessing of 4-n-propylguaiacol to n-propylphenols and ethylphenols: metal, loading, and support effects. (A) Influence of different metal catalysts (285 oC and 4.5 h-1 WHSV, unless indicated otherwise). (B) Influence of different nickel catalysts (285 oC). (C) Gas chromatograms of conversion of 4-n-propylguaiacol over 17 wt.% Ni/Al2O3 and 64 wt.% Ni/SiO2 (at similar conversion, ca. 62%). Reaction conditions: 1 bar of total pressure (0.4 bar of H2 partial pressure). The data are taken at time-on-stream (TOS) of 5 h.

Page 41: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

41

Fig. S9. Gas chromatograms of conversion of 4-n-propylguaiacol (PG) over different catalysts. (A) 5 wt.% Rh/Al2O3, (B) 5 wt.% Pt/Al2O3, (C) 17 wt.% Ni/Al2O3, (D) 64 wt.% Ni/SiO2. The small signals between the methane and n-propylbenzene peak are related to (alkyl)benzenes (toluene, ethylbenzene, etc.) whereas the small signals between the n-propylbenzene and n-propylphenols peak are related to some alkylphenols (cresols, ethylphenols, etc.). Reaction conditions: WHSV=4.5 h-1, 285 oC, 1 bar of total pressure (0.4 bar of H2 partial pressure).

Page 42: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

42

Fig. S10. The influence of WSHV in the conversion of propylguaiacol. (A) Conversion of propylguaiacol as a function of WHSV over different catalysts; (B) The selectivity to n-propylphenols (and ethylphenols) as a function of conversion of propylguaiacol. The conversion was changed by varying the WHSV (in A). In (B), hollow symbols: n-propylphenols, solid symbols: n-propylphenols and ethylphenols. Reaction conditions: 285 oC, 1 bar of total pressure (0.4 bar of H2 partial pressure).

Page 43: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

43

Fig. S11. The influence of H2 partial pressure in the conversion of propylguaiacol over 64 wt.% Ni/SiO2. Conversion of propylguaiacol and selectivity to n-propylphenols (and ethylphenols) as a function of H2 partial pressure. Reaction conditions: WHSV= 9.0 h-1, 285 oC, 1 bar of total pressure.

Page 44: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

44

Fig. S12. The influence of reaction temperature in the conversion of propylguaiacol over 64 wt% Ni/SiO2. (A) Conversion of propylguaiacol as a function of time-on-stream (TOS) in the conversion of propylguaiacol over 64 wt.% Ni/SiO2 at 285 oC or 365 oC; (B) Conversion of propylguaiacol and selectivity to n-propylphenols (and ethylphenols) over 64 wt.% Ni/SiO2 at different temperatures, the inset highlights coke formation after 5 h. In (B), the data are obtained at time-on-stream (TOS) of 5 h. Reaction conditions: WHSV=18 h-1, 1 bar of total pressure (0.4 bar of H2 partial pressure).

Page 45: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

45

Fig. S13. Stability test and catalyst regeneration for the conversion of propylguaiacol over 64 wt.% Ni/SiO2. (A) Regeneration test. Reaction conditions: 285 °C, 1 bar of total pressure (0.4 bar of H2 partial pressure). The reduction was conducted at 285 oC for 3 h in the reactor by changing gas feedstock. (B) Peak area of substrate and products as analyzed by GC over a 64 wt.% Ni/SiO2 catalyst as a function of time-on-stream in different conditions (the labels indicate the flows present). The total flows of 100% N2 and N2+H2 are same. T = 285 oC. (C) H2-TPR test for the spent 64 wt.% Ni/SiO2 catalyst.

In fig. S13B, 64 wt.% Ni/SiO2 was first evaluated in the conversion of propylguaiacol. After 20 h, the reactor was flushed with nitrogen (no hydrogen) for several hours at 285 oC. During this period, the products and substrate reduced almost to zero confirming that all weakly-adsorbed species were removed. Then, hydrogen was introduced into the stream, which induced an immediate appearance of methane (likely due to methanation of coke precursors). The result of H2-TPR of spent catalyst also shows the evolution of CH4 (fig. S13C) above 250°C. These results indicate that catalyst deactivation is caused by fouling.

Page 46: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

46

Fig. S14. Gas chromatograms of different substrates conversion over 64 wt.% Ni/SiO2. (A) 4-ethylguaiacol (EG), (B) 4-n-propylguaiacol (PG), (C) isoeugenol, (D) 4-n-propylsyringol (PS). The small signals between the methane and n-propylbenzene peak are related to (alkyl)benzenes (toluene, ethylbenzene, etc.), whereas the small signals between the n-propylbenzene and n-propylphenol peak are related to alkylphenols (cresols, ethylphenols, etc.). Reaction conditions: EG, PG, and isoeugenol at 285 oC and 8.2, 6.0, and 4.4 h-1 WHSV, respectively. PS(WHSV=6.4 h-1) and PS(WHSV=5.3 h-1) at 305 oC and 6.4 h-1, and 5.3 h-1 WHSV, respectively. 1 bar of total pressure (0.4 bar of H2 partial pressure).

Page 47: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

47

Fig. S15. Hydroprocessing of 4-n-propylguaiacol over 64 wt.% Ni/SiO2 at different WHSVs. Evolution of the conversion and product selectivity with increasing WHSV. Reaction conditions: 285 oC, 1 bar of total pressure (0.4 bar of H2 partial pressure). The data are taken at time-on-stream (TOS) of 5 h.

Page 48: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

48

Fig. S16. Hydroprocessing of different model compounds (4-n-propylcatechol, 4-n-propylanisole, and 4-n-propylpheol) over 64 wt.% Ni/SiO2 to understand the main reaction pathway of hydroprocessing step. Reaction conditions: 285 oC, 1 bar total pressure (0.4 bar of H2 partial pressure). The data are taken at time-on-stream (TOS) of 5 h. WHSV is indicated in the figure.

Page 49: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

49

Fig. S17. Hydroprocessing of 4-n-propylsyringol over 64 wt.% Ni/SiO2 at different WHSVs. Evolution of the conversion and product selectivity with increasing WHSV. Reaction conditions: 305 oC, 1 bar total pressure (0.4 bar of H2 partial pressure). The data are taken at time-on-stream (TOS) of 5 h.

Page 50: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

50

Fig. S18. The reaction network of hydroprocessing of (A) 4-n-propylguaiacol and (B) 4-n-propylsyringol over 64 wt.% Ni/SiO2.

Page 51: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

51

Fig. S19. Gas chromatograms of the crude unseparated mixture of pine (A) and birch (B) wood lignin monomers before and after hydroprocessing over 64 wt.% Ni/SiO2. Reaction conditions: for pine: 285 oC, WHSV=6.0 h-1; for birch: 305 oC, WHSV=5.3 h-1, 1 bar of total pressure (0.4 bar of H2 partial pressure).

Page 52: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

52

Fig. S20. Gas chromatograms of light fraction (gas, incondensable at 0-5 oC) in the conversion of 4-n-propylguaiacol over 64 wt.% Ni/SiO2. (A) GC-FID; (B) GC-TCD. Reaction conditions: 285 oC, WHSV=6.0 h-1, 1 bar of total pressure (0.4 bar of H2 partial pressure).

Page 53: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

53

Fig. S21. Gas chromatogram of the condensate from hydroprocessing products. It is analyzed by offlline GC.

Page 54: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

54

Fig. S22. Schematic illustration of access and diffusion characteristics of small (compared to the micropores) and bulky molecules in the microporous (left) and hierarchical (right) ZSM-5. Hierarchization of ZSM-5 improves the number of pore mouths to (i) access the active sites for bulky molecules, and (ii) diffuse in and out zeolite crystal for small molecules. Besides, the diffusion path of small molecules in the hierarchical ZSM-5 is shorter compared to that in the original microporous ZSM-5. A detailed description of hierarchical zeolite can be found in review (80).

Page 55: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

55

Fig. S23. Uptake of 4-n-propylphenol and 4-isopropyl-3-methylphenol at 40 oC on ferrierite (FER), ZSM-5 (Z40-P, Z140-P and Z140-H) and USY (USY-40) zeolites.

Page 56: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

56

Fig. S24. The N2 isotherms (A), DFT pore size distribution (B), and NH3-TPD profiles (C) of Z140-P and Z140-H.

Page 57: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

57

Fig. S25. The Gibbs free energy change (ΔG) of dealkylation of n-propylphenols as a function of temperature. This calculation assumes that the ΔS of reaction is independent on the temperature (namely, ΔS=ΔS0 (20 oC)), and ΔH=ΔH0 (20 oC). ΔG=ΔH-TΔS. The parameters of all chemicals are adapted from MOL-Instincts DATABASES (81). The themodynamic study indicates that the continuous catalytic dealkylation should be conducted at sufficiently high temperature to favor the equilibrium conversion at atmospheric pressure, as dealkylation is a reversible and endothermic reaction.

Page 58: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

58

Fig. S26. Dealkylation of the hydroprocessing products from the extracted monomers of pine wood lignin oil over Z140-H at different temperatures. Reaction conditions: WHSV=3.7 h-1, 1 bar, ramping rate is 1 oC min-1. C-mol yield represents the carbon molar yield in the product stream.

Page 59: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

59

Fig. S27. Dealkylation of the hydroprocessing products from the extracted monomers of birch wood lignin oil over Z140-H. Reaction conditions: 410 oC, WHSV=3.7 h-1, 1 bar. C-mol yield represents the carbon molar yield in the product stream.

Page 60: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

60

Fig. S28. Gas chromatograms of pine-derived crude alkylphenols before (red) and after (black) dealkylation over Z140-H at 470 oC.

Page 61: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

61

Fig. S29. Gas chromatogram of the condensate from dealkylation of the hydroprocessing products of birch wood lignin crude monomers. It is analyzed by offline GC.

Page 62: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

62

Fig. S30. Catalytic conversion of cresols over zeolites. (A) Conversion of cresols and selectivity to phenol in the conversion of cresols over Z140-H; (B) The catalytic stability of Z140-H in the conversion of cresols. Isomerization was used as the criterion for measuring the stability; (C) The conversion of cresols and selectivity to phenol in the conversion of cresols over Z40-P (Si/Al=40); (D) Gas chromatgram of cresols conversion over Z40-P; (E) The conversion of cresols and selectivity to phenol in the conversion of cresols over USY-40 (Si/Al=40); (F) Gas chromatgram of cresols conversion over USY-40. Selectivity (%)=yield of products/theoretical yield ×100%. Ramping rate=1 oC min-1 in (A), (C), and (E). Temperature in (B) is 410 oC, temperature in (D) and (F) is 470 oC. Reaction conditions: WHSV=2.9 h-1, molar ratio of H2O to 4-methylphenol=6, 1 bar. Since methylphenol (cresol) undergoes isomerization, the reported conversion in (A, C, and E) is the conversion of all isomers.

Page 63: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

63

Fig. S31. The reaction network of dealkylation of 4-n-propylphenol over acid zeolites. The isomers of 4-n-propylphenol, other alkylphenols (like ethylphenols and propylcresols) in the crude alkylphenols can undergo similar reactions. For ZSM-5 zeolites: the reactions are monomolecular reactions (i.e. dealkylation, isomerisation, cracking). Both monomolecular and bimolecular reactions are favoured in the micropores of large micropore zeolites such as USY. However, the bimolecular reactions (i.e. disproportionation and transalkylation) are strongly inhibited over ZSM-5 zeolites as the result of the confinement effect of micropores. The occurrence of bimolecular reactions will decrease the selectivity to phenol and propylene as the formation of large amount of cresols, which cannot be dealkylated into phenol (21).

OH

IsomerisationOH

DealkylationOH OH

+

Dealkylation

OH

OHOH

Disproportionation Dealkylation

OH

OH

+

Transalkylation

Catalytic cracking

OH+ Thermal cracking

OH

OHOH

+

OHOH

+

Dea

lkyl

atio

n

OHOH

Disproportionation/C-C cracking

Transalkylation/C-C cracking

OH

Transalkylation/C-C crackingsame reactions

of n-propylphenol

Monomolecular reactions: Dealkylation, Isomerisation, Catalytic cracking, Thermal cracking

Bimolecular reactions: Disproportionation (and C-C cracking), Transalkylation (and C-C cracking)

OHOH

+

Transalkylation

Disproportionation

Page 64: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

64

Fig. S32. Dealkylation of 4-isopropyl-3-methylphenol over Z140-H and Z140-P. (A) The products distribution as a function of temperature over Z140-H, ramping rate=1 oC min-1; (B) Gas chromatograms of the sterically demanding 4-isopropyl-3-methylphenol conversion over Z140-H at low and high temperature (conversion); (C) The products distribution as a function of temperature over Z140-P, ramping rate=1 oC min-1; (D) Conversion rate and selectivity to cresols and propylene as a function of temperature in the dealkylation of propylcresols over Z140-H and Z140-P; (E) The stability of Z140-H and Z140-P in dealkylation of propylcresols. In (A) and (C) others include phenol, xylenols, small amounts of alkylbenzenes, and some unidentified products. Reaction conditions: 4.1 h-1 WHSV, molar ratio H2O to 4-isopropyl-3-methylphenol is 6, 1 bar. Propylcresols include all isomers. In (E) the temperatures are 305 and 395 oC for Z140-H and Z140-P, respectively. Since 4-isopropyl-3-methylphenol undergoes not only dealkylation but also isomerization (A), and all isomers can be dealkylated, the reported conversion rate and conversion in (D) and (E) are based on the conversion of all isomers.

Page 65: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

65

Fig. S33. Dealkylation of 4-n-propylphenol over Z140-H. (A) The products distribution as a function of temperature, ramping rate=1 oC min-1; (B) Gas chromatograms at low and high temperature (conversion); (C) Conversion rate and selectivity to phenol and propylene as a function of temperature in the dealkylation of n-propylphenols over Z140-H; (D) Stability of Z140-H for dealkylation of n-propylphenols (395 oC). In (A), others include cresols, alkylbenzenes, and some unidentified products. Reaction conditions: WHSV=3.7 h-1, molar ratio of H2O to 4-n-propylphenol is 6, 1 bar. Since 4-n-propylphenol undergoes not only dealkylation but also isomerization (A), and all isomers can be dealkylated, the reported conversion rate and conversion in (C) and (D) are based on the conversion of all isomers.

Page 66: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

66

Fig. S34. Dealkylation of 4-ethylphenol over Z140-H. (A) The products distribution as a function of tempeature, ramping rate=1 oC min-1; (B) Gas chromatograms at low and high temperature (conversion); (C) Conversion of ethylphenols and selectivity to phenol and ethylene as a function of time-on-stream (TOS) at 420 oC; In (A) others include cresols, alkylbenzenes, and some unidentified products. Reaction conditions: 3.3 h-1 WHSV, molar ratio of H2O to 4-ethylphenol is 6, 1 bar. Since 4-ethylphenol undergoes not only dealkylation but also isomerization (A), and all isomers can be dealkylated, the reported conversion in (C) is the conversion of all isomers.

Page 67: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

67

Fig. S35. Dealkylation of n-propylbenzene over Z140-H. (A) Conversion rate of n-propylbenzene and selectivity to benzene and propylene as a function of temperature (ramping rate=1 oC min-1, no water); (B) The products distribution as a function of tempeature (ramping rate=1 oC min-1, no water); (C) Gas chromatograms at low and high temperature (conversion); Conversion of n-propylbenzene as a function of time-on-stream (TOS) over (D) Z140-H (410 oC) and (E) Z12-P (350 oC) without water. In (B) others include toluene, ethylene, and some unidentified products. Reaction conditions: WHSV=3.2 h-1, 1 bar. The results reported in this figure were obtained in absence of water in the reaction stream. However, presence of water gives the same results (not shown here).

Page 68: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

68

Fig. S36. Gas chromatogram of light fraction (gas, incondensable fraction at 0-5 oC) in the Poraplot Q column for dealkylation of n-propylphenols over Z140-H in the presence of H2 and water. Reaction conditions: 450 oC, 3.7 h-1 WHSV, 1 bar, conversion=100%. Note that: Though high-pressure hydrogenation over metal-free ZSM-5 was reported (82), only trace amounts of propane (<0.5%) were found by deliberately using H2 instead of N2 in our work.

Page 69: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

69

Fig. S37. Mass flows of lignin valorization. (A) Mass flow of valorization of 1000 kg birch wood native lignin; (B) Mass flow of valorization of 1000 kg pine wood native lignin. 5128 kg birch wood (lignin content: 19.5 wt.%) and 3623 kg pine wood (lignin content: 27.6%) correspond to 1000 kg of the birch wood lignin and 1000 kg of pine wood lignin, respectively. The hydrogen flow is the consumed hydrogen; for instance, valorization of 1000 kg birch wood lignin requires 41.2 kg H2 (20.5 kg for RCF, 20.7 kg for hydroprocessing). Hydrogenolysis of methanol can also occur during RCF, but is not taken into account here. The methanol here is the mass incorporated into lignin oil, which is not quantified for pine wood lignin. Therefore, the mass of methanol and water in the RCF is not listed here.

Page 70: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

70

Fig. S38. General scheme of the step-wise synthesis of ink from phenolics. Phenolics includes para-nonylphenol, acetosolv spruce wood lignin, methanolsolv birch wood lignin, RCF birch wood lignin oligomers in this work.

Page 71: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

71

Fig. S39. Chemical structure of rosin-modified resin, partly based on reference (41). The resole part is synthesized from para-nonylphenol, cardanol, and formaldehyde. In this work, the substitution of para-nonylphenol, a known endocrine disruptor (25), by the RCF birch wood lignin oligomers was assessed.

Page 72: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

72

Fig. S40. Gel permeation chromatograms of (A) birch wood lignin oligomers (from RCF) and methanosolv birch wood lignin and (B) commercial acetosolv spruce wood lignin.

Page 73: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

73

Fig. S41. Resins synthesized from para-nonylphenol (left), self-prepared methanosolv birch wood lignin (middle), and RCF birch wood lignin oligomers (right).

Page 74: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

74

Fig. S42. Gel permeation chromatograms of resins based on (i) para-nonylphenol, (ii) self-prepared methanosolv birch wood lignin, and (iii) RCF birch wood lignin oligomers.

Page 75: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

75

Fig. S43. Assessment of emulsification capacity (EC) of varnishes based on different resins. The amount of water added (EC, x-axis) is expressed as wt.% relative to the amount of varnish (25 g) in the sample holder.

Page 76: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

76

Fig. S44. Varnishes prepared from different resins (commercial resin, self-prepared para-nonylphenol based resin, methanosolv birch wood based resin, RCF birch wood lignin oligomers based resin).

Page 77: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

77

Fig. S45. Graphical representation of the process flow diagram (PFD) used to model the conversion process of birch wood toward phenol, propylene, lignin phenolic oligomers and carbohydrate pulp.

Page 78: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

78

Fig. S46. Sankey diagram providing a graphical representation of all major process streams. The mass of pulp is determined by weighing residue (excluding the catalyst) after the reductive catalytic fractionation of birch wood. Note that: the size of the arrows is not exactly representing the amount of mass flow.

Page 79: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

79

Fig. S47. (A) Mass balance of this integrated biorefinery (assuming conversion of 1 ton dry and extracted birch wood); (B) Mass balance of monomers conversion (on the basis of 1 ton dry and extracted birch wood). The amount of methanol in the scheme is the consumed amount, which is around 1.4% of total amount of methanol in the reaction. The hydrogen is the loaded hydrogen (including consumed hydrogen and excess hydrogen used for energy, details can be found in table S8). The amount of cellulose, hemicellulose, and lignin before and after reductive catalytic fractionation was quantified by compositional analysis (see experimental section).

Page 80: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

80

Fig. S48. CAPEX (Capital expenditure) distribution for the different subunits of the biorefinery.

Page 81: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

81

Fig. S49. Tornado diagram illustrating the results from the sensitivity analysis on several design and cost assumptions. The horizontal bars show the deviation of the internal rate of return to the basic case (∆internal rate of return/basic internal rate of return×100%). The internal rate of return of basic case is 23.33%. Parameters applied for the basic scenario are provided in table S9 (feedstock price, product price, catalyst price) or in the design basis (methanol/wood ratio, extraction solvents, catalyst life time, RCF reaction time). Not profitable represents the negative internal rate of return or negative income.

Page 82: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

82

Fig. S50. System boundary of LCA of the production process of phenol from crude fossil oil.

Page 83: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

83

Fig. S51. System boundary of LCA of the production process of phenol from lignocellulose (birch wood).

Page 84: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

84

Fig. S52. GWPs [in kg of CO2-equivalent per kg of product (kg CO2e per kg product)] of phenol, propylene, oligomers, and carbohydrate pulp production from lignocellulose (birch wood) with different hydrogen sources and without/with sustainable forest management strategies in this integrated biorefinery. Without sustainable forest management is based on an average over the European continent. (A) GWP of phenol production from lignocellulose with different hydrogen sources and without sustainable forest management. (B) GWPs of propylene, oligomers, and carbohydrate pulp production from lignocellulose with different hydrogen sources without sustainable forest management. (C) GWP of phenol production from lignocellulose with/without sustainable forest management and with non-renewable hydrogen or renewable hydrogen. (D) GWPs of propylene, oligomers, and carbohydrate pulp production from lignocellulose with/without sustainable forest management and with non-renewable hydrogen or renewable hydrogen. The GWPs of non-renewable H2 I and II are 11.89 and 8.20 kg of CO2-equivalent per kg of H2, respectively. The renewable H2 III is obtained by electrolysis of water with electricity from wind energy, which has a GWP of 0.97 kg of CO2-equivalent per kg of H2. The GWPs of wood without and with sustainable forest management are -0.453 and -1.78 kg of CO2-equivalent per kg of wood, respectively.

Page 85: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

85

Fig. S53. Sensitivity analysis for GWP of phenol. (A) The sensitivity ratio (SR) with -10% variation; (B) Sensitivity ratio (SR) with changing the price of products.

Page 86: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

86

Tables Table S1. Monomer yield and distribution obtained from RCF of birch wood and pine wood.

Feed

stoc

k

Con

vers

ion

of li

gnin

/ %

a

Monomers yield / wt.%

Olig

omer

s yie

ld /

wt.%

4-n-

prop

ylgu

aiac

ol

isoe

ugen

ol

4-(3

-met

hoxy

prop

yl)-g

uaia

col

4-n-

prop

anol

guai

acol

4-et

hylg

uaia

col

4-pr

opyl

syrin

gol

4-pr

op-1

-eny

lsyr

ingo

l

4-n-

prop

anol

syrin

gol

syrin

gol

4-m

ethy

lsyr

ingo

l

4-et

hyls

ryin

gol

4-(3

-met

hoxy

prop

yl)-s

yrin

gol

ot

hers

Tota

l mon

omer

s

Birch 80.69 9.71 0.49 <0.19 0.89 0.30 33.85 0.32 2.21 0.43 0.28 1.03 0.79 <0.02 50.51 30.18

Pine 37.30 9.97 0.83 0.00 1.96 0.21 0.02 0.40 0.01 0.00 0.02 0.21 0.00 <0.42 14.05 23.25

a It is same as delignification.

OHO

OHO

OHO

O

OHO

OH

OHO

OHOO

OHOO

OHOO

OH

OHOO

OHOO

OHOO

OHOO

O

Page 87: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

87

Table S2. Semi-quantitative integration results of the 2D HSQC NMR spectra shown in fig. S5, for the lignin sub-units, inter-unit linkages and end-groups from the RCF birch wood lignin phenolic oligomers. % in RCF birch wood lignin phenolic oligomersa Lignin sub-units Syringyl (S) 75.0% Guaiacyl (G) 25.0% S/G 3 Inter-unit linkages β-O-4 (A) 4.1% β-5 phenylcoumarane (C) 0.0% β-β resinol (B) 1.6% End-groups 4-propyl (P) 12.3% 4-propanol (POH) 15.0% Cinnamyl alcohol (I) 0.9% % γ-OHb 56.4% a The quantity of the sub-units, inter-unit linkages and end-groups is expressed per 100 aromatic units, based on comparison to the G + 1/2S aromatic integral. b Relative content of γ-OH: (POH+I)/(POH + P + I).

Page 88: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

88

Table S3. Hydroxyl group content in RCF birch wood lignin phenolic oligomers as determined by 31P-NMR.

a Measurements in triplicate, standard deviation between parentheses.

OH content (mmol OH g-1)a Carboxylic OH 0.08 (0.001) p-Hydroxyphenyl OH 0.14 (0.011) Guaiacyl OH 0.8 (0.01) C5 substituted OH 2.53 (0.004) Aliphatic OH 2.48 (0.006) Total phenolic OH 3.46 (0.024) Aliphatic OH/Phenolic OH 0.71 Total OH 6.03 (0.028)

Page 89: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

89

Table S4: The condensate composition of hydroprocessing products.a

Substrate Propylbenzene Cresols Ethylphenols Propylphenolsb Propylcresols Propylguaiacol Othersd

Propylguaiacol 5.6 1.8 3.9 80.0 2.5 1.8 4.5

Pine lignin monomers 5.7 2.3 3.6 78.8 3 2.5 4.2

Birch lignin monomers 6.8 3.1 4.4 77.2 2.7 3.1c 2.7

a The condensate contains 10-17 wt.% water, depending on feedstock. The composition is on carbon molar basis. As the 64 wt.% Ni/SiO2 is not highly stable, propylguaiacol is presented in the products. b Propylphenols include 4-n-propylphenol and 3-n-propylphenol. c 5-methoxy-3-propylphenol is also included. d Others include propylanisole, benzene, ethylbenzene, some unidentified products.

Page 90: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

90

Table S5. Properties of Z140-P and Z140-H.

Materials Smeso

a /

m2 g-1

Vmesoa /

cm3 g-1

Vmicroa /

cm3 g-1

150 oCb 250 oCb 350 oCb

B /

µmol g-1

L /

µmol g-1

B /

µmol g-1

L /

µmol g-1

B /

µmol g-1

L /

µmol g-1

Z140-P 17 0.04 0.18 53 7 49 4 42 3

Z140-H 71 0.28 0.17 57 18 55 14 52 11

a Determined by N2 sorption, the Smeso and Vmeso were determined by t-plot, Vmicro=Vpore-Vmeso. b Determined by pyridine FTIR, B: Brønsted acid sites, L: Lewis acid sites. 150 oC, 250 oC, and 350 oC are the desorption temperature of pyridine before measurement at 150 oC.

Page 91: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

91

Table S6. Overview of the evaluation of resins synthesized from different phenol sources.

Phenol sourcea para-Nonylphenol (reference)

Methanosolv birch wood ligninb

RCF birch wood lignin oligomers

Vacuum time / minc - 15 + 15 + 20 5 + 5 Viscosity at 25 Hz d / Pa.s 19.8 20.1 20.3

Flow behaviour indexd 0.861 0.893 0.937 Residue on filtere / ppm 82 314 160

Acid valued / mg(KOH) g-1 33.8 20.1 19.9 Cloud pointe / °C 82 105 120

Mn / g mol-1 1295 1069 1063 Mw / g mol-1 142401 43245 24336

a Comprises 60 wt.% cardanol and 40 wt.% other (para-nonylphenol, methanosolv birch wood lignin or RCF birch wood lignin oligomers).

b Methanosolv birch wood lignin was prepared by processing birch wood under the same RCF conditions, in the absence of Ru/C and H2. The same work-up protocol was applied.

c During the synthesis of resin, vacuum can be applied to adjust the viscosity. The more difficult the incorporation of lignin in the resole chain, the more vacuum is required to increase the viscosity. For industrial application, the vacuum time and number of cycles is preferably kept to a minimum.

d Because resin is a solid, the viscosity and flow behaviour index were measured for a 40 wt.% solution in mineral oil PKWF 6/9 AR Blend.

e Determined for a 10 wt.% solution in mineral oil PKWF 6/9 AFN.

Page 92: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

92

Table S7. Overview of the evaluation of ink varnishes synthesized from different resins.

Resin type Commercial resin para-Nonylphenol based resin

Methanosolv birch wood lignin based

resin

RCF birch wood lignin oligomers

based resin Compositiona

Lineseed oil / wt.% 15 15 15 15 Rapeseed oil methyl ester / wt.% 33 35 31 29

Resin / wt.% 26 (ER 9520) 26 (SP 2330) 50 54 56

Evaluation Viscosity at 25 Hz / Pa.s 117 88.7 95.6 97.6

Flow behaviour index 0.821 0.797 0.882 0.888 a The relative ratio of the varnish constituents was set to ensure a viscosity of approximately 100 Pa.s.

Page 93: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

93

Table S8. Overview of the mass and energy flows in the integrated (birch wood-to-phenol, propylene, lignin phenolic oligomers, and carbohydrate pulp) biorefinery.

a Dry and extracted wood. b Composition: 277 kg h-1 hexane, 28973 kg h-1 monomers (it is assumed that the n-hexane is not converted in the hydroprocessing and dealkylation). c Composition: 19830 kg h-1 methyl acetate, 767 kg h-1 methanol, 767 kg h-1 water, 1414 kg h-1 H2, and 1419 kg h-1 CH4. d Composition: 1414 kg h-1 H2, 1419 kg h-1 CH4. e Composition: 19830 kg h-1 methyl acetate, 767 kg h-1 methanol, 767 kg h-1 water. f The products from hydroprocessing without separation (including water, methane, hydrogen, alkylphenols, etc.). g Composition: 2251 kg h-1 H2, 6634 kg h-1 CH4, 232 kg h-1 C2H4, 101 kg h-1 C3H6 (lost in the separation). h Composition: 7200 kg h-1 water, 3445 kg h-1 DCM (dichloromethane), 20392 kg h-1 methyl-xylose, 18169 kg h-1 other sugars (it is assumed that the density of sugars and DCM is 1.4 g ml-1). i Composition: 8894 kg h-1 water, 277 kg h-1 hexane [density=0.66 g ml-1],3005 kg h-1 other organics (including 1043 kg h-1 (alkyl) benzene [density=0.87 g ml-1], 985 kg h-1 cresols [density=1.03 g ml-1] , 976 kg h-1 others [density=0.95 g ml-1]). j Incineration of off gas and H2 gas bypass generates 272003 kWh of energy. This energy is converted in the trigeneration unit toward electric, heating and cooling energy with an overall efficiency of 85%.

RCF Separation Hydroprocessing Dealkylation/distillation Trigeneration Balance

In Out In Out In Out In Out In Out - generation + consumption

Wood (kg h-1) 273364a 273364

Water (kg h-1) 5078 7200 3911 3852 9885

Methanol (kg h-1) 16713 16713

Hydrogen (kg h-1) 2773 2037 4810

Nitrogen (kg h-1) 172 172

DCM (kg h-1) 3445 3445

n-Hexane (kg h-1) 277 277

Combustion air (kg h-1) 378450 378450

Crude Oil (kg h-1) 85063 85063 0

Pulp (kg h-1) 178512 -178512

Monomers (kg h-1) 29250b 29250 0

Oligomers (kg h-1) 17529 -17529

H2 gas bypass (kg h-1) 24197c 2833d 21364e 0

Hydroprocessing Product (kg h-1) 34120f 34120 0

Phenol (kg h-1) 11408 -11408

Propylene (kg h-1) 5230 -5230

Off gas (kg h-1) 172 (N2) 9218g 9218 -172

Wastewater (m³ h-1) 37.2h 12.5i 4 -53.7

Flue gas (kg h-1) 408264 -408264

Cool water (m³ h-1) 3135 3135 2406 2406 2801 2801 5541 5541 0

HP Steam (kW) 15628 4792 3902 6934 0

MP Steam (kW) 12802 33472 40736 61406 0

HP Condensate (kW) 6866 2105 1714 3047 0

MP Condensate (kW) 3829 10011 6182 0

Electric power (kWh) 1718 4428 36 5669 78515j -75520

Heating (kWh) 5223 18013 33483 45051 3907 61837j 0

Cooling (kWh) 24445 27991 41117 2703 90850j 0

Catalyst Ru (kg h-1) 2.524 2.524

Catalyst Ni (kg h-1) 0.312 0.312

Catalyst Zeolite (kg h-1) 0.376 0.376

Page 94: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

94

Table S9. List of economic parameters.

Component Price / € tonne-1 Wild birch wooda 158 Watera 0.5 Methanola 393 Hydrogena 1050 Nitrogena 102 Dichloromethanea 313 n-Hexanea 712 5 wt.% Ru/C catalystb 418000 64 wt.% Ni/SiO2b 100000 Hierarchical ZSM-5c 15000 Wastewater treatmenta 9238 Monomersd 650 Oligomerse 1750 Carbohydrate pulpf 400 Phenolg 1300 Propyleneg 830 Excess electricityh 0.1 € kWh-1

a Estimated based on the value reported in Intratec Solutions (accessed 10/2018). Since the birch wood is dry and extracted in this model, the birch wood is around 1.5 times the price of literature (7) used. b Obtained from chemistry industry. c Estimated based on the value obtained from Zeopore. d Estimated based on the value of phenol, propylene, and the process cost. e Estimated based on the value obtained from LAWTER. f Conservatively estimated based on the price for birch wood (2.5 times of birch wood) and sugars. g Estimated based on the average price between 2014-2017 from ICIS. h Estimated based on the average price for non-household consumers between 2008-2017 in EU28.

Page 95: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

95

Table S10. Manufacture cost.

Manufacture cost / MM€ yr-1 649 Raw Materials / MM€ yr-1 492 Utilities / MM€ yr-1 24 Labor / MM€ yr-1 3.3 Supplies / MM€ yr-1 10 Maintenance / MM€ yr-1 47 Office & Service Overhead / MM€ yr-1 1.7 Property Taxes / MM€ yr-1 24 Depreciation (10 years) / MM€ yr-1 47

Page 96: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

96

Table S11. Summary of economics for the integrated biorefinery design using birch wood.

Total plant cost / MM€ 477 Working capital / MM€ 339 Start-up expense / MM€ 36 Manufacture cost / MM€ yr-1 649 Cost of selling goods / MM€ yr-1 56 SAREa / MM€ yr-1 113 Tax on income (50%) / MM€ yr-1 155 Total revenue / MM€ yr-1 1129 Income after tax / MM€ yr-1 155 Internal rate of return / % 23.33 Payout time / yr 4.19

a Sales administration and research expenses.

Page 97: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

97

Table S12. Global warming potentials (GWPs) of phenol, propylene, and nonylphenol production from crude fossil oil through Hock process, steam cracking of naphtha, and alkylation of phenol with nonenes, respectively.

Product Process GWP [in kg of CO2-equivalent per kg of product]

Phenol Hock process 1.73E0

Propylene Steam cracking 1.47E0

Nonylphenol Alkylation of phenol with nonenes >1.58E0a a The GWP of nonylphenol was not obtained in the database or literature. Since it is industrially obtained by alkylation of phenol with fossil-based nonenes, the GWP of nonylphenol is roughly calculated. Theoretically (100% selectivity for each step), production of 1 kg of nonylphenol needs 0.427 kg of phenol and 0.573 kg of nonenes (i.e. 0.573 kg of propylene as nonenes are trimers of propylene). Therefore, GWP of nonylphenol=0.427×1.73+1.47×0.573=1.581 kg of CO2-equivalent per kg of nonylphenol. The GWP of nonylphenol should be higher than this value due to lower selectivity (<100%) for each step and energy consumption.

Page 98: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

98

Table S13. Global warming potential (GWP) of phenol production from birch wood in this integrated biorefinery.a

Process GWP [in kg of CO2-equivalent per kg of phenol]

Reductive catalytic fractionation -7.72E-1

Separation 1.26E-1

Hydroprocessing/dealkylation/distillation 1.38E0

Total -7.36E-1 a The GWP of H2 (via steam reforming) is 8.20 kg of CO2-equivalent per kg of H2 in this calculation.

Page 99: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

99

Table S14. Global warming potentials (GWPs) of propylene, oligomers, carbohydrate pulp production from birch wood in this integrated biorefinery.a

Product GWP [in kg of CO2-equivalent per kg of product]

Propylene 4.69E-1

Oligomers -9.49E-1

Carbohydrate pulp -2.17E-1 a The GWP of H2 (via steam reforming) is 8.20 kg of CO2-equivalent per kg of H2 in this calculation.

Page 100: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

References and Notes

1. J.-F. Bastin, Y. Finegold, C. Garcia, D. Mollicone, M. Rezende, D. Routh, C. M. Zohner, T.

W. Crowther, The global tree restoration potential. Science 365, 76–79 (2019).

doi:10.1126/science.aax0848 Medline

2. K. Van Meerbeek, B. Muys, M. Hermy, Lignocellulosic biomass for bioenergy beyond

intensive cropland and forests. Renew. Sustain. Energy Rev. 102, 139–149 (2019).

doi:10.1016/j.rser.2018.12.009

3. T. Werpy, G. Petersen, “Top value added chemicals from biomass: Volume I—Results of

screening for potential candidates from sugars and synthesis gas” (National Renewable

Energy Laboratory, 2004); www.osti.gov/biblio/15008859.

4. J. E. Holladay, J. F. White, J. J. Bozell, D. Johnson, “Top value-added chemicals from

biomass: Volume II—Results of screening for potential candidates from biorefinery

lignin” (Pacific Northwest National Laboratory, 2007); www.osti.gov/biblio/921839.

5. A. Kätelhön, R. Meys, S. Deutz, S. Suh, A. Bardow, Climate change mitigation potential of

carbon capture and utilization in the chemical industry. Proc. Natl. Acad. Sci. U.S.A. 116,

11187–11194 (2019). doi:10.1073/pnas.1821029116 Medline

6. International Energy Agency, “The future of petrochemicals: Towards more sustainable

plastics and fertilisers” (IEA Publications, 2018);

www.connaissancedesenergies.org/sites/default/files/pdf-

actualites/the_future_of_petrochemicals.pdf.

7. D. M. Alonso, S. H. Hakim, S. Zhou, W. Won, O. Hosseinaei, J. Tao, V. Garcia-Negron, A. H.

Motagamwala, M. A. Mellmer, K. Huang, C. J. Houtman, N. Labbé, D. P. Harper, C.

Maravelias, T. Runge, J. A. Dumesic, Increasing the revenue from lignocellulosic

biomass: Maximizing feedstock utilization. Sci. Adv. 3, e1603301 (2017).

doi:10.1126/sciadv.1603301 Medline

8. A. J. Ragauskas, G. T. Beckham, M. J. Biddy, R. Chandra, F. Chen, M. F. Davis, B. H.

Davison, R. A. Dixon, P. Gilna, M. Keller, P. Langan, A. K. Naskar, J. N. Saddler, T. J.

Tschaplinski, G. A. Tuskan, C. E. Wyman, Lignin valorization: Improving lignin

processing in the biorefinery. Science 344, 1246843 (2014). doi:10.1126/science.1246843

Medline

9. C. Li, X. Zhao, A. Wang, G. W. Huber, T. Zhang, Catalytic transformation of lignin for the

production of chemicals and fuels. Chem. Rev. 115, 11559–11624 (2015).

doi:10.1021/acs.chemrev.5b00155 Medline

10. Z. Sun, B. Fridrich, A. de Santi, S. Elangovan, K. Barta, Bright side of lignin

depolymerization: Toward new platform chemicals. Chem. Rev. 118, 614–678 (2018).

doi:10.1021/acs.chemrev.7b00588 Medline

11. J. G. Linger, D. R. Vardon, M. T. Guarnieri, E. M. Karp, G. B. Hunsinger, M. A. Franden, C.

W. Johnson, G. Chupka, T. J. Strathmann, P. T. Pienkos, G. T. Beckham, Lignin

valorization through integrated biological funneling and chemical catalysis. Proc. Natl.

Acad. Sci. U.S.A. 111, 12013–12018 (2014). doi:10.1073/pnas.1410657111 Medline

Page 101: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

12. I. Kumaniaev, E. Subbotina, J. Sävmarker, M. Larhed, M. V. Galkin, J. S. M. Samec, Lignin

depolymerization to monophenolic compounds in a flow-through system. Green Chem.

19, 5767–5771 (2017). doi:10.1039/C7GC02731A

13. E. M. Anderson, M. L. Stone, R. Katahira, M. Reed, G. T. Beckham, Y. Román-Leshkov,

Flowthrough reductive catalytic fractionation of biomass. Joule 1, 613–622 (2017).

doi:10.1016/j.joule.2017.10.004

14. W. Schutyser, T. Renders, S. Van den Bosch, S.-F. Koelewijn, G. T. Beckham, B. F. Sels,

Chemicals from lignin: An interplay of lignocellulose fractionation, depolymerisation,

and upgrading. Chem. Soc. Rev. 47, 852–908 (2018). doi:10.1039/C7CS00566K Medline

15. Q. Song, F. Wang, J. Cai, Y. Wang, J. Zhang, W. Yu, J. Xu, Lignin depolymerization (LDP)

in alcohol over nickel-based catalysts via a fragmentation–hydrogenolysis process.

Energy Environ. Sci. 6, 994–1007 (2013). doi:10.1039/c2ee23741e

16. S. Van den Bosch, W. Schutyser, R. Vanholme, T. Driessen, S.-F. Koelewijn, T. Renders, B.

De Meester, W. Huijgen, W. Dehaen, C. Courtin, B. Lagrain, W. Boerjan, B. F. Sels,

Reductive lignocellulose fractionation into soluble lignin-derived phenolic monomers and

dimers and processable carbohydrate pulps. Energy Environ. Sci. 8, 1748–1763 (2015).

doi:10.1039/C5EE00204D

17. A. Vuori, J. B. Bredenberg, Hydrogenolysis and hydrocracking of the carbon-oxygen bond.

5. hydrogenolysis of 4-propylguaiacol by sulfided CoO-MoO3/γ-Al2O3. Holzforschung

38, 253–262 (1984). doi:10.1515/hfsg.1984.38.5.253

18. N. Joshi, A. Lawal, Hydrodeoxygenation of 4-propylguaiacol (2-methoxy-4-propylphenol) in

a microreactor: Performance and kinetic studies. Ind. Eng. Chem. Res. 52, 4049–4058

(2013). doi:10.1021/ie400037y

19. H. L. Chum, D. K. Johnson, S. Black, M. Ratcliff, D. W. Goheen, in Advances in Solar

Energy: An Annual Review of Research and Development, K. W. Böer, Ed. (Springer,

1988), pp. 91–200.

20. X. Liu, C. Wang, Y. Zhang, Y. Qiao, Y. Pan, L. Ma, Selective Preparation of 4-Alkylphenol

from Lignin-Derived Phenols and Raw Biomass over Magnetic Co-Fe@N-Doped Carbon

Catalysts. ChemSusChem 12, 4791–4798 (2019). doi:10.1002/cssc.201901578 Medline

21. D. Verboekend, Y. Liao, W. Schutyser, B. F. Sels, Alkylphenols to phenol and olefins by

zeolite catalysis: A pathway to valorize raw and fossilized lignocellulose. Green Chem.

18, 297–306 (2016). doi:10.1039/C5GC01868D

22. Y. Liao, R. Zhong, E. Makshina, M. d’Halluin, Y. van Limbergen, D. Verboekend, B. F.

Sels, Propylphenol to phenol and propylene over acidic zeolites: Role of shape selectivity

and presence of steam. ACS Catal. 8, 7861–7878 (2018). doi:10.1021/acscatal.8b01564

23. H. Sixta, Handbook of Pulp (Wiley, 2006).

24. S. Schiavoni, F. D’Alessandro, F. Bianchi, F. Asdrubali, Insulation materials for the building

sector: A review and comparative analysis. Renew. Sustain. Energy Rev. 62, 988–1011

(2016). doi:10.1016/j.rser.2016.05.045

Page 102: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

25. A. Soares, B. Guieysse, B. Jefferson, E. Cartmell, J. N. Lester, Nonylphenol in the

environment: A critical review on occurrence, fate, toxicity and treatment in wastewaters.

Environ. Int. 34, 1033–1049 (2008). doi:10.1016/j.envint.2008.01.004 Medline

26. T. Renders, W. Schutyser, S. Van den Bosch, S. F. Koelewijn, T. Vangeel, C. M. Courtin, B.

F. Sels, Influence of acidic (H3PO4) and alkaline (NaOH) additives on the catalytic

reductive fractionation of lignocellulose. ACS Catal. 6, 2055–2066 (2016).

doi:10.1021/acscatal.5b02906

27. C. W. Dence, in Methods in Lignin Chemistry, S. Y. Lin, C. W. Dence, Eds. (Springer, 1992),

pp. 33–61.

28. J. Snelders, E. Dornez, B. Benjelloun-Mlayah, W. J. J. Huijgen, P. J. de Wild, R. J. A.

Gosselink, J. Gerritsma, C. M. Courtin, Biorefining of wheat straw using an acetic and

formic acid based organosolv fractionation process. Bioresour. Technol. 156, 275–282

(2014). doi:10.1016/j.biortech.2014.01.069 Medline

29. C. M. Courtin, H. Van den Broeck, J. A. Delcour, Determination of reducing end sugar

residues in oligo- and polysaccharides by gas-liquid chromatography. J. Chromatogr. A

866, 97–104 (2000). doi:10.1016/S0021-9673(99)01064-X Medline

30. C. Gourson, R. Benhaddou, R. Granet, P. Krausz, B. Verneuil, P. Branland, G. Chauvelon, J.

F. Thibault, L. Saulnier, Valorization of maize bran to obtain biodegradable plastic films.

J. Appl. Polym. Sci. 74, 3040–3045 (1999). doi:10.1002/(SICI)1097-

4628(19991220)74:13<3040:AID-APP5>3.0.CO;2-M

31. A. M. Silvestre-Albero, J. M. Juárez-Galán, J. Silvestre-Albero, F. Rodríguez-Reinoso, Low-

pressure hysteresis in adsorption: An artifact? J. Phys. Chem. C 116, 16652–16655

(2012). doi:10.1021/jp305358y

32. C. A. Emeis, Determination of integrated molar extinction coefficients for infrared-

absorption bands of pyridine adsorbed on solid acid catalysts. J. Catal. 141, 347–354

(1993). doi:10.1006/jcat.1993.1145

33. L. A. Colón, L. J. Baird, in Modern Practice of Gas Chromatography, R. L. Grob, E. F.

Barry, Eds. (Wiley, ed. 4, 2004), pp. 275–337.

34. X. Meng, C. Crestini, H. Ben, N. Hao, Y. Pu, A. J. Ragauskas, D. S. Argyropoulos,

Determination of hydroxyl groups in biorefinery resources via quantitative 31P NMR

spectroscopy. Nat. Protoc. 14, 2627–2647 (2019). doi:10.1038/s41596-019-0191-1

Medline

35. P. Korntner, I. Sumerskii, M. Bacher, T. Rosenau, A. Potthast, Characterization of technical

lignins by NMR spectroscopy: Optimization of functional group analysis by 31P NMR

spectroscopy. Holzforschung 69, 807–814 (2015). doi:10.1515/hf-2014-0281

36. W. Schutyser, S. Van den Bosch, J. Dijkmans, S. Turner, M. Meledina, G. Van Tendeloo, D.

P. Debecker, B. F. Sels, Selective nickel-catalyzed conversion of model and lignin-

derived phenolic compounds to cyclohexanone-based polymer building blocks.

ChemSusChem 8, 1805–1818 (2015). doi:10.1002/cssc.201403375 Medline

Page 103: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

37. N. Dowe, J. McMillan, “SSF experimental protocols: Lignocellulosic biomass hydrolysis and

fermentation” (National Renewable Energy Laboratory, 2001);

www.nrel.gov/docs/gen/fy08/42630.pdf.

38. X. Ju, M. Bowden, M. Engelhard, X. Zhang, Investigating commercial cellulase

performances toward specific biomass recalcitrance factors using reference substrates.

Appl. Microbiol. Biotechnol. 98, 4409–4420 (2014). doi:10.1007/s00253-013-5450-4

Medline

39. M. M. Demeke, H. Dietz, Y. Li, M. R. Foulquié-Moreno, S. Mutturi, S. Deprez, T. Den Abt,

B. M. Bonini, G. Liden, F. Dumortier, A. Verplaetse, E. Boles, J. M. Thevelein,

Development of a D-xylose fermenting and inhibitor tolerant industrial Saccharomyces

cerevisiae strain with high performance in lignocellulose hydrolysates using metabolic

and evolutionary engineering. Biotechnol. Biofuels 6, 89 (2013). doi:10.1186/1754-6834-

6-89 Medline

40. M. M. Demeke, F. Dumortier, Y. Li, T. Broeckx, M. R. Foulquié-Moreno, J. M. Thevelein,

Combining inhibitor tolerance and D-xylose fermentation in industrial Saccharomyces

cerevisiae for efficient lignocellulose-based bioethanol production. Biotechnol. Biofuels

6, 120 (2013). doi:10.1186/1754-6834-6-120 Medline

41. Y.-B. Ha, M. Y. Jin, S.-S. Oh, D.-H. Ryu, Synthesis of an environmentally friendly phenol-

free resin for printing ink. Bull. Korean Chem. Soc. 33, 3413–3416 (2012).

doi:10.5012/bkcs.2012.33.10.3413

42. R. B. Bird, R. C. Armstrong, O. Hassager, Dynamics of Polymeric Liquids, Volume 1: Fluid

Mechanics (Wiley, 1987).

43. J. R. Couper, W. R. Penney, J. R. Fair, Chemical Process Equipment-Selection and Design

(Revised 2nd Edition) (Gulf Professional Publishing, 2009).

44. M. Weber, M. Weber, M. Kleine-Boymann, in Ullmann's Encyclopedia of Industrial

Chemistry (Wiley, 2004), pp. 503–519.

45. Ecoinvent, version 3.3; www.ecoinvent.org/.

46. ThinkStep GaBi Professional Database 2016; www.gabi-

software.com/america/databases/gabi-databases/professional/.

47. Plastics Europe, “Phenol and acetone CEFIC petrochemicals Europe –Phenol and acetone

sector group September 2016” (2016); www.petrochemistry.eu/wp-

content/uploads/2018/01/EPD-Phenol-Acetone-10-16.pdf.

48. P. Nuss, M. J. Eckelman, Life cycle assessment of metals: A scientific synthesis. PLOS ONE

9, e101298 (2014). doi:10.1371/journal.pone.0101298 Medline

49. F. Zhu, J. A. Johnson, D. W. Ablin, G. A. Ernst, Efficient Petrochemical Processes:

Technology, Design and Operation (Wiley, 2019).

50. R. J. Schmidt, Industrial catalytic processes—phenol production. Appl. Catal. A. 280, 89–103

(2005). doi:10.1016/j.apcata.2004.08.030

Page 104: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

51. R. Bal, M. Tada, T. Sasaki, Y. Iwasawa, Direct phenol synthesis by selective oxidation of

benzene with molecular oxygen on an interstitial-N/Re cluster/zeolite catalyst. Angew.

Chem. Int. Ed. 45, 448–452 (2006). doi:10.1002/anie.200502940 Medline

52. C. O. Tuck, E. Pérez, I. T. Horváth, R. A. Sheldon, M. Poliakoff, Valorization of biomass:

Deriving more value from waste. Science 337, 695–699 (2012).

doi:10.1126/science.1218930 Medline

53. R. Rinaldi, R. Jastrzebski, M. T. Clough, J. Ralph, M. Kennema, P. C. A. Bruijnincx, B. M.

Weckhuysen, Paving the way for lignin valorisation: Recent advances in bioengineering,

biorefining and catalysis. Angew. Chem. Int. Ed. 55, 8164–8215 (2016).

doi:10.1002/anie.201510351 Medline

54. G. J. Snell, D. T. Huibers, U.S. Patent US4409416A (1983).

55. M. Wang, M. Liu, H. Li, Z. Zhao, X. Zhang, F. Wang, Dealkylation of lignin to phenol via

oxidation-hydrogenation strategy. ACS Catal. 8, 6837–6843 (2018).

doi:10.1021/acscatal.8b00886

56. G. J. Snell, D. T. Huibers, U.S. Patent US4420644A (1983).

57. P. Ferrini, R. Rinaldi, Catalytic biorefining of plant biomass to non-pyrolytic lignin bio-oil

and carbohydrates through hydrogen transfer reactions. Angew. Chem. Int. Ed. 53, 8634–

8639 (2014). doi:10.1002/anie.201403747 Medline

58. L. Shuai, M. T. Amiri, Y. M. Questell-Santiago, F. Héroguel, Y. Li, H. Kim, R. Meilan, C.

Chapple, J. Ralph, J. S. Luterbacher, Formaldehyde stabilization facilitates lignin

monomer production during biomass depolymerization. Science 354, 329–333 (2016).

doi:10.1126/science.aaf7810 Medline

59. L. D. Eltis, R. Singh, in Lignin Valorization: Emerging Approaches, G. T. Beckham, Ed.

(Royal Society of Chemistry, 2018), pp. 290–313.

60. T. Nimmanwudipong, R. C. Runnebaum, D. E. Block, B. C. Gates, Catalytic conversion of

guaiacol catalyzed by platinum supported on alumina: Reaction network including

hydrodeoxygenation reactions. Energy Fuels 25, 3417–3427 (2011).

doi:10.1021/ef200803d

61. M. B. Griffin, F. G. Baddour, S. E. Habas, D. A. Ruddy, J. A. Schaidle, Evaluation of silica-

supported metal and metal phosphide nanoparticle catalysts for the hydrodeoxygenation

of guaiacol under ex situ catalytic fast pyrolysis conditions. Top. Catal. 59, 124–137

(2016). doi:10.1007/s11244-015-0512-8

62. M. Saidi, F. Samimi, D. Karimipourfard, T. Nimmanwudipong, B. C. Gates, M. R.

Rahimpour, Upgrading of lignin-derived bio-oils by catalytic hydrodeoxygenation.

Energy Environ. Sci. 7, 103–129 (2014). doi:10.1039/C3EE43081B

63. W. Schutyser, G. Van den Bossche, A. Raaffels, S. Van den Bosch, S.-F. Koelewijn, T.

Renders, B. F. Sels, Selective conversion of lignin-derivable 4-alkylguaiacols to 4-

alkylcyclohexanols over noble and non-noble-metal catalysts. ACS Sustain. Chem. Eng.

4, 5336–5346 (2016). doi:10.1021/acssuschemeng.6b01580

Page 105: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

64. G.-Y. Xu, J.-H. Guo, Y.-C. Qu, Y. Zhang, Y. Fu, Q.-X. Guo, Selective hydrodeoxygenation

of lignin-derived phenols to alkyl cyclohexanols over a Ru-solid base bifunctional

catalyst. Green Chem. 18, 5510–5517 (2016). doi:10.1039/C6GC01097K

65. Y. Nakagawa, M. Ishikawa, M. Tamura, K. Tomishige, Selective production of cyclohexanol

and methanol from guaiacol over Ru catalyst combined with MgO. Green Chem. 16,

2197–2203 (2014). doi:10.1039/C3GC42322K

66. L. Beránek, M. Kraus, Catalytic dealkylation of alkylaromatic compounds. XIV. The effect

of structure of monoalkylbenzenes on their reactivity in hydrodealkylation on a nickel

catalyst. Collect. Czech. Chem. Commun. 31, 566–575 (1966).

doi:10.1135/cccc19660566

67. V. N. Bui, D. Laurenti, P. Afanasiev, C. Geantet, Hydrodeoxygenation of guaiacol with

CoMo catalysts. Part I: Promoting effect of cobalt on HDO selectivity and activity. Appl.

Catal. B 101, 239–245 (2011). doi:10.1016/j.apcatb.2010.10.025

68. V. N. Bui, D. Laurenti, P. Delichère, C. Geantet, Hydrodeoxygenation of guaiacol: Part II:

Support effect for CoMoS catalysts on HDO activity and selectivity. Appl. Catal. B 101,

246–255 (2011). doi:10.1016/j.apcatb.2010.10.031

69. X. Zhang, Q. Zhang, T. Wang, L. Ma, Y. Yu, L. Chen, Hydrodeoxygenation of lignin-

derived phenolic compounds to hydrocarbons over Ni/SiO2-ZrO2 catalysts. Bioresour.

Technol. 134, 73–80 (2013). doi:10.1016/j.biortech.2013.02.039 Medline

70. J. Bredenberg, R. Ceylan, Hydrogenolysis and hydrocracking of the carbon-oxygen bond. 3.

Thermolysis in tetralin of substituted anisoles. Fuel 62, 342–344 (1983).

doi:10.1016/0016-2361(83)90093-5

71. A. Demirbaş, Mechanisms of liquefaction and pyrolysis reactions of biomass. Energy

Convers. Manage. 41, 633–646 (2000). doi:10.1016/S0196-8904(99)00130-2

72. R. N. Olcese, M. Bettahar, D. Petitjean, B. Malaman, F. Giovanella, A. Dufour, Gas-phase

hydrodeoxygenation of guaiacol over Fe/SiO2 catalyst. Appl. Catal. B 115–116, 63–73

(2012). doi:10.1016/j.apcatb.2011.12.005

73. S. Song, J. Zhang, G. Gözaydın, N. Yan, Production of terephthalic acid from corn stover

lignin. Angew. Chem. Int. Ed. 131, 4988–4991 (2019). doi:10.1002/ange.201814284

Medline

74. L. Dong, Y. Xin, X. Liu, Y. Guo, C.-W. Pao, J.-L. Chen, Y. Wang, Selective

hydrodeoxygenation of lignin oil to valuable phenolics over Au/Nb2O5 in water. Green

Chem. 21, 3081–3090 (2019). doi:10.1039/C9GC00327D

75. R. Srivastava, M. Choi, R. Ryoo, Mesoporous materials with zeolite framework: Remarkable

effect of the hierarchical structure for retardation of catalyst deactivation. Chem.

Commun. 2006 4489–4491 (2006). doi:10.1039/b612116k Medline

76. F. Imbert, M. Guisnet, S. Gnep, Comparison of Cresol Transformation on USHY and HZSM-

5. J. Catal. 195, 279–286 (2000). doi:10.1006/jcat.2000.2984

77. H. Fiege, in Ullmann's Encyclopedia of Industrial Chemistry (Wiley, 2000), pp. 419–461.

Page 106: Supplementary Materials for · 2/12/2020  · Life-cycle assessment .....19 . Supplementary text ... lignin phenolic oligomers .....87 . 5. Table. S3 Hydroxyl group content in RCF

78. P. L. Spath, M. K. Mann, “Life cycle assessment of hydrogen production via natural gas

steam reforming” (National Renewable Energy Laboratory, 2000);

www.energy.gov/sites/prod/files/2014/03/f12/27637.pdf.

79. E. Cetinkaya, I. Dincer, G. Naterer, Life cycle assessment of various hydrogen production

methods. Int. J. Hydrogen Energy 37, 2071–2080 (2012).

doi:10.1016/j.ijhydene.2011.10.064

80. J. Pérez-Ramírez, C. H. Christensen, K. Egeblad, C. H. Christensen, J. C. Groen, Hierarchical

zeolites: Enhanced utilisation of microporous crystals in catalysis by advances in

materials design. Chem. Soc. Rev. 37, 2530–2542 (2008). doi:10.1039/b809030k Medline

81. The Mol-Instincts Database; www.molinstincts.com.

82. J. Kanai, J. A. Martens, P. A. Jacobs, On the nature of the active sites for ethylene

hydrogenation in metal-free zeolites. J. Catal. 133, 527–543 (1992). doi:10.1016/0021-

9517(92)90259-K