Graphitization Behavior of Loblolly Pine Wood ... - ACS Publications

May 29, 2018 - ray diffractometer (PANalytical, Westborough, MA, USA) with a. HTK2000N ... as a standard to determine the degree of graphitization (Th...
0 downloads 0 Views 12MB Size
Subscriber access provided by NORTH CAROLINA A&T UNIV

Article

Graphitization Behavior of Loblolly Pine Wood investigated by in situ High Temperature X-ray Diffraction Seunghyun Yoo, Ching-Chang Chung, Stephen Kelley, and Sunkyu Park ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.8b01446 • Publication Date (Web): 29 May 2018 Downloaded from http://pubs.acs.org on May 30, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 39 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

1

Graphitization Behavior of Loblolly Pine Wood investigated by in situ High Temperature X-ray

2

Diffraction

3 4

Seunghyun Yooa, Ching-Chang Chungb, Stephen S. Kelleya, Sunkyu Parka,*

5 6

a

7 8

b

Department of Forest Biomaterials, North Carolina State University, Raleigh, North Carolina, 27695, United States Department of Materials Science and Engineering, North Carolina State University, Raleigh, North Carolina, 27695, United States

9 10

*

11

University, 431 Dan Allen Drive, Campus Box 8005, Raleigh, NC 27607-8005, USA

12

E-mail: [email protected]

Corresponding author: Sunkyu Park, Department of Forest Biomaterials, North Carolina State

13 14

ABSTRACT Graphitization is a complex process involving chemical and morphological

15

changes, although the detailed mechanism for different starting materials is not well understood. In this

16

work, in situ high temperature X-ray diffraction (XRD) and differential scanning calorimetry (DSC)

17

were used to examine the phase transition occurring between 1,000 and 1,500oC in loblolly pine wood

18

derived carbon materials. Electron energy loss spectroscopy (EELS) was also used to study these wood

19

derived carbon materials. XRD data showed the disappearance of disordered carbon phase between

20

1,300 and 1,400oC, followed by the formation of a crystalline graphitic phase between 1,400 and

21

1,500℃. Lattice parameters and crystal structure of the loblolly pine wood derived graphite were

22

systematically calculated from the empirical data. The presence of a large endothermic peak at 1,500oC

23

in the DSC thermogram supported this observation. Selected area electron diffraction patterns showed

24

the growth of graphitic crystallites after heat treatment. EELS spectra also supported the presence of a

25

well-developed graphite structure.

1 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 39

26 27

Keywords: Biomass graphite, Graphitization behavior, In situ X-ray diffraction, Electron energy loss

28

spectroscopy, High temperature differential scanning calorimetry

29 30

Graphitization behavior of pine biomass. This in situ

31

observation will bring a deeper understanding of sustainable

32

material conversion to versatile carbon materials.

33 34 35 36 37 38

Introduction Biomass is a naturally available carbon-rich material that exists in many forms e.g., wood, plant,

39

crop, fruit, and animal waste. When biomass is thermally treated under an inert environment, the labile

40

oxygen-rich structure decomposes to produce a carbon-rich residue. Biomass has been used as a source

41

of carbon materials since ancient time due to the abundance and the easiness in conversion.1 While

42

biomass derived carbon has been widely available for centuries, its detailed structure has not been

43

clearly elucidated due to its structural complexity. It is generally accepted that the structure of biomass

44

derived carbon material is determined by the structure of precursor biomass and the highest treatment

45

temperature (HTT). Yet, the exact pathway for the conversion of biomass to carbon is not well

46

understood. Recent work on biomass-derived carbon has shown the potential for producing high-end 2 ACS Paragon Plus Environment

Page 3 of 39 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

47

electronic materials such as graphene quantum dot and electrodes for energy storage devices.2 To

48

develop advanced biomass-derived carbon for value-added electronic materials, it is imperative to

49

understand both the chemical and morphological changes that occur during the thermal treatment.

50

Biomass is a complex mixture of cellulose, hemicellulose, lignin, and inorganic elements. The

51

ratio among semi-crystalline cellulose, amorphous hemicellulose, polyaromatic lignin, minor

52

extractives, and mineral components depends on the specific biomass source. As a result, it is difficult to

53

elucidate the specific chemical and morphological changes which take place during the thermal

54

conversion of biomass to an ordered carbon material.3-6 During the thermal conversion, long cellulose

55

and hemicellulose chains are broken into smaller molecules that are released as bio-oil and pyrolysis

56

vapors.7 The thermal conversion process of biomass is thought to be a 3-step process; 1) at a temperature

57

range between 300 to 900℃, biomass is converted into a porous-disordered biochar,8-9 2) above 900℃,

58

the disordered biochar is turned into turbostratic carbon, which has randomly distributed 2 to 3 layers of

59

small graphitic stacking, 10 and 3) finally at temperatures above 1,500oC more ordered graphitic carbon

60

is formed.

61

The graphitic carbon has a long-range crystalline structure with multiple layers of graphitic

62

stacking. Emmerich et al. showed a systematic ex situ X-ray diffraction (XRD) pattern observation

63

during the graphitization of babassu nut.11 This work tracked the biomass graphitization process at

64

temperatures up to 2,200℃, followed by cooling to room temperature. The ex situ XRD data was to

65

develop a graphitization model, butdue to recrystallization of graphitized material during the slow

66

cooling process, the ex situ observation cannot represent the real phase change during the graphitization

67

of biomass at high temperature.12

68 69

To address thelimitations of ex situ observations, the current work used in situ high temperature XRD and DSC to simultaneously observe the structural changes during the thermal treatment of 3 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 39

70

biomass. Both XRD and DSC results revealed simultaneous chemical and morphological changes at

71

temperature between 1,200 and 1,600oC. Lattice parameters and crystal structure of the loblolly pine

72

wood derived graphite were numerically calculated from the XRD data. The electron diffraction pattern

73

allowed for calculation of the reciprocal spacing distance in loblolly pine derived carbon. Finally, EELS

74

showed the evolution of electronic structure during the graphitization.

75 76

Experimental

77

High temperature in situ X-ray diffraction analysis

78

20-mesh size Loblolly Pine (Pinus taeda) particles were pre-carbonized at 800℃ for 15 minutes

79

under nitrogen gas flow (1L/min) by using OTF-1200X quartz tube furnace (MTI Corporation,

80

Richmond, CA, USA). This pre-carbonization prevents extensive mass loss during the in situ XRD

81

measurements graphitization that reduce the intensity of XRD pattern. The pre-carbonized sample was

82

labeled N800 biochar. The graphitization of the N800 biochar was analyzed using an Empyrean X-ray

83

diffractometer (PANalytical, Westborough, MA, USA) with a HTK2000N heating stage (Anton Paar,

84

Graz, Austria) and a Cu-Kα X-ray source (0.15415 nm) in a temperature range of 25 to 1,600℃. To

85

avoid chemical reactions between biochar and tungsten heating stage, a thin platinum foil was placed in

86

between biochar and tungsten heating stage, and the entire system remained in vacuum. XRD patterns

87

were acquired in a 2θ range of 10o to 35o every 100℃ on the heating cycle. The step size and count time

88

for the in situ XRD measurements were 0.026 o and 93.8 sec/step, respectively. The heating rate was

89

20K/min and the temperature was held constant for 5 minutes prior to each XRD data collection. After

90

cooling the sample to room temperature, sample was retained for the further analysis and labeled as

4 ACS Paragon Plus Environment

Page 5 of 39 1 2 3 91 4 5 6 92 7 8 9 93 10 11 12 94 13 14 95 15 16 17 96 18 19 97 20 21 98 22 23 24 99 25 26 100 27 28 101 29 30 31 102 32 33 34 103 35 36 37 104 38 39 105 40 41 106 42 43 107 44 45 46 108 47 48 49 109 50 51 52 110 53 54 55 111 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

biomass graphite. A polycrystalline natural graphite was used as a standard to determine the degree of graphitization (Thermo Fischer Scientific, Waltham, MA, USA).

Room temperature x-ray diffraction analysis Room temperature XRD patterns of N800 biochar, biomass graphite, and natural graphite were collected by using a SmartLab X-ray diffractometer (Rigaku, Woodlands, TX, USA) equipped with a Cu X-ray source. in the range of 10o – 50o 2θ at ambient conditions. The step size and count time for the in situ XRD measurements were 0.05o and 3 sec/step, respectively. Background subtraction of the XRD patterns was processed with HighScore Plus 3.0 software (PANalytical, Westboroough, MA, USA). Additional peak fitting and analyses were conducted using MAUD software (Department of Industrial Engineering, University of Trento, Trento, Italy). Thermal analysis with simultaneous thermogravimetry/differential scanning calorimetry High temperature DSC was conducted with a NETZSCH model STA 449 F1 Jupiter® simultaneous thermal analyzer. The sample was heated from 25 to 1,550℃ at heating rate of 20K/min and was remained at 1,550℃ for 10 min, under a constant stream of dry argon. During the heating, sample mass change (thermogravimetry, TG) and energy transformation (differential scanning calorimetry, DSC) were simultaneously recorded.

Transmission electron microscopy imaging and selected area electron diffraction analysis Samples for the transmission electron microscopy (TEM) were prepared by using UC7 Ultramicrotome (Leica Microsystems Inc. Buffalo Grove, IL, USA). N800 biochar, biomass graphite, 5 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 112 4 5 6 113 7 8 114 9 10 115 11 12 116 13 14 15 117 16 17 18 118 19 20 21 119 22 23 24 120 25 26 121 27 28 122 29 30 31 123 32 33 124 34 35 125 36 37 126 38 39 40 127 41 42 43 128 44 45 46 129 47 48 49 130 50 51 52 131 53 54 132 55 56 57 58 59 60

Page 6 of 39

and natural graphite were sliced into samples of approximately 100 nm thick. Sliced specimens were transferred to the copper TEM grid (Protochips, Morrisville, NC, USA). TEM images were taken by Titan 80-300 Probe Aberration Corrected Scanning Transmission Electron Microscope (FEI, Hillsboro, OR). TEM images were taken at high magnification using diffracted beam at a voltage of 200kV to provide point-to-point resolution of 0.20nm. Selected area diffraction patterns (SAED) were taken at TEM mode. The reciprocal space distance was measured by ImageJ to determine diffraction planes.

Electron energy loss spectroscopy analysis (EELS) The EELS were collected with a Gatan (Pleasanton, CA, USA) using STEM mode with a high voltage of 200 kV and energy resolution of 0.15 eV. Sample zero energy loss, low energy loss, and carbon Kedge core energy loss spectra were collected, and zero loss peak subtraction, background subtraction, and electron multiple scattering effect removal were conducted using the Gatan Digital Micrograph software.13 Surface plasmon and bulk plasmon energy were obtained from extracted low energy loss spectrum. Corrected carbon K-edge energy loss spectrum was further deconvoluted into three Gaussian spectra to calculate the π to π* transition ratio.14-15

Results and Discussion Graphitization behavior of loblolly pine investigated by in situ X-ray diffraction Figure 1 shows the XRD pattern for the development of graphitic structure from biochar to turbostratic carbon to primitive graphitic carbon. As the temperature increases from 800 to 1,300℃, the intensity of the disordered peak (located at around 24° 2θ) decreases with temperature. The strong and

6 ACS Paragon Plus Environment

Page 7 of 39 1 2 3 133 4 5 6 134 7 8 135 9 10 136 11 12 137 13 14 15 138 16 17 139 18 19 140 20 21 22 141 23 24 142 25 26 143 27 28 29 144 30 31 145 32 33 146 34 35 147 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

broad peak at low 2θ angle is the combination of (002) reflection of few-layer graphitic stacking and disordered carbon structure in biochar. Importantly and unexpectedly, the diffraction peak of disordered carbon diminishes between 1,300 to 1,500℃. Previous ex situ study of babassu nut graphitization at room temperature observed continuous graphitic structure development during the thermal treatment.11 This suggests a model for structural development where the thermally treated biomass is continuously converted from disordered carbon to organized graphitic carbon layers.10 However, the ex situ model ignores the potential for recrystallization of carbon phases during the cooling process that it does not directly represent the biomass graphitization phenomenon.12 As a result, in this work in situ observations suggest a different pathway for the biomass graphitization process. These results suggest that the graphitic structure does not develop continuously from the disordered carbon structure, but the process includes an intermediate process where no solid crystalline graphitic structure exists. XRD pattern suggests that the disordered carbon was nearly disappeared at 1,400℃. After reaching 1,500℃, a minor tiny peak indicative of graphite (002), appears at 26.67°. These observations suggest that the intermediate turbostratic carbon structure re-organizes to create the final graphite structure. These changes are consistent with the DSC results shown below.

7 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 148 33 149 34 35 150 36 37 38 151 39 40 41 152 42 43 153 44 45 154 46 47 155 48 49 50 156 51 52 53 54 55 56 57 58 59 60

Page 8 of 39

Figure 1. In situ X-ray diffraction patterns at the temperature range of 800 to 1,600℃.

Room temperature XRD patterns of N800 biochar, biomass graphite, and natural graphite are presented in Figure 2. N800 biochar shows a broad pattern at low 2θ angle around 24°. Biomass graphite produced at the highest treatment temperature (HTT) of 1,600℃ shows a sharp (002) reflection with a weak and broad background pattern. Natural graphite shows a strong (002) reflection. These XRD patterns were background subtracted and refined to calculate lattice parameters, layer coherence length, and degree of graphitization (Table 1).

8 ACS Paragon Plus Environment

Page 9 of 39 1 2 3 157 4 5 6 158 7 8 159 9 10 11 12 160 13 14 15 161 16 17 18 162 19 20 21 163 22 23 164 24 25 26 165 27 28 29 166 30 31 167 32 33 34 168 35 36 37 169 38 39 170 40 41 42 171 43 44 172 45 46 173 47 48 49 174 50 51 175 52 53 176 54 55 177 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Lattice parameters were calculated by Bragg’s law and reciprocal lattice vector notation of hexagonal crystal system (equations (1) and (2)) 16-17 where dhkl is the interlayer spacing of (hkl), λ is the wavelength of X-ray radiation, and θ is the Bragg angle.

1

2 𝑑𝑑ℎ𝑘𝑘𝑘𝑘

dℎ𝑘𝑘𝑘𝑘 = =

λ 2 sin(𝜃𝜃)

(1)

4 ℎ2 + ℎ𝑘𝑘 + 𝑘𝑘 2 𝑙𝑙 2 � � + 3 𝑎𝑎2 𝑐𝑐 2

(2)

Two layer coherence lengths, La (orthogonal to c-direction) and Lc (parallel to c-direction), were calculated by Scherrer equation (equation (3)) to show the graphitic ordering 9, 17-18 where β is the FWHM (in radians of theta) of each d spacing, λ is the wavelength of X-ray, and θ is the Bragg angle. 𝐿𝐿𝑎𝑎 =

1.84𝜆𝜆 , 𝛽𝛽cos(𝜃𝜃)

𝐿𝐿𝑐𝑐 =

0.91𝜆𝜆 𝛽𝛽cos(𝜃𝜃)

(3)

The degree of graphitization is calculated by the d002 fractional ratio (equation (4)) of ideal graphite (3.354Å) and non-ideal graphite (3.440Å).17 g� =

3.440 − 𝑑𝑑002 3.440 − 3.354

(4)

Lattice parameters of biomass graphite were calculated as a = 2.619Å and c = 6.735Å. Compared to lattice parameters of natural graphite, lattice parameter a is 6.25% longer and lattice parameter c is 0.02% longer. Both La and Lc dramatically increased by 1,470% and 2,020% from N800 biochar to biomass graphite. The (002) interlayer spacing distance is the largest at N800 biochar. The degree of graphitization value of biomass graphite is similar to that of natural graphite. Furthermore, the thermal expansion in crystal structure is observed from biomass graphite XRD pattern obtained at 1,600℃. Lattice parameter a expanded 2.0% and lattice parameter c expanded 4.7% when compared to lattice parameters obtained at room temperature. Elemental composition and weight yield of N800 biochar, biomass graphite, and natural graphite are given at Table S1. 9 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 178 16 179 17 180 18 181 19 20 182 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 183 37 184 38 39 185 40 41 42 186 43 44 187 45 46 188 47 48 49 189 50 51 190 52 53 191 54 55 56 192 57 58 59 60

Page 10 of 39

Figure 2. Room temperature XRD patterns of (a) N800 biochar, (b) biomass graphite (patterns collected after cool down to room temperature), and (c) natural graphite obtained at room temperature. Table 1. Quantified lattice parameters of N800 biochar, biomass graphite, and natural graphite.

N800

Biomass Graphite

Measurement Temperature

25℃

a (Å) c (Å) Lc (Å)

N/A N/A 8.3

2.658 7.020 158.3

La (Å) d002 (Å) g̅ (%)

23.5 4.046 N/A

331.5 3.510 N/A

Natural Graphite

Cooled to 25℃

25℃

2.671 7.049 175.9

2.619 6.735 176.3

2.465 6.734 235.8

368.0 3.524 N/A

369.0 3.368 84.12

316.6 3.367 84.83

1500℃ 1600℃

Thermal analysis with simultaneous thermogravimetry/differential scanning calorimetry analysis Figure 3 shows heat flows and mass changes during the high temperature thermal analysis of biochar. From the DSC curve (black line), four phases are distinguished depending on heat input/output status. Below 350℃ (17 min), there is an endothermic peak in the DSC and extensive mass loss in TGA due to the initial degradation of biomass components.19 From 350 to 800℃ (36 min), the DSC shows a broad exothermic region, which is a typical feature of biomass carbonization

20-21

and limited mass loss

supports the formation of the disordered biochar structure above 500℃.7 The XRD in this region is dominated by the broad (002) XRD pattern, which represents a disordered carbon structure. The third 10 ACS Paragon Plus Environment

Page 11 of 39 1 2 3 193 4 5 6 194 7 8 195 9 10 196 11 12 197 13 14 15 198 16 17 199 18 19 200 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 201 202 44 45 203 46 47 204 48 49 205 50 51 206 52 53 54 207 55 56 208 57 58 59 60

ACS Sustainable Chemistry & Engineering

phase is initiated at about 800℃, where turbostratic carbon structure starts to develop 8. As the turbostratic carbon structure develops, the DSC heat flow changes from exothermic reaction to endothermic. The slope of DSC curve is modest until the temperature reaches 1,300℃ (62min), where (002) XRD pattern disappears. At about 1,300℃, the slope of DSC curve becomes steeper and a major endothermic peak occurs at 1,530℃ (75min) consistent with the conversion of disordered biochar to ordered crystalline graphite, which requires significant energy input. As mentioned above the XRD pattern also shows a significant increase in order between 1,500 and 1,600oC. Both the DSC and XRD results are consistent with the formation of ordered graphite at temperatures between 1,500 and 1,600oC.

Figure 3. Simultaneous TG/DSC curve of loblolly pine wood.

Transmission electron microscopy image and selected area electron diffraction pattern analysis Figure 4(a, b, c) shows SAED patterns of N800 biochar, biomass graphite, and natural graphite, respectvely. Natural graphite shows clear polycrystalline diffraction rings while N800 biochar and natural graphite exhibit poorly defined diffraction rings. From SAED patterns of N800 biochar and biomass graphite, a strong diffraction ring from (002) reflection and weak diffraction rings from (100) and (011) 11 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 209 4 5 6 210 7 8 211 9 10 212 11 12 213 13 14 15 214 16 17 215 18 19 216 20 21 22 217 23 24 218 25 26 219 27 28 29 220 30 31 221 32 33 222 34 35 36 37 38 39 40 41 42 43 44 45 46 47 223 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 39

are observed. Existence of three diffraction rings also corresponds to the observation from XRD patterns. A comparison between N800 biochar and biomass graphite shows that as the N800 biochar is heated the (112) reflection develops and the radius of diffraction ring becomes larger. Overall SAED pattern becomes clearer at biomass graphite. TEM images taken at high magnification show the development of graphitic structure within samples. N800 biochar shows a relatively large piece of carbon material without any layered structure (Figure 4(d)). Biomass graphite shows a primitive layered structure related topology, which is consistent with the formation of graphitic stacking (Figure 4(e)). Natural graphite clearly shows single graphene layer and few layers of graphitic stacking (Figure 4(f)). Together the SAED patterns and TEM images suggest that the biomass graphite produced at 1,600℃ has a significant portion of crystalline graphite that coexists with disordered carbon. Combined with the XRD analysis result, the crystalline part of the biomass graphite has a similar structural property as natural graphite. However, the disordered part of biomass graphite confirmed by the XRD pattern at low 2θ angle region and TEM analysis still differentiates the biomass graphite from natural graphite.

12 ACS Paragon Plus Environment

Page 13 of 39 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 224 16 225 17 226 18 227 19 20 228 21 22 229 23 24 25 230 26 27 231 28 29 232 30 31 233 32 33 34 234 35 36 235 37 38 236 39 40 41 237 42 43 238 44 45 239 46 47 48 49 50 51 52 53 54 55 56 240 57 58 59 60

ACS Sustainable Chemistry & Engineering

Figure 4. SAED patterns (a-c) and TEM images (d-f) of N800 biochar (a, d), biomass graphite produced at 1,600℃ (b, e), and natural graphite (c, f). TEM images were obtained at 50 nm magnification. Electron energy loss spectroscopy analysis (EELS) EELS is a powerful analytical technique that reveals details of both the electronic and chemical structure of material at the molecular orbital level. Figure 5 shows low energy loss region (a, b, c) and carbon K-edge energy loss (d, e, f) spectra of the three samples. At low energy loss region, plasmonic features, collective oscillations of electrons in material, are observed. At 5 to 6 eV region, a surface plasmon peak is observed. At 22 to 27 eV region, a bulk plasmon peak is observed. As the material has more ordered graphitic structure, the peak position of surface and bulk plasmon shift to higher energy region (Table 2). The bulk plasmon is an indicator of different carbon structure

13, 22

. Diamond bulk

plasmon peak appears at 33 eV, graphite bulk plasmon peak appears at 27 eV, and amorphous and disordered carbon bulk plasmon peaks appear from 22 to 25 eV.9, 13, 23 Quantified bulk plasmon peak values found in this work correspond to values from these references.

13 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 241 13 242 14 243 15 244 16 17 18 245 19 20 246 21 22 247 23 24 248 25 26 27 249 28 29 250 30 31 251 32 33 34 252 35 36 253 37 38 39 254 40 41 42 255 43 44 256 45 46 47 257 48 49 258 50 51 259 52 260 53 54 55 56 57 58 59 60

Page 14 of 39

Figure 5. Low energy loss and Carbon K-edge core energy loss spectra of (a, d) N800 biochar, (b, e) biomass graphite, and (c, f) natural graphite.

From the carbon K-edge electron energy loss spectrum, an electronic transition from carbon 1s (core shell orbital) to π* (valence shell orbital) is detected (Figure 5(d, e, f)). This transition is closely related to the aromatic sp2 ratio in the carbon materials (Add a reference). Specifically, when an electron beam passes through a thin carbon specimen, three different electronic transitions occur and these three transitions can be fitted with Gaussian curves (shown in Figure S1, S2, S3) 13. The transition at 285.0eV (G1) represents the electronic transition from carbon 1s core orbital to C=C π* bonding orbital.14-15, 22-25 The second transition at 292.0eV (G2) represents the electronic transition from carbon 1s core orbital to C-C σ* bonding orbital and the third transition at 298.0eV (G3) represents the electronic transition from carbon 1s core orbital to C=C σ* bonding orbital.14-15, 22-25 The ratio of sp2 bonding is given as, C 1s to π∗ transition ratio =

𝑎𝑎𝑎𝑎𝑎𝑎𝑎𝑎(𝜋𝜋 ∗ ) 𝐺𝐺1 = 𝑎𝑎𝑎𝑎𝑎𝑎𝑎𝑎(𝜋𝜋 ∗ + 𝜎𝜎 ∗ ) 𝐺𝐺1 + 𝐺𝐺2 + 𝐺𝐺3

C 1s to π* transition ratio of N800 biochar is 0.093 and as the sample is heat treated at 1,600℃, the transition ratio increases to 0.100 (Table 2). The transition ratio of biomass graphite is about 80% of that of natural graphite.

Table 2. Quantified parameters from electron energy loss spectra

14 ACS Paragon Plus Environment

Page 15 of 39 1 2 3 4 5 6 7 8 9 10 11 12 13 261 14 262 15 263 16 264 17 18 19 265 20 21 266 22 23 267 24 25 26 268 27 28 269 29 30 270 31 32 271 33 34 35 272 36 37 273 38 39 274 40 41 42 275 43 44 276 45 46 47 277 48 49 50 278 51 52 53 279 54 55 280 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

N800 0.093 ± 0.002

C1s to π* Transition Ratio Surface Plasmon 5.10 ± 0.57 Energy (eV) Bulk Plasmon 22.94 ± 0.34 Energy (eV) C-C Bond Length (Å) 1.4242 ± 0.0004

Biomass Graphite 0.100 ± 0.001

Natural Graphite 0.125 ± 0.003

5.66 ± 0.33

6.35 ± 0.06

23.68 ± 0.44

26.78 ± 0.42

1.4232 ± 0.0013

1.4207 ± 0.0003

* 5 samples measurements were done for N800 and biomass graphite and 4 sample measurements were done for natural graphite

At carbon K-edge electron energy loss spectrum, a broad feature known as a multiple scattering resonance (MSR) structure appears at 325 to 327 eV (Figure 5(d, e, f)). The MSR structure is generated by the local electron resonance in carbon-carbon bonding.22 When an atom is excited by electron beam, core electrons are ejected and backscatter from first and second nearest neighbor atoms. The backscattering in electron energy loss near-edge structure (ELNES) is effective within 1nm from the excited core atom and appears as MSR structure.26 The wave number of ejected electron k follows the below relationship, kR = constant

(5)

where R is bond length 27. The MSR energy (EMSR) is proportional to the inverse square of R as shown in equation (6).22, 28-29 The distance from excited core carbon atom to the second nearest neighbor carbon atom, RMSR, is given as 2.467 Å and the constant, KMSR, is given as 1980.8904 eVÅ2.22 The average C-C bond length is the radius of the primary shell, so RMSR value from equation (6) fits into equation (7). 𝐸𝐸𝑀𝑀𝑀𝑀𝑀𝑀 =

𝐾𝐾𝑀𝑀𝑀𝑀𝑀𝑀 2 𝑅𝑅𝑀𝑀𝑀𝑀𝑀𝑀

C − C bond length =

𝑅𝑅𝑀𝑀𝑀𝑀𝑀𝑀 2 sin(60°)

(6)

(7)

Calculated average C-C bond lengths of N800 biochar, biomass graphite, and natural graphite are given at Table 2. Average C-C bond length is the longest at N800 biochar which value is 1.424Å. Biomass 15 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 281 4 5 6 282 7 8 283 9 10 284 11 12 285 13 14 15 16 286 17 18 19 287 20 21 22 288 23 24 289 25 26 290 27 28 29 291 30 31 32 292 33 34 35 293 36 37 38 39 294 40 41 295 42 43 44 296 45 46 47 297 48 49 50 51 298 52 53 299 54 55 300 56 57 58 59 60

Page 16 of 39

graphite has shorter average C-C bond length of 1.423Å. Natural graphite has the shortest average C-C bond length of 1.421Å which is identical to known graphite C-C bond length.30 The plasmonic features can be interpreted by the quantum mechanical approach. By using the Drude model, the displacement of a quasi-free electrons in a local electric field can be modeled. Equations for the resonance frequency of plasma oscillation and the plasmon energy are given as, 𝑛𝑛𝑒𝑒 2 𝜔𝜔𝑝𝑝 = � 𝜀𝜀0 𝑚𝑚

𝐸𝐸𝑝𝑝 = ħ𝜔𝜔𝑝𝑝

(10)

where n is the electron density (valence electrons per unit volume), e is the elementary charge of electron, m is the mass of electron, and ε0 is the vacuum permittivity 13. The resonance frequency of plasma directly notates that the bulk plasmon energy is proportional to the square root of the electron density. However, plasma resonance in a real solid is strongly confined by the damping factor occurred by the single electron transition 13. As a result, oscillator strength term f is added to the equation 31-32. f=

2𝑚𝑚𝐸𝐸𝑔𝑔 𝑒𝑒 2 ħ2

2 𝑎𝑎𝑛𝑛𝑛𝑛

(11)

𝑓𝑓𝑓𝑓𝑒𝑒 2 𝐸𝐸𝑝𝑝 = ħ� 𝜀𝜀0 𝑚𝑚

(12)

Eg is an energy gap and ani is an atomic dipole matrix element for the excitation. Then, equation of changing energy gap (ΔEg, equation (13)) induces equation (14).32 2𝜋𝜋 2 ħ2 ∆E𝑔𝑔 = ∗ 2 𝑚𝑚 𝑑𝑑

∆𝐸𝐸𝑝𝑝

(𝑏𝑏)

𝐸𝐸𝑃𝑃

𝜋𝜋 2 ħ2 1 = (𝑏𝑏) 2 𝑚𝑚∗ 𝐸𝐸 𝑑𝑑 𝑔𝑔

(13)

(14)

m* is the effective mass, d is the average distance between carbon atoms, and the subscript (b) remarks the properties of bulk. Equation (14) implies a physical relationship between molecular bond length and plasmonic features. The inverse square of average C-C bond length is plotted as a function of bulk plasmon 16 ACS Paragon Plus Environment

Page 17 of 39 1 2 3 301 4 5 6 302 7 8 303 9 10 304 11 12 305 13 14 15 306 16 17 307 18 19 308 20 21 22 309 23 24 310 25 26 311 27 28 29 30 31 32 33 34 35 36 37 38 39 312 40 313 41 314 42 43 315 44 45 316 46 47 317 48 49 318 50 51 52 319 53 54 320 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

energy in Figure 6. The bulk plasmon energy is physically related to the average C-C bond length. The inverse square of average C-C bond length values shows linear correlation with increasing bulk plasmon energy. N800 biochar with its disordered structure and longer average C-C bond length reflects the smaller value of bulk plasmon energy. As the degree of graphitization increases, the average C-C bond length decreases and the bulk plasmon energy increases. It is known that a theoretically ‘perfect’ graphitic structure has 27.0 eV of the bulk plasmon energy with 0.496Å-2 of the inverse square of the average C-C bond length.22 In this work the average C-C bond length and bulk plasmon energy of natural graphite are almost the same as that of theoretically perfect graphite. This quantum mechanical relationship can be used as an indicator of the carbon structure development during the carbonization and graphitization of biomass.

Figure 6. Correlation between bulk plasmon excitation energy and inverse square of the average C-C bond length

Conclusions Phase transitions occurring during the loblolly pine wood graphitization was quantitatively analyzed based on empirical observations from a series of complimentary techniques. From the in situ XRD measurement, disappearance of disordered carbon phase was observed between 1,300 to 1,400℃ where no crystalline structure exists. But with additional heating to 1,500 to 1,600oC this non-ordered

17 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 321 4 5 6 322 7 8 323 9 10 324 11 12 325 13 14 15 326 16 17 327 18 19 328 20 21 22 329 23 24 330 25 26 331 27 28 332 29 30 333 31 32 33 334 34 35 335 36 37 336 38 39 40 337 41 42 338 43 44 339 45 46 340 47 48 49 341 50 51 342 52 53 343 54 55 344 56 57 58 59 60

Page 18 of 39

state is rapidly converted into a highly order graphite carbon. Previous ex situ studies suggested a continuous structure development model for the progressive formation of graphitic structure. However, the observations from this work strongly suggest a new model, where there is a non-continuous process for the formation of graphitic structure. The endothermic nature of graphite formation was confirmed by the DSC measurement. The results from analysis of graphitic lattice parameters, electron diffraction, electron microscopy imaging, sp2 ratio, average C-C bond length, and plasmonic features all provide compelling evidence for the formation of graphitic structure from loblolly pine wood. This combination of tools, and the results of this work help to clarify the detailed mechanisms involved in the formation of biomass derived graphitic materials.

Acknowledgement. This project is supported by the USDA National Institute of Food and Agriculture (Award 2011-68005-30410) and the US-DOE Office of Energy Efficiency and Renewable Energy (Award Number DE-EE0006639). The characterization was performed at the Analytical Instrumentation Facility (AIF), supported by the State of North Carolina and the National Science Foundation (Award ECCS-1542015). A special appreciation to the NETZSCH Instruments Applications Laboratory (Burlington, MA, USA) for running the high temperature DSC measurement.

Supporting Information Supplementary Table S1: Yield and elemental composition of samples Supplementary Figure S1: EELS spectrum and deconvolution of N800 biochar Supplementary Figure S2: EELS spectrum and deconvolution of biomass graphite Supplementary Figure S3: EELS spectrum and deconvolution of natural graphite

References 18 ACS Paragon Plus Environment

Page 19 of 39 1 2 3 345 4 346 5 347 6 7 348 8 349 9 350 10 351 11 352 12 353 13 354 14 355 15 356 16 17 357 18 358 19 359 20 360 21 361 22 362 23 363 24 364 25 365 26 27 366 28 367 29 368 30 369 31 370 32 371 33 372 34 35 373 36 374 37 375 38 376 39 377 40 378 41 379 42 380 43 381 44 45 382 46 383 47 384 48 385 49 386 50 387 51 388 52 389 53 54 390 55 391 56 392 57 58 59 60

ACS Sustainable Chemistry & Engineering

(1) Jorio, A.; Ribeiro-Soares, J.; Cancado, L. G.; Falcao, N. P. S.; Dos Santos, H. F.; Baptista, D. L.; Ferreira, E. H. M.; Archanjo, B. S.; Achete, C. A., Microscopy and spectroscopy analysis of carbon nanostructures in highly fertile Amazonian anthrosoils. Soil Tillage Res. 2012, 122, 61-66. (2) Dong, Y. Q.; Zhou, N. N.; Lin, X. M.; Lin, J. P.; Chi, Y. W.; Chen, G. N., Extraction of Electrochemiluminescent Oxidized Carbon Quantum Dots from Activated Carbon. Chem. Mater. 2010, 22, 58955899. (3) Yaman, S., Pyrolysis of biomass to produce fuels and chemical feedstocks. 2004, 45, 651-671. (4) Kim, D. Y.; Nishiyama, Y.; Wada, M.; Kuga, S., Graphitization of highly crystalline cellulose. Carbon. 2001, 39, 1051-1056. (5) Wang, Y.; Yang, R.; Li, M.; Zhao, Z. J., Hydrothermal preparation of highly porous carbon spheres from hemp (Cannabis sativa L.) stem hemicellulose for use in energy-related applications. Ind Crop Prod. 2015, 65, 216-226. (6) RodriguezMirasol, J.; Cordero, T.; Rodriguez, J. J., High-temperature carbons from kraft lignin. Carbon. 1996, 34, 43-52. (7) Kim, K. H.; Eom, I. Y.; Lee, S. M.; Choi, D.; Yeo, H.; Choi, I. G.; Choi, J. W., Investigation of physicochemical properties of biooils produced from yellow poplar wood (Liriodendron tulipifera) at various temperatures and residence times. J. Anal. Appl. Pyrolysis. 2011, 92, 2-9. (8) Keiluweit, M.; Nico, P. S.; Johnson, M. G.; Kleber, M., Dynamic Molecular Structure of Plant BiomassDerived Black Carbon (Biochar). Environ. Sci. Technol. 2010, 44, 1247-1253. (9) Yoo, S.; Kelley, S. S.; Tilotta, D. C.; Park, S., Structural Characterization of Loblolly Pine Derived Biochar by X-ray Diffraction and Electron Energy Loss Spectroscopy. 2018, 6, 2621-2629. (10) Lehmann, J.; Joseph, S., Biochar for environmental management : science, technology and implementation. Second ed.; Routledge, Taylor & Francis Group: London ; New York, 2015. (11) Emmerich, F. G.; Desousa, J. C.; Torriani, I. L.; Luengo, C. A., Applications of a Granular Model and Percolation Theory to the Electrical-Resistivity of Heat-Treated Endocarp of Babassu Nut. Carbon. 1987, 25, 417424. (12) Guet, J. M.; Tchoubar, D., Insitu X-Ray-Diffraction Performed on Mesophase Formation during the Carbonization of Acenaphtylene. Carbon. 1985, 23, 273-280. (13) Egerton, R. F., Electron energy-loss spectroscopy in the electron microscope. Third ed.; Springer: New York, 2011. (14) Zhang, Z. L.; Brydson, R.; Aslam, Z.; Reddy, S.; Brown, A.; Westwood, A.; Rand, B., Investigating the structure of non-graphitising carbons using electron energy loss spectroscopy in the transmission electron microscope. Carbon. 2011, 49, 5049-5063. (15) Marriott, A. S.; Hunt, A. J.; Bergstrom, E.; Wilson, K.; Budarin, V. L.; Thomas-Oates, J.; Clark, J. H.; Brydson, R., Investigating the structure of biomass-derived non-graphitizing mesoporous carbons by electron energy loss spectroscopy in the transmission electron microscope and X-ray photoelectron spectroscopy. Carbon. 2014, 67, 514-524. (16) DeGraef, M.; McHenry, M. E., Structure of Materials: An Introduction to Crystallography, Diffraction and Symmetry, 2nd Edition. 2012, 1-739. (17) Seehra, M. S.; Pavlovic, A. S., X-Ray-Diffraction, Thermal-Expansion, Electrical-Conductivity, and Optical Microscopy Studies of Coal-Based Graphites. Carbon. 1993, 31, 557-564. (18) Kercher, A. K.; Nagle, D. C., Microstructural evolution during charcoal carbonization by X-ray diffraction analysis. Carbon. 2003, 41, 15-27. (19) Yang, H. P.; Yan, R.; Chen, H. P.; Lee, D. H.; Zheng, C. G., Characteristics of hemicellulose, cellulose and lignin pyrolysis. Fuel. 2007, 86, 1781-1788. (20) Rautiainen, M. Torrefied and carbonized wood, fuel properties and turn of exothermic reaction. Helsingin yliopisto, 2014. (21) Brenes, M. D., Biomass and bioenergy: new research. Nova Publishers: 2006. 19 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 393 4 394 5 395 6 396 7 8 397 9 398 10 399 11 400 12 401 13 402 14 403 15 16 404 17 405 18 406 19 407 20 408 21 409 22 410 23 411 24 412 25 26 413 27 414 28 415 29 30 416 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 39

(22) Daniels, H.; Brydson, R.; Rand, B.; Brown, A., Investigating carbonization and graphitization using electron energy loss spectroscopy (EELS) in the transmission electron microscope (TEM). Philos. Mag. 2007, 87, 4073-4092. (23) Berger, S. D.; Mckenzie, D. R.; Martin, P. J., Eels Analysis of Vacuum Arc-Deposited Diamond-Like Films. Philos. Mag. Lett. 1988, 57, 285-290. (24) Galvan, D.; Pei, Y. T.; De Hosson, J. T. M.; Cavaleiro, A., Determination of the sp(3) C content of a-C films through EELS analysis in the TEM. Surf. Coat. Technol. 2005, 200, 739-743. (25) Muller, J. O.; Su, D. S.; Wild, U.; Schlogl, R., Bulk and surface structural investigations of diesel engine soot and carbon black. Phys. Chem. Chem. Phys. 2007, 9, 4018-4025. (26) Wang, F.; Egerton, R. F.; Malac, M.; McLeod, R. A.; Moreno, M. S., The spatial resolution of electron energy loss and x-ray absorption fine structure. J. Appl. Phys. 2008, 104. (27) Bianconi, A.; Incoccia, L.; Stipcich, S., EXAFS and near edge structure : proceedings of the international conference, Frascati, Italy, September 13-17, 1982. Springer-Verglag: Berlin; New York, 1983. (28) Craven, A. J.; Garvie, L. A. J., Electron-Energy-Loss near-Edge Structure (Elnes) on the Carbon K-Edge in Transition-Metal Carbides with the Rock-Salt Structure. Microsc., Microanal., Microstruct. 1995, 6, 89-98. (29) McCulloch, D.; Brydson, R., Carbon K-shell near-edge structure calculations for graphite using the multiple-scattering approach. J. Phys.: Condens. Matter. 1996, 8, 3835. (30) Stankovich, S.; Piner, R. D.; Chen, X. Q.; Wu, N. Q.; Nguyen, S. T.; Ruoff, R. S., Stable aqueous dispersions of graphitic nanoplatelets via the reduction of exfoliated graphite oxide in the presence of poly(sodium 4styrenesulfonate). J Mater Chem. 2006, 16, 155-158. (31) Hummel, R. E., Electronic properties of materials. 4th ed.; Springer: New York, 2011. (32) Mitome, M.; Yamazaki, Y.; Takagi, H.; Nakagiri, T., Size Dependence of Plasmon Energy in Si Clusters. J. Appl. Phys. 1992, 72, 812-814.

20 ACS Paragon Plus Environment

Page 21 of 39 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

276x266mm (96 x 96 DPI)

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

248x190mm (150 x 150 DPI)

ACS Paragon Plus Environment

Page 22 of 39

Page 23 of 39 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

396x278mm (96 x 96 DPI)

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

299x190mm (150 x 150 DPI)

ACS Paragon Plus Environment

Page 24 of 39

Page 25 of 39 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

297x297mm (96 x 96 DPI)

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

297x297mm (96 x 96 DPI)

ACS Paragon Plus Environment

Page 26 of 39

Page 27 of 39 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

297x297mm (96 x 96 DPI)

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

297x297mm (96 x 96 DPI)

ACS Paragon Plus Environment

Page 28 of 39

Page 29 of 39 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

297x297mm (96 x 96 DPI)

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

297x297mm (96 x 96 DPI)

ACS Paragon Plus Environment

Page 30 of 39

Page 31 of 39 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

158x119mm (96 x 96 DPI)

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

158x119mm (96 x 96 DPI)

ACS Paragon Plus Environment

Page 32 of 39

Page 33 of 39 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

158x119mm (96 x 96 DPI)

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

158x119mm (96 x 96 DPI)

ACS Paragon Plus Environment

Page 34 of 39

Page 35 of 39 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

158x119mm (96 x 96 DPI)

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

158x119mm (96 x 96 DPI)

ACS Paragon Plus Environment

Page 36 of 39

Page 37 of 39 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

414x241mm (96 x 96 DPI)

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

146x60mm (96 x 96 DPI)

ACS Paragon Plus Environment

Page 38 of 39

Page 39 of 39 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

165x42mm (96 x 96 DPI)

ACS Paragon Plus Environment