Concurrent Cellulose Hydrolysis and Esterification to Prepare a

Dec 28, 2015 - Thus, they can be used for melt-processing operations performed at moderately high temperatures without thermal decomposition. Nanocomp...
0 downloads 12 Views 6MB Size
Subscriber access provided by CMU Libraries - http://library.cmich.edu

Article

Concurrent Cellulose Hydrolysis and Esterification to Prepare SurfaceModified Cellulose Nanocrystal Decorated with Carboxylic Acid Moieties Stephen Spinella, Anthony Maiorana, Qian Qian, Nathan J. Dawson, Victoria Hepworth, Scott A McCallum, Manoj Ganesh, Kenneth D. Singer, and Richard A Gross ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.5b01489 • Publication Date (Web): 28 Dec 2015 Downloaded from http://pubs.acs.org on January 2, 2016

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 free 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 accessible to all readers and 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.

ACS Sustainable Chemistry & Engineering 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

Concurrent Cellulose Hydrolysis and Esterification to Prepare Surface-Modified Cellulose Nanocrystal Decorated with Carboxylic Acid Moieties Stephen Spinella†,‡ Anthony Maiorana,† Qian Qian,† Nathan J. Dawson,§ Victoria Hepworth,† Scott A. McCallum, † Manoj Ganesh,‡ Kenneth D. Singer,§ and Richard A. Gross† †

Department of Chemistry and Biology, Rensselaer Polytechnic Institute (RPI), 4005B BioTechnology Bldg., 110 8th Street, Troy, N.Y. 12180, USA ‡

Department of Chemical and Bimolecular Engineering, NYU Polytechnic School of Engineering, Six Metrotech Center, Brooklyn, New York 11201, USA

§

Department of Physics, Case Western Reserve University, 2076 Adelbert Road, Cleveland, Ohio 44106, USA.

*Corresponding Author Richard A. Gross; Email: [email protected]; Ph: 518-276-3734; Fax: 518-276-3405

Key Words: Cellulose nanocrystals, Fischer esterification, cellulose hydrolysis, carboxylic acid, nanocomposite, polyvinyl alcohol

ACS Paragon Plus Environment

1

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

ABSTRACT

Cellulose nanocrystals (CNCs) were modified with natural di- and tricarboxylic acids using two concurrent acid-catalyzed reactions including hydrolysis of amorphous cellulose segments and Fischer esterification, resulting in the introduction of free carboxylic acid functionality onto CNC surfaces. CNC esterification was characterized by Fourier Transform Infrared Spectroscopy,

13

C solid state magic-angle spinning (MAS) and conductometric titration

experiments. Average degree of substitution values for malonate, malate, and citrate CNCs are 0.16, 0.22 and 0.18, respectively. Despite differences in organic acid pKa, optimal HCl cocatalyst concentrations were similar for malonic, malic and citric acids. After isolation of modified CNCs, residual cellulose co-products were identified that are similar to microcrystalline cellulose based on SEM and XRD analysis. As proof of concept, recycling experiments were carried to increase the yield of citrate CNCs. The by-product was then recycled by subsequent citric acid/HCl treatments that resulted in 55% total yield of citrate CNCs. The crystallinity, morphology, and substitution of citrate CNCs from recycled cellulose coproduct is similar to modified citrate CNCs formed in the first reaction cycle. Thermal stability of all modified CNCs under air and nitrogen resulted in T10% and T50% values above 256 °C and 323 °C, respectively. Thus, they can be used for melt-processing operations performed at moderately high temperatures without thermal decomposition. Nanocomposites of polyvinyl alcohol with modified CNCs (1 wt% malonate-, malate-, citrate and unmodified CNCs) were prepared. An increase in the thermal decomposition temperature by almost 40 °C was obtained for PVOH-citrate modified CNC nanocomposites.

ACS Paragon Plus Environment

2

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

INTRODUCTION Cellulose nanocrystals (CNCs) are of great interest due to their unique physical and chemical properties, intrinsic abundance in nature, renewability and sustainability.1,2 Controlled acid hydrolysis of macroscopic cellulose effectively removes amorphous domains thereby liberating highly crystalline rod-like CNC’s.2 Since CNC’s are derived from the most abundant polymer on earth, developing efficient routes for their production and modification can lead to low-cost nanocrystals that: i) are biodegradable, ii) have a high axial elastic modulus (110 to 180 GPa),3 iii) low density (approximately 100 g/m3), and iv) possess highly reactive surfaces due to the presence of hydroxyl groups.4 The aspect ratio (length/width) of uniaxial CNCs can further be controlled by selecting the appropriate cellulose source. Aspect ratios range from 10-30 for CNCs extracted from cotton5 to 70 for CNCs extracted from tunicates.6 The unique structure and morphology of CNCs have stimulated investigation of these materials for: reinforcement agents in hydrogels7 and hydrophobic polymers,8 components in tissue engineering materials9 and drug delivery systems,10 supports for enzyme immobilization,11 and as building blocks for electronic materials.12 Despite vast progress in CNC modification and scalability,13 many challenges remain which prohibit realization of their large scale production and commercial use. Acid hydrolysis of cellulose is usually performed with strong mineral acids (e.g. sulfuric and hydrochloric acid) and the resulting CNCs are inherently hydrophilic due to the large number of hydroxyl groups present on CNC surfaces. Drying unmodified CNCs results in irreversible agglomeration due to strong hydrogen bonds.14 Cellulose hydrolysis with sulfuric acid results in functionalization of CNCs with sulfate ester moieties that form stable colloidal solutions.15 However, these sulfate

ACS Paragon Plus Environment

3

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

bonds have poor chemical stability and are cleaved under mild alkaline conditions or at high temperatures.2 The modification of CNC surfaces has been extensively explored. The following are examples of modification methods currently used, all of which suffer from one or more of the following problems: i) the need for multi-step processes, ii) use of toxic reagents, iii) use of large solvent volumes and iv) reaction conditions that lead to low modified CNC volumetric productivity. For example, many papers report passing aqueous suspensions of CNCs through a series of solvent exchanges so they can reside in a non-polar solvent to accommodate reactions that require nonaqueous conditions.13,16 Silylation of CNC surfaces has been performed in citrate buffer to introduce amino and methacrylate groups onto CNC surfaces.17 Oxidation of C6 hydroxyl groups at CNC surfaces to carboxylic acids with 2,2,6,6-tetramethylpiperidine-1-oxyl (TEMPO) is conducted in low solid loading suspensions of unmodified CNCs and requires a primary oxidant such as sodium hypochlorite.18 Carboxylic acid groups obtained by TEMPO oxidation have been used for subsequent modifications such as grafting of amine terminated polymers (e.g. PEG).19 Furthermore, CNC surfaces have been decorated with double and triple bonds, ATRP initiators and mercapto groups by two step processes such as CNC synthesis by acid hydrolysis followed by Fischer esterification.20,21 Our research team reported that cellulose can be hydrolyzed and modified in a single step by a one-pot procedure that combines concurrent cellulose acid hydrolysis and organic acid-catalyzed Fischer Esterification resulting in surface-modified CNCs.22,23 For example, surface modification of purified cotton cellulose was performed using a catalytic quantity of HCl and an organic acid (e.g. acetic-, butyric- or lactic acid) that both serves as the reaction solvent and reagent for esterification reactions.23,24 The organic acid used can be selected to introduce the desired

ACS Paragon Plus Environment

4

Page 5 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

functionality, hydrophilicity or hydrophocity to the modified CNCs. For instance, CNCs were modified with lactate groups that facilitated the dispersion of corresponding lactate-CNCs within a polylactide matrix by direct melt compounding.23 Despite the different organic acids and functional groups that have been used in the concurrent cellulose acid hydrolysis and Fischer esterification process, the relationship of organic acid pKa on cellulose hydrolysis has been largely neglected. Modifying CNCs with the acetic acid derivatives (i.e. phenylacetic acid and 2-bromopropanoic acid) with pKa values of 4.31 and 4.0, respectively, required a two-step process due to high organic acid pKa values.25 Additionally, it was shown that CNC acid hydrolysis conducted in the presence of only high pKa organic acids required up to 10 days at 100 °C.20,26 In fact, the majority of concurrent cellulose acid hydrolysis and Fischer esterification reactions have been conducted with organic acids with pKa values ≥ 4.0, thus necessitating the addition of an HCl catalyst or a two-pot process.23,24 It is well known that total proton concentration is critical for cellulose hydrolysis.27 Strong mineral acids (e.g. sulfuric and hydrochloric acid) dissociate completely in water to H+ and A-. Consequently, these acids have been the focus of most work on acid-catalyzed cellulose hydrolysis. For example, acid hydrolysis of cellulose using sulfuric acid at 70 °C was studied by Zhu et al. who showed that acid concentration was the most critical factor for CNC formation. Furthermore, a H2SO4 concentration of 58% resulted in CNC’s in superior yield with rod-like morphologies and high crystallinities.28–30 A 20 wt% solution of formic acid has similar cellulose hydrolysis reactivity to 0.5 wt% sulfuric acid at 180°C under elevated pressure.31 Hou et al showed that formic acid can be used for the one step acid hydrolysis of cellulose/Fischer esterification to produce formate surface modified CNC’s.32 However, a high proton concentration consisting of 90% formic acid and 3 M HCl was used at 90°C for 3 h to prepare

ACS Paragon Plus Environment

5

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 6 of 39

formate modified CNC’s and the relative contributions of formic acid and HCl to the reaction was not investigated. Organic dicarboxylic acids have been used to hydrolyze cellulose at high temperatures (>200°C) under pressure.27,33 Oxalic acid (pKa = 1.25) and maleic acid (pKa = 1.9) showed little cellulose degradation after reactions conducted with α-cellulose performed at 105°C for 3 h.33 Most of the above mentioned studies were performed with the goal of producing glucose and other biofuels and not producing modified CNCs and thus the effect of dicarboxylic acid on the overall morphology and crystallinity is unknown. To date, no work has addressed the opportunity to modify cellulose nanocrystals via concurrent acid hydrolysis/Fischer esterification with diacids to introduce carboxcylic acid functionality on CNC surfaces. In the present work, green, solvent-less, one-pot concurrent acid hydrolysis/Fischer esterification reactions were investigated using biobased di- and trifunctional organic acids (citric, malonic or malic). The success of these reactions would provide a simple route to decorate CNC surfaces with acid functionalities. A key focus of this study was to determine the effects of reaction variables (e.g. HCl concentration and organic acid pKa) on modified CNC yield, CNC morphology and surface substitution degree. We hypothesized that lower pKa diand trifunctional organic acids would possess an enhanced ability to hydrolyze cellulose and, consequently, would require less of the HCl co-catalyst for concurrent acid hydrolysis/Fischer esterification reactions. Furthermore, the influence of HCl co-catalyst concentration on the corresponding modified CNC yield and crystallinity was investigated. A co-product of the reaction which is similar to micro-crystalline cellulose was recovered and re-used for subsequent reaction cycles to increase cellulose conversion to modified CNCs. The substitution degree of modified CNCs (malonate-, malate- and citrate-CNCs) was characterized by FTIR and solid state

ACS Paragon Plus Environment

6

Page 7 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

13

C-NMR experiments. Furthermore, nanocomposites of polyvinyl alcohol (PVOH) and citrate-

modified CNCs were prepared and improvements in PVOH’s thermal properties were determined. EXPERIMENTAL AND MATERIALS Materials: Ramie Cellulose was purchased from the Woolery (http://www.woolery.com/) and purified according to previous reports.34 Citric acid (ACS reagent, >99.5%), malic acid (>99%), malonic acid (>99%), polyvinyl alcohol (PVOH, Mw = 90 kg/mol, >99% hydrolyzed) and HCl (ACS reagent grade, 37% HCl) were purchased from Sigma Aldrich and used as received. Preparation of HCl-CNCs, citrate, malonate or malate cellulose nanocrystals The method for preparation of HCl-CNCs from ramie cellulose described below followed a literature method used to prepare the CNCs from cotton linters.24 Ramie fibers were purified by two successive washing cycles with 4% NaOH solutions that resulted in removal of residual lignin and hemi-cellulose that equaled 3% of the initial ramie fiber weight. In all preparations of HCl and organic acid modified CNCs, the concentration of ramie was fixed at 4 wt%. The organic acids were dissolved in water by heating (between 80-90°C) to prepare 80 wt-% aqueous solutions of citric, malonic and malic acids. To prepare HCl CNCs, 13 g of ramie cellulose pieces (2cm x 2cm) were mixed with 310 g of 0.4 M HCl and soaked overnight at room temperature (20 °C). For organic acid modified CNCs and HCl CNCs, ramie pieces (2 cm x 2 cm) were soaked overnight at room temperature (20 °C) in 80 wt-% concentrated aqueous solutions of citric-, malonic- and malic acid. The reaction mixtures (HCl, organic acid and ramie pieces) were transferred to 1 L round bottom flasks which were maintained with magnetic stirring under reflux (oil bath temperature 140 °C) for 3 h. Organic acid solutions refluxed at

ACS Paragon Plus Environment

7

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 8 of 39

120 °C, whereas HCl solutions refluxed at 100 °C. Subsequently, reaction suspensions were mechanically agitated in a Waring laboratory blender for 5 minutes. Then, the reaction mixture was diluted one-fold by addition of DI water and the modified and HCl CNCs were isolated by repeated centrifugation (15 cycles) at 14,000 RPM for 8-10 min with replacement of the initially clear supernatant with an equal volume of DI water. Once the pH reached about 5, the supernatant remained turbid after centrifugation, indicating the presence of CNCs. After collection of the first turbid supernatant, the pellet was re-suspended with a shear homogenizer at 25,000 rpm for 1 minute and again centrifuged to obtain additional CNCs in the form of a turbid supernatant. The ten turbid supernatants (i.e. CNC aqueous suspensions) were recovered, freeze dried and analyzed. CNC yields were obtained gravimetrically taking the initial weight as that of purified ramie and the final weight as the product on a cellulose basis (i.e. not taking into account the increase in molecular weight due to modification) by determining the recovered CNC amount and subtracting the weight due to esterification by the organic acid (from degree of substitution values).

Characterization Thermal Gravimetric Analysis (TGA): The thermal stability of CNCs and PVOH nanocomposites were analyzed using a Thermal Analysis Q50 with an alumina pan. All samples were about 10 mg. CNC powder was compacted at the bottom of alumina pans. For PVOH/CNC nanocomposites, small pieces of film (c.a. 2 µm in thickness) were cut and weighed (c.a. 10 mg). For all experiments, heating was carried out at 20 °C/min under nitrogen or air from room temperature to 800 °C.

ACS Paragon Plus Environment

8

Page 9 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

Fourier Transform Infrared Spectroscopy (FTIR): Attenuated total reflectance FTIR were performed using a Perkin Elmer Spectrum One with a ZnSe Plate 45° HATR sampling accessory. Spectra were recorded from 450 to 4000 cm-1. 13

C Cross-Polarization, Magic-Angle Spinning, Solid-State Nuclear Magnetic Resonance

(13C CP MAS ssNMR):

13

C CP MAS ssNMR experiments were performed following R.E.

Taylor’s method.35,36 Spectra were obtained using a 600 MHz 89 mm wide-bore Bruker Advance III spectrometer equipped with a 4 mm HXY solid-state MAS probe configured with channels for 1H and

13

C. Prior to addition of pre-dried samples, polytetrafluoroethylene ("10 mg”) and a

teflon plug was added to the bottom of the rotor to position samples. The rotor was weighed before and after sample addition. Experimental parameters for 13C CP MAS ssNMR experiments were as follows: 7000 scans, spinning rate of 11.3 KHz, acquisition time 0.02 s, and temperature of 278 k. CNC crystallinity indexes were calculated by dividing the area under the C4 crystalline portion by the total area of C4 resonances from residual amorphous and crystalline domains.37 Transmission Electron Microscopy (TEM): A Zeiss Libra 200EF transmission electron microscope (TEM) was used to take bright field images of the CNCs. We used 400 mesh copper grids with carbon (3-4 nm) and formvar (25-50 nm) thin film coatings purchased from Electron Microscopy Sciences. TEM grids were prepared by placing them in a Novascan PSDP-UV8T UV-ozone system for 30 minutes at room temperature. A suspension of CNCs in deionized water was prepared at a concentration of 0.4 mg/mL. The suspension was sonicated in a bath for two hours at room temperature and then set aside for another hour where any remaining aggregates were allowed to form a precipitate at the bottom of the vial. A drop of the aqueous suspension of CNCs and two individual drops of 2% solutions of uranyl acetate in deionized water were placed on Parafilm. The UV-ozone treated grids were

ACS Paragon Plus Environment

9

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 10 of 39

floated on the drop with suspended CNCs for 60 seconds. The grid was removed from the drop and dried by placing filter paper at the edge of the grid. The grid was washed by floating it on the first drop of uranyl acetate solution for 5 seconds followed by a drying step with filter paper. Then, the grid was stained by floating it on the second drop of uranyl acetate solution for 60 seconds and dried with filter paper. The grids were then placed in the TEM for imaging with a CCD camera. Scanning Electron Microscopy (SEM): Experiments were performed on modified cellulose with 3 nm platinum sputter coating using a Versa3D SEM operating at 20 kV. Wide Angle X-Ray Spectroscopy: Wide angle X-Ray Spectroscopy (WAXS) patterns were obtained using a Panalytical X’Pert Pro X-Ray diffractometer with Cu Kα radiation (0.154 nm) at 45 kV and 40 mA. The scan rate was 10°/min. The diffraction angle was measured from 5° to 60°. The crystallinity index (IC) was calculated according to the equation IC = 1 – (I1/I2) where I1 is the intensity at 2θ = 18.8° and I2 is the intensity when 2θ = 22.8°.38 Conductometric titration: Modified CNCs were dispersed in DI water at 2 mg/mL at room temperature for 30 minutes using a tip sonicator fitted with a QSonica Q125 tip probe. Subsequently, the pH was adjusted to 10 with 0.1 M NaOH. The resulting solution was titrated with 0.1 M HCl and the change in voltage was monitored using a handheld conductivity meter (VWR product # 89094-958). Average degree of substitution values were calculated according to ATSM D1439.39 This method involves dispersing modified CNCs in a 0.1 M NaOH solution and titrating with and 0.1 M HCl solution. Degrees of substitution were calculated using the following equation. G=0.162A/(1-XA) where A=(BC-DE)/F: with B = mL of NAOH solution added, C = E = 0.1 (Molarities of HCl and NaOH), F = amount of modified CNCs used in g, 162 = the molar mass of the AGU unit and X = increase in molecular mass for the modified CNC (i.e.

ACS Paragon Plus Environment

10

Page 11 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

X = 58 for malonate- , 118 for malate- and 176 for citrate modified CNCs. Values obtained for citrate CNCs were divided by two, due to the fact that each citrate unit introduces two free carboxylic acids. Polyvinyl Alcohol/CNC Preparations: PVOH (Mw = 90 kg/mol, >99% hydrolyzed) was dissolved in DI water (10% by wt. PVOH) by stirring at 90°C for 1 hour. CNCs (2 mg/mL) were dispersed in the PVOH DI water solution by sonication performed for 20 minutes at room temperature or until the solution was transparent. Sonication was carried out using a sonicator fitted with a QSonica Q125 tip probe. Nanocomposites were then prepared by solution casting 10-wt-% solutions of PVOH containing 1wt% suspended CNCs. Samples were dried under ambient conditions for 48 hours before being transferred to a vacuum oven for water removal at 60°C and 10-1 mbar for one week Karl-Fischer Titration of Polyvinyl Alcohol Films: Karl-Fischer titration experiments were performed using a Denver instrument (model 725) Karl-Fisher titrator. All polyvinyl alcohol films were dried in a vacuum oven ( malate > citrate CNCs. Corresponding values for T50% are 366°C, 365°C, 350°C and 345°C, respectively. The amount of residual char at 600°C for HCl, malonate, malate and citrate CNCs is 8, 13, 16 and 20%, respectively. Hence, it follows that increasing the T50% for di- and tri-acid modified CNCs results in correspondingly lower char formation.

Increased char amounts is likely due to CNC

functionalizations that lead to relatively larger number of cross-linking events at elevated temperatures. Since the exclusion of air during melt processing is generally not practical, the effect of CNC modification on thermal stability in air was also determined and the corresponding TGA thermograms are displayed in Figure S3. Values for T10% and T50% are listed in the supporting information (see Table S8). The thermal stability of modified CNCs, determined from the peak of the DTG (Figure S3, right panel), are as follow: HCl CNCs ≈ Malonate CNCs > Malate CNCs ≈ Citrate CNCs. Corresponding values of T50% are 335, 334, 323 and 324 °C, respectively. The general trend of thermal stabilities as a function of CNC modification chemistry is almost identical for analyses conducted under N2 (g) and air. Values of T10% in air for HCl, malonate, malate and citrate modified CNCs are 311, 273, 305 and 267 °C. Higher T10% values for malate modified CNCs may be due to the presence of malate oligomers attached to CNC surfaces.70

ACS Paragon Plus Environment

23

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 24 of 39

Similarly, the potential formation of lactate oligomers for lactate-modified CNCs is believed to increase their thermal stability relative to modification by acetic acid.23 The T10% values under N2 (g) relative to air differ by +6, +53, +21 and -49 oC for HCl, malonate, citrate and maloante CNCs, respectively. Unlike TGA performed under nitrogen, cellulose tends to combust at high temperatures.71 Consequently, under air, there is little char remaining after heating to 600 °C.72 From the above, all modified CNCs prepared herein have T10% and T50% values under air and nitrogen that are greater than 256 °C and 323 °C, respectively. Hence, the relatively high thermal stability of modified CNCs indicates they can be used for melt-processing operations performed at moderately high temperatures without thermal decomposition to produce polymer CNC nanocomposites. Repeated Hydrolysis cycles on Recovered Cellulose Reactions performed to prepare modified CNCs also resulted in a non-hydrolyzed cellulose co-product that was separated by centrifugation from modified CNCs suspended in the supernatant. Based on SEM and WAXS analysis (see Figures S4 and S5), recovered cellulose is comprised of crystalline particles (crystallinity index about 70%) that are similar to MCC. Furthermore, it is well known that CNCs can be isolated from MCC using controlled acid hydrolysis.73 This led us to explore whether we could convert residual cellulose to modified CNCs in order to improve the overall yield of citrate CNCs. Hydrolysis reactions on recovered cellulose were conducted using 80% citric acid solutions with 0.05 M HCl for 1 hour and the results from three hydrolysis cycles are listed in Table 1. The reaction time of 1 hour was selected since longer times (e.g. 3 h) had deleterious effects on CNC crystallinity without leading to improved yields. For example, WAXS showed that 2 hour reactions gave lower crystallinity

ACS Paragon Plus Environment

24

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

(by about 20%) CNCs. Hence, hydrolysis on recovered cellulose from the first and second reaction cycles led to yields of 25% and 10%, respectively. Consequently, the overall yield obtained over 3 reaction cycles is 55%. However, this increase in overall yield comes at the cost of expending additional reagents and energy-input. Future studies are planned to circumvent the Table 1: Yield of HCl-CNCs and Acid modified CNCs a,b Organic acid

Hydrolysis cycle

CNC yield (%)b

Yield (%) non- Yield (%) of soluble solubilized byproductc co-productsd

Citric Acid

First

20 ± 5.2

72 ± 2

7 ± 3.7

Citric Acid

Second

25 ± 3.0

68.7 ± 2.0

6.3 ± 3.0

Citric Acid

Third

10 ± 6.1

83 ± 5.1

6.7 ± 23.2

a) Values reported are the mean from three independent experiments and the corresponding standard deviation. b) Determined gravimetrically from isolated modified CNCs after freeze drying. c) Determined gravimetrically from isolated non-solubilized material. d) Presumed to consist of glucose and corresponding oligomers. The %-yield is determined from the following equation: (100-[% yields of CNC + non-solubilized cellulose]).

use of additional reagents by recovering and re-using those from the previous reaction cycle. Figure 7 shows TEM images of CNCs recovered from three reaction cycles. For each cycle, citrate-CNCs have nearly identical average lengths and widths (approximately 240 nm by 10 nm). Consequently, there is no significant difference in the overall morphology of CNCs recovered from the first, second and third hydrolysis cycles. Analysis of the crystallinity index from WAXS performed on citrate CNCs isolated from the first, second and third hydrolysis cycles (Figure S6 and table S9) give crystallinity index values of 80%, 76% to 75%, respectively. Furthermore,

ACS Paragon Plus Environment

25

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 26 of 39

FTIR and CP-MAS NMR experiments performed on citrate CNCs from each reaction cycle show that each has similar degrees of modification.

Figure 7: TEM of citrate modified CNCs from re-hydrolysis experiments: (a) first hydrolysis, (b) second hydrolysis and (3) third hydrolysis. For the first hydrolysis, the reaction was conducted for 3 hours at 140°C with 0.05 M HCl and 80% organic acid. The second and third reaction cycles were performed under identical conditions as the first except that the reaction time was reduced to one hour. Scale bar = 100 nm Polyvinyl Alcohol/Acid Modified CNC Nanocomposites Malonate-, malate- and citrate CNCs after sonication form stable aqueous colloidal suspensions (2 mg/mL) with no sedimentation after 7 days. In contrast, HCl CNCs precipitate over hours (data not shown). Increased colloidal stability is attributed to repulsion of negatively charged carboxylic acid functionalities present on CNC surfaces. One potential application of stable aqueous colloidal CNC solutions is their incorporation in water-soluble polymers to improve their physico-mechanical properties. Polyvinyl alcohol (PVOH) is an industrially relevant water soluble polymer. Despite PVOH’s numerous advantages, it’s low thermal stability results in degradation prior to melting. This hinders PVOH melt processing and applications at

ACS Paragon Plus Environment

26

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

high temperatures.74 One potential route to enable PVOH melt processing is by the addition of plasticizers such as polyols75 and urea.76 While these plastizers effectively depress the melting point, they can leach from the processed material over time causing changes in material properties and problems for use in certain applications such as for food-contact. Thus, improving PVOH heat stability will allow PVOH to be used in melt blending applications while potentially reducing plasticizer amount. In an attempt to circumvent the low thermal stability of PVOH, PVOH/CNC nanocomposites were prepared by casting films from aqueous solutions containing 10% by wt. PVOH solution and 1% by-wt. with respect to PVOH of HCl-, malonate-, malateand, citrate CNCs. Subsequently, films were first dried under ambient conditions and then in vacuo (60°C, 10-1 mbar, one week). Karl-Fischer titrations were performed to ascertain the effect of modified CNCs on the amount of residual water present in films.77 PVOH films both with and without modified CNC after drying have less than 0.1% residual water. Thermal stability of PVOH nanocomposites was studied by TGA under N2 (g) (Figure 8). Values of T10% and T50% listed in Table S10 show that PVOH thermal stability is dependent on the di- or triacid that was used for CNC modification. Neat PVOH has T10% and T50% values of 288°C and 346°C, respectively. In comparison, T10% and T50% values for PVOH blended with 1%-by-wt citrate, malate, HCl, and malonate CNCs are 306, 299, 289 and 283 oC, respectively, and 385, 360, 353, and 345 oC, respectively. Hence, the largest improvement in thermal stability (18 and 39 oC for T10% and T50%, respectively) was achieved by incorporating citrate CNCs in PVOH. PVOH degradation is known to occur by a multi-step process that includes elimination and dehydration reactions. Subsequent degradation phenomena occur by radical and cyclization reactions.74 Study of DTG plots (Figure 8, right panel) reveal a shift in the degradation steps for PVOH by incorporation of 1% citrate CNCs.

ACS Paragon Plus Environment

27

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 28 of 39

Figure 8: Thermogravimetric analysis of Polyvinyl alcohol/CNC nanocomposites under N2

Figure 9: DSC of polyvinyl alcohol/CNC (PVOH/CNC) nanocomposites with 1% modified CNCs under N2 flow (second scan) at a heating rate of 10 °C/min from 0 to 250 °C. We hypothesize that the additional carboxyl group of CNC citrate ester moieties and the resulting enhanced hydrogen bonding with PVOH hydroxyl groups is responsible for improving

ACS Paragon Plus Environment

28

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

PVOH thermal stability. Also, incorporation of rigid CNCs in PVOH can restrict diffusion in the matrix which can further contribute to increasing PVOH thermal stability. Differential Scanning Calorimetry (DSC) was performed on as-cast dried PVOH/CNC nanocomposites and the DSC thermograms are shown in Figure 9. Values of the peak melting transition (Tm), enthalpy of melting (∆H1, J/g) and %-crystallinity are listed in Table S11. The presence of modified and un-modified CNCs has little effect on PVOH Tm (variation of up to 4 °C). The fact that the Tm did not decrease by addition of the modified CNCs indicates they do not plasticize PVOH.

By incorporating 1% malate and citrate modified CNCs into PVOH,

calculated %-crystallinity from experimentally determined ∆H1 values increased to 45 and 48%, respectively, relative to 34% for neat PVOH.40 Hence, it appears that malate and citrate-CNCs act as nucleating agents for PVOH. Similar effects have been observed by Jorfi et al. who showed that sulfonated CNCs act as nucleating agents for PVOH, albeit with slight higher concentrations of CNCs (4 % loading w/w).78 However, unlike Jorfi et al. who showed a significant increase in PVOHs Tg, little difference was observed in nanocomposites with modified CNCs in this study. These results show that fine-tuning the surface chemistry of modified CNCs can have a profound effect on crystallization phenomena of corresponding nanocomposites. Conclusion CNCs were modified with pendant malonate-, malate- and citrate groups by a one-pot, solventless, concurrent acid catalyzed cellulose hydrolysis and Fischer esterification reaction. Modified CNCs were compared to CNCs produced by HCl hydrolysis (i.e. produced in the absence of organic acid). We hypothesized lower pKa acids (i.e. malic vs. citric acid) would require different amounts of HCl co-catalyst for optimal CNC yield. However, little effect of acid pKa

ACS Paragon Plus Environment

29

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 30 of 39

was found on the overall modified CNC yield, as malic-, malonic and, citric acids all had optimal yields between 0.025 and 0.1M HCl. All modified CNCs were shown to be highly crystalline with little effect on overall CNC morphology. That is, the TEM of modified and unmodified CNCs had similar lengths, widths and shape factors. WAXS studies showed that modification had little effect on CNC crystallinity, as unmodified and modified CNCs had similar crystallinity indexes. Thus, modifying CNCs with bio-based poly organic acids proved to be an efficient way to introduce carboxylic acid functionality unto CNC surfaces. This provides strong evidence that CNC modifications primarily occur on CNC surfaces. Structural analysis by FTIR and

13

C solid state CP-MAS NMR studies is consistent with the

expected structures of modified CNCs. Quantitative direct detection carbon NMR experiments gave average degree of substitution (d.s.) values of malonate, citrate and malate CNCs of 0.16, 0.18 and 0.22, respectively. Conductometric titration showed that the acid value of modified CNCs can be tuned, with acid values ranging from 1108±94, 1617±170 and, 1884±1224 mmol COOH/mg cellulose for malate-, malonate- and, citrate CNCs, respectively. Yields of citrate modified CNCs were increased to 55% for citrate CNCs by reusing the residual cellulose byproduct. CNCs isolated from re-hydrolysis cycles have similar degrees of modification and crystallinity as that obtained from the first cycle. Analysis of modified CNC thermal stability under air and nitrogen showed that T10% and T50% values are above 256 °C and 323 °C, respectively. These results demonstrate that the modified CNC’s prepared herein can be used for melt-processing operations performed at moderately high temperatures without thermal decomposition. One potential application of CNCs modified with free acids is the fabrication of nanocomposites with water-soluble polymers. For this purpose, nanocomposites were prepared

ACS Paragon Plus Environment

30

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

by solution casting from water PVOH with 1% CNC loading. With 1% citrate and malate CNCs, the T50% of PVOH increased by 14 and 49 °C relative to neat PVOH films. Incorporation of modified CNCs in PVOH films lead to increases in PVOH crystallinity 34% for neat PVOH to 45% and 48% for malate- and citrate-CNC nanocomposites. While we have not quantitatively assessed the extent that the 12 principles of Green Chemistry79 were followed by the one-pot CNC esterification with organic acids bearing two or three acid moieties, the following provides a qualitative discussion. The CNCs prepared herein use readily renewable feedstocks, toxic reagents are not used or generated, circumvents the need for solvent exchanges which has often been used to pass aqueous CNC suspensions from water to non-polar solvents for ester bond forming reactions (see the introduction for examples and references). Furthermore, the modified CNCs are expected to be fully degradable when disposed in a bioactive environment. While we have improved the conversion of cellulose to CNC’s by recycling of recovered cellulose, this approach increases energy and chemical utilization. To improve the sustainability of this multicycle process it is critically important that future work address the development of a practical method for recovery and re-use of the large excess of organic acid that is not used in the first reaction cycle. Finally, future studies are underway with a series of water-soluble bio-sourced polymers to gain insight into effects of modified CNC structure on nanocomposite mechanical properties. ASSOCIATED CONTENT Supporting Information. Figures supplied in the supporting information include quantitative solid state NMR spectra, conductometric titration, TGA thermograms for neat CNCs and polyvinyl alcohol/CNC nanocomposites, SEM of un-hydrolyzed cellulose byproducts and WAXS of CNCs from residual cellulose particles. Tables include literature values for organic

ACS Paragon Plus Environment

31

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 32 of 39

acid pKa values, tabulated modified CNC yields, CNC crystallinity indexes, CNC acid values from titration, CNC thermal stability under nitrogen and air.

Corresponding author * e-mail: [email protected] Acknowledgments The authors are grateful for funding received from the National Science Foundation Partnerships for International Research and Education (PIRE) Program (Award #1243313). References: (1)

Eichhorn, S. J. Cellulose Nanowhiskers: Promising Materials for Advanced Applications Soft Matter 2011, 7, 303.

(2)

Habibi, Y.; Lucia, L. A.; Rojas, O. J. Cellulose Nanocrystals: Chemistry, Self-Assembly, and Applications. Chem. Rev. 2010, 110, 3479–3500.

(3)

Iwamoto, S.; Kai, W.; Isogai, A.; Iwata, T. Elastic Modulus of Single Cellulose Microfibrils from Tunicate Measured by Atomic Force Microscopy Biomacromolecules 2009, 10, 2571–2576.

(4)

Moon, R. J.; Martini, A.; Nairn, J.; Simonsen, J.; Youngblood, J. Cellulose nanomaterials review: structure, properties and nanocomposites.; 2011; Vol. 40.

(5)

De Souza Lima, M. M.; Wong, J. T.; Paillet, M.; Borsali, R.; Pecora, R. Translational and Rotational Dynamics of Rodlike Cellulose Whiskers Langmuir 2003, 19, 24–29.

(6)

Anglès, M. N.; Dufresne, A. Plasticized Starch/Tunicin Whiskers Nanocomposites. 1. Structural Analysis Macromolecules 2000, 33, 8344–8353.

(7)

Yang, J.; Han, C. R.; Duan, J. F.; Xu, F.; Sun, R. C. Mechanical and Viscoelastic Properties of Cellulose Nanocrystals Reinforced Poly(ethylene glycol) Nanocomposite Hydrogels ACS Appl. Mater. Interfaces 2013, 5, 3199–3207.

(8)

Lin, N.; Huang, J.; Dufresne, A. Preparation, Properties and Applications of Polysaccharide Nanocrystals in Advanced Functional Nanomaterials: a Review Nanoscale 2012, 4, 3274.

ACS Paragon Plus Environment

32

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

(9)

Domingues, R. M. A; Gomes, M. E.; Reis, R. L. The Potential of Cellulose Nanocrystals in Tissue Engineering Strategies Biomacromolecules 2014, 15, 2327–2346.

(10)

Lin, N.; Dufresne, A. Nanocellulose in Biomedicine: Current status and Future Prospect Eur. Polym. J. 2014, 59, 302–325.

(11)

Cao, S.; Li, X.-H.; Lou, W.-Y.; Zong, M.-H. Preparation of Novel Magnetic Cellulose Nanocrystal and Its Efficient Use for Enzyme Immobilization J. Mater. Chem. B 2014, 2, 5522–5530.

(12)

Zhou, Y.; Khan, T. M.; Liu, J. C.; Fuentes-Hernandez, C.; Shim, J. W.; Najafabadi, E.; Youngblood, J. P.; Moon, R. J.; Kippelen, B. Efficient Recyclable Organic Solar Cells on Cellulose Nanocrystal Substrates With a Conducting Polymer Top Electrode Deposited by Film-Transfer Lamination Org. Electron. physics, Mater. Appl. 2014, 15, 661–666.

(13)

Habibi, Y. Key Advances in the Chemical Modification of Nanocelluloses. Chem. Soc. Rev. 2014, 43, 1519–1542.

(14)

Dufresne, A. Processing of Polymer Nanocomposites Reinforced with Polysaccharide Nanocrystals Molecules 2010, 15, 4111–4128.

(15)

Dong, X. M.; Gray, D. G. Effect of Counterions on Ordered Phase Formation in Suspensions of Charged Rodlike Cellulose Crystallites Langmuir 1997, 13, 2404–2409.

(16)

Eyley, S.; Thielemans, W. Surface Modification of Cellulose Nanocrystals Nanoscale 2014, 6, 7764–7779.

(17)

Raquez, J.-M.; Murena, Y.; Goffin, A.-L.; Habibi, Y.; Ruelle, B.; DeBuyl, F.; Dubois, P. Surface-Modification of Cellulose Nanowhiskers and Their Use as Nanoreinforcers into Polylactide: A Sustainably-Integrated Approach Compos. Sci. Technol. 2012, 72, 544– 549.

(18)

Habibi, Y.; Chanzy, H.; Vignon, M. R. TEMPO-Mediated Surface Oxidation of Cellulose Whiskers Cellulose 2006, 13, 679–687.

(19)

Cha, R.; Wang, C.; Cheng, S.; He, Z.; Jiang, X. Using Carboxylated Nanocrystalline Cellulose as an Additive in Cellulosic Paper and Poly (vinyl alcohol) Fiber Paper Carbohydr. Polym. 2014, 110, 298–300.

(20)

Boujemaoui, A.; Mongkhontreerat, S.; Malmström, E.; Carlmark, A. Preparation and Characterization of Functionalized Cellulose Nanocrystals Carbohydr. Polym. 2015, 115, 457–464.

(21)

Espino-Pérez, E.; Domenek, S.; Belgacem, N.; Bras, J. Green Process for Chemical Functionalization of Nanocellulose with Carboxylic Acids Biomacromolecules 2014, 15, 4551–4560.

ACS Paragon Plus Environment

33

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 34 of 39

(22)

Sobkowicz, M. J.; Braun, B.; Dorgan, J. R. Decorating in Green: Surface Esterification of Carbon and Cellulosic Nanoparticles Green Chem. 2009, 11, 680.

(23)

Spinella, S.; Lo Re, G.; Liu, B.; Dorgan, J.; Habibi, Y.; Leclère, P.; Raquez, J.-M.; Dubois, P.; Gross, R. A. Polylactide/Cellulose Nanocrystal Nanocomposites: Efficient Routes for Nanofiber Modification and Effects of Nanofiber Chemistry on PLA Reinforcement Polymer, 2015, 65, 9–17.

(24)

Braun, B.; Dorgan, J. R. Supra-Molecular Ecobionanocomposites Based on Polylactide and Cellulosic Nanowhiskers: Synthesis and Properties. Biomacromolecules 2009, 10, 334–341.

(25)

Dippy, J. F. J.; Hughes, S. R. C.; Rozanski, A. Disubstituted Succinic Acids J. Chem. Soc. 1959.

(26)

Espino-Pérez, E.; Bras, J.; Ducruet, V.; Guinault, A.; Dufresne, A.; Domenek, S. Influence of Chemical Surface Modification of Cellulose Nanowhiskers on Thermal, Mechanical, and Barrier Properties of Poly(lactide) Based Bionanocomposites Eur. Polym. J. 2013, 49, 3144–3154.

(27)

Kupiainen, L.; Ahola, J.; Tanskanen, J. Kinetics of Formic Acid-catalyzed Cellulose Hydrolysis BioResources 2014, 9, 2645–2658.

(28)

Chen, L.; Wang, Q.; Hirth, K.; Baez, C.; Agarwal, U. P.; Zhu, J. Y. Tailoring the Yield and Characteristics of Wood Cellulose Nanocrystals (CNC) Using Concentrated Acid Hydrolysis Cellulose 2015, 1753–1762.

(29)

Wang, Q. Q.; Zhu, J. Y.; Reiner, R. S.; Verrill, S. P.; Baxa, U.; McNeil, S. E. Approaching Zero Cellulose Loss in Cellulose Nanocrystal (CNC) Production: Recovery and Characterization of Cellulosic Solid Residues (CSR) and CNC Cellulose 2012, 19, 2033– 2047.

(30)

Wang, Q.; Zhao, X.; Zhu, J. Y. Kinetics of Strong Acid Hydrolysis of a Bleached Kraft Pulp for Producing Cellulose Nanocrystals (CNCs). Ind. Eng. Chem. Res. 2014, 53, 11007–11014.

(31)

Kupiainen, L.; Ahola, J.; Tanskanen, J. Comparison of Formic and Sulfuric acids as a Glucose Decomposition Catalyst Ind. Eng. Chem. Res. 2010, 49, 8444–8449.

(32)

Yu, H. Y.; Qin, Z. Y.; Sun, B.; Yan, C. F.; Yao, J. M. One-Pot Green Fabrication and Antibacterial Activity of Thermally Stable Corn-Like CNC/Ag Nanocomposites J. Nanoparticle Res. 2014, 16, 1–12.

(33)

vom Stein, T.; Grande, P.; Sibilla, F.; Commandeur, U.; Fischer, R.; Leitner, W.; Domínguez de María, P. Salt-Assisted Organic-Acid-Catalyzed Depolymerization of Cellulose Green Chem. 2010, 12, 1844.

ACS Paragon Plus Environment

34

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

(34)

Habibi, Y.; Goffin, A.-L.; Schiltz, N.; Duquesne, E.; Dubois, P.; Dufresne, A. Bionanocomposites Based on Poly(ε-caprolactone)-Grafted Cellulose Nanocrystals by Ring-Opening Polymerization J. Mater. Chem. 2008, 18, 5002.

(35)

Fu, L.; Mccallum, S. A.; Miao, J.; Hart, C.; Tudryn, G. J.; Zhang, F.; Linhardt, R. J. Rapid and accurate Determination of the Lignin Content of Lignocellulosic Biomass by SolidState NMR Fuel 2015, 141, 39–45.

(36)

Taylor, R. E. Setting Up 13C CP/MAS Experiments Concept Concepts Magn. Reson. Part A Bridg. Educ. Res. 2004, 22, 37–49.

(37)

Park, S.; Baker, J. O.; Himmel, M. E.; Parilla, P. a; Johnson, D. K. Cellulose Crystallinity Index: Measurement Techniques and Their Impact on Interpreting Cellulase Performance Biotechnol. Biofuels 2010, 3, 10.

(38)

Buschle-Diller, G.; Zeronian, S. H. J. Enhancing the Reactivity and Strength of Cotton Fibers Appl. Polym. Sci. 1992, 45, 967–979.

(39)

Feddersen, R. L.; Thorp, S. N. Standard Test Methods for Sodium Carboxymethylcellulose Ind. Gums 1993, 03, 537–578.

(40)

Ricciardi, R.; Auriemma, F.; Gaillet, C.; De Rosa, C.; Lauprêtre, F. Investigation of the Crystallinity of Freeze/Thaw Poly(Vinyl Alcohol) Hydrogels by Different Techniques Macromolecules 2004, 37, 9510–9516.

(41)

Hebeish, A.; Guthrie, J. T. The Chemistry and Technology of Cellulosic Copolymers; The Chemistry and Technology of Cellulosic Copolymers; 1981.

(42)

Yan, L.; Greenwood, A. a.; Hossain, A.; Yang, B. A Comprehensive Mechanistic Kinetic Model for Dilute Acid Hydrolysis of Switchgrass Cellulose to Glucose, 5-HMF and Levulinic Acid RSC Adv. 2014, 4, 23492.

(43)

Xiang, Q.; Kiim, J. S.; Lee, Y. Y. Comprhensive Kinetic Model for Dilute-Acid Hydrolysis of Cellulose Appl. Biochem. Biotechnol. 2003, 105, 337–352.

(44)

Zhbankov, R. G. Infrared Spectra of Cellulose and its Derivatives; 1st ed.; Consultants Bureau: New York, 1966.

(45)

Azzam, F.; Heux, L.; Putaux, J. L.; Jean, B. Preparation by Grafting Onto, Characterization, and Properties of Thermally Responsive Polymer-Decorated Cellulose Nanocrystals Biomacromolecules 2010, 11, 3652–3659.

(46)

Salajková, M.; Berglund, L. A.; Zhou, Q. Hydrophobic Cellulose Nanocrystals Modified With Quaternary Ammonium Salts J. Mater. Chem. 2012, 22, 19798.

(47)

Yang, C. Q.; Wang, X. The Formation of Five Membered Cyclic Anhydride Intermediates by Polycarboxylic Acids Studied by the Combination of Thermal Analysis and FT-IR Spectroscopy J. Appl. Polym. Sci. 1998, 70, 2711–2718.

ACS Paragon Plus Environment

35

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 36 of 39

(48)

Yang, C. Q.; Xu, Y.; Wang, D. FT-IR Spectroscopy Study of the Polycarboxylic Acids Used for Paper Wet Strength Improvement Ind. Eng. Chem. Res. 1996, 5885, 4037–4042.

(49)

Harifi, T.; Montazer, M. Past, Present and Future Prospects of Cotton Cross-Linking: New Insight into Nano Particles Carbohydr. Polym. 2012, 88, 1125–1140.

(50)

Kim, B. J.; White, J. L. J. Durable Press Finish of Cotton Fabric Using Malic Acid as a Crosslinker Appl. Polym. Sci. 2003, 88, 1429–1437.

(51)

Wu, W.; Huang, F.; Pan, S.; Mu, W.; Meng, X.; Yang, H.; Xu, Z.; Ragauskas, A. J.; Deng, Y. J. Thermo-Responsive and Fluorescent Cellulose Nanocrystals Grafted with Polymer Brushes Mater. Chem. A Mater. energy Sustain. 2015, 3, 1995–2005.

(52)

Gårdebjer, S.; Bergstrand, A.; Idström, A.; Börstell, C.; Naana, S.; Nordstierna, L.; Larsson, A. Solid-state NMR to Quantify Surface Coverage and Chain Length of Lactic Acid Modified Cellulose Nanocrystals, Used as Fillers in Biodegradable Composites Compos. Sci. Technol. 2015, 107, 1–9.

(53)

Kono, H.; Erata, T.; Takai, M. CP / MAS 13 C NMR Study of Cellulose and Cellulose Derivatives. 2. Complete Assignment of the 13 C Resonance for the Ring Carbons of Cellulose Triacetate Polymorphs J. Am. Chem. Soc. 2002, 124, 7506–7511.

(54)

Isogai, A.; Usuda, M. Solid-state CP/MAS 13C NMR Study of Cellulose Polymorphs Macromolecules 1989, 22, 3168–3172.

(55)

Roy, D.; Semsarilar, M.; Guthrie, J. T.; Perrier, S. Cellulose Modification by Polymer Grafting: a Review. Chem. Soc. Rev. 2009, 38, 2046–2064.

(56) Maunu, S.; Liitiä, T.; Kauliomäki, S.; Hortling, B.; Sundquist, J. 13C CPMAS NMR Investigations of Cellulose Polymorphs in Different Pulps Cellulose Cellulose 2000, 7, 147–159. (57)

Carlsson, D. O.; Lindh, J.; Strømme, M.; Mihranyan, A. Susceptibility of Iα - and Iβ Dominated Cellulose to TEMPO-Mediated Oxidation Biomacromolecules 2015, 1643– 1649.

(58)

Ramos, A.; Jordan, K. N.; Cogan, T. M.; Santos, H. Appl. Env. Microbiol. 601739-1748 1994, 60, 1739–1748.

(59)

Gout, E.; Bligny, R.; Pascal, N.; Douce, R. 13C Nuclear Magnetic Resonance Studies of Malate and Citrate Synthesis and Compartmentation in Higher Plant Cells J Biol Chem 1993, 268, 3986–3992.

(60)

Garg, B.; Bisht, T.; Ling, Y.-C. Sulfonated Graphene as Highly Efficient and Reusable Acid Carbocatalyst for the Synthesis of Ester Plasticizers RSC Adv. 2014, 4, 57297– 57307.

ACS Paragon Plus Environment

36

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

(61)

Borchert, R.; Everett, G. W. Plant Physiol. Nuclear Magnetic Resonance Study of Acetate Incorporation into Malate During Ca-uptake by Isolated Leaf Tissues 1987, 84, 944–949.

(62)

Niwayama, S.; Cho, H.; Lin, C. Highly Efficient Selective Monohydrolysis of Dialkyl Malonates and Their Derivatives Tetrahedron Lett. 2008, 49, 4434–4436.

(63)

Tang, J.-S.; Guo, C.-C. Palladium-Catalyzed Alcoholysis of 3-Iodopropynamides: Selective Synthesis of Carbamoylacetates Synthesis 2014, 47, 108–112.

(64)

Wei, Y.; Lee, D.-K.; Ramamoorthy, A. solid-state (13)C NMR Chemical Shift Anisotropy Tensors of Polypeptides J Am Chem Soc 2001, 123, 6118–6126.

(65)

Johnson, R. L.; Schmidt-Rohr, K. Quantitative Solid-State 13C NMR with Signal Enhancement by Multiple Cross Polarization J. Magn. Reson. 2014, 239, 44–49.

(66)

Ziarelli, F.; Caldarelli, S. Solid-State NMR as an Analytical Tool: Quantitative Aspects Solid State Nucl. Magn. Reson. 2006, 29, 214–218.

(67)

Fu, R.; Hu, J.; Cross, T. A. Towards Quantitative Measurements in Solid-State CPMAS NMR: A Lee-Goldburg Frequency Modulated Cross-Polarization Scheme. J. Magn. Reson. 2004, 168, 8–17.

(68)

Ureña-Benavides, E. E.; Kitchens, C. L. Wide-angle X-ray Diffraction of Cellulose Nanocrystal-Alginate Nanocomposite Fibers Macromolecules 2011, 44, 3478–3484.

(69)

Way, A. E.; Hsu, L.; Shanmuganathan, K.; Weder, C.; Rowan, S. J. PH-responsive Cellulose Nanocrystal Gels and Nanocomposite ACS Macro Lett. 2012, 1, 1001–1006.

(70)

Kajiyama, T.; Kobayashi, H.; Taguchi, T.; Kataoka, K.; Tanaka, J. Improved Synthesis with High Yield and Increased Molecular Weight of Poly(Alpha,Beta-Malic Acid) by Direct Polycondensation Biomacromolecules 2004, 5, 169–174.

(71)

Rachini, A.; Le Troedec, M.; Peyratout, C.; Smith, A. Effects of Expandable Graphite and Modified Ammonium Polyphosphate on the Flame-Retardant and Mechanical Properties of Wood Flour-Polypropylene Composites J. Appl. Polym. Sci. 2008, 112, 226–234.

(72)

Ramiah Mv. J. Appl. Polym. Sci. Thermogravimetric and Differential Thermal Analysis of Cellulose, Hemicellulose, and Lignin 1970, 14, 1323–1337.

(73)

Çetin, N. S.; Tingaut, P.; Özmen, N.; Henry, N.; Harper, D.; Dadmun, M.; Sèbe, G. Acetylation of Cellulose Nanowhiskers with Vinyl Acetate Under Moderate Conditions Macromol. Biosci. 2009, 9, 997–1003.

(74)

Gilmanl, J. W. Thermal Decomposition Chemistry of Poly (Vinyl Alcoh Fire Polym. II Mater. Test Hazard Prev. ACS 1995, 599, 161.

(75)

Wu, W.; Tian, H.; Xiang, A. Influence of Polyol Plasticizers on the Properties of Polyvinyl Alcohol Films Fabricated by Melt Processing J. Polym. Environ. 2012, 20, 63– 69.

ACS Paragon Plus Environment

37

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 38 of 39

(76)

Jang, J.; Lee, D. K. Plasticizer Effect on the Melting and Crystallization Behavior of Polyvinyl Alcohol Polymer, 2003, 44, 8139–8146.

(77)

Badgujar, K. C.; Dhake, K. P.; Bhanage, B. M. Immobilization of Candida Cylindracea Lipase on Poly(Lactic Acid), Polyvinyl Alcohol and Chitosan Based Ternary Blend Film: Characterization, Activity, Stability and its Application for N-a Process Biochem. 2013, 48, 1335–1347.

(78)

Jorfi, M.; Roberts, M. N.; Foster, E. J.; Weder, C. Non-isothermal Crystallisation Kinetics of Self-Assembled Polyvinyl Alcohol/Silica Nano-Composite ACS Appl. Mater. Interfaces 2013, 5, 1517–1526.

(79) Anastas, P. T.; Warner, J. C. Green Chemistry: Theory and Practice, Oxford University Press: New York, 1998, p.30.

ACS Paragon Plus Environment

38

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

For Table of Contents Use Only

Concurrent Cellulose Hydrolysis and Esterification to Prepare Surface-Modified Cellulose Nanocrystal Decorated with Carboxylic Acid Moieties Stephen Spinella, Anthony Maiorana, Qian Qian, Nathan J. Dawson, Victoria Hepworth, Scott A. McCallum, Manoj Ganesh, Kenneth D. Singer, and Richard A. Gross Modification of cellulose nanocrystals using bio-based organic acids was carried out to afford CNCs functionalized with carboxylic acids in a single step.

ACS Paragon Plus Environment

39