Water Diffusion in Amorphous Hydrophilic Systems: A Stop and Go

Sep 21, 2015 - A picture of H2O molecule motion comprising alternating steps of being bound at an adsorption site (“stop”) and moving (“go”) e...
0 downloads 9 Views 6MB Size
Subscriber access provided by Stockholm University Library

Article

Water diffusion in amorphous hydrophilic systems: a stop and go process Karol Kulasinski, Robert Guyer, Dominique Derome, and Jan Carmeliet Langmuir, Just Accepted Manuscript • DOI: 10.1021/acs.langmuir.5b03122 • Publication Date (Web): 21 Sep 2015 Downloaded from http://pubs.acs.org on September 22, 2015

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.

Langmuir 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 28

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

Langmuir

Figure 1. a) Snapshots of the modelled MD structures in dry state: (a) amorphous cellulose, (b) galactoglucomannan, a hemicellulose, (c) cellulose microfibril, composed of crystalline cellulose (grey) and hemicellulose (red). 76x92mm (300 x 300 DPI)

ACS Paragon Plus Environment

Langmuir

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

Figure 1. b) Snapshots of the modelled MD structures in dry state: (a) amorphous cellulose, (b) galactoglucomannan, a hemicellulose, (c) cellulose microfibril, composed of crystalline cellulose (grey) and hemicellulose (red). 238x289mm (96 x 96 DPI)

ACS Paragon Plus Environment

Page 2 of 28

Page 3 of 28

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

Langmuir

Figure 1. c) Snapshots of the modelled MD structures in dry state: (a) amorphous cellulose, (b) galactoglucomannan, a hemicellulose, (c) cellulose microfibril, composed of crystalline cellulose (grey) and hemicellulose (red). 264x289mm (96 x 96 DPI)

ACS Paragon Plus Environment

Langmuir

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

Figure 3. a) Output of the model: (a) microscopic diffusion coefficient and (b) tortuosity. 998x879mm (96 x 96 DPI)

ACS Paragon Plus Environment

Page 4 of 28

Page 5 of 28

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

Langmuir

Figure 3. b) Output of the model: (a) microscopic diffusion coefficient and (b) tortuosity. 998x879mm (96 x 96 DPI)

ACS Paragon Plus Environment

Langmuir

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

Figure 4. a) (a) Porosity and (b) free volume as related to the total volume, as a function of H2O content. 70x61mm (600 x 600 DPI)

ACS Paragon Plus Environment

Page 6 of 28

Page 7 of 28

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

Langmuir

Figure 4. b) (a) Porosity and (b) free volume as related to the total volume, as a function of H2O content. 70x61mm (600 x 600 DPI)

ACS Paragon Plus Environment

Langmuir

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

Figure 5. a) Model of a binding site: (a) Fourier transform of MSD during waiting periods, after background removal at s→0, (b) estimated binding energy at a binding site. 998x879mm (96 x 96 DPI)

ACS Paragon Plus Environment

Page 8 of 28

Page 9 of 28

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

Langmuir

Figure 5. b) Model of a binding site: (a) Fourier transform of MSD during waiting periods, after background removal at s→0, (b) estimated binding energy at a binding site. 998x879mm (96 x 96 DPI)

ACS Paragon Plus Environment

Langmuir

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

TOC For graphical table of content only 1723x929mm (96 x 96 DPI)

ACS Paragon Plus Environment

Page 10 of 28

Page 11 of 28

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

1 2 3 4 5 6 7 8 9 10

Langmuir

Water diffusion in amorphous hydrophilic systems: a stop and go process Karol Kulasinski1,2, Robert Guyer3,4, Dominique Derome2, Jan Carmeliet1,2 * 1

Chair of Building Physics, Swiss Federal University of Technology ETH Zurich, Stefano-Franscini-Platz 5, 8093 Zürich, Switzerland 2 Laboratory for Multiscale Studies in Building Physics, Swiss Federal Laboratories for Materials Science and Technology Empa, Überlandstrasse 129, 8600 Dübendorf, Switzerland 3 Solid Earth Geophysics Group, Los Alamos National Laboratory, MS D446, Los Alamos, New Mexico 87545, United States 4 Department of Physics, University of Nevada, Reno, Nevada 89557, United States * corresponding author: [email protected]

Abstract

11 12 13 14 15 16 17 18 19

The diffusion of H2O in three amorphous polymer-H2O systems is studied as a function of H2O content using molecular dynamics. A picture of H2O molecule motion comprising alternating steps of being bound at an adsorption site (‘stop’) and moving (‘go’) emerges. This picture is made quantitative. The bound time, frequency of stop-go steps and tortuosity all decrease with H2O content. Fourier analysis of particle motion during bound time segments provides a measure of an attempt frequency that is connected quantitatively to the bound time and an activation energy of a hydrogen bond. On increase in H2O content, the polymer-H2O systems swell leading to an increase in diffusion coefficient and porosity, and a decrease in activation energy.

20 21 22 23 24 25 26 27 28

Cellulosic materials are the subject of intense investigation. They are basic to numerous applied physical systems, e.g., the paper of Leonardo's self-portrait 1, a T-shirt 2, a hydro-actuated plant response 3. Consequently these materials are studied in the laboratory at the thermodynamic level (adsorption isotherms, elastic constants, etc. 4,5), at the mesoscopic level (n-scattering from a "stick" of a wood cell wall 6) and at the microscopic level (X-ray scattering, NMR, etc. 7–9). These materials are ideal systems for exploration by modeling with molecular dynamics. Cellulosic materials are of interest in the crystalline or amorphous state, dry or in contact with H2O, e.g. the adsorption of water causes the swelling of wood, leading to an increase in the ability of wood to transport water 10–13.

29 30 31 32 33

Phenomena like hygro-actuated plant response, e.g. the timely opening of a Pine cone for seed dispersal that depends on the coupling of H2O to a cellulosic-like material, the closing of the resurrection plant to protect the plant during a drought 14, focus attention on the dynamics of H2O in these materials. Several analytical models for the diffusion of H2O in various polymeric system have been proposed 15–17.

34 35 36 37 38 39 40

Molecular dynamics (MD) studies of H2O diffusion in a number of complex systems, e.g. amorphous polyamide, have been undertaken 18–22. These studies, modest in scope, yield results in qualitative accord with a number of phenomenological models. There is a number of attempts that explain the diffusion of fluid at the molecular scale, e.g. 23–27. Jepps et al. 24 consider two diffusive regimes, at low and high saturation, and attribute to them two different flow models. A similar approach has recently been used to describe the behavior of adsorbed molecules at a larger scale 28 or, by a subdivision of molecular trajectories, to quantify the generalized fast-exchange diffusion equation 27.

41 42 43 44 45 46 47

In this paper, we describe the results of an extensive MD study of the transport of H2O in several amorphous cellulosic systems. The systems we study have no simple geometry where concepts like substrate, interface, vicinal would apply. They are polymer-H2O systems in which the polymer, fully engaged with the H2O, evolves in structure as the H2O molecules move within it at fixed H2O content and evolves further in structure as the H2O content evolves. We observe a stop-and-go behavior and introduce a model to organize the analysis of our findings that is tied quantitatively to a microscopic model of H2O motion. This model involves a picture of H2O movement that holds promise for

Introduction

1 ACS Paragon Plus Environment

Langmuir

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 12 of 28

48 49

application beyond the domain of our study, e.g. a modification of diffusion constants. The efficacy of this model justifies the expectation that the style of analysis described here has general validity.

50 51 52 53 54 55 56 57

Materials and Methods In this paper we investigate MD models of H2O in three polymeric systems: amorphous cellulose 29 (AC), amorphous hemicellulose 29 (HC) and a heterogeneous structure comprising crystalline cellulose and galactoglucomannan, a model for a microfibril (μF) 8,30 (Fig. 1). Both polymers are present in large quantities in wood tracheids, being building units of the S2 layer structure of a wood cell wall. Amorphous cellulose and hemicellulose are strongly hydrophilic due to their open porous structure and the presence of three exposed hydroxyl groups per glucose unit with which H2O molecules easily form hydrogen bonds 31,32.

58 59 60 61 62 63 64 65 66 67 68 69 70

The MD simulations are carried out using Gromos 53a6 united-atom force field 33,34 with leap-frog algorithm for integration of Newton’s equations of motion using Single Point Charge water molecules in isobaric-isothermal ensemble at 300 K and zero pressure, controlled by Velocity-Rescaling thermostat 35 and anisotropic Berendsen barostat 36. The cut-off radius of the Van der Waals interactions between atoms, implemented as Lennard-Jones potential, is set to 1 nm, the long-range electrostatic interactions are implemented with Particle-Mesh Ewald summation, the energy and pressure are corrected for dispersion forces, and the distance between oxygen and hydrogen in OH bonds is constrained. The simulations are integrated with a time step of 1 fs and periodic boundary conditions are applied in all directions. Prior to relaxation runs, the systems are energy-minimized with steepest-descent and then conjugated-gradient algorithms. The obtained structures, nominally having volume 37, 95, 94 nm3, respectively, and containing 180, 432, 470 basic units, respectively, are formed using the temperature/annealing protocol described previously 29,37,38. As a result, they are characterized by a density of 1.33 (AC), 1.22 (HC) and 1.35 g cm-3 (microfibril) 29.

71

(a)

(b)

(c)

72 73 74 75

Figure 1. Snapshots of the modelled MD structures in dry state: (a) amorphous cellulose, (b) galactoglucomannan, a hemicellulose 29, (c) cellulose microfibril, composed of crystalline cellulose (grey) and hemicellulose (red).

76 77 78 79 80 81 82

Molecules of H2O are introduced one by one into the void space between the polymer chains, each single insertion being followed by a 10 ps relaxation run, as described in details in 38. For each system (AC, HC, μF), the water content (number of H2O molecules) is characterized by s=N/Nmax where N is the number of adsorbed water molecules and Nmax is the maximum number of the molecules that can be adsorbed in each system. Nmax is determined from experimentally measured adsorption isotherms and equal to 800, 2900 and 1000, for AC, HC and μF, respectively. As a preface to the studies of H2O diffusion, the systems have been studied in the range 0≤s≤1 as equilibrium 2 ACS Paragon Plus Environment

Page 13 of 28

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

Langmuir

83 84 85 86 87 88 89 90 91

thermodynamic systems; the chemical potential, elastic constants and porosity have been determined 29,37,38. Porosity is understood as the volume of the voids, when all water is removed, divided by the total volume, whereas the free volume is the remaining void volume when water molecules are still present. Both quantities are determined as the acceptance ratios, measured using an H20 molecule as the probe, an obvious choice in a H20-polymer interactions study. Typically, a water molecule is inserted 106 times per structure and per s, therefore the free space is properly sampled. This common approach was used for example by Lourenco et al. 39 using the Van der Waals radii suggested by Bondi 40. Additionally, the behavior of the hydrogen bond structure has been studied 29,37.

92 93 94

To calculate the diffusion coefficient a system is brought to thermal equilibrium at s by energy minimization and equilibration and the Mean Square Displacement (MSD) of N(s) H2O molecules, followed over the subsequent tmax=2 ns, is averaged

95

  = lim →  ∑





   

(1)

96 97 98 99 100 101 102 103 104 105 106 107 108 109

The diffusion coefficient is obtained from the slope of the linear part of an MSD curve, typically after 600 ps. We note that within 2 ns the molecules travel on average a distance of 2 nm which is almost 2/3 of the size of the simulation box and, by running longer 10 ns simulations, it was found that 2 ns is sufficient to extract a valid diffusion coefficient (Fig. 2a). We complement the determination of D with the examination of the motion of individual molecules. We observe that the motion of an H2O molecule is built up of alternating segments of moving (‘go’) and ‘standing still’ (‘stop’ or ‘waiting’). The relative amount of time spent in the two states evolves as s evolves. We quantify the description of H2O motion using these two states. We define a criterion for ‘stop’: an H2O molecule is waiting at time t, ti≤t≤ti+1, if the deviation of the displacement from its average value is smaller than 0.1 nm (OH bond length) and (b) the rate of change of its squared displacement is smaller than 8×10-8 nm2 ns-1, in order to eliminate very slow but progressive movement. This value is selected by an observation of H2O molecules movement at low saturation (Fig. 2b). The waiting time is then τwi= ti+1-ti. As waiting and moving succeed one another, a trajectory is a sequence of alternating segments (Fig. 2b). We define two average times, waiting and moving, at each s:

110

     =  ∑   ∑    , !  =  ∑  ∑  !  (2)

111 112

where Pi is the number of changes of state of an H2O molecule. The frequency of switching between states during the simulation time is ν, given by

113

"   =

114

We necessarily have

115

$ = $ + $! =  ∗ "$ + ! ∗ "$, "  =  + !

116 117

We define a microscopic diffusion coefficient, Dµ, related to the time during which a particle is moving, tm=τmνt, schematically,

118

 = lim

119

and the related tortuosity ξ as

120

+=



 



∑ 



,-./0 ,1

.

#

'  



,



,

(3)

= lim

' 

→ ( )

(4)

! " = * ! " (5)

(6) 3 ACS Paragon Plus Environment

Langmuir

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

121 122 123 124 125 126 127 128

where Dbulk is the diffusion coefficient of bulk liquid H2O. The microscopic diffusion coefficient describes the molecular diffusion in the porous material as if the molecules did not wait at the sorption sites, by “removing” the waiting time. We employ the definition of tortuosity as related to the complexity of the path an H2O molecule follows during its trajectory, as defined for example by Johnson et al. 41. With this definition, at large water content a water molecule is thought of simply be in bulk water. The separation of particle motion into two distinct states is qualitatively similar to the treatment of H2O in the rather different context of bulk/vicinal water in a pore studied by Levitz et al.

129 130

Finally we suggest that a quantitative understanding of τw can be provided by the transition-state formula 42 for escape over a barrier

131

  = 23 4

132 133 134 135 136 137

where fa is an attempt frequency and ϵB is an activation energy of an H2O to an adsorption site. The adopted physical picture is that the motion of a molecule through the system has no reason to be characterized by a particular frequency except during those periods when it is effectively stopped at a binding site where a molecule experiences locally an approximately harmonic potential. Occurrences of this harmonic potential are then found in the analysis of time series of particles displacement.

138 139 140 141 142 143 144 145 146 147 148 149 150 151 152 153 154

Results We present the results for three different hydrophilic polymers in order to clarify that the model is not case-dependent. The diffusion coefficient, which we show in Fig. 2c for each case, although differing slightly among the systems, displays a similar behavior for all, i.e. an increase of about one order of magnitude as s evolves from 0 to 1. We note that the associated relative error on the data ranges between 5-10% and decreases with s. The values of D at s=1 are in agreement with measurements 6,7,10,11,18 and a similar exponential relation is reported by Topgaard and Söderman 26 that used NMR to characterize the water adsorbed in cellulose fibers. For ν, τw and τm we find the results shown in Fig. 2d-f. At s→0 the amount of time spent at an adsorption site is comparable to the amount of time spent moving, τw/τm≈1, whereas as s→1 the waiting time is small compared to the moving time, τm/τw≈40. Qualitatively, a similar picture has been found by Malani and Ayappa 43 where, by studying the dynamics of rotational jumps of water molecules confined in a hydrophilic pore system, they demonstrated that the water in the contact layer possesses a ‘residence time’ that is approximately 10 times larger than that of bulk water. The longer waiting time is a consequence of the stronger hydrogen bonds and stronger electrostatic interactions. The waiting time at saturation obtained in this study is larger than the measured residence time of a bulk water, 2-4 ps 43,44, as bulk water is obviously devoid of solid-water interactions.

155 156 157 158 159 160 161

We note that the nature of sorption sites of cellulose and hemicellulose, that causes molecules to wait, is their electrostatic attractiveness to water molecules as the sorption sites are strongly polarized and form hydrogen bonds. For example, in the applied force field, the difference of the charge between the oxygen and the hydrogen of a hydroxyl group is 1.051e, whereas, for comparison, the corresponding difference in a SPC model of water molecules amount to 1.23e. These large charge shifts result in a strong attractive forces between the water molecules and sorption sites, which leads to formation of hydrogen bonds.

28

5  6

06 7

, 89  = :9 ; ln23   (7)

162 163 4 ACS Paragon Plus Environment

Page 14 of 28

Page 15 of 28

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

164

Langmuir

(a)

(b)

165 166

(c)

(d)

(e)

(f)

167 168

169

5 ACS Paragon Plus Environment

Langmuir

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

170 171 172

Figure 2. (a) MSD curves in log-log plot for different s. (b) Example of a single water molecule trajectory with highlighted waiting periods. Basic results: (c) diffusion coefficient, (d) average frequency of stop&go periods, (e) average waiting time, (f) average moving time.

173 174 175 176 177 178 179 180 181 182 183

After an extraction of Dµ from D, we find the result shown in Fig. 3a and the associated tortuosity shown in Fig. 3b. We find Dµ to monotonically increase with s. The tortuosity, a measure that uses D in bulk H2O as reference (determined by MD, Dbulk=4.17), decreases by a factor of about 10 in going from s=0 to s=1. We note that we documented in 29 that the corresponding porosity of AC and HC increases by about a factor of 10 in going from s=0 to s=1. We note that Topgaard and Söderman attributed the difference between the measured diffusion coefficient and that of bulk water to tortuosity and that they found tortuosity to decrease with H2O content 26. In a similar way as used in the fast-exchange diffusion model 27, the obtained diffusion coefficient D(s) may be interpreted as the fraction of H2O molecules found instantaneously in the mobile state multiplied by the microscopic diffusion coefficient Dµ. (a) (b)

184 185

Figure 3. Output of the model: (a) microscopic diffusion coefficient and (b) tortuosity.

186 187 188 189 190 191 192 193 194 195 196 197

The decrease in tortuosity can be understood by analyzing the system porosity (Fig. 4a). The porosity at a given H2O content is estimated by probing the pore space with a water molecule and calculating the corresponding acceptance ratio 29. As a result of adsorption, porosity increases in a linear manner (s>0.2 for μF) and this increase is accompanied by a decreasing tortuosity, as the pores merge and expand 29. A diffusion coefficient increasing with pore size has been also found by Jepps et al. 23. The free volume, i.e. the pore volume minus the volume occupied by H2O molecules is estimated at different H2O content and presented in Fig. 4b. The free volume related to the total volume decreases as the adsorbed amount increases, in particular for s0.5, the percentage of the free volume remains at a constant level, below 10%. For this reason, the increasing diffusion coefficient cannot be explained on the basis of the free volume theory 45, attributing an increase in diffusion to an increase in the free volume available on average for a water molecule, which justifies the choice of our approach.

198

(a)

(b)

6 ACS Paragon Plus Environment

Page 16 of 28

Page 17 of 28

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

Langmuir

199 200 201

Figure 4. (a) Porosity and (b) free volume as related to the total volume, as a function of H2O content.

202 203 204 205 206

To employ the transition state formula, Eq. (7), with τw in hand, we need fa or ϵB. In order to find fa, we go to s→0 and examine the Fourier transform of the displacement of H2O molecules during the times that are defined as waiting times. This Fourier transform has a single peak against a background that we remove by considering the motion of a H2O molecule in bulk water, i.e. a system without adsorption sites. What results from this procedure is the Fourier spectrum shown in Fig. 5a.

207 208 209 210 211 212 213 214 215 216 217 218

We take fa to be the frequency of oscillatory motion of an H2O molecule at an adsorption site to be 2.4×1013 s-1 for AC, 1.5×1013 s-1 for HC and 2.6×1013 s-1 for μF. Frequencies of this order are sensibly related to the expected behavior of H2O, e.g. the Debye temperature of water corresponds to a frequency of order 1013 s-1 46. Employing these frequencies at all s leads to the "barrier height" or activation (binding) energy, ϵB, as a function of s shown in Fig. 5b. The activation energy decreases slightly as s increases. The activation energy at s→0, e.g. ϵB(0)=22.6 kJ·mol-1 for AC, also found by other methods 18, is close to the corresponding difference in chemical potential at s→0, which amounts to 21.4 kJ mol-1 38. This implies that, in the limit s→0, the change in chemical potential is the activation energy of hydrogen bonds. It is assumed that at a given saturation the water molecules are uniformly distributed in the system and the adsorption sites have similar energies. However, the calculated energy is an average energy of the binding sites at a given saturation and does not have to be identical for all sorption sites.

219 220 221 222

The experience of a molecule in the system is approximated as made up of two states. The molecule is trapped at a binding site or free to move diffusively in the bulk. The energy ϵB is the activation energy that must be provided by thermal fluctuations to promote the molecule from a binding site (where it waits) into the bulk (where it is able to move diffusively).

223

(a)

(b)

7 ACS Paragon Plus Environment

Langmuir

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

224 225 226

Figure 5. Model of a binding site: (a) Fourier transform of MSD during waiting periods, after background removal at s→0, (b) estimated binding energy at a binding site.

227 228 229 230 231 232 233 234 235 236 237 238

Dry amorphous cellulosic material comprises a "backbone structure", whose integrity is partially enforced by H-bonds, and that is entangled with itself, the entanglement being partially supported also by H-bonds. The introduction of H2O molecules, having an affinity for H-bonds, into such a system results in breaking of the H-bonds leading to an unfolding of the self-entanglement and a modification of the structure. As H2O molecules are introduced, the system swells. The pore space, that necessarily accompanies the disorganization associated with amorphous packing, evolves, becomes larger, as the porosity of HC and AC increases by an order of magnitude in going from a degree of saturation s=0 to s=1, and is re-arranged leading to a decrease in tortuosity. The resulting re-arrangement feeds back on the H2O molecules offering them different/additional opportunities to bond. Added to the dynamics conferred by this geometry is that from thermal fluctuations. Thus the H2O molecules, whose diffusion coefficient we have measured, move through a porosity that is a dynamic entity changing with degree of saturation.

239 240 241 242 243 244 245 246 247 248 249 250 251 252 253 254 255 256 257 258

Conclusion This qualitative picture is made quantitative with the MD simulation. The changes in the diffusion coefficient, as s increases, result from changes in the geometry of the system and in the encounters of H2O molecules with this geometry. At s→0 the H2O molecules spend a large fraction of their time bound, in essential isolation, and when liberated move in a tortuous environment characterized by: small porosity (φ≈0.1), large waiting time (τm/(τw+τm)≈0.4), activation energy close to the chemical potential (ϵB≈∆µ(0)) and high tortuosity (ξ≈100). The H2O molecules bound at adsorption sites are not diffusing and therefore the diffusion coefficient at low s is low. On increase in s, the H2O molecules, in competition with one another for binding sites that are less attractive due to superposed Coulomb interactions, spend more time moving in an environment opened up by swelling and less tortuous. At s→1, a system is characterized by large porosity (φ≈10·φ(0)), negligible waiting time (τm/(τw+τm)≈1), smaller activation energy (ϵB≈0.85ϵB(s→0)) and lower tortuosity (ξ≈10). As a consequence the diffusion coefficient increases with a factor 100 compared to that of the dry material. We note that, although the presented model does not involve porosity directly, it is implicitly involved in the tortuosity, as the increase in porosity is accompanied by an opening of the pore network. The decrease in tortuosity is a result of unfolding chains, an increase in porosity and pore size, as well as merging of water clusters 29,38. This quantification description rests importantly on the two-state picture of the H2O molecules experience in the system. This picture is given support by the connection we have established between the waiting time τw, the attempt frequency fa and the activation energy ϵB. We have shown that water diffusion in hydrophilic systems can be 8 ACS Paragon Plus Environment

Page 18 of 28

Page 19 of 28

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

Langmuir

259 260 261 262

understood by the notion of ‘stop&go’ periods. This model can be used to predict a change in diffusion coefficient resulting from a potential modification of the activation energy of adsorption sites. It may lead to an improved design of functionalized polymer systems, tracking their structural change, swelling and diffusion upon increase of H2O content.

263

References

264 265 266

(1)

Conte, A. M.; Pulci, O.; Misiti, M. C.; Lojewska, J.; Teodonio, L.; Violante, C.; Missori, M. Visual Degradation in Leonardo Da Vinci’s Iconic Self-Portrait: A Nanoscale Study. Appl. Phys. Lett. 2014, 104 (22), 224101.

267

(2)

Langmuir Shirts & T-Shirts http://www.cafepress.co.uk/+langmuir+t-shirts.

268 269 270

(3)

Harrington, M. J.; Razghandi, K.; Ditsch, F.; Guiducci, L.; Rueggeberg, M.; Dunlop, J. W. C.; Fratzl, P.; Neinhuis, C.; Burgert, I. Origami-like Unfolding of Hydro-Actuated Ice Plant Seed Capsules. Nat. Commun. 2011, 2, 337.

271 272

(4)

Nishino, T.; Takano, K.; Nakamae, K. Elastic Modulus of the Crystalline Regions of Cellulose Polymorphs. J. Polym. Sci. Part B Polym. Phys. 1995, 33 (11), 1647–1651.

273 274

(5)

Tremaine, P. R.; Gray, D. G. Adsorption of Non-Swelling Vapours on the Surface of Cellulose. J. Chem. Soc. Faraday Trans. 1 Phys. Chem. Condens. Phases 1975, 71, 2170.

275 276

(6)

Kollmann, F. F. P.; Cote Jr, W. A. Principles of Wood Science and Technology. Vol. I. Solid Wood; Springer, 1968.

277 278

(7)

Froix, M. F.; Nelson, R. The Interaction of Water with Cellulose from Nuclear Magnetic Resonance Relaxation Times. Macromolecules 1975, 8 (6), 726–730.

279 280

(8)

Nishiyama, Y. Structure and Properties of the Cellulose Microfibril. J. Wood Sci. 2009, 55 (4), 241–249.

281 282

(9)

Stevanic, J. S.; Salmén, L. Orientation of the Wood Polymers in the Cell Wall of Spruce Wood Fibres. Holzforschung 2009, 63 (5), 497–503.

283 284

(10)

Tremblay, C.; Cloutier, A.; Fortin, Y. Determination of the Effective Water Conductivity of Red Pine Sapwood. Wood Sci. Technol. 2000, 34 (2), 109–124.

285 286

(11)

Tremblay, C.; Cloutier, A.; Fortin, Y. Experimental Determination of the Convective Heat and Mass Transfer Coefficients for Wood Drying. Wood Sci. Technol. 2000, 34 (3), 253–276.

287 288 289

(12)

Sonderegger, W.; Vecellio, M.; Zwicker, P.; Niemz, P. Combined Bound Water and Water Vapour Diffusion of Norway Spruce and European Beech in and between the Principal Anatomical Directions. Holzforschung 2011, 65 (6), 819–828.

290 291 292

(13)

Araujo, C. D.; Avramidis, S.; MacKay, A. L. Behaviour of Solid Wood and Bound Water as a Function of Moisture Content. A Proton Magnetic Resonance Study. Holzforschung 1994, 48 (1), 69–74.

293 294

(14)

Rafsanjani, A.; Brulé, V.; Western, T. L.; Pasini, D. Hydro-Responsive Curling of the Resurrection Plant Selaginella Lepidophylla. Sci. Rep. 2015, 5, 8064. 9 ACS Paragon Plus Environment

Langmuir

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

295

(15)

Do, D. D. Adsorption Analysis: Equilibria and Kinetics; 1998; Vol. 2.

296 297 298

(16)

Angelico, R.; Palazzo, G.; Colafemmina, G.; Cirkel, P. A.; Giustini, M.; Ceglie, A. Water Diffusion and Headgroup Mobility in Polymer-like Reverse Micelles: Evidence of a Sphere-to-Rod-toSphere Transition. J. Phys. Chem. B 1998, 102 (16), 2883–2889.

299 300

(17)

Peppas, N. A.; Brannon-Peppas, L. Water Diffusion and Sorption in Amorphous Macromolecular Systems and Foods. J. Food Eng. 1994, 22 (1-4), 189–210.

301

(18)

Nelson, R. M. Diffusion of Bound Water in Wood. Wood Sci. Technol. 1986, 20 (4), 309–328.

302 303 304

(19)

Levitt, M.; Hirshberg, M.; Sharon, R.; Laidig, K. E.; Daggett, V. Calibration and Testing of a Water Model for Simulation of the Molecular Dynamics of Proteins and Nucleic Acids in Solution. J. Phys. Chem. B 1997, 101 (25), 5051–5061.

305 306 307

(20)

Roberts, C. J.; Debenedetti, P. G. Structure and Dynamics in Concentrated, Amorphous Carbohydrate−Water Systems by Molecular Dynamics Simulabon. J. Phys. Chem. B 1999, 103 (34), 7308–7318.

308 309

(21)

Müller-Plathe, F. Diffusion of Water in Swollen Poly(vinyl Alcohol) Membranes Studied by Molecular Dynamics Simulation. J. Memb. Sci. 1998, 141 (2), 147–154.

310 311 312

(22)

Kotelyanskii, M. J.; Wagner, N. J.; Paulaitis, M. E. Molecular Dynamics Simulation Study of the Mechanisms of Water Diffusion in a Hydrated, Amorphous Polyamide. Comput. Theor. Polym. Sci. 1999, 9 (3-4), 301–306.

313 314

(23)

Jepps, O. G.; Bhatia, S. K.; Searles, D. J. Wall Mediated Transport in Confined Spaces: Exact Theory for Low Density. Phys. Rev. Lett. 2003, 91 (12), 126102.

315 316

(24)

Jepps, O. G.; Bhatia, S. K.; Searles, D. J. Modeling Molecular Transport in Slit Pores. J. Chem. Phys. 2004, 120 (11), 5396–5406.

317 318 319

(25)

Petropoulos, J. H.; Papadokostaki, K. G. May the Knudsen Equation Be Legitimately, or at Least Usefully, Applied to Dilute Adsorbable Gas Flow in Mesoporous Media? Chem. Eng. Sci. 2012, 68 (1), 392–400.

320 321

(26)

Topgaard, D.; Söderman, O. Diffusion of Water Absorbed in Cellulose Fibers Studied with 1 HNMR. Langmuir 2001, 17 (9), 2694–2702.

322 323 324

(27)

Zeigermann, P.; Naumov, S.; Mascotto, S.; Kärger, J.; Smarsly, B. M.; Valiullin, R. Diffusion in Hierarchical Mesoporous Materials: Applicability and Generalization of the Fast-Exchange Diffusion Model. Langmuir 2012, 28 (7), 3621–3632.

325 326 327

(28)

Levitz, P.; Bonnaud, P. A.; Cazade, P.-A.; Pellenq, R. J.-M.; Coasne, B. Molecular Intermittent Dynamics of Interfacial Water: Probing Adsorption and Bulk Confinement. Soft Matter 2013, 9 (36), 8654.

328 329 330

(29)

Kulasinski, K.; Guyer, R.; Keten, S.; Derome, D.; Carmeliet, J. Impact of Moisture Adsorption on Structure and Physical Properties of Amorphous Biopolymers. Macromolecules 2015, 48 (8), 2793–2800.

10 ACS Paragon Plus Environment

Page 20 of 28

Page 21 of 28

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

Langmuir

331 332 333

(30)

Fernandes, A. N.; Thomas, L. H.; Altaner, C. M.; Callow, P.; Forsyth, V. T.; Apperley, D. C.; Kennedy, C. J.; Jarvis, M. C. Nanostructure of Cellulose Microfibrils in Spruce Wood. Proc. Natl. Acad. Sci. U. S. A. 2011, 108 (47), E1195–E1203.

334 335

(31)

Nishiyama, Y.; Chanzy, H.; Langan, P. Neutron and Synchrotron X-Ray Fiber Diffraction Studies of Cellulose Polymorphs. Abstr. Pap. Am. Chem. Soc. 2002, 223, D26–D26.

336 337 338

(32)

Nishiyama, Y.; Langan, P.; Chanzy, H. Crystal Structure and Hydrogen-Bonding System in Cellulose I Beta from Synchrotron X-Ray and Neutron Fiber Diffraction. J. Am. Chem. Soc. 2002, 124 (31), 9074–9082.

339 340 341

(33)

Oostenbrink, C.; Villa, A.; Mark, A. E.; Van Gunsteren, W. F. A Biomolecular Force Field Based on the Free Enthalpy of Hydration and Solvation: The GROMOS Force-field Parameter Sets 53A5 and 53A6. J. Comput. Chem. 2004, 25 (13), 1656–1676.

342 343

(34)

Lins, R. D.; Hünenberger, P. H. A New GROMOS Force Field for Hexopyranose-based Carbohydrates. J. Comput. Chem. 2005, 26 (13), 1400–1412.

344 345

(35)

Bussi, G.; Donadio, D.; Parrinello, M. Canonical Sampling through Velocity Rescaling. J. Chem. Phys. 2007, 126 (1), 14101.

346 347

(36)

Berendsen, H. J. C.; Postma, J. P. M.; van Gunsteren, W. F.; DiNola, A.; Haak, J. R. Molecular Dynamics with Coupling to an External Bath. J. Chem. Phys. 1984, 81 (8), 3684–3690.

348 349 350

(37)

Kulasinski, K.; Keten, S.; Churakov, S. V.; Derome, D.; Carmeliet, J. A Comparative Molecular Dynamics Study of Crystalline, Paracrystalline and Amorphous States of Cellulose. Cellulose 2014, 21 (3), 1103–1116.

351 352 353

(38)

Kulasinski, K.; Keten, S.; Churakov, S. V; Guyer, R.; Carmeliet, J.; Derome, D. Molecular Mechanism of Moisture-Induced Transition in Amorphous Cellulose. ACS Macro Lett. 2014, 3, 1037–1040.

354 355 356

(39)

Lourenço, T. C.; Coelho, M. F. C.; Ramalho, T. C.; van der Spoel, D.; Costa, L. T. Insights on the Solubility of CO2 in 1-Ethyl-3-Methylimidazolium Bis(trifluoromethylsulfonyl)imide from the Microscopic Point of View. Environ. Sci. Technol. 2013, 47 (13), 7421–7429.

357

(40)

Bondi, A. Van Der Waals Volumes and Radii. J. Phys. Chem. 1964, 68 (3), 441–451.

358 359

(41)

Johnson, D. L.; Koplik, J.; Dashen, R. Theory of Dynamic Permeability and Tortuosity in FluidSaturated Porous Media. J. Fluid Mech. 1987, 176, 379–402.

360 361

(42)

Kramers, H. A. Brownian Motion in a Field of Force and the Diffusion Model of Chemical Reactions. Physica 1940, 7 (4), 284–304.

362 363

(43)

Malani, A.; Ayappa, K. G. Relaxation and Jump Dynamics of Water at the Mica Interface. J. Chem. Phys. 2012, 136 (19), 194701.

364 365

(44)

Laage, D.; Hynes, J. T. A Molecular Jump Mechanism of Water Reorientation. Science 2006, 311 (5762), 832–835.

11 ACS Paragon Plus Environment

Langmuir

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

366 367 368

(45)

Ramesh, N.; Davis, P. K.; Zielinski, J. M.; Danner, R. P.; Duda, J. L. Application of Free-Volume Theory to Self Diffusion of Solvents in Polymers below the Glass Transition Temperature: A Review. J. Polym. Sci. Part B Polym. Phys. 2011, 49, 1629–1644.

369

(46)

Ashcroft, N. W.; Mermin, N. D. Solid State Physics; 1976; Vol. 2.

370 371

FOR GRAPHICAL TABLE OF CONTENT ONLY

372

373

12 ACS Paragon Plus Environment

Page 22 of 28

Page 23 of 28

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

Langmuir

Figure 2. (a) MSD curves in log-log plot for different s. (b) Example of a single water molecule trajectory with highlighted waiting periods. Basic results: (c) diffusion coefficient, (d) average frequency of stop&go periods, (e) average waiting time, (f) average moving time. 69x61mm (600 x 600 DPI)

ACS Paragon Plus Environment

Langmuir

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

Figure 2. (a) MSD curves in log-log plot for different s. (b) Example of a single water molecule trajectory with highlighted waiting periods. Basic results: (c) diffusion coefficient, (d) average frequency of stop&go periods, (e) average waiting time, (f) average moving time. 69x61mm (600 x 600 DPI)

ACS Paragon Plus Environment

Page 24 of 28

Page 25 of 28

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

Langmuir

Figure 2. (a) MSD curves in log-log plot for different s. (b) Example of a single water molecule trajectory with highlighted waiting periods. Basic results: (c) diffusion coefficient, (d) average frequency of stop&go periods, (e) average waiting time, (f) average moving time. 998x879mm (96 x 96 DPI)

ACS Paragon Plus Environment

Langmuir

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

Figure 2. (a) MSD curves in log-log plot for different s. (b) Example of a single water molecule trajectory with highlighted waiting periods. Basic results: (c) diffusion coefficient, (d) average frequency of stop&go periods, (e) average waiting time, (f) average moving time. 998x879mm (96 x 96 DPI)

ACS Paragon Plus Environment

Page 26 of 28

Page 27 of 28

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

Langmuir

Figure 2. (a) MSD curves in log-log plot for different s. (b) Example of a single water molecule trajectory with highlighted waiting periods. Basic results: (c) diffusion coefficient, (d) average frequency of stop&go periods, (e) average waiting time, (f) average moving time. 998x879mm (96 x 96 DPI)

ACS Paragon Plus Environment

Langmuir

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

Figure 2. (a) MSD curves in log-log plot for different s. (b) Example of a single water molecule trajectory with highlighted waiting periods. Basic results: (c) diffusion coefficient, (d) average frequency of stop&go periods, (e) average waiting time, (f) average moving time. 998x879mm (96 x 96 DPI)

ACS Paragon Plus Environment

Page 28 of 28