• 沒有找到結果。

The violation of the Stokes–Einstein relation in supercooled water

N/A
N/A
Protected

Academic year: 2021

Share "The violation of the Stokes–Einstein relation in supercooled water"

Copied!
5
0
0

加載中.... (立即查看全文)

全文

(1)

The violation of the Stokes–Einstein relation

in supercooled water

Sow-Hsin Chen*†, Francesco Mallamace*, Chung-Yuan Mou§, Matteo Broccio*, Carmelo Corsaro, Antonio Faraone‡¶, and Li Liu*

*Department of Nuclear Science and Engineering, Massachusetts Institute of Technology, Cambridge, MA 02139;‡Dipartimento di Fisica and Consorzio Nazionale Interuniversitario per le Scienze Fisiche della Materia, Universita` di Messina, I-98166 Messina, Italy; and§Department of Chemistry, National Taiwan University, Taipei 106, Taiwan

Edited by H. Eugene Stanley, Boston University, Boston, MA, and approved June 27, 2006 (received for review April 21, 2006) By confining water in nanopores, so narrow that the liquid cannot

freeze, it is possible to explore its properties well below its homogeneous nucleation temperature TH⬇ 235 K. In particular, the

dynamical parameters of water can be measured down to 180 K, approaching the suggested glass transition temperature Tg⬇ 165

K. Here we present experimental evidence, obtained from Nuclear Magnetic Resonance and Quasi-Elastic Neutron Scattering spec-troscopies, of a well defined decoupling of transport properties (the self-diffusion coefficient and the average translational relax-ation time), which implies the breakdown of the Stokes–Einstein relation. We further show that such a non-monotonic decoupling reflects the characteristics of the recently observed dynamic cross-over, at⬇225 K, between the two dynamical behaviors known as

fragile and strong, which is a consequence of a change in the hydrogen bond structure of liquid water.

decoupling of transport properties兩 dynamic crossover 兩 MCM-41

D

espite its fundamental importance in science and

tech-nology, the physical properties of water are far from completely understood. The liquid state of water is unusual, especially at low temperatures (1–3). For example, contrary to other liquids, water behaves as if there exists a singular temperature toward which its thermodynamical properties, such as compressibility, thermal expansion coefficient, and specific heat, diverge (1). The efforts of scientists from many disciplines to seek a coherent explanation for this unusual behavior, in combination with its wide range of impacts, make water one of the most important open questions in science today. On the other hand, the nature of the glass transition (GT) of water represents another challenging subject for current research (4). Dynamical measurements in glass-forming liquids have shown a dramatic slowdown of both

macroscopic (viscosity␩ and self-diffusion coefficient D) and

microscopic (average translational correlation time␶)

observ-ables, as temperature is lowered toward the GT temperature

Tg. Accordingly, a comprehension of the GT has been sought

through the study of the dynamics at the molecular level, which, despite all efforts, has not yet been completely under-stood (5– 8). Keeping in mind the ‘‘complexities’’ of both low-temperature water and its GT, we present here direct measurements of two dynamical parameters of water: the self-diffusion coefficient and the average translational relax-ation time, in the temperature range from 280 to 190 K, obtained by NMR and quasi-elastic neutron scattering (QENS) experiments, respectively.

Bulk water can be supercooled below its melting

tempera-ture (TM) down to⬇235 K, below which it inevitably

crystal-lizes; it is just in such supercooled metastable state that the anomalies in its thermodynamical properties are most pro-nounced, showing a power law divergence toward a singular

temperature TS⫽ 228 K. At ambient pressure, water can exist

in a glassy form below 135 K. Depending on T and P, glassy water has two amorphous phases with different structures: a

low (LDA) and a high (HDA) density amorphous ice; thus it shows a polymorphism. LDA can be formed from HDA and vice versa; LDA, if heated, undergoes a glass-to-liquid transi-tion transforming into a highly viscous f luid, then crystallizes

into cubic ice at TX ⬇ 150 K. Thus, an experimentally

inaccessible T region exists in bulk water between THand TX.

Experiments performed within this interval could be of fun-damental interest for understanding the many open questions on the physics of water. For example, the presence of a first order liquid–liquid transition line (LLTL), the precise location

of its Tg, recently suggested at⬇165 K (4, 9), and the existence

of a fragile-to-strong dynamic crossover (FSC) on approaching

Tgfrom the liquid side (10). The existence of a LLTL leads to

conjecture that liquid water possesses a low-temperature

second critical point (predicted to be located at Tc⬇ 220 K, Pc

⬇ 1 Kbar) (2), below which it can switch from one phase, a high-density liquid (HDL), to another phase, a low-density liquid (LDL), whose corresponding vitreous forms are the HDA and LDA, respectively. The difference between the two liquid phases lies in the water structure: in the HDL, the local tetrahedrally coordinated hydrogen-bond network is not fully developed, whereas in the LDL, a more open, locally ice-like,

hydrogen-bond network is fully developed (11). Thus, near Tc,

water is a mixture of both LDL and HDL phases associated with a diverging density f luctuation. At higher temperatures, the two liquid phases are indistinguishable. Lowering temper-ature or increasing pressure will result in an increase of the LDL phase with respect to the HDL phase. The FSC can be identified by the temperature at which transport properties,

like the shear viscosity␩ or the inverse self-diffusion

coeffi-cient 1兾D, cross over from a non-Arrhenius (fragile) to an

Arrhenius (strong) behavior on approaching Tg.

A possibility to enter this inaccessible temperature range of water, named ‘‘no-man’s-land,’’ is now shown by confining water in nano-size pores (12–15). When contained within these pores,

water does not crystallize, and can be supercooled well below TH.

Vycor pores (14, 15) (a porous hydrophilic silica glass), micellar systems, or layered vermiculite clay (12) are examples of con-fining nanostructures. The latter systems have been used to

Conflict of interest statement: No conflicts declared. This paper was submitted directly (Track II) to the PNAS office.

Abbreviations: GT, glass transition; QENS, quasi-elastic neutron scattering; HDL, high-density liquid; LDL, low-high-density liquid; MD, molecular dynamics; SER, Stokes–Einstein relation; VTF, Vogel–Fulcher–Tamman

See Commentary on page 12955.

To whom correspondence should be addressed. E-mail: sowhsin@mit.edu.

Present addresses: Department of Material Science and Engineering, University of

Mary-land, College Park, MD 20742, and National Institute of Standards and Technology Center for Neutron Research, Gaithersburg, MD 20899.

储Present address: Department of Mechanical, Aerospace, and Nuclear Engineering, Rens-selaer Polytechnic Institute, Troy, NY 12180.

(2)

explore the Arrhenius behavior of the dielectric relaxation time

(␶D) of very deeply supercooled water.

The FSC was recently confirmed by a QENS experiment, which measured the T and P dependences of the average

translational relaxation time具␶T典 for water confined in

nanop-ores of silica glass (16, 17). In particular, as the temperature is

lowered, a 具␶T典 versus 1兾T plot exhibits a cusp-like crossover

from a non-Arrhenius to an Arrhenius behavior at a temperature

TL(P). This crossover temperature decreases steadily upon

in-creasing P, until it intersects the THline of bulk water at P⬇ 1.6

Kbar. Beyond this point, the FSC can no longer be identified. These results, suggestive of the existence of the two liquid phases, have been explained in a molecular dynamics (MD) simulation study by considering the existence of a critical point. The MD study shows that the FSC line coincides with the line of

specific heat maxima Cpmax, called Widom line. The Widom line

is the critical isochore above the critical point in the one-phase region (18). Moreover, it is observed that crossing this line corresponds to a change in the T dependence of the dynamics. More precisely, the calculated water diffusion coefficient, D(T), changes according to a FSC, whereas the structural and ther-modynamic properties change from those of HDL to those of LDL.

Results and Discussion

In this report, we present a detailed study done by using two different experimental techniques, neutron scattering and NMR, to probe dynamical properties of confined water at low temper-atures, well inside the inaccessible region of bulk water. Our main aim is to clarify the properties of water as a glass forming

Fig. 1. The temperature dependence of the inverse of self-diffusion coefficient of water and its average translational relaxation time. (A) For the fully hydrated MCM-41-S samples with diameters of 14 and 18 Å, the inverse of the self-diffusion coefficient of water D measured by NMR as a function of 1兾T in a log-linear scale. The solid line denotes the fit of the data to a VFT relation. The short dotted line denotes the fit to an Arrhenius law with the same prefactor 1兾D0. (B) The average translational relaxation time具␶T典 obtained from QENS spectra in the same experimental conditions of the NMR experiment as a function of 1兾T. The dashed line denotes the VFT law fit, and the dotted line denotes the Arrhenius law fit with the same prefactor␶0. The values of fitting parameters are shown.

APPLIED

PHYSICAL

SCIENCES

SEE

(3)

material, measuring directly, with NMR spectroscopy, the self-diffusion coefficient D as a function of temperature, and com-paring the obtained results with the translational relaxation time

具␶T典 measured by QENS. 具␶T典 is a quantity proportional to the

viscosity ␩. These measurements enable one to compare the

proportionality of the transport coefficient 1兾D and viscosity␩,

and can provide a test for the theoretical description of dynamics at the molecular scale of glass-forming materials. According to Ito et al. (10), the FSC can be intimately connected to the presence of a thermodynamic event in liquid water: i.e., the temperature dependence of the inverse self-diffusion coefficient does not follow that of viscosity or inverse mutual diffusion coefficient. In other words, supercooled water must show, on

approaching Tg, the marked decoupling of translational diffusion

coefficient from viscosity or rotational correlation time, as recently observed in some supercooled liquids (19–21). Here we confirm by means of NMR data the existence of a FSC in supercooled water, as proposed by the QENS and MD studies (16–18) and show that the Stokes–Einstein relation (SER) breaks down in different ways on both the fragile and strong sides of FSC.

Fig. 1A shows a log-linear plot of the inverse of the

self-diffusion coefficient of water 1兾D measured by NMR as a

function of 1兾T for the fully hydrated MCM-41-S samples with

pore diameters of 14 and 18 Å. Fig. 1B shows the average

translational relaxation time具␶T典, obtained by analyzing QENS

spectra of the same samples, versus 1兾T, using the

Relaxing-Cage Model (RCM) as shown in refs. 16 and 17. As can be

observed from Fig. 1, the measured values of D and具␶T典 are

independent of the pore size of the samples. This observation indicates that NMR field-gradient measurements, having a length scale larger than the size of the pores, are insensitive to the system geometry. In Fig. 1 A, the solid line denotes the fit of

the data to a Vogel–Fulcher–Tamman (VFT) law 1兾D ⫽

1兾D0exp(BT0兾(T ⫺ T0)), where 1兾D0⫽ 2.4 ⫻ 107(s兾m2), B

1.775, and T0⫽ 187 K. B is a constant providing a measure of

the system fragility, and T0is the ideal GT temperature. The

short dotted line denotes the fit to an Arrhenius law 1兾D ⫽

1兾D0exp(EA兾kBT), where we keep the same 1兾D0value as in the

VFT law fit, and EA⫽ 3.98 Kcal兾mol. Fig. 1B shows the 具␶T典 data

at ambient pressure. The dashed lines denote the VFT law fit,

and the dotted lines denote the Arrhenius law fit, with the same

prefactor ␶0. We obtained the following values: EA ⫽ 5.4

Kcal兾mol, T0⫽ 200 K. The consequence of insisting on the same

prefactor in both the VFT and the Arrhenius laws results in an

equation determining the crossover temperature TL in the

following form: 1兾TL⫽ 1兾T0⫺ BkB兾EA. We obtained TL⫽ 224.5

K from the 1兾D data and TL⫽ 225.8 K from the 具␶T典 data. The

agreement between NMR and QENS results is thus satisfactory,

especially regarding the two relevant quantities EAand TL. Fig.

1 also shows that within the size range of 10–20 Å, the crossover

temperature TL is independent of pore size. As previously

mentioned, a FSC occurring at 228 K has been proposed by Ito et al. (10) for water, which is fragile at room and moderately supercooled temperatures, but has been shown to be a strong liquid by dielectric relaxation measurements near the GT tem-perature (12). The interpretation of this transition as a variant of the structural arrest transition (as predicted by the ideal mode coupling theory) was the essence of the recent QENS study of the structural relaxation time and MD study of the self-diffusion coefficient (16–18). These NMR results presented above con-stitute, by means of a direct measurement of the self-diffusion coefficient of supercooled water, an independent confirmation of the existence of FSC in water.

Let us now focus on the SER that relates the self-diffusion

coefficient D, viscosity␩, and temperature T as D ⬀ T兾␩, which,

as it is well known, is usually accurate for normal- and

high-temperature liquids. Because具␶T典 is proportional to the viscosity,

we examine the relationship between D and具␶T典 in Fig. 2. In Fig.

2, the quantity D具␶T典兾T is reported as a function of T. Dots and

squares represent its values coming from the experimental data of samples with pore diameters of 14 and 18 Å, respectively, whereas the dotted line represents the same quantity obtained by using the corresponding fitted lines reported in Fig. 1. The

temperature dependence of D具␶T典兾T shows that this quantity is

constant at higher T, but increases steeply as T goes below the FSC temperature. Furthermore, it shows a small bump at the FSC temperature, in accordance with the predictions of a recent theoretical study (22). Therefore, in the supercooled region the

temperature behavior of D and具␶T典 is inconsistent with SER,

signaling a marked decoupling between these two transport parameters, on decreasing T. In recent experimental studies on

Fig. 2. The quantity D具␶T典兾T as a function of T. Dots and squares represent its values coming from the experimental data of D and 具␶T典 in samples with diameters of 14 and 18 Å, respectively. The dotted line represents same quantity obtained by using the fitting values obtained from the data reported in Fig. 1.

(4)

some supercooled liquids, it has been reported that SER breaks down as the GT is approached. The self-diffusion coefficient shows, as far as water in the present experiment is concerned, an enhancement of orders of magnitude from what expected from SER (19–21, 23, 24). These decouplings of the transport coef-ficients, observed as a SER violation, have been attributed to the occurrence of dynamical heterogeneities in structural glass form-ers (19, 21, 25, 26). Thus, in supercooled liquids there exist regions of varying dynamics, i.e., fluctuations that dominate their transport properties near the GT. The extent of such decouplings may depend on the material and the microscopic details of the specific transport parameters.

The SER breakdown can be described by using a scaling

concept, in particular, the law D⬇␶⫺␰, where ␰ ⫽ ␣(T)兾␤(T),

with␣ and ␤ being temperature-dependent scaling exponents of

D and␶, respectively (27). Recently, it has been shown that, for

Tris-naphthylbenzene (a fragile glass former), ␰ ⫽ 0.77 (21),

whereas an MD simulation of Lennard–Jones binary mixture has

given␰ ⫽ 0.75 (28). By using such an approach, we will discuss

our SER results for confined supercooled water. Fig. 3 shows the D vs.具␶T典 plot in a log–log scale. The red dots represent data

corresponding to temperatures above TL, where water behaves as

a fragile glass former, and the blue dots pertain to the strong Arrhenius region. As it can be observed, the data clearly show two different scaling behaviors above and below the FSC

tem-perature, in particular␰ ⯝ 0.74 on the fragile side (dotted line)

and ⬇2兾3 on the strong side (solid line). The dashed line

represents the situation in which SER holds, D ⬇␶⫺1. These

results agree with those obtained in Tris-naphthylbenzene (21) and, more specifically, with those of a recent theoretical study in which the decoupling of transport coefficients in supercooled liquids was investigated by using two class of models, one describing diffusion in a strong glass former, and the other in a fragile one (27). The main result of this study is that, while in the fragile case, the SER violation is weakly dependent on the

dimensionality d, with␰ ⫽ 0.73, in the strong case the violation

is sensitive to d, going as D⬇␶⫺2/3for d⫽ 1, and as D ⬇␶⫺0.95

for d⫽ 3. On considering the geometry of the system that we

have used in our experiment to confine water (1D cylindrical tubes, with a length of some micrometers and pore diameters of 14 and 18 Å), the scaling plot shown in Fig. 3 compares

remarkably well with the findings of the theoretical investigation (27) on both fragile and strong sides.

In summary, we explore dynamical properties of water in a deeply supercooled regime (well inside the ‘‘no-man’s land’’) by means of NMR and QENS experiments, which separately give a conclusive proof of the existence of a FSC. This finding supports the hypothesis that liquid water is consisting of a mixture of two different liquid structures (the LDL and HDL phases). Accord-ingly, a liquid–liquid phase separation line exists in the P–T plane with a liquid–liquid critical point as its end point. Remarkably, we give experimental proof of the existence of a violation of SER above and below the FSC in water, i.e., in both the fragile and strong regimes of supercooled water. This result clearly reflects the decoupling of transport coefficients of the liquid when

temperature is lowered toward Tg. This latter result certainly

constitutes an element that serves to clarify one of the most intriguing properties of water.

Methods

Water was confined in micellar-templated mesoporous silica matrices MCM-41-S, which have 1D cylindrical pores with a length of some micrometers arranged in 2D hexagonal arrays, synthesized after a similar method for the previous synthesis of MCM-48-S (29). The MCM-41-S materials are the same as those used in the QENS study of confined water (17). Pore size was determined by using a nitrogen absorption–desorption

tech-nique (16, 17). Investigated samples have hydration levels of H⯝

0.5 (0.5 g of H2O per g of MCM-41-S), obtained by exposing dry

powder samples to water vapor at room temperature in a closed chamber. This water-confining system can be regarded as one of the most suitable adsorbent models currently available (30, 31). High-resolution QENS spectroscopy method was used to

determine the temperature dependence of 具␶T典 for confined

water. Because neutrons can easily penetrate the wall of sample cell and because they are predominantly scattered by hydrogen atoms in water, rather than by the matrices containing them, incoherent QENS is an appropriate tool for our study. Using two separate high-resolution QENS spectrometers, we were able to measure the translational-relaxation time from 0.2 to 10,000 ps over the whole temperature range under study.

Fig. 3. The scaling plot in a log–log scale of D vs.具␶T典. Red dots are data corresponding to temperatures above TL, i.e., when water is in the fragile glass phase, whereas blues dots correspond to the strong Arrhenius region. Two different scaling behaviors exist above and below the temperature of the FST. In the fragile region, the scaling exponent is␰ ⯝ 0.74 (dotted line) and ⬇2兾3 in the strong side (solid line). Dashed line represents the situation in which the SER holds, D ⬇ ␶⫺1.

APPLIED

PHYSICAL

SCIENCES

SEE

(5)

The experiments were performed at both the High-Flux Backscattering (HFBS) and the Disk-Chopper Time-of-Flight (DCS) spectrometers in the National Institute of Standards and Technology Center for Neutron Research (NIST NCNR). The two spectrometers have two widely different dynamic ranges (for the chosen experimental setup), one with an energy resolution

of 0.8␮eV (HFBS) and a dynamic range of ⫾11 ␮eV (32), and

the other with an energy resolution of 20 ␮eV (DCS) and a

dynamic range of⫾0.5 meV (33) to be able to extract the broad

range of relaxation times from the measured spectra. The experiment was done at a series of temperatures, covering both below and above the transition temperature, and the data were analyzed by using Relaxing-Cage Model to extract the average

translational relaxation time具␶T典.

The NMR experiments on fully hydrated MCM-41-S samples with pore diameters of 18 and 14 Å were performed at ambient pressure by using a Bruker AVANCE NMR spectrometer, operating at 700-MHz proton resonance frequency. The self-diffusion coefficient of water D was measured with the pulsed gradient spin–echo technique (PGSE) (34, 35) in the

tempera-ture range 190–298 K (with an accuracy of ⫾0.2 K). The T

dependence of the chemical shift of methanol was used as a T standard. All details about the NMR experiment and the sample properties are reported elsewhere (36). The reported D values were derived from the measured mean square displacement

具r2(t)典 of molecules diffusing along the NMR pulsed-field

gra-dients direction r, during the time interval t.

We thank Chun-Wan Yen (National Taiwan University) for preparing MCM-41-S samples. Technical supports in QENS measurements from E. Mamontov and J. R. D. Copley (National Institute of Standards and Technology Center for Neuron Research) are greatly appreciated. The research at MIT is supported by Department of Energy Grants DE-FG02-90ER45429 and 2113-MIT-DOE-591. The research in Messina is supported by the MURST-PRIN2004. The research at National Taiwan University is supported by the National Science Council of Taiwan. This work used facilities supported in part by the National Science Foundation under Agreement No. DMR-0086210 and facilities of the Messina SCM-HR-NMR Center of Consiglio Nazionale delle Ricerche– L’Istituto Nazionale per la Fisica della Materia. We benefited from affiliation with European Union Marie-Curie Research and Training Network on Arrested Matter.

1. Angell, C. A. (1982) in Water: A Comprehensive Treatise, ed. Franks, F. (Plenum, New York), Vol. 7, pp. 1–81.

2. Debenedetti, P. G. & Stanley, H. E. (2003) Phys. Today 56, 40–46. 3. Mishima, O. & Stanley, H. E. (1998) Nature 396, 329–335.

4. Velikov, V., Borick, S. & Angell, C. A. (2001) Science 294, 2335–2338. 5. Sokolov, A. P. (1996) Science 273, 1675–1676.

6. Go¨tze, W. & Sjo¨lander, L. (1992) Rep. Prog. Phys. 55, 241–376.

7. Angell, C. A. (1997) in Complex Behavior of Glassy Systems, eds. Rubı´, M. & Pe´res-Vicente, C. (Springer, Berlin).

8. Angell, C. A. (1995) Science 267, 1924–1935.

9. Poole, P. H., Sciortino, F., Essmann, U. & Stanley, H. E. (1992) Nature 360, 324–328.

10. Ito, K., Moynihan, C. T. & Angell, C. A. (1999) Nature 398, 492–495. 11. Soper, A. K. & Ricci, M. A. (2000) Phys. Rev. Lett. 84, 2881–2884. 12. Bergman, R. & Swenson, J. (2000) Nature 403, 283–286. 13. Koga, K., Tanaka, H. & Zeng, X. C. (2000) Nature 408, 564–567. 14. Webber, B. & Dore, J. (2004) J. Phys. Condens. Matt. 16, S5449–S5470. 15. Dore, J. (2000) Chem. Phys. 258, 327–347.

16. Faraone, A., Liu, L., Mou, C.-Y., Yen, C.-W. & Chen, S.-H. (2004) J. Chem.

Phys. 121, 10843–10846.

17. Liu, L., Chen, S.-H., Faraone, A., Yen, C.-W. & Mou, C.-Y. (2005) Phys. Rev.

Lett. 95, 117802-1–117802-4.

18. Xu, L. Kumar, P., S. Buldyrev, V., Chen, S.-H., Poole, P. H., Sciortino, F. & Stanley, H. E. (2005) Proc. Natl. Acad. Sci. USA 102, 16558–16562.

19. Ediger, M. D. (2000) Annu. Rev. Phys. Chem. 51, 99–128.

20. Fujara, F., Geil, B., Sillescu, H. & Fleishcer, G. (1992) Z. Phys. B 88, 195–204. 21. Swallen, S. F., Bonvallet, P. A., McMahon, R. J. & Ediger, M. D. (2003) Phys.

Rev. Lett. 90, 015901-1–015901-4.

22. Pan, A. C., Garrahan, J. P. & Chandler, D. (2005) Chem. Phys. Chem. 6, 1783–1785.

23. Chang, J. & Sillescu, H. (1997) J. Phys. Chem. B 101, 8794–8801. 24. Cicerone, M. T. & Ediger, M. D. (1996) J. Chem. Phys. 104, 7210–7218. 25. Xia, X. Y. & Wolynes, P. G. (2001) J. Phys. Chem. B 105, 6570–6573. 26. Ngai, K. L., Magill, J. H. & Plazek, D. J. (2000) J. Chem. Phys. 112 1887–1892. 27. Jung, Y.-J., Garrahan, J. P. & Chandler, D. (2004) Phys. Rev. E 69, 061205-1–

061205-7.

28. Yamamoto, R. & Onuki, A. (1998) Phys. Rev. Lett. 81, 4915–4918. 29. Shih, P.-C., Lin, H.-P. & Mou, C.-Y. (2003) Stud. Surf. Sci. Catal. 146, 557–560. 30. Schreiber, A., Ketelsen, I. & Findenegg, G. H. (2001) Phys. Chem. Chem. Phys.

3,1185–1195.

31. Morishige, K. & Nobuoka, K. (1997) J. Chem. Phys. 107, 6965–6969. 32. Meyer, A., Dimeo, R. M., Gehring, P. M. & Neumann, D. A. (2003) Rev. Sci.

Instrum. 74, 2759–2777.

33. Copley, J. R. D. & Cook, J. C. (2003) Chem. Phys. 292, 477–485. 34. Stejskal, E. O. & Tanner, J. E. (1965) J. Chem. Phys. 42, 288–292. 35. Price, W. S. (1998) Concepts Mag. Reson. 10, 197–237.

36. Mallamace, F., Broccio, M., Corsaro, C., Faraone, A., Wanderlingh, U., Liu, L., Mou, C.-Y. & Chen, S.-H. (2006) J. Chem. Phys. 124, 161102-1–161102-4.

數據

Fig. 1. The temperature dependence of the inverse of self-diffusion coefficient of water and its average translational relaxation time
Fig. 1A shows a log-linear plot of the inverse of the self- self-diffusion coefficient of water 1 兾D measured by NMR as a function of 1 兾T for the fully hydrated MCM-41-S samples with pore diameters of 14 and 18 Å
Fig. 3. The scaling plot in a log–log scale of D vs. 具␶ T 典. Red dots are data corresponding to temperatures above T L , i.e., when water is in the fragile glass phase, whereas blues dots correspond to the strong Arrhenius region

參考文獻

相關文件

The first row shows the eyespot with white inner ring, black middle ring, and yellow outer ring in Bicyclus anynana.. The second row provides the eyespot with black inner ring

A cylindrical glass of radius r and height L is filled with water and then tilted until the water remaining in the glass exactly covers its base.. (a) Determine a way to “slice”

You are given the wavelength and total energy of a light pulse and asked to find the number of photons it

好了既然 Z[x] 中的 ideal 不一定是 principle ideal 那麼我們就不能學 Proposition 7.2.11 的方法得到 Z[x] 中的 irreducible element 就是 prime element 了..

Wang, Solving pseudomonotone variational inequalities and pseudocon- vex optimization problems using the projection neural network, IEEE Transactions on Neural Networks 17

volume suppressed mass: (TeV) 2 /M P ∼ 10 −4 eV → mm range can be experimentally tested for any number of extra dimensions - Light U(1) gauge bosons: no derivative couplings. =>

For pedagogical purposes, let us start consideration from a simple one-dimensional (1D) system, where electrons are confined to a chain parallel to the x axis. As it is well known

Define instead the imaginary.. potential, magnetic field, lattice…) Dirac-BdG Hamiltonian:. with small, and matrix