Hostname: page-component-8448b6f56d-wq2xx Total loading time: 0 Render date: 2024-04-24T03:32:55.968Z Has data issue: false hasContentIssue false

On the pressure dependence of the rate factor in Glen’s flow law

Published online by Cambridge University Press:  10 July 2017

Ralf Greve
Affiliation:
Institute of Low Temperature Science, Hokkaido University, Sapporo, Japan E-mail: greve@lowtem.hokudai.ac.jp
Thomas Zwinger
Affiliation:
CSC – IT Center for Science, Espoo, Finland
Yongmei Gong
Affiliation:
Arctic Centre, University of Lapland, Rovaniemi, Finland
Rights & Permissions [Opens in a new window]

Abstract

Type
Correspondence
Copyright
Copyright © International Glaciological Society 2014

1. Introduction

Using the terminology and notation of Reference Greve and BlatterGreve and Blatter, (2009), Glen’s flow law, which describes secondary creep of polycrystalline glacier ice, reads

(1)

where D = sym grad v is the strain-rate or stretching tensor, v is the three-dimensional velocity vector, tD is the deviatoric part of the Cauchy stress tensor, T is the absolute temperature, p is the pressure that their exponential arguments are equal, is the effective stress, A (T, p)is the rate factor (softness) and fe) is the creep function. The latter is commonly expressed as a power law,

(2)

and for the stress exponent n most frequently the value n = 3 is used (e.g. Hutter,1983; Reference Cuffey and PatersonCuffey and Paterson, 2010; Reference Van der VeenVan der Veen, 2013, and references therein).

As for the rate factor, it is usually expressed in the form of an Arrhenius law,

(3)

where A 0 is the pre-exponential constant, Q is the activation energy, V is the activation volume and R = 8.314 J mol–1 K–1 is the universal gas constant.

It has often been stated (e.g. Reference RigsbyRigsby, 1958; Reference PatersonPaterson, 1994; Reference HookeHooke, 2005; Reference Greve and BlatterGreve and Blatter, 2009; Reference Cuffey and PatersonCuffey and Paterson, 2010) that the pressure dependence in Eqn (3) can be properly accounted for by dropping the activation volume V, while replacing the absolute temperature T by the temperature relative to pressure melting T ,

(4)

In this expression, T 0 = 273.15 K, and T m is the pressure-dependent melting temperature of ice,

(5)

where β = 7.42 × 10-8 K Pa-1 is the Clausius–Clapeyron constant for pure ice (Reference HookeHooke, 2005). Combining Eqns (4) and (5) yields

(6)

and the simplified rate factor with the dependence on T takes the form

(7)

2. Consistency of Computed and Measured Activation Volumes

We now investigate whether replacing the rate factor in the form of Eqn (3) by that in the simplified form of Eqn (7) is consistent with measured values of the activation volume V. We start by demanding that Eqns (3) and (7) are equal, so that their exponential arguments are equal,

(8)

Using Eqn (6), this is equivalent to

(9)

A short scaling analysis with β= 7.42 ×10-8 K Pa-1, p = 400 bar=4×107Pa (appropriate for the thickest ice in recent ice sheets) and T≈250 K gives

(10)

so that, to a very good approximation (relative error ˜ (βp/T)2 = O(10-4)), Eqn (9) simplifies to

(11)

Solving this for the activation volume V yields

(12)

We compute numerical values of the activation volume found in Eqn (12) for the temperature range 220–273 K that is relevant for ice in terrestrial ice sheets and glaciers. For the activation energy Q, we follow Reference Cuffey and PatersonCuffey and Paterson, (2010) and use Q = 6 × 104 J mol-1 for temperatures below 263 K, and Q = 1: 15 × 105 J mol-1 for temperatures above 263 K. This yields a range of activation volumes (units of m3 mol-1)

(13)

and

(14)

These results are illustrated in Figure 1.

Fig. 1. Computed (Eqn (12), with both values of the Clausius– Clapeyron constant for pure and air-saturated ice) and experimentally determined (Reference WeertmanWeertman, 1973; Reference Durham, Kirby and SternDurham and others, 1997; error margins also shown) values of the activation volume V as a function of the temperature T

Laboratory experiments on polycrystalline ice reported by Reference Durham, Kirby and SternDurham and others, (1997) for the temperature range 195–240 K gave a value for the activation volume V of (–1.3 ± 0.3) × 10-5 m3 mol-1. Reference WeertmanWeertman, (1973) analysed experiments for the temperature range –8.1˚C to –5.9˚C (265.05–267.25 K) and gave a value of (–2 ±0.5) × 10-5 m3 mol-1. Therefore, the computed values of Eqns (13) and (14) are ~50% larger than the experimentally determined values. When the computation is done with the larger value β = 9.8×10-8 K Pa-1 of the Clausius–Clapeyron constant for air-saturated rather than pure ice (Reference HookeHooke, 2005), this mismatch increases to about a factor 2 (see also Fig. 1). Nevertheless, taking into account the great difficulty in measuring this small effect, we consider this a reasonable agreement.

3. Conclusion

We have demonstrated that expressing the pressure dependence of the rate factor (Eqn (3)) of Glen’s flow law only by a dependence on the temperature relative to pressure melting is consistent, within the rather large uncertainties, with measured values of the activation volume. Our findings include the possibility that this very common practice may over-predict the impact of pressure on ice deformability to some degree. However, until more comprehensive experimental data become available that will allow a more robust assessment, we conclude that it is a valid assumption to employ the rate factor in the simplified form of Eqn (7).

Acknowledgements

We thank K. Hutter and H. Blatter for useful discussions. Constructive comments by the reviewers E. Bueler and K.M. Cuffey and the scientific editor, T.H. Jacka, helped to improve the manuscript. R.G. was supported by a Grant-in-Aid for Scientific Research A (No. 22244058) from the Japan Society for the Promotion of Science, JSPS. T.Z. and Y.G. acknowledge funding from the Nordic Centre of Excellence, SVALI

9 March 2014

References

Cuffey, KM and Paterson, WSB (2010) The physics of glaciers, 4th edn. Butterworth-Heinemann, Oxford Google Scholar
Durham, WB, Kirby, SH and Stern, LA (1997) Creep of water ices at planetary conditions: a compilation. J. Geophys. Res., 102(E7), 16 29316 302 (doi: 10.1029/97JE00916)Google Scholar
Greve, R and Blatter, H (2009) Dynamics of ice sheets and glaciers. Springer, Dordrecht CrossRefGoogle Scholar
Hooke, RLeB (2005) Principles of glacier mechanics, 2nd edn. Cambridge University Press, Cambridge CrossRefGoogle Scholar
Hutter, K (1983) Theoretical glaciology; material science of ice and the mechanics of glaciers and ice sheets. D. Reidel, Dordrecht/Terra Scientific, Tokyo Google Scholar
Paterson, WSB (1994) The physics of glaciers, 3rd edn. Elsevier, Oxford Google Scholar
Rigsby, GP (1958) Effect of hydrostatic pressure on velocity of shear deformation on single ice crystals. J. Glaciol., 3(24), 273278/271 Google Scholar
Van der Veen, CJ (2013) Fundamentals of glacier dynamics, 2nd edn. CRC Press, Boca Raton, FL Google Scholar
Weertman, J (1973) Creep of ice. In Whalley E, Jones SJ and Gold LW eds. Physics and chemistry of ice. Royal Society of Canada, Ottawa, Ont., 320337 Google Scholar
Figure 0

Fig. 1. Computed (Eqn (12), with both values of the Clausius– Clapeyron constant for pure and air-saturated ice) and experimentally determined (Weertman, 1973; Durham and others, 1997; error margins also shown) values of the activation volume V as a function of the temperature T