Hostname: page-component-76fb5796d-skm99 Total loading time: 0 Render date: 2024-04-28T04:22:31.364Z Has data issue: false hasContentIssue false

Glacier Sliding Down an Inclined Wavy Bed With Friction

Published online by Cambridge University Press:  30 January 2017

L. W. Morland*
Affiliation:
School of Mathematics and Physics, University of East Anglia, Norwich NR 47TJ, England
Rights & Permissions [Opens in a new window]

Abstract

The effects of frictional tangential traction combined with regelation on the basal sliding of a temperate glacier down an inclined wavy bed are examined. Two friction models are treated. First, a Coulomb law model having the assumptions that sliding occurs everywhere and that the tangential traction is proportional to the normal pressure. Secondly, a velocity power law in which the tangential traction is proportional to a power of the slip velocity. The ice motion is approximated by steady slow Newtonian flow and the bed undulation about a mean bed-line has a maximum slope ∊ ≪I. Flow solutions are constructed as perturbations (in powers off ∈) of the plane laminar flow corresponding to non-slip on the mean bed-line assuming that the ice remains everywhere in contact with the bed; that is, no cavitation takes place. If the normal traction is predicted to be tensile over part of the bed, implying that cavitation has occurred, then a new solution is needed in which the ice base over cavities is traction-free. Since the cavity sections and profile of the free ice base are then part of the overall solution, an intricate mixed boundary-value problem is set up for the flow and the present analysis is inadequate.

For a sinusoidal bed the perfect-slip (zero tangential traction) solution predicts compressive normal traction everywhere on the bed provided that the mean bed-line inclination α (to the horizontal) is less than a critical value αe which is of order ε. For greater values of α including a range of order ∊, the normal traction is tensile on some parts of the bed, and a solution with cavitation is needed. If the tensile sections are relatively small it is expected that the resulting cavitation will not change the overall solution significantly. However, the Coulomb friction solution has extensive zones of tensile traction for all values of α, so that extensive cavitation would occur. In contrast, the velocity-power friction solution has compressive traction everywhere on the bed for α ⩽αe=0(I) provided that the ice depth is not too large, and also for deep glaciers for α ⩽αe=O∊ Furthermore, the predicted basal sliding velocity varies much less with the length scale of the bed undulation than in the perfect-slip solution, and is smaller.

Résumé

Résumé

On a examiné les effets de la traction tangentielle de frottement combinée avec le regel sur le lit d'un glacier tempéré glissant sur un fond incliné avec des ondulations. Deux modèles de frottement sont envisaés le premier est une loi de Coulomb dans laquelle la traction tangentielle est proportionnelle à la pression normale, le second est une loi-puissance en fonction de la vitesse dans laquelle la traction tangentielle est proportionnelle à une puissance de la vitesse de glissement. Le mouvement de la glace est assimilé à un écoulement stationnaire lent Newtonien et les ondulations du lit autour d'une direction moyenne, ont une pente maximum ∈ ≪ I. Les solutions pour l'écoulement sont construites comme des perturbations en puissances de ∈ autour de l'écoulement plan laminaire correspondant au non-glissement sur le lit plan moyen, avec l'hypothèse que la glace reste partout au contact du lit, c'est- à -dire qu'il n'y a pas cavitation. Si l'on prévoit que l'effort normal sur une partie du lit sera une traction, ce qui implique que la cavitation peut se produire, alors il faut une nouvelle solution dans laquelle la glace de la base au-dessus des cavéuts n'est pas soumise à une traction. Dès lors que les profil et sections des cavités de la glace libre de la base font partie de la solution générale un problème inextricable de valeurs aux limites est soulevé par la détermination de l'écoulement et la présente analyse est inadéquate.

Pour un lit sinusoïdal, la solution du glissement parfait (traction tangentielle nulle) prévoit un effort normal de compression partout sur le lit pourvu que la pente moyenne ∊∊ (sur l'horizontale) soit inférieure à une valeur critique αc qui est de l'ordre de ∈. Pour des valeurs supérieures de α, y compris un ordre de grandeur voisin de ∈, il y a un effort de traction normal quelque part sur le lit, et il est nécessaire de chercher une solution avec cavitation. Si les sections avec traction sont relativement petites, on s'attend à ce que la cavitation résultante ne change pas la solution générale de manière significative. Cependant, la solution de la friction de Coulomb présente de larges zones avec efforts normaux de traction pour toutes les valeurs de α si bien qu'une cavitation importante peut s'y produire. Par contre, la solution avec frottement en puissance de la vitesse est valable partout si α ⩽ αc avec une valeur αc=O (I). pourvu que la profondeur du glacier ne soit pas excessive, et aussi pour les glaciers profonds lorsque α ⩽ αc avec αc = O(∈). En conséquence, la vitesse de glissement àl a base qui est prédite, varie beaucoup moins avec léchelle des longueurs d'onde des irrégularités; du lit que dans la solution à glissement parfait et est plus faible.

Zusammenfassung

Zusammenfassung

Untersucht wird die Auswirkung tangentialen Reibungszuges in Kombination mit Regelation auf das Gleiten eines temperierten Gletschers über ein geneigtes und gewelltes Bett. Zwei Reibungsmodelle werden herangezogen: erstens ein Coulomb-Gesetz, bei dem die tangentiale Reibung proportional zum Normaldruck anwächst; zweitens ein Geschwindigkeitspotenzgesetz, bei dem die tangentiale Reibung proportional zu einer Potenz der Gleitgcsch-wtndigkeit ist.Die Eisbewegung wird durch stationäres, langsames Newtonsches Fliessen angenähert; die Bettundulation um ein mittleres Bettprofil hat eine Maximalneigung von ∈ ≪ I. Die Lösungen für das Stömungsfeld werden als Störungen der ebenen laminaren Strömung, bei der kein Gleiten auf dem mittleren Bettprofil stattfindet, nach Potenzen von ∈ entwickelt; dabei wird angenommen, dass das Eis überall mit dem Bett in Berührung bleibt, d.h. keine Kavitation stattfindet. Für den Fall, dass sich eine Normalkomponcnte der Kraft über einem Teil des Bettes in Zugrichtung ergibt, braucht man eine neue Lösung, die Kavitation einschliesst, und bei der die Eisuntergrenze über den Hohläumen reibungsfrei ist. Da die Hohlraum-abschnittc und das Profil der freien Eisuntergrenze dann einen Teil der Gesamtlösung bilden, erhält man für das Fitessen ein verwickeltes vermischtes Randwertproblem und die vorliegende Untersuchung ist unzulänglich.

Für ein sinusförmiges Bettprofil liefert die Lösung für vollkommenes Gleiten (ohne tangcntialc Reibung) überall im Ben eine Normalkrai'i in Druekrichtung, vorausgesetzt, die Neigung $ des mittleren Bettprofils gegenüber der Waagrechten ist kleiner als ein kritischer Wert αc. (in der Grösse von ∈). Fur höhere Werte Von α, einschliesslich eines Bereiches in der Grösse von ∈, ist die Normalkrafi an einigen Stellen des Bettes in Zugrichtung, und man braucht eine Lösung mil Kavitation. Wenn die Abschnitte unter Zug verhalt-nisässig klein sind, wird angenommen, dass die entstehende Kavitation die Gesamtösung nicht entscheidend verädert. Die Lösung für die Coulomb-Reibung hat jedoch für alle Werte von a ausgedehnte Zonen unter Zug, so dass entspechende ausgedehnte Kavitation auftreten wird. In Gegensatz dazu gilt die Lösung mit Reibung nach dem Geschwindigkeitspotenzgesetz überall, wenn α ⩽ αc = 0(I), vorausgesetzt, dass die Eisdicke nicht zu gross ist, und auch für mächtige Gletscher, wenn α ⩽ αc = 0(∈). Des weiteren schwankt die vorausgesagte Gleiigeschwindigkeit weit weniger mit der Längenausdehnung der Bettundulation als in der Lösung des vollkommenen Gleitens und ist niedriger.

Type
Research Article
Copyright
Copyright © International Glaciological Society 1976

I. Introduction

A thin water layer produced by the melting and refreezing of basal ice in a temperate glacier on up- and down-stream faces of bed protuberances provides lubrication for the basal sliding which may be a significant part of the overall motion. The shear stress in such a thin layer is negligible and so there is no resistive tangential traction to the ice motion over the bed. The bed drag is the resultant, along the mean bed line, of the pressure distribution over the protuberances. This perfect-slip model is the basis of flow solutions obtained by Reference NyeNye(1969,1970), Reference KambKamb (1970), and more recently, Reference MorlandMorland (1976). It is assumed that the ice can be approximated as an incompressible Newtonian fluid of high viscosity in slow steady flow, and that the bed profile is periodic with small maximum slope ∊∊ relative to the bed line.Solutions are obtained as power series expansions in ∊∊ assuming that the ice base remains everywhere in contact with the bed. For a given glacier depth h an inclination ∊∊ of the mean bed line to the horizontal, and a profile shape, the plane flow solution determines the basal-sliding velocity, which is defined as the tangential velocity along the mean bed line.A calculation for a sinusoidal bed shows that the basal-sliding velocity is sensitive to the length scale of the bed undulation. Furthermore, the normal traction on the bed remains compressive everywhere only if ∊∊ for some critical angle ∊∊ which is of order ∊∊, and hence the solution predicts tensile tractions on part of the bed for ∊∊ of the order of one and for some range of ∊∊ of order ∊∊. In these situations cavitation must occur, and a valid formulation must incorporate cavity sections over which the ice base is traction free, these sections and the profile of the ice base being part of the solution. This intricate problem involving, as it does, such mixed boundary values, has not yet been attempted. It is, however, expected that when the tensile sections are relatively small, the resulting cavitation will not have a significant effect on the overall solution.

The condition of zero tangential traction requires the existence of a continuous water layer. Pinching-out of the water layer at the crests of protuberances will cause local failure of this condition, but debris protruding from the basal ice makes frictional contact with the bed across the large areas over which it is carried (Reference Boulton and CoatesBoulton, 1974, p. 41;1975, p. 7). The interaction of such debris with the bed is complex, as is its dependence on bed profile (which reveals how it is transported and deposited). Data obtained by G. S. Boulton, A. Armstrong and E. M. Morris from field work carried out on the Glacier d'Argentière in 1973 and 1975 are being analysed in an attempt to construct "simple" friction laws exhibiting the main features observed in different situations. To ensure a manageable analysis one must assume that the mean effect of the individual debris contacts can be described by friction laws which apply continuously over the bed surface. It is with such an analysis in mind that any possible qualitative effects of bed friction on basal sliding are explored here, by solving the plane flow problem for two conventional friction laws.

First, the Coulomb law

(1)

where ts and t n are the tangential.and normal tractions on the ice base, with s defining a tangential coordinate in the direction of flow and n a normal coordinate directed into the ice,v is the friction coefficient assumed to be of order unity. The limiting friction form of Equation (1) presupposes that sliding occurs everywhere. It is reasonable to assume that any non-slip zones will be of limited extent and should not influence significantly the overall basal sliding. Secondly, a velocity power law

(2)

where V s is the tangential velocity of the base ice, and E and m are constants. The form of

Equation (2) 2 has been used by Reference NyeNye (1959) in a different context, and is also inferred by Reference WeertmanWeertman (1957,1964) as a global relation between mean drag and basal-sliding velocity in his regelation theory which takes a non-linear viscous law for the ice, Weertman suggested that m is approximately 2 or a little larger, but gave no explicit value for E. In the present analysis, however, Equation (2) 2 is considered as a possible qualitative relation arising from debris friction, in contrast with Equation (I), and also in an exploration of its effects on basal-sliding velocity and cavitation. A value of m ∊∊ 2 is used in the calculations, and a range of values for E compatible with steady flow down the inclined bed is considered.

If part of the work done by the basal friction is released as a surface heat flux, and not used solely for surface crushing or other mechanical effects, it contributes to the thermal balance in the regelation mechanism and a non-linear thenno-mechanical coupling is introduced into the boundary conditions. When only a sma∊∊ part is released as heat this coupling may be neglected, and it is also shown here that, under moderate restrictions on the glacier depth and bed profile slope, the entire work contribution is small compared with the latent-heat terms. The solutions are therefore derived neglecting any friction contribution to the thermal balance, so that the surface distribution of heat sources at the ice base per unit area, is LV n (Reference NyeNye, 1969), where L = 2.8 × 108 J m -3 is tin latent heat and V n is the normal velocity of the basal ice. Thus, at the bed

(3)

where T and S denote temperature in the ice and bedrock respectively, and k i and k b are the

thermal conductivities of ice and bedrock respectively, k i ∊∊ 2.1 J m-1 K-1 and k b = rk i

where r ranges from 1 to 2 for typical bed rocks; the value r = 1.6 (appropriate to granite) is

used in later calculations. In the regelation model the basal ice is everywhere al its melting

point, and if (P o, T o) is a pressure-melting point on the bed-line, then at the bed

(4)

Such a linear relation presupposes that T and S remain close to T 0 at the bed, and C — 0.7 × 107 K/(N m2). Following Reference KambKamb (1970) the firsl-order dependence is on pressure p only with no contribution from deviatoric stress changes.Figure I shows the overall plane-flow problem, with coordinates (x,y) respectively along and normal to a bed line which is inclined at angle ∊∊ to the horizontal. U s is the surface velocity at y = h, in the x-direction, and the basal-sliding velocity U b is defined as the leading-order term of the x-velocity in a flow continued onto y — o. The bed profile is

(5)

Where f(x) is smooth and extends periodically as x ∊∊ It is assumed that the bed slope is everywhere small, thus

(6)

Fig. 1. Glacier flow over a wavy bed.

The bed conditions (Equations (I) or (2)) and Equations (3) and (4) are applied y ∊∊ yb(x).

On the upper free surface the stress ∊∊ satisfies

(7)

where p a is atmospheric pressure, and where it is anticipated that the normal velocity Vy on y = h is zero to the required order in ∊∊. The exponential decay of Vy with height demonstrated in the solutions justifies the surface prescription y ∊∊h under a very weak restriction on h. A geothermal heat flux Q normal to the bed line, with a typical value of 4 × 10-2 J m-2 s-1 (Reference PatersonPaterson, 1969) is included, requiring

(8)

in the half-plane solutions determined for T and S.

2.Flow Equations

The construction of velocity and temperature fields, in terms of complex potentials, and the development of boundary-condition expansions in powers of ∊∊, have been described by Reference MorlandMorland (1976). Equations from this earlier paper are prefixed by the letter "M", only the main steps are repeated here.

A length scale of the bed undulation λ is defined by

(9)

and dimensionless coordinates (X, Y) are defined by

(10)

x = ∊∊X, y = ∊∊Y,

in which the bed profile becomes

(11)

The following velocity, pressure and temperature decompositions are formulated:

(12)

(13)

(14)

(15)

A value for ∊∊ of 3 × 1012 N m-2 s is used in the calculations. For given values of h,Equation (15) 15 determines Us once k is known. The velocities u and v are dimensionless with a scale unit Us P is a dimensionless pressure with unit 2μU s/λ (∑ will denote a dimensionless stress tensor with the same unit), ∊∊ and ∊∊ dimensionless temperatures with unit ∊∊ The free surface conditions of Equation (7) 7 and the (flux conditions of Equation (8) 8 are satisfied exactly if

(16)

The condition of momentum balance is satisfied for slow viscous flow if

(17)

so that P, u and v are the pressure and velocity fields for viscosity 0.5 in the absence of any body force (Reference LangloisLanglois, 1964). Vz is the two-dimensional Laplacian in (X,Y) coordinates. Steady heat conduction in the ice and bed respectively requires that

(18)

if we neglect the motion of the ice (Reference KambKamb, 1970). Solutions of the biharmonic relations (Equation (17) 17) and the harmonic relations (Equations (18)) can be expressed in terms of analytic functions of a complex variable z ∊∊ X∊∊ir (equations (M43) to (M47)). Let the

values of u, v,∊∊ and ∊∊ on r = o (continued analytically where r = o lies outside the domain) be denoted by

(19)

and also assume that U, V, ∊∊ and ∊∊ vanish or behave sinusoidally as X ∊∊ 00. Then, assuming that the complex potentials vanish at infinity in a way consistent with Equation (16) 16, the potentials may be represented as Cauchy integrals U, V, ∊∊ and ∊∊(equations (M50)-(M53)).The boundary values at Yr b which arise in the bed conditions (Equations (I)-(4)), can be approximated by a truncated Taylor scries in r about r = o. The series appears as a power expansion in ∊∊ by virtue of Equation (II). Evaluation of the various quantities and derivatives on r = o is made from equations (M50)-(M53), these involve Hubert transforms Erdélyi 1954) If W(t) is continuous and vanishes or behaves sinusoidally at infinity, then its Hubert transform is

(20)

Where ∊∊ denotes the Cauchy principal value. Also

(21)

the latter being the inversion theorem. The useful results when F(x) is a truncated Fourier series are

(22)

the constant function does not satisfy the inversion theorem.

Now the bed conditions represented by Equations (3) and (4) are given by equations (M65), (M63) and (M62), respectively. Neglecting terms of order ∊∊2 (U, V,∊∊, ∊∊)), these are

(23)

(24)

(25)

The argument X is omitted for brevity. Λ natural length ∊∊ and later a length ∊∊ occur, these are defined by

(26)

with previous values, and the ratios

(27)

are introduced for convenience. The constants B and D in Equations (24) and (25) are given by

(28)

The bed conditions given by Equations (I) and (2) require expansions for the dimension-less tractions Σs and Σn, and the dimensionless tangential velocity V 8/U8 To the same approximation as Equations (23)-(25), these expansions are:

(29)

(30)

Where

(31)

(32)

(33)

It is supposed that

(34)

So that Equation (16) 16 is satisfied to any order in ∊∊ with the exponential decay of u, v and P in r. Also

(35)

In order to balance the boundary conditions in powers of ∊∊, let

(36)

(37)

A determination of the coefficients ∊∊∊∊ gives k in terms of ∊∊ and, in turn, U s and U b It is expected that k ∊∊ O(i)so y o = o, but this will follow from the balance of the boundary conditions.

Finally, if a proportion j of he work done against basal friction is released as heat, then in Equation (3) 3 V n is replaced by F"" ∊∊ For the Coulomb law (Equation (I)) t s is of order pgh (and possibly also for the velocity power law (Equation (2) 2) when ∊∊ is of order I)

and generally (Vn/Vs ) is of order ∊∊, so neglect of the coupling term requires

(38)

This is satisfied for small j, or j approximately one, provided h is moderate. In fact, for Equation

(2) the condition for no cavitation when ∊∊ is of order one also requires that h be not too

large. When ∊∊ is of order ∊∊ and non-coupling only requires∊∊;

3.Coulomb Law

The bed conditions are given by Equations (23) to (25) together with Equation (1) 1 written in the form

(39)

where Σ3 and Σn are given by Equations (29) and (30). First, consider

(40)

and suppose ∊∊ is of order I in the series balance. Values of ∊∊ which are smaller or larger may be treated directly, or obtained as limits to the present solution.

The term in ∊∊ gives immediately from Equation (39) 39

(41)

and the ∈0 terms give

(42)

setting constant terms to zero in order to obtain the required behaviour at infinity. Now the ∊∊ terms of Equations (23) to (25) give

(43)

leading to

(44)

A search for a periodic solution without constant term,

(45)

and the use of the results of Equation (22) 22 shows that both αn and bn are zero, so

(46)

Similarly, the ∊ terms of Equations (39) show that

(47)

and Equations (24) and (25) show that

(48)

Using these results in the ∊ balance of Equation (23) 23 gives

(49)

where

(50)

It was shown in Reference MorlandMorland (1976) that A has a maximum value of 0.05 for extreme values of the glacier parameters, so contributions to the geothermal heat flux which arise solely from this coefficient are small. The bounded complementary solution of Equation (49) 49 is zero, analogous to Equation (44) 44. The determination of a particular periodic integral needs F to be specified. As an illustration of this consider a sinusoidal bed whose shape is described by the equation

(51)

The solution of Equation (49) 49 is now

(52)

(53)

As v tends to zero, U1 tends also to zero, and

(54)

thus we recover the solution for perfect slip (Reference MorlandMorland, 1976).

The leading velocity terms UI and VI , are now determined but, in contrast with the perfect slip solution, y2 = o and so k, Us and Ub are still unknown. From the «2 terms of Equation (39) 39

(55)

Thus, the second-order velocity coefficient U2 is given by the balance of periodic terms whereas ßly3 is equal to the constant term of

(56)

For the profile represented by Equation (51) 51

(57)

which, since κ arid £, are positive, implies thai [a is less than zero. This contradictory result stems from the application of Equation (39) 39 to the regions of negative pressure (Σ" > o) given in this solution, when the friction (Σ, > o) is directed up-glacier. From Equation (30) 30 the leading normal pressure term is

(58)

which oscillates equally between positive and negative values. Thus there is no balance without cavitation for Equations (39) and (40).

Now, a is of order one as are both β and β 0. If we assume that vβo is not equal to one, then the ∊∊0terms of Equation (39) 39 give

(59)

and so Equations (23) to (25) lead to Equations (43) to (46) again. Similarly, the ∊∊ terms give

(60)

and also Equations (48) to (54), while the «r1 terms of Equation (39) 39 give Equation (56) 56 for a left-hand side of y2( β ov I).Thus

(61)

which tends to (Ik) ∊∊2/(I + ∊∊2)) as v tends to zero; we recover the perfect slip solution in this case also. Provided v is less than I /β then we are able to predict a positive surface velocity Us But, again, Equation (58) 58 determines the leading normal pressure term, showing that there is no balance without cavitation. The case where ∊∊ is equal to one allows a balance only if Us is zero.

In conclusion, the solutions constructed for the Coulomb law, assuming no cavitation, are invalid for all ranges of bed-line inclination ∊∊

4.Velocity Power Law

The law represented by Equation (2) 2 can be rewritten as

(62)

but we lack data for the physical constant E. However, if we suppose that tt is less than or equal to pgh sin « (a condition compatible with the gravity driving force) then

since K is of the order of one for Ub greater than zero. This is satisfied by the equation

(63)

where e is less than or equal to the order of one provided that k is of order one. If any solution with e of order one predicts a smaller k and hence a Vs of the order of Ub and Us , then larger values of ts ,, occur, presumably with compensating bed tensions as with the Coulomb law, so that cavitation occurs. However, the form of Equation (63) 63 includes all possible valid situations. By construction, the equation

(64)

is independent of the solution variable Us , but depends explicitly on ∊∊ and on h if m is not equal to 2. The perfect slip law ∑8 = 0 is given by the limit as e tends to zero.

Now Vs is of order of Ub , and if

then Equation (62) 62 approaches the perfect slip law. These conditions do not, in general, follow Reference MorlandMorland (1976), so we assume that

(65)

and hence from Equation (33) 33

(66)

With the restriction of Equation (34) 34 the ∊∊0 terms of Equation (63) 63 give

(67)

and then, the ∊∊0 terms of Equations (23) to (25) are identical with the perfect slip balance implying that

(68)

Next, the t terms of Equation (63) 63 give

(69)

One solution to these equations is

(70)

thus determining k,and Us independent of F(X) Again, the ∊∊ balance of Equations (23) to (25) is identical with the perfect slip solution, so

(71)

and, for the sinusoidal bed (Equation (51) 51)

(72)

In the limit as e tends to zero, Equation (69) 69 implies that ∊∊1 = O as in the perfect slip solution. The alternative solution of Equation (69) 69 is simply

(73)

and the perfect-slip results (Equations (71) and (72)) again follow. Now, the ∊∊2 balance in Equation (63) requires that

(74)

The product FH[V”1] is in general the sum of a constant I’ and a periodic term W X), as in the perfect slip solution, so

(75a)

(75b)

where a non-zero Γ implies that y2 , is not equal to zero. As

(which is equivalent to y1 , being non-zero), Equation (75) 75 approaches the result. Equation (70) 70). Thus, the first solution is given by taking small values for he2 in Equation (75b). For the sinusoidal bed (by equation (M88)).

(76)

and U2 and V2 are given by the perfect slip solutions (equations (M90) and (M94)).Thus the velocity perturbation is changed only by the scale factor Us , which changes with (Equation (8) 8). Now Equation (75b) becomes

(77)

When e = o, Equation (77) 77 reduces to the perfect slip result (equation (M89)). The basal-sliding velocity is given by

(78)

and, for fixed values of A and ?, as e increases k/(I —k) increases and so Ub , decreases. As expected, friction decreases the sliding velocity. Calculations have been made for m — 2. Figures 2 and 3 show the variation of Ub with ? for values for h∊∊2 of o. 1, 0.4. I, and 4, for e = 0.5 (Fig. 2) and e = I (Fig. 3), compared with Ub of the perfect-slip solution (e = o).

In the perfect-slip case the ratio U b∊∊2/h is independent of h∊∊2. For ∊∊.0.1, this range of h∊∊2 covers values of h running from 10-400 m. If e, ∊∊ and ? are fixed, we see that U b decreases as h decreases. Also, for fixed e, the variation of U b with ? (the length scale of the undulation) decreases as h decreases, and, in particular, U b is very insensitive to changes in ω provided h is small. (As the graph ordinates are proportional to U b/h (Figs 2 and 3), so these effects are more significant than the Figures appear to suggest. Figure 4 shows the variation of the ratio U b/U with ∊∊ for the values of he2 used in the previous Figures. This ratio decreases with increase of e for each values of he2 and, including the case where e = o (figure 3, Reference MorlandMorland (1976)). At fixed Ub/Us becomes less sensitive to a change in w as he2 decreases. At both fixed e and the ratio increases as h∊∊2 decreases.

Finally, the relation for normal bed pressure (Equation (30) 30) has leading terms (recall Equation (40) 40)

(79)

Now if X/h is of order ∊∊ Equation (77) 77 reduces to Equation (70) 70 with y1O, k = O(I), and,for such relatively thin glaciers. Equation (79) 79 shows that there is no cavitation if ∊∊ is of order ∊∊ or if ∊∊ is of order one and

For the sinusoidal bed the latter condition becomes

(80)

Fig. 2. Variation of basal sliding velocity Ub with undulation length λ (SI-units) for e — 0.5. The perfect-slip solution (e = ) is shown by the dotted line.

Fig. 3. Variation of baml sliding velocity Vb with undulation length A (SI-units ) for e = 1.0. The perfect-slip solution (e = 0) is shown by the dotted line.

Fig. 4. Variation of the ratio of basal-sliding velocity to surface velocity with undulation length λ for different values of h∈2(SI-units). These values are shown against the curves. The dashed lines are for e = 0 .5, the full lines are for e = 1.0.

For λ/λ of order ∊∊ zero and y 2 non-zero, Equation (79) 79 shows that Σn oscillates about zero to the order of ∊∊ when ∊∊ is of the order of one, so cavitation always occurs then. But for ∊∊ of the order of ∊∊, cavitation does not occur if ∊∊ is greater than or equal to (H[V’1]) max For the sinusoidal bed. neglecting the small coefficient A, the condition becomes

(81)

As with the perfect-slip result of Reference MorlandMorland (1976) (equation (M102)), there is a critical limit ∊∊ of the order $ which increased with the friction coefficient e through the factor k/(I — k)

Thus, the law represented by Equation (2) 2 allows non-cavitation solutions for "thin glaciers" for all ∊∊ less than or equal to ∊∊e (of order one), and for "thick glaciers" for ∊∊ less than or equal to ∊∊c (of order ∊∊).

5.Concluding Remarks

The flow solutions which have been established for both friction laws exhibit features which differ from each other and from the perfect-slip solution. With the Coulomb law. (Equation (I)), the normal bed traction becomes a tensile stress over finite sections of the profile however small the inclination ∊∊; this implies the onset of cavitation and the failure of the solution which assumes contact everywhere. In the perfect-slip solution the bed pressure remains positive everywhere provided that ∊∊ is less than or equal to ∊∊e. where ∊∊e is of the order of ∊∊, Since the mean tangential traction must be less than the mean down-plane gravity force, and by Equation (I) ts is of the same order as the normal pressure —tn in the mathematical solution, regions of large normal pressure are counterbalanced by regions of negative pressure accompanied by negative values of ts (the traction driving the glacier). Thus, the Coulomb law is possible only if a significant amount of cavitation takes place. In contrast, the power-velocity law (Equal (2)) gives a bed pressure which is everywhere positive if ∊∊ is less than or equal to ∊∊ (of order one), provided the glacier depth h is not too large, and if ∊∊ is less than or equal to ∊∊e (∊∊ee ∊∊) for deep glaciers. Furthermore, the predicted basal-sliding velocity for a sinusoidal bed is smaller and varies much less with the length scale of the undulation than it does in the perfect-slip solution, both the magnitude and variation decreasing as h decreases. These broad features of the different friction laws may be helpful in the construction of friction models from empirical data.

Acknowledgment

This investigation was pursued in connection with a Natural Environment Research Council Grant GR3/2680 "Flow of glaciers over deformable materials" held jointly with Dr G. S. Boulton in the School of Environmental Sciences, University of East Anglia.

References

Boulton, G.S. 1974 Processes and patterns of glacial erosion, (In Coates, D.R., ed. Glacial geomorphology. Binghamton. State University or New York. p. 4187. (Publica tions in Geomorphology)) Google Scholar
Boulton, G 1975 Processes and patterns of subglacial sedimenentation; a theoretical approach, (In Wright, A. E. Moseley, F ed. Ice ages : a1lcient alld modem Liverpool. Seel House Press. p. 742.) Google Scholar
Erdélyi, A. 1954 Tables of integral transforms. Vol. 2. Based, in part, on notes left by Harry Bateman. New York etc., McGraw–Hill Book Co., Inc. Google Scholar
Kamb, W.B. 1970 Sliding motion of glaciers: theory and observation. Reviews of Geophysics and Space Physics, Vol. 8, No. 4, p. 673728. Google Scholar
Langlois, W.E. 1964 Slow viscous flow. New York, Macmillan. Google Scholar
Morland, L.W. 1976. Glacier sliding down an inclined wavy bed. Journal of Glaciology. Vol. 17, No. 77. p. 44762. CrossRefGoogle Scholar
Nye, J.F. 1959 The motion of ice sheets and glaciers. Journal of Glaciology. Vol. 3, No. 26, p. 493507. Google Scholar
Nye, J.F. 1969 A calculation on the sliding of ice over a wavy surface using a Newtonian viscous approximation. Proceedings of the Royal Society of London, Ser. A, Vol. 311, No. 1506 p. 44567 Google Scholar
Nye, J.F. 1970 Glacier sliding without cavitation in a linear viscous approximation, Proceedings of the Royal Society of London, Ser. A. Vol. 315, No. 1522 p. 381403. Google Scholar
Paterson, W.S.B. 1969 The physics of glaciers. Oxford, Pergamon Press. (The Commonwealth and International Library. Geophysics Division.) Google Scholar
Weertman, J. 1957 On the sliding of glaciers. Journal of Glaciology, Vol. 3. No. 21, p. 3338. CrossRefGoogle Scholar
Weertman, J. 1964 The theory of glacier sliding. Journal of Glaciology, Vol. 5. No. 39. p. 287303 CrossRefGoogle Scholar
Figure 0

Fig. 1. Glacier flow over a wavy bed.

Figure 1

Fig. 2. Variation of basal sliding velocity Ub with undulation length λ (SI-units) for e — 0.5. The perfect-slip solution (e = ) is shown by the dotted line.

Figure 2

Fig. 3. Variation of baml sliding velocity Vb with undulation length A (SI-units ) for e = 1.0. The perfect-slip solution (e = 0) is shown by the dotted line.

Figure 3

Fig. 4. Variation of the ratio of basal-sliding velocity to surface velocity with undulation length λ for different values of h∈2(SI-units). These values are shown against the curves. The dashed lines are for e = 0 .5, the full lines are for e = 1.0.