Hostname: page-component-8448b6f56d-gtxcr Total loading time: 0 Render date: 2024-04-19T19:41:20.271Z Has data issue: false hasContentIssue false

Comments on the use of chemically based mixing models in glacier hydrology

Published online by Cambridge University Press:  20 January 2017

Martin Sharp
Affiliation:
Department of Geography, University of Alberta, Edmonton, Alberta T6G 2H4, Canada
Giles H. Brown
Affiliation:
Centre for Glaciology, Institute of Earth Studies, University of Wales, Aberystwyth, Dyfed SY23 3DB, Wales
Martyn Tranter
Affiliation:
Department of Geography, University of Bristol, Bristol BS8 1SS, England
Ian C. Willis
Affiliation:
Department of Geography, University of Cambridge, Cambridge CB2 3EN, England
Bryn Hubbard
Affiliation:
Centre for Glaciology, Institute of Earth Sciences, University of Wales, Aberystwyth, Dyfed Sϒ23 3DB, England
Rights & Permissions [Opens in a new window]

Abstract

The assumptions involved in the use of chemically based mixing models for analysis of flow routing of meltwaters in glacierized basins are critically evaluated. The assumption that glacial drainage systems consist of only two primary flow components is arbitrary and must be supported by independent evidence. Recent studies of the processes by which meltwaters acquire solute indicate that the assumption that flow components have unique and constant chemical compositions is unlikely to be correct. Source-water composition and weathering potential will vary over the course of a melt season, and the extent of subglacial weathering is strongly dependent upon such factors as meltwater residence time and the availability of reactive sediment, both of which are known to vary on diurnal to seasonal time-scales. Mixing of flow components does not appear to be confined to the terminal regions of glaciers and is therefore unlikely to be conservative as assumed. A multi-parameter mixing model is applied to the analysis of data on the chemistry of waters sampled from boreholes drilled through Haut Glacier d’Arolla, Switzerland, to demonstrate the range of dissolved species for which the assumption of conservative mixing is violated. The consequences of this violation for quantitative hydrograph separation are shown to be highly significant. The utility of mixing models as a tool for the investigation of glacier hydrological systems is questionable and the results of previous studies are unreliable.

Type
Research Article
Copyright
Copyright © International Glaciological Society 1995

Introduction

A recurrent theme in research on glacier hydrology has been the use of measurements of meltwater quality to derive information about the character of subglacial and englacial drainage systems (Reference Collins,Collins, 1978, Reference Collins,1979; Reference Oerter,, Behrens,, Hibsch,, Rauert, and Stichler,Oerter and others, 1980; Reference Collins, and Young,Collins and Young, 1981; Reference Gurnell, and Fenn,Gurnell and Fenn, 1984; Reference Sharp,Sharp, 1991, Reference Tranter, and Raiswell,Tranter and itaiswell, 1991; Reference Lecce,Lecce, 1993). One technique which has been widely adopted in these studies and used to make quantitative inferences about meltwater fluxes through the different elements of such drainage systems is the chemically based mixing model (Reference Pinder, and Jones,Pinder and Jones, 1969). This technique is used to separate meltwater run-off hydrographs into components which allegedly arise from flow through two (or sometimes three) elements of the drainage system, The technique makes a number of assumptions about the nature of englacial/subglaciai drainage systems and the manner in which meltwaters acquire their solute load, most of which have not been tested in previous studies. Recent developments in our understanding of the nature and behaviour of glacier drainage systems (Reference Nienow,, Sharp, and Willis,Nienow and others, in press), and of chemical weathering processes in glacial environments (Reference Tranter,, G,, Raiswell,, Sharp, and Gurnell,Tranter and others, 1993), cast considerable doubt on the validity of these assumptions and thus on the application of the method. The purpose of this note is to draw attention to the nature of these assumptions, to present some of the data which challenge validity and to examine their implications for quantitative hydrograph separation. It is prompted by a recent application of the mixing-model method (Reference Lecce,Lecce, 1993), which, like previous studies, draws conclusions which we do not believe are sustainable in the light of the information presented.

Assumptions of Chemically Based Mixing Models

Reference Collins,(Collins 1978, Reference Collins,1979) argued that a simple mixing model of the form:

(1)

could be used to separate meltwater-discharge hydro-graphs into two flow components. Here Q refers to meltwater discharge and C to solute concentration. The subscripts t, s and e refer to the total flow, and to what Collins termed the “subglacial” and “englacial” components. Collins assumed that waters which pass more slowly through the system (“subglacial” waters) experience considerable enrichment as a result of contact with subglacial bedrock and finely comminuted rock debris. He also assumed that these two types of water mixed in a region close to the glacier snout, the englacial waters diluting the subglacial waters to produce bulk meltwaters with a chemical composition dependent upon the initial chemistries of the two components and their mixing ratio (the relative proportions of the two components in bulk waters). If the chemical composition of the components is known, the mixing-model equation can be rewritten to allow solution for the component discharges:

(2)

Models of this form make a number of assumptions which we now examine.

Two drainage components

An immediate problem concerns the definition of the flow components. Collins envisaged two principal drainage routeways through a glacier — a system of major channels which allows rapid transit and isolates waters from potential solute sources, and a system which permits much slower transit and brings waters into contact with solute sources. The first he defined as the “englacial” component, even though it was said to drain via a system which includes “arterial conduits located at the bed” (Reference Collins,Collins, 1979, p. 349), and the second as the “subglacial” component, which was said to drain via “channels which may be moraine-walled or cut in bedrock, and in which sediment transport occurs” (Reference Collins,Collins, 1979, p. 349). This use of the terms “subglacial” and “englacial” to describe the flow components is misleading, since both can apparently drain through channels located at the glacier bed. The actual basis for distinguishing between the two components is the residence time of meltwaters at the glacier bed rather than their location or morphology. In this sense, it would seem more logical to differentiate the components using terms such as delayed flow and quick flow (cf. Reference Tranter,, G,, Raiswell,, Sharp, and Gurnell,Tranter and others, 1993). Nevertheless, we retain Collins’ terminology for the purposes of the present discussion.

Collins’ definition of the hydrological pathways followed by the two components is vague, since icewalled conduits at the glacier bed will presumably also have bedrock or morainic floors, and it is not clear why, given the definitions presented, the two pathways would result in systematic variations in residence time within the glacier. Equally, whilst it is clear that Collins envisaged the source of “englacial” waters to be surface melting, it is not clear what source he envisaged for “subglacial” waters. The source would presumably be surface melt, internal/subglacial melt or ground water. Which of these is actually the source has major implications for flux magnitudes, patterns of flux variation in time and component chemistries. A more rigorous definition of the morphology of the hydrological pathways followed by the two flow components is required if the results of mixing-model calculations are to be interpreted. Theoretical analyses of glacier hydrology now identify at least seven possible components to suhglacial and englacial drainage systems. These include englacial conduits, subglacial water films (Reference Weertman,Weertman, 1972), R-channels incised upwards into the glacier sole (Reference Röthlisberger,Röthlisberger, 1972), Nye channels incised into bedrock (Reference Nye,Nye, 1973), linked-cavity systems (Reference Walder,Walder, 1986), permeable subglacial sediments (Reference Shoemaker,Shoemaker, 1986) and broad canals incised into the surface of such sediments (Reference Walder, and Fowler,Walder and Fowler, 1994). Given this diversity, it seems arbitrary and unreasonable simply to assume that the drainage system of a given glacier contains only two (or perhaps three) primary components. There is clearly a need for applications of the mixing-model technique to be supported by independent evidence (such as is provided by intensive dye-tracing or borehole-based studies) for the character of the subglacial drainage system.

Where multiple components do exist within a drainage system, it is important to understand how they are arranged in space and how they are interconnected. Failure of the drainage sytem to conform to the assumption that the hydrological pathways represented in the mixing model are arranged in such a way that mixing is confined to the terminal regions of the glacier has important hydrochemical implications. Where two distinct drainage pathways do exist, field evidence suggests that the slower one may be located up-glacier of the faster, so that “subglacial” waters actually pass along the same path as “englacial” waters in the lower glacier (Reference Humphrey,, Raymond, and Harrison,Humphrey and others, 1986; Reference Nienow,, Sharp, and Willis,Nienow and others, in press). If this is so, there would be considerable scope for continued chemical evolution of the combined waters after mixing. Such evolution would violate the assumption of conservative mixing (all solute transported by the bulk meltwaters is derived directly from the flow components) made by mixing models (see below).

Where mixing models have been applied to the interpretation of run-off data collected over whole melt seasons (Reference Gurnell, and Fenn,Gurnell and Fenn, 1984), or large parts thereof, the use of a model with a single form is tantamount to assuming that the structure of the drainage system is constant over time. This assumption is directly contradicted by the results of dye-tracing studies which indicate significant changes in drainage-system structure over individual melt seasons (Reference Willis,, Sharp, and Richards,Willis and others, 1990; Reference Hock, and Hooke,Hock and Hooke, 1993; Reference Nienow,Nienow, 1994). Since these structural changes are often manifested most clearly by changes in meltwater residence time, they have profound implications for the processes and magnitude of solute acquisition by meltwaters, and thus for the assumption made in most studies that component chemistries are unique and conrtant over time (see below).

Unique component chemistries

In most applications of the mixing model published to date (e.g. Reference Collins,Collins, 1978, Reference Collins,1979; Reference Gurnell, and Fenn,Gurnell and Fenn, 1984), it has been assumed that the chemical composition of the “subglacial” and “englacial” components is constant, and that variations in bulk meltwater chemistry reflect variations in the mixing ratio of the two components (an exception to this is the work of Reference Tranter, and Raiswell,Tranter and Raiswell (1991)). Reference Collins,Collins (1979) justified this assumption on the grounds that (a) the drainage pathways followed by “englacial” waters are essentially free of solute sources, and (b) that solute acquisition in the pathways followed by “subglacial” waters is sufficiently rapid to bring them to a uniform equilibrium concentration. Neither argument is convincing. Collins’ “englacial” waters are derived largely from melting of surface snow and ice. It is well known that impurities held within a snowpack are not released in meltwater at a uniform concentration corresponding to their bulk concentration within the snow, but that they are released at significantly higher concentrations within the first meltwater produced (Reference Johannessen,, Dale,, Gjessing,, Henriksen, and Wright,Johannessen and others, 1977). The chemistry of such waters will therefore vary over a melt season in response to this leaching effect. Seasonal changes in the acidity of source waters will also impact on their ability to acquire solute by weathering of subglacial sediment, as demonstrated by laboratory weathering experiments (Reference Brown,, Tranter, and Sharp,Brown and others, in press).

Experimental studies of solute acquisition from glacial flour by dilute aqueous solutions (Reference Tranter,, Raiswell,, Mills, and Miles,Tranter and others, 1989; Reference Brown,, Tranter, and Sharp,Brown and others, in press) indicate that, although rates of acquisition, decrease exponentially over time, they are nevertheless still significant after periods of time comparable with meltwater residence times in glaciers indicated by dye-tracing experiments (Reference Collins,Collins, 1982; Reference Burkimsher,Burkimsher, 1983; Reference Seaberg,, Seaberg,, LeB, Hooke, and Wiberg,Seaberg and others, 1988; Reference Willis,, Sharp, and Richards,Willis and others, 1990; Reference Fountain,Fountain, 1993; Reference Hock, and Hooke,Hock and Hooke, 1993; Reference Nienow,Nienow, 1994). Thus, if meltwaters come into contact with suspended sediment or other rock material during their transit through a glacier, they are likely to experience significant enrichment in solutes. Since both residence time and suspended-sediment concentration vary with meltwater discharge (Reference Burkimsher,Burkimsher, 1983; Reference Gurnell,, Gurnell, and Clark,Gurnell, 1987), the degree of enrichment is likely to vary over time, producing variations in meltwater chemistry which are not directly attributable to the mixing ratio (Reference Brown,, Sharp,, Tranter,, Gurnell, and Nienow,Brown and others, 1994). The kinetics of the weathering processes are such that the impact of such variations on solute acquisition will be greatest in major subglacial channels where significant suspended-sediment When dilute waters are placed in contact with suspended sediment, solute-acquisition rates are initially rapid but they decline exponentially over time (Reference Brown,, Tranter, and Sharp,Brown and others, in press). Thus, a given change in water-sediment contact time will have maximum impact on solute acquisition when the average contact time is short. Dye-tracing studies indicate that this is usually the case in major subglacial channels.

The consequence of assuming constant-component chemistries, when these are in fact variable, is that variations in component chemistry will be expressed as variations in the mixing ratio in the mixing-model results. For instance, increased solute concentrations in bulk waters, such as occur under recession-flow conditions, would normally be interpreted as resulting from an increased proportion of “subglacial” waters in bulk runoff. In reality, they may be at least partly attributable to “subglacial” waters becoming more concentrated over time as discharges decrease and residence times consequently increase. If such artefacts, which may involve bath the magnitude of the component discharges and their pattern of variation over time, are to he avoided, it is necessary to determine accurately the chemistries of the components and their temporal variability. Reference Tranter, and Raiswell,Tranter and Raiswell (1991) suggested a method for doing this using of the properties of bulk meltwaters, but direct measurements in the subglacial environment would be preferable.

A further problem with the assumption of constant meltwater chemistry relates to the choice of appropriate concentration values for each component. Various suggestions have been made concerning which values should be used. For the “englacial” component, these include the concentration of a species in supraglacial waters and the minimum concentration in bulk meltwaters recorded during a melt season. For the “subglacial” component, they include the concentration in ground-water springs, winter base flow or summer recession flows (see Reference Fenn,, Gurnell, and Clark,Fenn (1987) for a review). Each of these possibilities will result in a different value for the component chemistry, which will affect both the magnitude and temporal pattern of flow calculated for each component (Reference Brown, and Tranter,Brown and Tranter, 1990). Higher values of Cs and/or Ce will reduce Qs, though the effect of varying Ce is less marked because Ce occurs on both sides of Equataion (2). Choice of an appropriate value for Ce is, however, critical at times when Qe comprises the majority of bulk meltwaters. We have particular concerns about the use of maximum and minimum solute concentrations in bulk run-off to characterize the compositions of the flow components because, during the summer of 1993, we regularly sampled waters from the base of boreholes at Haut Glacier d’Arolla, Switzerland, which were up to three times more concentrated than any which drained from the glacier during that melt season.

Conservative mixing

As defined by Equation (1), the mixing model assumes that the mass of solutes transported by the two flow components remains unchanged after they mix. Reference Collins,Collins (1979) defended this assumption on the grounds that sedimentary particles in suspension retain significant quantities of ions adsorbed to their surfaces after they exit from the glacier portal. This is said to imply that solution and surface exchange between suspended load and meltwaters play only a minor role in modifying the composition of meltwaters once flow is in large channels. This conclusion is, however, contradicted by the fact that bulk meltwaters are commonly undersaturated with respect to reactive minerals such as calcite (Reference Raiswell,Raiswell, 1984; Reference Brown,, Sharp,, Tranter,, Gurnell, and Nienow,Brown and others, 1994), as are waters sampled at the glacier bed from boreholes (Table 1). In addition, laboratory weathering experiments involving suspended sediment which has passed through the drainage system of Haut Glacier d’Arolla (Reference Brown,, Tranter, and Sharp,Brown and others, in press) show significant and continuing solute acquisition by deionized water placed in contact with sediment concentrations comparable to those recorded in melt streams over at least the first 3 h of contact (Fig. 1 ). Other experiments, which simulate the mixing of dilute and concentrated meltwaters, also reveal significant solute acquisition from suspended sediment after mixing (Reference Tranter,, Raiswell,, Mills, and Miles,Tranter and others, 1989). If, as suggested earlier, mixing is not confined to the extreme terminal regions of the glacier, it is extremely unlikely to be conservative. Reference Brown,, Tranter, and Sharp,Brown and others (in press) estimated that as much as 70% of the Ca2+ load transported by bulk meltwaters draining from Haut Glacier d’Arolla in August 1990 may have been acquired by post-mixing reactions.

Table 1. Chemical composition (mean and standard error of the mean) of the three water types sampled from boreholes at Haut Glacier d’Arolla, Switzerland, in 1992. Concentrations of cations and anions are measured in micro-equivalents per litre. EC0 is the electrical conductivity of the waters at 0°C (μS cm−1). SIcc is the saturation index for calcite and p(CO2) is the partial pressure of carbon dioxide (atmospheres) with which the water appears to be in equilibrium. The latter two parameters are calculated from group mean data rather than individual sample analyses

To illustrate the significance of this point, we present the results of analyses of waters from 22 boreholes drilled through Haut Glacier d’Arolla in August-September 1992. Over 3000 measurements of electrical conductivity (EC) were made on these waters, and the resulting frequency distribution shows three clear modes. Samples of each water type were analyzed for their pH and major cation and anion composition (Table 1). The conventional mixing-model approach to glacier hydrology would make the assumption that waters of intermediate composition (mode 2) were produced by conservative mixing of more dilute (mode 1) and more concentrated (mode 3) waters. To determine whether our chemical data are consistent with this assumption, we used the multivariate mixing model SOLMINEQ88 (Reference Kharaka, and Barnes,Kharaka and Barnes, 1973) to determine whether or not mixing was conservative for the full range of dissolved species analyzed. A mixture of nine parts mode 1 and one part mode 3 waters would have concentrations of Mg2+, Cl, NO3 and SO4 2− which are comparable with those of mode 2 waters. However, such a mixture would contain significantly less Ca2+, K+ and HCO3 , and have a lower pH, EC and calcite-saturation index, and higher p(CO2) than mode 2 waters. To reproduce these properties of mode 2 waters, it was necessary to simulate the dissolution of an additional 5 × 10−5 moll−1 of calcite and 8 × 10−6 moll−1 of potash feldspar in the mixture. This clearly indicates that, for many dissolved species, it would be incorrect to assume that mixing of dilute and concentrated waters is conservative. It may also be the case that mode 2 waters are not, in any event, produced by mixing of mode 1 and mode 3 waters, but that the three water types evolve independently along different flow paths with different residence times.

Fig. 1. Concentrations of Ca2+ and HCO3 as a function of time in a laboratory dissolution experiment using, a water:rock ratio of 4 g l−1. (b) Concentrations of Mg2+, Na+ and K+ as a function of time in a laboratory dissolution experiment using a water: rock ratio of 4 g l−1. The sediment used in the experiment was collected from the margins of the main meltwater stream draining from Haut Glacier d’Arolla; the solvent was deionized water left in free contact with the atmosphere and stirred continuously for the duration of the experiment.

Nevertheless, it is useful for our present purposes to try to quantify the significance of violation of the assumption of conservative mixing. If we take the mean EC of mode 1 (Ce ) and mode 3 (Cs ) waters as a measure of their solute content (Table 1), and use Equation (1) to calculate the EC of mode 2 waters (Ct ), we find that a 9:1 mixture of mode 1 (Qe ) and mode 3 (Qs ) waters would give Ct = 6.3 μS cm−1. Mode 2 waters have a measured mean EC of 10.85μS cm. If we then use this measured value as an estimate of Ct, and take the mean values for mode 1 and mode 3 waters as estimates of Ce and CS we can use Equation (2) to estimate Qs. We obtain a value of 2.1 instead of the true value of 1. The results of similar calculations performed using concentrations of other dissolved species are presented in Table 2. Estimates of Qs range from 1 (for SO4 2−) to 7.9 (K+), when the true value (assuming conservative mixing for SO4 2−) is 1. Thus, even if one accepts the validity of the basic mixing-model approach to glacier hydrology, hydrograph separations which are based upon dissolved species for which mixing is non-conservative may be very seriously in error.

Table 2. Results of mixing-model calculations performed using water compositions equivalent to the mean composition of each of the three water types sampled front boreholes at Haut Glacier d’Arolla in 1992. Mode 2 concentration refers to the observed solute concentration (μeq l−1) in waters of intermediate composition believed to form by mixing of dilute (mode 1) and concentrated (mode 3) waters with additional solute furnished by post-mixing reaction. Predicted Ct is the expected concentration of intermediate waters formed by mixing nine parts mode 1 with one part mode 3 waters with no post-mixing reaction. The percentage of solute derived by post-mixing reaction is given by: ((Mode 2 concentration - Predicted Ct)/ Mode 2 concentration) × 100. Predicted Qs refers to the estimate of Qs that would be obtained using Equation (2) if Qt = 10 and the measured mean compositions of the three borehole-water types given in Table 1 are used as estimates of Ce, Ct and Cs

Most previous mixing-model studies have based separations on the EC of meltwaters (e.g. Reference Collins,Collins, 1978, Reference Collins,1979; Reference Gurnell, and Fenn,Gurnell and Fenn, 1984). The results presented above suggest that EC is not a conservative property of glacial meltwaters and that this practice is therefore unreliable. Better estimates are more likely to be obtained if separations are based upon measurements of the concentration of conservative species such as Mg2+, SO4 2−, Cl and NO3 . Of these species, Mg2+ and SO4 2− are likely to be more useful than Cl and NO3 because the latter two species are normally derived primarily from the snowpack and will be preferentially leached from the catchment in the early part of the melt reason (Reference Tranter,, G,, Raiswell,, Sharp, and Gurnell,Tranter and others, 1993). This conclusion is consistent with the argument put forward by Reference Tranter, and Raiswell,Tranter and Raiswell (1991) that SO4 2− may be the best species to use for mixing-model calculations.

Summary

The application of mixing models to the investigation of flow routing of meltwaters in glaciers demands:

  • a) Independent confirmation that the number of flow components included in the mixing model is appropriate for the glacier in question.

  • b) More rigorous definition of the hydrological pathways associated with each flow component, so that interpretation of mixing-model results can be facilitated.

  • c) Confirmation that the flow components identified have unique chemistries and selection of appropriate values for these. If component chemistries are not unique, it is necessary to evaluate how they vary over time.

  • d) Determination of an appropriate solute species on which to base hydrograph separations in order to minimize the impact of violating the assumption of conservative mixing. The commonly used parameter, EC, is not conservative and is therefore inappropriate for this purpose.

At present, it is not clear that these requirements can actually be met and we therefore advocate restraint in the application of mixing models in glacier hydrology. We also believe that conclusions drawn from past applications are unreliable and that they do not provide a satisfactory basis for discussion of the functioning of glacier drainage systems.

Acknowledgements

We are grateful for grants from the U.K. Natural Environment Research Council (GR3 7004a and GR3 8114). G.H.B, holds a NERC Fellowship (GT5/F/91/AAPS/13). M. Nielsen, M. Tully, C. Smart and S. Gordon made major contributions to our borehole studies in 1992.

References

Brown,, G. H. and Tranter,, M. 1990. Hydrograph and chemograph separation of bulk meltwaters draining the Upper Arolla glacier, Valais, Switzerland. International Association of Hydrological Sciences Publication 193 (Symposium at Lausanne 1990— Hydrology in Mountainous Regions I. Hydrological Measurements; the Water Cycle), 429437.Google Scholar
Brown,, G. H. Sharp,, M. Tranter,, M. Gurnell,, A. M. and Nienow,, P. W. 1994. Impact of post-mixing chemical reactions on the major ion chemistry of bulk meltwater draining the Haut Glacier d’Arolla, Valais, Switzerland. Hydrological Processes, 8, 465480.Google Scholar
Brown,, G.H. Tranter,, M. and Sharp,, M. In press. Experimental. investigations of solute acquisition from suspended sediment by Alpine glacial meltwaters. Hydrological Processes.Google Scholar
Burkimsher,, M. 1983. Investigations of glacier hydrological systems using dye tracer techniques: observations at Pasterzengletscher, Austria. J. Glaciol., 29 (103), 403416.CrossRefGoogle Scholar
Collins,, D.N. 1978. Hydrology of an Alpine glacier as indicated by the chemical composition of meltwater. Z. Gletscherkd. Glazialgeol., 13 (1/2),1977, 219230.Google Scholar
Collins,, D.N. 1979. Quantitative determination of the subglacial hydrology of two Alpine glaciers. J. Glaciol., 23 (89), 347362.CrossRefGoogle Scholar
Collins,, D.N. 1982. Flow-routing of meltwater in an alpine glacier indicated by dye tracer tests. Beitr. Geol. Schweiz. Hydrol., 28 (2), 523534.Google Scholar
Collins,, D. N. and Young,, G. J. 1981. Meltwater hydrology and hydrochemistry in snow- and ice-covered mountain catchments. Nord. Hydrol., 12 (4-5), 319334.CrossRefGoogle Scholar
Fenn,, C.R. 1987. Electrical conductivity, In Gurnell,, A. M. and Clark,, M. J. eds. Glacio-fluvial sediment transfer; an alpine perspective.Chichester, etc., John Wiley and Sons, 377414.Google Scholar
Fountain,, A. G. 1993. Geometry and flow conditions of subglacial water at South Cascade Glacier, Washington State, U.S.A.; an analysis of tracer injections. J. Glaciol., 39 (131), 143156.CrossRefGoogle Scholar
Gurnell,, A.M. 1987. Suspended sediment. In Gurnell,, A. M. and Clark,, M. J. eds. Glacio-fluvial sediment transfer: an alpine perspective. Chichester, etc., John Wiley and Sons, 305354.Google Scholar
Gurnell,, A. M. and Fenn,, C. R. 1984. Flow separation, sediment source areas and suspended sediment transport in a pro-glacial stream. Catena Supplement 5, 109119.Google Scholar
Hock,, R. and Hooke,, R. LeB 1993. Evolution of the internal drainage system in the lower part of the ablation area of Storglaciären, Sweden. Geol. Soc. Ann. Bull., 105 (4), 537546.Google Scholar
Humphrey,, N. Raymond,, C. F. and Harrison,, W. D. 1986. Discharges of turbid water during mini-surges of Variegated Glacier, Alaska, U.S.A. J. Glaciol., 32 (111), 195207.Google Scholar
Johannessen,, M. Dale,, T. Gjessing,, E. T. Henriksen,, A. and Wright,, R. F. 1977. Acid precipitation in Noway: the regional distribution of contaminants in snow and the chemical concentration processes during snowmelt. International Association of Hydrological Sciences Publication 118 (Symposium at Grenoble 1975 — Isotopes and Impurities in Snow and Ice. 116120.Google Scholar
Kharaka,, Y. K. and Barnes,, I. 1973.SOLMNEQ solution-mineral equilibrium computations. U. S. Geol. Surv. Water. Resources Paper 73002. Google Scholar
Lecce,, S.A. 1993. Flow separation and diurnal variability in the hydrology of Conness Glacier, Sierra Nevada, California, U.S.A. J. Glaciol., 39 (132), 216222.CrossRefGoogle Scholar
Nienow,, P. 1994. Dye tracer investigations of glacier hydrological systems. (Ph. D. thesis., University of Cambridge.)Google Scholar
Nienow,, P. Sharp,, M. and Willis,, I. In press. Dye. tracer investigations at the Haut Glacier d’Arolla, Switzerland. II. Seasonal changes in the morphology of the subglacial drainage system. Earth Surface Processes and Landforms.Google Scholar
Nye,, J.F. 1973. Water at the bed of a glacier, International Association of Scientific Hydrology Publication 95 (Symposium at Cambridge 1969 — Hydrology of Glaciers). 189194.Google Scholar
Oerter,, H. Behrens,, H. Hibsch,, G. Rauert,, W. and Stichler,, W. 1980. Combined environmental isotope and electrical conductivity investigations of the runoff of Vernagtferner (Oetztal Alps). Mater. Glyatsiol. Issted. Khron. Obsuzhdeniya 39, 157161.Google Scholar
Pinder,, G. F. and Jones,, J. F. 1969. Determination of the ground-water component of peak discharge from the chemistry of total runoff. Water Resour. Res. 5 (2), 438145.CrossRefGoogle Scholar
Raiswell,, R. 1984. Chemical models of solute acquisition in glacial melt waters. J. Glaciol., 30 (104), 4957.Google Scholar
Röthlisberger,, H. 1972. Water pressure in intra- and subglacial channels. J. Glaciol., 11 (62), 177203.Google Scholar
Seaberg,, S. Z. Seaberg,, J. Z. LeB, R. Hooke,, and Wiberg,, W. 1988. Character of the englacial and subglacial drainage system in the lower part of the ablation area of Storglaciären, Sweden, as revealed by dye-trace studies. J. Glaciol., 34 (117). 217227.CrossRefGoogle Scholar
Sharp,, M. 1991. Hydrological influences from meltwater quality data: the unfulfilled potential. Proceedings of the British Hydrological Society 3rd National Symposium. Southampton. 16–18 September 1991, 5.15.8. Google Scholar
Shoemaker,, E. M. 1986. Subglacial hydrology for an ice sheet resting on a deformable aquifer. J. Glaciol., 32 (110), 2030.CrossRefGoogle Scholar
Tranter,, M. and Raiswell,, R. 1991. The composition of the englacial and subglacial component in bulk meltwaters draining the Cornergletscher. J. Glaciol., 37 (125), 5966.Google Scholar
Tranter,, M. Raiswell,, R. and Mills,, R.A. 1989. Chemical weathering reactions in Alpine glacial meltwaters. In Miles,, D.L. ed. Proceedings. of the 6th International Symposium on Water-Rock Interaction, Malvern, 1989. Rotterdam, Balkema, 687690.Google Scholar
Tranter,, M. G,, Brown. Raiswell,, R. Sharp,, M. and Gurnell,, A. 1993. A conceptual model of solute acquisition by alpine glacial meltwaters. J. Glaciol., 39 (133), 573581.Google Scholar
Walder,, J.S. 1986. Hydraulics of subglacial cavities. J. Glaciol., 32 (112), 439445.CrossRefGoogle Scholar
Walder,, J. S. and Fowler,, A. 1994. Channelized subglacial drainage over a deformable bed. J. Glaciol., 40 (134), 315.CrossRefGoogle Scholar
Weertman,, J. 1972. General theory of water flow at the base of a glacier or ice sheet. Rev. Geophys., Space Phys., 10 (1), 287333.Google Scholar
Willis,, I. C. Sharp,, M. J. and Richards,, K. S. 1990. Configuration of the drainage sytem of Midtdalsbreen, Norway, as indicated by dye-tracing experiments. J. Glaciol., 36 (122), 89101.CrossRefGoogle Scholar
Figure 0

Table 1. Chemical composition (mean and standard error of the mean) of the three water types sampled from boreholes at Haut Glacier d’Arolla, Switzerland, in 1992. Concentrations of cations and anions are measured in micro-equivalents per litre. EC0 is the electrical conductivity of the waters at 0°C (μS cm−1). SIcc is the saturation index for calcite and p(CO2) is the partial pressure of carbon dioxide (atmospheres) with which the water appears to be in equilibrium. The latter two parameters are calculated from group mean data rather than individual sample analyses

Figure 1

Fig. 1. Concentrations of Ca2+ and HCO3 as a function of time in a laboratory dissolution experiment using, a water:rock ratio of 4 g l−1. (b) Concentrations of Mg2+, Na+ and K+ as a function of time in a laboratory dissolution experiment using a water: rock ratio of 4 g l−1. The sediment used in the experiment was collected from the margins of the main meltwater stream draining from Haut Glacier d’Arolla; the solvent was deionized water left in free contact with the atmosphere and stirred continuously for the duration of the experiment.

Figure 2

Table 2. Results of mixing-model calculations performed using water compositions equivalent to the mean composition of each of the three water types sampled front boreholes at Haut Glacier d’Arolla in 1992. Mode 2 concentration refers to the observed solute concentration (μeq l−1) in waters of intermediate composition believed to form by mixing of dilute (mode 1) and concentrated (mode 3) waters with additional solute furnished by post-mixing reaction. Predicted Ct is the expected concentration of intermediate waters formed by mixing nine parts mode 1 with one part mode 3 waters with no post-mixing reaction. The percentage of solute derived by post-mixing reaction is given by: ((Mode 2 concentration - Predicted Ct)/ Mode 2 concentration) × 100. Predicted Qs refers to the estimate of Qs that would be obtained using Equation (2) if Qt = 10 and the measured mean compositions of the three borehole-water types given in Table 1 are used as estimates of Ce, Ct and Cs