• Aucun résultat trouvé

TOPIC 2 Geochemistr y dur ing inf iltration and f low

N/A
N/A
Protected

Academic year: 2022

Partager "TOPIC 2 Geochemistr y dur ing inf iltration and f low"

Copied!
84
0
0

Texte intégral

(1)

TOPIC 2

Geochemistr y

dur ing inf iltrat ion and f low

(2)
(3)

Abstract

Hydrogeologic calculations, transient tracers and deliberate tracer experiments are established methods of deter- mining ground water travel times. A two-year long deliberate tracer experiment using sulfur hexafluoride (SF6) was conducted at the Montebello Forebay (LA County, CA) artificial recharge site to determine travel time to wells within 150 m of spreading ponds. SF6was detected at seven monitoring wells, all screened within 50 m of the ground surface and at nine of the eighteen production wells. Travel time was best correlated with screen depth. At four of the nine wells with SF6detections, hydrogeologic travel times were less than 0.3 years (16 weeks) and are in basic agreement with the SF6results. However, at the other five wells, the hydrogeologic travel times were estimated to be more than 4 years, significantly longer than indicated by the tracer data. Transport through low permeability layers due to gaps or leakage must be occurring. At all wells where SF6was not detected, the hydro- geologic travel times were greater than 3 years. At the production wells, tritium/3He apparent ages were greater than 10 years, indicating mixing of young and old ground water caused by the long well screens.

Keywords

Artificial recharge, geochemical tracer, ground water travel time, SF6, Tritium/3He dating.

I N T R O D U C T I O N

Quantification of subsurface residence times and flow paths are important criteria for managing artificial recharge sites. This basic hydrologic information is needed for evaluating subsurface water quality changes that may result from in situ biogeochemical reactions and mixing of water from different sources. It is also needed for validation of numerical models of flow (e.g., Thomson et al., 1999). Ground water travel time near artificial recharge sites is not simple to define as flow paths to wells can be numerous and convoluted. This especially true at facilities, such as the Montebello Forebay in Los Angeles, California, where percolation from large spreading basins is the principle method of recharge. Ground water travel times can be determined using hydrogeologic calculations or modeling, transient tracers, and deliberate tracer experiments. Each of these methods determines the flow of ground water dif- ferently and thus, provides complimentary information.

Ground water travel times can be estimated using Darcy’s law and inferred flow paths. These flow paths are deter- mined based on the depth of well perforation, aquifer stratigraphy, pump capacity, and distance to recharge location.

Probable travel times are estimated using local hydrogeologic data, these inferred flow paths, and known hydraulic gradients (Bookman-Edmonston, 1994).

Determining residence time for shallow ground water is also possible using environmental tracers such as tritium/3He (Schlosser et al., 1989; Cook and Solomon, 1997). Tritium, a radioactive isotope of hydrogen, was released in large quantities in the late 1950s and early 1960s during above ground nuclear bomb testing. Although there is a natural (cosmogenic) source, most tritium in the environment was produced during these tests. The tri- tium concentration in precipitation reached a maximum in the mid-1960s, 2 to 3 orders of magnitude higher than

A comparison of three methods for determining travel times

near a large artificial recharge facility

Dror Avisar, Jordan F. Clark, Jeni McDermott and G. Bryant Hudson

VV

(4)

prior to the tests. Since then, its concentration has decreased quasi-exponentially. The age of the ground water can be calculated using the radioactive decay relationship between tritium and its daughter isotope, 3He:

(1)

where tis time (apparent ground water age), t1/2is the half-life of tritium (12.43 years), [T] is the measured tritium content, and [3He]triis the concentration of the 3He derived from the decay of tritium. [3He]triis calculated from a helium mass balance that considers other helium sources (Schlosser et al., 1989; Cook and Solomon, 1997). The typical analytical uncertainty of this method is ±2 years. However, the uncertainty of the apparent age can be much larger if mixing between flow paths of different ages occurs.

With deliberate tracer experiments, tracer is added to the recharge water prior to percolation and its arrival at wells is determined directly by periodic sampling. The tracer breakthrough curves are characterized by a number of times including the initial, peak, and mean arrivals. In a homogenous aquifer, the initial and mean arrival times represent, respectively, the fastest and mean flow paths in the aquifer. The former cannot be determined with a high degree of certainty from either geochemical dating techniques or numerical flow models and is the most important time scale when evaluating the potential transport of reactive contaminants near artificial recharge facilities. In aquifers with preferential flow, the tracer breakthrough curves are more complicated, often showing multiple peaks (Clark et al., 2004). The first breakthrough will be representative of transport through the fastest flow path.

Sulfur hexafluoride (SF6), a non-toxic, odorless, colorless, non-reactive gas tracer, is commonly used during delib- erate tracer experiments at artificial recharge sites (Gamlin et al., 2001; Clark et al., 2004; 2005). SF6is less expen- sive than fluorescent dyes and ionized substances but has similar properties, making it possible and cost-effective to tag large bodies of water (>106m3) without the interference of density-induced flow (Istok and Humphrey, 1995).

Laboratory and field experiments have shown that its transport is not retarded within porous media (Wilson and Mackay, 1996; Gamlin et al., 2001). The loss of SF6at the air-water interface from gas exchange can be problematic in recharge facilities that rely on infiltration from spreading basins or rivers. In order to define the input function of tracer to the ground water at these settings, careful monitoring of the surface water is required.

This paper presents an application and comparison of these three methods for determining travel times near the Montebello Forebay recharge site in North Los Angeles County, California (Figure 1). This artificial recharge site is

⎟⎟⎠⎞

⎜⎜⎝⎛ +

= [T]

] He 1 [ ) 2 ln(

3 2

/

1 tri

t t

100832

100089-90 200095

200096 100831

200013 200018 200087

100068-69 200099

100829

200089

200088 100833 200011

200012 100834

200061 200052

200065 200063

200055-59

200004 Production Well

Monitoring Well Rivers

Surface Road Freeway

200062 100830

Kilometer

0 1 Figure 1. Map of the MontebelloForebay

spreading basins and wells sampled during the study.

The Rio Hono basins and river lie to the west of the San Gabriel basins and river.

The light and heavy dashed lines are, respectively, major surface roads

and freeways(I-5 and I-605).

(5)

within the principle recharge area for the Los Angeles ground water basin. This alluvial basin is filled with layers of sands, gravels, silts and clays occurring in discrete aquifer or aquitard units, however the spatial extent of any given layer is poorly known (Bookman-Edmonston, 1994). During the last decade, approximately 1.6 x 108 m3/yr (129,000 acre-ft/yr) of surface water were recharged annually from the Montebello Forebay recharge site. About 40% of this water is natural storm runoff, 25% is imported water, and 35% is reclaimed (Johnson, personal commu- nication).

The recharge site contains 23 shallow (<4 m deep) spreading basins adjacent to the Rio Hondo and San Gabriel River with a total wetted area of about 200 hectares. Additional recharge occurs through unlined portions of the San Gabriel River; the Rio Hondo is lined with concrete throughout the study area. Travel times to ten monitoring and eighteen production wells were examined. With the exception of only two, the wells studied are within 150 m of the spreading basins. Depth of production varies considerably. Tops of well screens are as shallow 8 m and bottoms are as deep as 270 m.

M E T H O D S

S

SFF

66

ttrra acce err e ex xp pe erriim me en ntt

Beginning on February 14, 2003 and continuing for a period of seven days, 99.99% pure SF6gas was injected into spreading basins by bubbling every one to two days. Surface water samples (5–8 samples per pond) were collected from a small boat during this time. Personnel from the Water Replenishment District of Southern California col- lected well samples every two to ten weeks after the injection period from ten monitoring and eighteen production wells. All the samples were collected in pre-weighed Vacutainers. All samples were analyzed using a gas chromato- graph equipped with an electron capture detector following the method developed by Clark et al. (2004). The pre- cision and detection limit were ± 3% and 0.04 pmol l–1, respectively.

TTrriittiiu um m/ /

33

H He e g grro ou un nd dw wa atte err a ag ge e d da attiin ng g

Tritium/3He samples were collected independent of the SF6 experiment as part of the State of California’s Groundwater Ambient Monitoring and Assessment (GAMA) program. All tritium and 3He samples were collected in 500 mL glass bottles and 10 mL copper tubes sealed with pinch-off clamps, respectively. At Lawrence Livermore National Laboratory, the 3He/4He isotope ratio and concentrations of 4He and the heavier noble gases were deter- mined using a noble gas mass spectrometer (Radamacher et al., 2001). Tritium was measured by a 3He accu- mulation method (Surano et al. 1992).

H

Hyyd drro og ge eo ollo og giicc ttrra avve ell ttiim me ess

Geochemical travel times will be compared to hydrogeologic travel times calculated by Bookman-Edmonston Engineering, Inc of Glendale, CA (Bookman-Edmonston, 1994). The calculations used inferred flow paths between the spreading basins and wells, known hydraulic gradients, an assumed uniform porosity of 20%, and mean hydraulic conductivities calculated from well test data for the highly permeable layers and estimates based on grain size for the impermeable layers.

R E S U LT S A N D D I S C U S S I O N

SF6was not detected in any background samples collected six and two months prior to the start of the tracer exper- iment. After injection, concentrations of SF6within the spreading basins ranged from 10 – 80 pmol/L with an

T O P I C 2 Geochemistr y dur ing inf iltration and flow 249

(6)

average concentration of 30 pmol/L The temporal and spatial variability is due to a variety of factors including wet- ted area, percolation rate, basin mean depth, the amount of SF6injected, and time since the last injection. Tracer was detected in the wet basins for about two weeks after the end of the injection period.

SF6arrived at seven of the ten monitoring wells sampled during the two year long experiment, indicating that this gas tracer was successfully transferred through the unsaturated zone to the water table (Figure 2). SF6 was not detected at one shallow well (#100832) which was more 1 km from the spreading basins. The maximum ground water SF6 concentration was 4.8 pmol/L, 15% of the mean concentration in the surface water, and 20% of the con- centration in the nearest spreading basin. All monitoring wells with tracer detection have depths of less than 45 m below the nearest spreading basin. In addition to #1000832, SF6was not detected at monitoring wells #100068 and 100089. Both of these wells have screen depths greater than 70 m below the closest spreading basin. There is a strong correlation between screen depth and travel time for all monitoring wells sampled (Figure 3). There is no relationship between horizontal distance and travel time. The lack of a correlation may reflect the proximity of the wells to the spreading basins (distance < 150 m).

SF6 was detected at nine of the eighteen production wells. The maximum concentrations observed at the pro- duction wells were much lower than at the monitoring wells. At seven out of the nine production wells with SF6 detections, the maximum concentrations were less than 0.4 pmol/L or an order of magnitude less than at the moni- toring wells and two orders of magnitude less than the spreading basins. With the exception of two wells (#200061 and #200065), SF6peaks were difficult to resolve as tracer concentrations were near the limits of detection and the breakthrough curves were not smooth. The lower concentrations and complex breakthrough curves are probably due to dilution of the tagged water with untagged ground water caused by the long screened intervals. Similar to the monitoring wells, there is a strong relationship between tracer arrival time and depth to the top of the perforation (Figure 3) and no correlation between distance and travel time. Depth may be the most important factor influ- encing travel time because the deeper the well perforation is, the more likely is it for the screen to be situated below layers with low hydraulic conductivity.

The hydrogeologic analysis conducted by Bookman-Edmonston (1994) divided the production wells into two main categories. The first consisted of four shallow wells (#200061, 200055, 200058, 200065) screened in the uncon- fined aquifer. For all of these wells, the hydrogeologic and SF6travel times are consistent, both indicating travel

0 1 2 3 4 5

0 0.5 1 1.5 2

#100829

#100834

SF6 Concentration (pmol/L)

Time (yrs)

Figure 2. SF6 breakthrough curves at two monitoring wells.

See Figure 1 for well locations.

Depth (m)

0

20

40

60

80

1000 0.5 1 1.5 2

Production Monitoring

Time (yrs)

Figure 3. Vertical travel times to production and monitoring wells located within 150 m

of the Montebello spreading basins during the SF6tracer experiment

(7)

times of 0.3 years (16 weeks) or less. The hydrogeologic study identifies all other wells as having perforations in a confined aquifer with little direct contact with surface water. At these wells, estimated travel times range between 2.5 to 14 years. The longer travel times were caused by slow flow through the confining clay layers. Within the time constraints of the two-year tracer experiment, a complete analysis of the accuracy of the hydrogeologic travel times cannot be completed. However, four of the nine production wells near the Rio Hondo spreading grounds (#200062, 200088, 200089, 200099) that were determined to be in the confined aquifer, received tracer. The five wells that did not receive tracer are all located on the western side of the Rio Hondo River while the four wells that did see tracer are located on the eastern side. There is no discernable difference in the well log geology between the eastern and western sides. Nevertheless, the comparison between the tracer and hydrogeologic travel times suggests that discon- tinuities (gaps, fractures, or inter-bedded layers of course material) within the clay layers exist on the eastern side.

Of the five production wells studied near the San Gabriel spreading basins, only one well (#200012) screened in the lower confined aquifer received tracer in less than a year. All other deep wells studied near this river did not receive tracer during the study.

The T/3He apparent ages at production wells were between 7 and 40 years, all significantly older than travel times determined by the other two methods. The older apparent ages reflect mixing of young and old ground water within the well. The old component must have a residence time of decades and could make up the largest fraction at some wells. Because T/3He ages do not mix linearly (the mixed age is weighted by each flow path’s initial tritium content) this technique leads to over-estimates of the mean travel time. Good agreement between travel times determined with deliberate tracer experiments and T/3He ages were found at three of the four monitoring wells where both techniques were used. Good agreement has also been found at other artificial recharge sites (Clark et al., 2004).

I M P L I C AT I O N S

Each method of determining the travel time between the recharge basins and wells used during this study is based on different principles and assumptions. A conceptually better model of the flow and travel times can be achieved by comparing the results of the three methods. With deliberate tracer experiments, such as the one performed with SF6here, travel along the fastest flow paths is quantified. Results of deliberate tracer experiments are limited by their duration and by dilution with untagged water which can lower tracer concentrations below the detect limit.

Hydrogeologic calculations require local geologic information usually extrapolated from well bore measurements to estimate the flow along specific paths. Because details about preferential flow paths are usually lacking, these calculations estimate mean travel times and rates. The use of a deliberate tracer experiment concurrently with hydrogeologic calculations can either validate or lead to better models of the hydro-stratigraphy. When significant mixing of young and old ground waters occurs as in most production wells, T/3He apparent ages are the most diffi- cult to interpret. If linear mixing occurs, then these ages would be equivalent to the mean age. However, because the initial tritium content of recharge water varied by orders of magnitude during the last few decades, T/3He apparent ages do not mix linearly. Nevertheless, the T/3He apparent ages from this study indicate that productions wells are drawing in an old component and that this component has a travel time greater than a decade.

The California State Department of Health recommends that ground water containing reclaimed water reside in the subsurface for a minimum of six months to allow for virus inactivation (Bookman-Edmonston, 1994). In order to ensure the six-month residence time, all wells must be farther than 150 m from the point of infiltration. This study suggests that using horizontal distances does not ensure a specified travel time within the ground water system; the depth of production is a better criterion to ensure desired residence time close to spreading ponds.

T O P I C 2 Geochemistr y dur ing inf iltration and flow 251

(8)

AC K N O W LE D G E M E N T S

The work could not have been completed without T. Johnson and B. Chong of the Water Replenishment District of Southern California who assisted with the planning of the SF6 experiment and the collection of well samples.

M. Ragland helped to collect and analyze the surface and ground water samples. The research was supported by Water Replenishment District of Southern California. This is contribution #0689 of the Institute for Crustal Studies at the University of California, Santa Barbara.

R E F E R E N C E S

Bookman-Edmonston (1994) Hydrogeologic assessment and background data for domestic wells within 500 feet of the Montebello Forebay recharge areas.Bookman-Edmonston Engineering, Inc., Glendale, CA.

Clark, J. F., Hudson G. B., Davison M. L., Woodside G., and Herndon R. (2004). Geochemical imaging of flow near an artificial recharge facility, Orange County, CA. Ground Water, 42, 167–174.

Clark, J. F., Hudson G. B., and Avisar D. (2005). Gas Transport Below Artificial Recharge Ponds: Insights from Dissolved Noble Gases and a Dual Gas (SF6and 3He) Tracer Experiment, Enivorn. Sci. Tech.,(in press).

Cook, P. G. and Solomon D. K. (1997). Recent advances in dating young ground water: chlorofluorocarbons, 3H/3He and 85Kr. J. Hydro.,191, 245–265.

Gamlin, J. D., Clark J. F., Woodside G., and Herndon R. (2001). Tracing groundwater flow patterns in an area of arti- ficial recharge using sulfur hexafluoride. J. Environ. Eng. 127, 171–174.

Istok, J. D., and Humphrey M. D. (1995). Laboratory investigation of buoyancy-induced flow (plume sinking) dur- ing two well tracer tests. Ground Water,33, 597–604.

Johnson, T. Personal communication (Feb 18, 2005).

Rademacher, L. K., Clark J. F., Hudson G. B., Erman D. C., and Erman N. A. (2001). Chemical evolution of shallow groundwater as recorded by springs, Sagehen Basin, Nevada County, California. Chem. Geol.,179, 37–51.

Schlosser, P., Stute M., Sonntag C., and Munnich K.O. (1989). Tritiogenic 3He in shallow groundwater. Earth Planet.

Sci. Lett.,94, 245–256.

Surano, K. A., Hudson G. B., Failor R. A., Sims J. M., Holland R. C., MacLean S. C., and Garrison J. C. (1992).

Helium-3 mas spectrometry for low-level tritium analysis of environmental samples. J. Radioanal. Nuclear Chem.

Art.,161, 443–453.

Tompson, A. F. B., Carle, S. F. Rosenberg N. D., and Maxwell R. M. (1999). Analysis of ground water migration from artificial recharge in a large urban aquifer: A simulation perspective. Water Resour. Res.,35, 2981–2998.

Wilson, R. D. and Mackay D. M. (1996). SF6as a conservative tracer in saturated media with high intragranular porosity or high organic carbon content. Ground Water, 34, 241–249.

(9)

Abstract

The source of groundwater recharge to alluvial aquifers of the Lockyer Valley is difficult to confirm, especially as water chemistry of bores is highly variable. To better define recharge over 100 bores have been analysed for major ions and stable isotopic ratios δ2H and δ18O. Preliminary identification of hydrological processes was by ionic ratios (e.g. Cl/HCO3) and isotope plots, however, although these methods can indicate the source of water, they cannot show the degree of mixing. A plot of log TDS (mg/L) versus δ2H ‰ was used with some success. Within the plot four ‘end members’ can be established, with typical values: A. sandstone bedrock; B. stream storm flow;

C. deep artesian basin; and D. strongly evaporated alluvial groundwater. Within the ABC triangle relative percent- ages of each member can be determined; within the ABD triangle relative percentages of stream recharge and sandstone water can be established, and the degree of evaporation is indicated. The plot confirms the primary process of recharge to alluvial aquifers is directly as storm flow in streams, and indirectly from surrounding basalts. The contribution to alluvium of bedrock sandstone water is also shown to be significant during periods of low stream flow and low groundwater levels.

Keywords

Alluvial aquifers, Australia, evaporation, groundwater, irrigation, recharge, stable isotopes.

I N T R O D U C T I O N

The Lockyer Valley of southeast Queensland is an important area of crop production based on intensive ground- water irrigation. The agricultural development has placed excessive demands on groundwater and there is a trend of increasing salinity in many parts of the valley. As a consequence various studies have reported aspects of water chemistry and isotope composition of groundwater in the Lockyer Valley (e.g. Talbot and Dickson, 1969; Stevens, 1990; Dixon and Chiswell, 1992; Dixon and Chiswell, 1994).

The Lockyer Valley supplies around 35% of the state’s irrigated vegetables as both winter and spring crops, plus fodder and small crops. Prior to 1936 irrigation in the Lockyer was limited, but by 1956 the area under irrigation had increased to 3,500 ha, and by 1969 to 12,000 ha. By the mid-1970s the area irrigated had stabilised at around 13,000 ha; an estimate is that on the order of 16,800 ha of the valley has potential for irrigation. More recently there has been rural subdivision and the establishment of hobby farms. The supply of irrigation water in the catchment is closely related to rainfall. During the period 1990 to 1995 the severe drought affecting much of eastern Australia had a marked impact on irrigation in the valley. This drought broke in November, 1995, and in January, 1996 the area was subject to flooding, but dry conditions have subsequently reoccurred. There are substantial variations in water quality in the Lockyer, both spatial and temporal, and these can be difficult to assess with the irregular data on file. Results of studies to date show that hydrological processes are complex, notably due to the mixing of various waters.

The source and character of recharge to the groundwaters of the alluvial aquifers have been difficult to confirm, especially as water chemistry of bores within both the alluvium and the underlying sandstone is highly variable

Use of geochemical and isotope plots to determine recharge to alluvial aquifers:

Lockyer Valley, Queensland, Australia

Malcolm E. Cox and Andrew S. Wilson

VV

(10)

typically increases during dry period irrigation. Also of significance is natural evaporation in streams of poor flowand of irrigation runoff, plus the discharge of deep artesian basin CO2–bearing waters. Groundwater in the basalt aquifers of the surrounding ranges is also chemically different.

P H Y S I C A L S E T T I N G

C

Clliim ma attiicc ffa acctto orrss

The Lockyer Valley is a roughly circular basin of approximately 2,500 km2and is formed of a sequence of sedimen- tary rocks which are dominantly sandstone. Central to the valley is the meandering Lockyer Creek, which drains eastward to the Brisbane River. The town of Gatton, population of 4,600, is the largest in the valley and is 90 km west of the state capital Brisbane (Fig. 1). The climate of the area is sub-tropical and humid, with relatively long hot summers and winters that are short and mild with occasional frosts. Rainfall central to the valley has an annual average of 720 mm, while the annual rainfall increases on the adjacent ranges to the west and south to around 1,200 mm. Stream flow for much of the drainage system is ephemeral in nature and related to summer storms (November-March). There is a marked rainfall deficit and the average annual rate of evaporation is estimated to be 1,900 mm (Queensland Bureau of Meteorology, unpub. data). Alluvial groundwater levels respond rapidly to storm stream flow, and are strongly influenced by prolonged dry periods as well as the associated increase in irrigation extraction; water tables are now commonly below stream bed levels.

M

Mo orrp ph ho ollo og gyy a an nd d h hyyd drro og ge eo ollo og gyy

The valley is bordered on the south and west by the Great Dividing Range which is typically 250–300 m higher than the valley floor. This range is flat-topped and covered by Tertiary age basalt flows which weather to form dark fertile soils. The headwaters of the larger tributaries of the Lockyer drainage system abutt the basaltic ranges and escarpments in the south and west. To the north, the drainage divide is less well defined, and streams flowing south to Lockyer Creek are smaller and more irregular in nature. In that area, the rock types are predominantly coarse sandstone and granite, and of undulating topography. These northern subcatchments have limited development of alluvium and as a consequence less irrigation.

Figure 1. Location of the Lockyer Valley in southeast Queensland and in relation to Brisbane

(11)

The sedimentary formations of the valley consist of a conformable sequence of Triassic and Jurassic sediments and are described in detail by Cranfield (1981) and Wells and O’Brien (1993). These sequences predominantly consist of flat-lying fluvial sandstones, siltstones and shales, with minor conglomerate units (Fig. 2). Here we use a gener- alised description for the sedimentary formations of, lower: Gatton Sandstone (exposed in central valley floor along Lockyer Creek), middle: Winwill Conglomerate, and upper: Koukandowie Formation (exposed in sides of valley and upper slopes).

The groundwater supplies used for irrigation in the area are largely obtained from the Quaternary alluvium, which within the Lockyer drainage system covers approximately 28,000 ha. Water bores show the alluvium to be between 20 and 30 m thick in the central drainage systems, and to directly overlie the sedimentary formations. Alluvial deposits are composed of well-graded gravel, sand and silt with a clayey matrix; cobbles also occur in channel deposits. Clay-rich layers are common in the broad alluvium of the middle sections of Lockyer Creek. Typically, the lower section of the alluvium is composed of coarser sands and is the main water-bearing layer; commonly this layer can be semi-confined by silt-clay layers. Hydraulic conductivities (K) of these coarse sand-gravel layers are typically 50 to 80 m/day (Wilson, 2004).

In Lockyer Creek and its larger tributaries the stream channels tend to be incised into the alluvium, commonly 3–5 m, are often narrow and without distinct alluvial terraces. Over time these channels have meandered back and forth across valley floors and have formed floodplains of rich soil which can be 2 to 6 km in width. In dry periods, very little water is seen in the drainage system, but after heavy storms, stream channels rapidly receive and transmit flow.

T O P I C 2 Geochemistr y dur ing inf iltration and flow 255

Figure 2. Location and geology map of the Lockyer Valley (after Department of Natural Resources and Mines).

Lockyer Creek runs west to east past Gatton in the central valley. Locations of selected groundwater bores are numbered; H is Helidon, a spa and spring area.

(12)

S

Su urrffa acce e a an nd d g grro ou un nd dw wa atte err ssu up pp pllyy

Stream water is the main source of recharge to the alluvial aquifers and the response of groundwater levels to flood flows in the stream is rapid. During dry periods, however, such stream flow may only be in upper parts of the catch- ment. Water tables typically rise 7 to 10 days after several days of heavy rain in the headwaters or on the escapment.

Available records confirm that in dry periods the water table is below the creek beds and that hydraulic gradients slope towards the edges of the valleys. Small weirs to increase alluvial recharge have been built on most tributaries and although subject to some siltation most of these weirs are achieving an annual recharge of around 700 ML (DPI-WR, 1993). There are now 16 artificial recharge weirs on streams throughout the valley.

State government licensing for use of surface water supplies has been a requirement for many years. However, except for the central section of the valley around Gatton the use of groundwater for irrigation, and other purposes, has not been subject to the same statutory controls. Most of the expansion of irrigation in the 1960s and 1970s was from groundwater, which now accounts for over 80% of irrigation supply, and corresponds to an average annual groundwater withdrawal of some 45,000 ML. Groundwater storage in the alluvial aquifers was estimated at a safe annual yield of around 25,000 ML (e.g. DPI-WR, 1993). These amounts indicate that the alluvial aquifers are being over-exploited, which is clearly evident during dry periods when many bores become dry. The demand for irrigation water therefore exceeds supply, except in the wettest seasons. A preliminary valley-wide assessment of groundwater data over a 30 year period (Tien et al., 2004) indicates that the long-term trend is of increasing salinity, which based on the sites tested is 60%, or approximately 2% per year. The overall trend during this period is for decreasing groundwater levels, particularly over the last 15 years.

Figure 3. Piper diagram of groundwaters in the southwestern Lockyer Valley collected during 2003 displaying chemical difference between different groundwaters.

(13)

R E C H A R G E T O A L L U V I A L A Q U I F E R S

B

Ba acck kg grro ou un nd d

The variation of surface and groundwater chemical composition throughout the valley reflects a variety of sources and mixing of waters. Although it is generally accepted that recharge is primarily derived from the surrounding ranges, the alternative sources and possibly proportions to the alluvial aquifers have not been determined.

The use of hydrogeochemistry and of stable isotope analyses have provided some insight to recharge and mixing, at least in a qualitative sense. Studies by Dixon and Chiswell (1992; 1994) in the southwest of the valley confirmed that bedrock groundwaters are identifiable and do discharge into the alluvial aquifers. Studies in Sandy Creek in the southeast (McMahon, 1995; McMahon and Cox, 1996) confirmed that for groundwater within the sedimentary pro- file there is both a variation of salinity and chemical composition for different formations. The study by Mcleod (1998) of Ma Ma Creek in the southwest confirmed that with continued irrigation under dry conditions a greater proportion of bedrock groundwater is drawn into the alluvial bores. All studies show that overall, groundwater within the lower bedrock formations has a higher total dissolved solids. A recent investigation (Picarel, 2004) in the lower (eastern) Lockyer Creek confirms that some very high saline groundwaters are present in bedrock depres- sions. That study also demonstrated that 2004 water levels have been drawndown into the lower sections of the alluvium.

T O P I C 2 Geochemistr y dur ing inf iltration and flow 257

Figure 4. Plot of δ2H ‰ versus Cl / HCO3for groundwaters in the southwestern Lockyer Valley collected during 2003 displaying processes such as mixing and evaporation.

(14)

C

Ch he em miicca all cco om mp po ossiittiio on n o off g grro ou un nd dw wa atte errss

Although most bores extract from alluvium some also extract a component of water from the underlying sandstones. Most bores in the valley are drilled 30–50 cm into bedrock, and are screened in the lower 1–2 m of the alluvium. The proportion of this latter groundwater can be also seen as baseflow and increases during periods of low river flow, and also downstream in many drainages. For example, the mainstream Lockyer is mostly Na,Mg,Ca- HCO3 water, but at low flows a Na,Mg-Cl water dominates. The highest salinity alluvial groundwaters are mainly in the centre of the valley, concentrated in the floodplains that support the most intensive agriculture. There are, how- ever, conflicting views as to whether the high salinity levels in certain areas are a natural occurrence, or whether they have been worsened by poor quality water entering the alluvial aquifers.

Recent investigations in the Tenthill-Ma Ma catchments in the southwest and the alluvial plain at their conflence with Lockyer Creek (Wilson, 2004; Wilson et al., 2004) were effective in identifying groundwaters from different aquifers. Alluvial aquifers are dominated by waters of Mg,Na(Ca)-Cl(HCO3) composition (Fig. 3). It can be seen in the Piper plot that in the central valley groundwater in the alluvium has a greater proportion of HCO3; although this can reflect a rainfall recharge influence, it can also be related to high HCO3deep waters. This plot clearly demonstrates the Na-Cl nature of groundwater in the sandstone bedrock; basalt groundwater from the surrounding ranges is commonly of a Mg-HCO3type, and similar to flowing streams (i.e. recharging waters).

Salinity variations also exist between these various waters and typically depend on, (a) bedrock formation, and (b) irrigation practice, or both. Typical salinities (mg/L) are, alluvium: Lockyer (800–3,700), Tenthill (1,000–2,500) and Ma Ma (5,400 –7,700); sandstone: Gatton, lower (5,700 –7,500) and Koukandowie, middle-upper (3,000–6,800).

Figure 5. Plot of δ2H versus δ18O for groundwaters from the eastern section of the Lockyer Valley (in northeast of Figure 2). Samples were collected in 2004 (after Picarel, 2004).

(15)

Both hydrogeochemical and isotope methods provide some understanding of the hydrological processes occuring.

The Piper plot indicates some degree of mixing between waters. These processes are displyed well in a plot of δ2H‰ versus Cl / HCO3 (Fig. 4), for which both parameters indicate the source of the water. Flowing streams represent recharging water and runoff or throughflow from basalt aquifers; alluvial and bedrock groundwaters are evident lower in the plot, as is some mixing between them. Typically, groundwater from bedrock aquifers has the most depleted values of δ2H‰ and higher Cl / HCO3 ratios. Evaporative processes in standing streams and of irri- gation water are evident in the upper part of the figure.

As an additional method of determining hydrological processes in the Lockyer Valley, groundwater samples from the lower (eastern) alluvial irrigation area were analysed for stable isotopes (Fig. 5). In this part of the valley alluvium is largely overlying the lower bedrock formation, the Gatton Sandstone. The isotopic compositions for most samples tend to fall along the Global Meteoric Water Line (Craig, 1961). Samples of flowing streams and alluvial groundwater immediate to streams are grouped on this MWL; groups with differing degrees of surface or near-surface evaporative concentration fall along a line with a steep slope. The shaded samples in the ‘stream recharge’ and ‘evaporation’ groups are stream waters; the latter is from a non-flowing pool. The lower trend in δ18O enrichment is groundwaters in alluvial bores that are distant from the stream, and contain re-circulated irrigation water. Alluvial bores with significant bedrock input (plus bores into sandstone) are shown to be the most depleted;

this supports the idea of recharge to sandstone bedrock aquifers as being distant (further to the west or south).

D E T E R M I N AT I O N O F S O U R C E O F R E C H A R G E

H

Hyyd drro ollo og giicca all p prro occe esssse ess

The plots above demonstrate the type of hydrological processes that occur within the valley, and the obvious con- nection between groundwater and surface waters. The main forms of recharge are also indicated. These processes are summarised in Figure 6, in which features A to D also represent endpoints in the ternary diagrams of Figure 7.

T O P I C 2 Geochemistr y dur ing inf iltration and flow 259

Figure 6. Main processes of the hydrological systems related to recharge of groundwater within alluvial aquifers of Lockyer Valley. Two examples of groundwater irrigation bores are shown; the deeper type is the most productive

(e.g. 5 – 8 L /sec) and produces the most drawdown.

(16)

These processes are as follow:

A: discharge from sandstone bedrock as (a) seepage into permeable sands and gravels at base of the alluvium, (b) surface discharge from springs in outcropping bedrock, and (c) seepage along upper sections of stream courses;

B: recharge of low salinity water from (a) streams during periods of flow, and (b) basalts on surrounding ranges as throughflow to stream headwaters;

C: discharge of deep basin waters (Great Artesian Basin) from faults at base of alluvium, or from exposed bedrock.

These waters are cool, but contain abundant CO2;

D: alluvial groundwaters that have experienced evaporative concentration, usually towards the edges of alluvial plains or at shallow depth.

Of note is the overall low permeability of the shalower profile of the alluvium, due to high silt content. Low infil- tration rates were determined for surface alluvial material in the central parts of the valley (e.g. Ellis and Dharmasiri, 1998). This finding supports the conclusion that recharge to the alluvium is dominantly along the drainage system. Flow of groundwater through the sandstones is relatively slow, and recharge to these foramtions is distant to the immediate valley (e.g. Dharmasiri et al., 1997).

D

De effiin niittiio on n o off p prro op po orrttiio on n o off rre ecch ha arrg ge e

To assist in the future management of the groundwater resources better definition of recharge processes will be of value. Although the forms of recharge are relatively well understood, no quantification of amount has been reported. As shown above water chemistry and isotope character can reflect many processes. Further consideration of over 100 samples from throughout the valley indicated some merit in a plot of log TDS (total dissolved solids, mg/L) and δ2H (‰).

Figure 7 (upper) displays a diamond-shape field consisting of four end points and composed of two ternary plots.

The end points are the representative values for different water types and are based on 2–4 samples. The left triangle ABC is the field for sandstone-stream-deep basin sources; the right triangle ADC is the field for sandstone-stream- evaporation. The two triangles are divided by lines of 10%; the position within each triangle enables the relative proportion of each source to be calculated. For the ADC triangle the endpoints are stream waters, sandstone waters and groundwaters influenced by evaporation; as the latter parameter is relative fields of low, medium and high are identified.

In Figure 7 (lower) eight examples of groundwaters from a variety of bores throughout the valley have been assessed to test the method.

Results are summarised in Table 1, showing the indicated percent recharge from each primary source, as well as TDS (mg / L). It is clear that bores with a larger component of sandstone groundwater input are more saline. These results are compared to Cl/HCO3: bores with a deep basin component typically have very low values, however, low Cl/HCO3values can also indicate samples more proximal to rainfall recharge areas near to stream headwaters (e.g.

sites 7 and 8).

(17)

T O P I C 2 Geochemistr y dur ing inf iltration and flow 261

Figure 7. Plots of log TDS (mg/L) and δ2H (‰) showing (upper) various fields, and (lower) examples of water samples summarised in Table 1.

(18)

C O N C L U S I O N S

In summary, the log TDS (mg/L) and δ2H (‰) plot shows the major forms of recharge are directly from streams and indirectly from basalts, and from sandstone bedrock. A wide range of degress of mixing occur, and temporal variation of rainfall is also a factor. The latter is especially the case during extended dry periods. In the ABC field, some bores are shown to have a component of deep basin water; although this is the case for some deep bores, a number of bores in the valley floor alluvium have a high HCO3 content, suggesting shallow dispersion of such water. For exmple, bore 1 near Laidley may have such an input from a concealed fault. Of note, the Helidon spa water falls outside the field, but is indicated to be 80% deep basin water; these waters spread over 100 metres along the drainage system. In the ADC field most alluvial bores display minor or medium evaporative concentration, but a number (e.g. 6) are shown to have experienced marked evaporation.

AC K N O W LE D G E M E N T S

We thank our colleagues for assistance in the field, with data and for much discussion: Dr Vivienne McNeil, Dr Gerard McMahon, Robert Ellis and Bruce Pearce (Department of Natural Resources & Mines), Dr Micaela Preda (QUT), and Julie Picarel (University of Montpellier, France).

Table 1. Parameters of selected groundwaters

(19)

R E F E R E N C E S

Cox, M.E., Hillier, J., Foster, L. and Ellis, R., 1996, Effects of a rapidly urbanising environment on groundwater, Brisbane, Queensland, Australia. Hydrogeology Journal, 4, 1, 30–47.

Craig, H., 1961. Isotopic variations in meteoric waters. Science, 133, 1702–1703.

Cranfield, L. C., 1981, Stratigraphic drilling report – GSQ Ipswich 24 and 25: Queensland Government Mining Journal, Brisbane, Australia, 82, 468–477.

Dharmasiri, J.K., Morawska, L. and Hillier, J. 1997. Application of stable isotopes to identify sources of recharge to an alluvial aquifer in Gatton, Queensland. Murray-Darling Workshop ’97, Toowoomba, 26–28 August, 1997, Queensland, Australia, 249–252.

Dixon, W. and Chiswell, B., 1992, The use of hydrochemical sections to identify recharge areas and saline intrusions in alluvial aquifers, southeast Queensland, Australia: Journal Hydrology,135, 259–274.

Dixon, W. and Chiswell, B., 1994, Isotopic study of alluvial groundwaters, south-west Lockyer Valley, Queensland, Australia: Hydrological Processes, 8, 359–367.

DPI-WR, 1993, State water conservation strategy. A discussion paper: Report Water Resources Group, Queensland Department of Primary Industries, Brisbane, Australia, 128 pp.

Ellis, R. and Dharmasiri, J.K. 1998. Chemical and stable isotope methods used in investigating groundwater quality deterioration in the Lockyer Valley. IAH International Groundwater Conference ’98, 8–13 February, 1998, Groundwater Sustainable Solutions, Melbourne, Australia.

Mcleod, K.A., 1998. A study of the groundwater in the Ma Ma Creek catchment, Lockyer Valley, southeast Queensland. Honours thesis (unpub), School of Natural Resource Sciences, Queensland University of Technology, Brisbane, Australia, 88 pp.

McMahon, G. A., 1995, Hydrochemistry of saline groundwater in the Sandy Creek catchment, Lockyer Valley, Southeast Queensland: Honours thesis (unpub), School of Geology, Queensland University of Technology, Brisbane, Australia, 62 pp.

McMahon, G.A. and Cox, M.E. 1996. The relationship between groundwater chemical type and Jurassic sedimentary formations: the example of the Sandy Creek catchment, Lockyer, southeast Queensland. Mesozoic 96, Geological Society of Australia, Extended Abstracts 43, 23–26 September, 1996, Brisbane, Australia, 374–382.

Picarel, J. M. 2004. Distribution of groundwater salinity within alluvial aquifers, lower Lockyer Valley, southeast Queensland. Honours thesis (unpub), (Polytech' Montpellier, France), School of Natural Resource Sciences, Queensland University of Technology, Brisbane, Australia, 55pp.

Stevens, N., 1990, Aspects of the geology and geochemistry of the catchment: in The Brisbane River, a source-book for the future. Editors, P. Davie, E. Stock and D. Low Choy, Queensland Museum, Brisbane, Australia, 17–27.

Talbot, R.J. and Dickson, T., 1969, Irrigation quality of some stream waters in the Lockyer Valley, south east Queens- land: Queensland Journal Agricultural Animal Science, 26, 565–580.

Tien, A.T., Jolly, P.B., McNeil, V.H., Preda, M. and Cox, M.E. 2004. A comparison of salinity trends in the irrigated Lockyer Valley, Queensland, and the Upper Roper Catchment, Northern Territory. ICID 2004, 2nd Asian Regional Conference, International Commission on Irrigation and Drainage, march 14–17, 2004, Moama, NSW, Australia, 9 pp.

Wells, A.T. and O'Brien, P.E., 1993, Fluvial architecture of Triassic-Jurassic sediments of the Bundamba Group in the northern part of the Clarence-Moreton Basin, Queensland: Record, Australian Geological Survey Organisation, Canberra, 1993/45, 24 pp.

Wilson, A.S. 2004. Hydrogeology, conceptual model and groundwater flow within alluvial aquifers of the Tenthill and Ma Ma catchments, Lockyer Valley, Queensland. Master of Applied Science thesis (unpub), School of Natural Resource Sciences, Queensland University of Technology, Brisbane, Australia, 141 pp.

Wilson, A.W. and Cox, M.E. 2004. Hydrochemistry and stable isotopes as tools to determine hydrological pro- cesses in alluvial aquifers of the tenthill and Ma Ma catchments, Lockyer Valley, Queensland. 8th Australasian Environmental Isotope Conference, University of Melbourne, 29 November–3 December, 2004, Melbourne, Australia, 4 pp.

T O P I C 2 Geochemistr y dur ing inf iltration and flow 263

(20)

Abstract

The aim of the study was to determine the acidity and nutrient status of the humus and uppermost mineral soil layers of a forest soil after two years of sprinkling infiltration, and one and two years after cessation of the treatment. Sprinkling infiltration was carried out in a Scots pine stand located on an esker in Central Finland.

Humus and mineral soil samples were taken from two control plots and two plots subjected to sprinkling infil- tration. As a result of sprinkling infiltration, the base saturation of the humus layer was close to 100%, and the pH above 6. The cation exchange capacity of the humus layer had increased considerably owing to the increase in pH.

The humus layer, as well as to a lesser extent the uppermost mineral soil layers, had retained large amounts of Ca and Mg from the lake water used in infiltration, but there was no corresponding increase in exchangeable K con- centrations. The treatment strongly reduced the concentrations of extractable P, presumably due to immobil- ization at elevated pH values. Sprinkling infiltration appears to have a relatively long-lasting, positive effect on the acidity and nutrient status of the surface soil on the forested esker.

Keywords

Base saturation; cation exchange capacity; forest soil; macronutrient; pH; sprinkling infiltration.

I N T R O D U C T I O N

Artificial recharge of groundwater is an important method for producing household water in Finland. One possible way to recharge groundwater artificially is to infiltrate surface water through the forest soil and unsaturated percol- ation water zone down into the saturated groundwater zone. This kind of infiltration method is called sprinkling infiltration. Special attention has to be paid to environmental aspects when planning artificial recharge groundwater plants in forested areas. In sprinkling infiltration, the forest ecosystem is subjected to the addition of extremely large amounts of water, the chemical composition of which differs considerably from that of natural precipitation.

Appreciable changes in the pH and exchangeable Ca and Mg concentrations of the forest soil have been reported in a study carried out on a forested esker subjected to sprinkling infiltration in southern Finland (Lindroos et al., 2001). Sprinkling infiltration has also caused marked changes in nitrogen transformation in the same study area (Lindroos et al., 1998; Paavolainen et al., 2000).

Although information is already available about possible changes in the chemical properties of the forest soil due to surface water infiltration at some sites in Finland, more data are needed about the behaviour of these processes in different kinds of boreal forest stand and podzolic soils. It is also important to obtain more information about the duration of the changes in the chemical properties of the forest soil after surface water infiltration has ceased, and about the possible recovery of the soil to its original condition. The aim of this study was to determine the acidity and nutrient status of the humus layer of a forest soil at the end of the treatment, and one and two years after the cessation of sprinkling infiltration. The studied artificial recharge (AR) groundwater area is located in central Finland, and the AR groundwater plant produces household water for the city of Jyväskylä.

on a forested esker in Central Finland

J. Derome, A.-J. Lindroos and H.-S. Helmisaari

(21)

M E T H O D S

The study site is located in a 110 year-old Scots pine stand growing on an esker near to the city of Jyväskylä in Central Finland. The sub-xeric site is of the Vaccinium vitis-idaea forest site type (Cajander, 1949), and the soil con- sists of relatively coarse, stratified sand deposits. The soil type is a humic podzol. Soil samples were taken in autumn 2001, 2002 and in 2003 from two control plots and from two plots that had been subjected to sprinkling infiltration during 1999 –2001. The organic layer and mineral soil (0 –5, 5 –10, 10 –20 and 20 – 40 cm depth layers) were sampled at 20 systematically located points on each plot (30 x 30 m in size). The samples were combined to give 2 samples / layer/plot. The samples were dried at 60oC, and the organic layer samples then milled to pass through a 1 mm sieve and the mineral soil samples passed through a 2 mm sieve to remove stones and large roots.

pH was determined in a soil/distilled water slurry (15 ml soil /25 ml H2O). Exchangeable Ca, Mg, K and Na were determined by extracting the samples with 1 M BaCl2 (3.75 g organic layer sample or 15 g mineral soil /150 ml extractant), and exchangeable acidity (EA) by titrating the extract to pH 7. The element concentrations were deter- mined by inductively coupled plasma atomic emission spectrometry (ICP/AES). Cation exchange capacity (CEC) was determined as the sum of equilavent concentrations of Ca, Mg, K, Na and EA. Base saturation (BS) was calcu- lated as the proportion of base cations (Ca, Mg, K, Na) out of CEC.

R E S U LT S A N D D I S C U S S I O N

A

Acciid diittyy

During the two-year sprinkling irrigation treatment, the pH of the organic and mineral soil layers (Fig. 1) had increased to ca. 6, which is relatively close to that of the lake water (mean pH 6.9) used for AR. These pH values are very high compared to the values normally measured in the Finnish forest soils (Tamminen, 2000). During the two- year period after the cessation of the treatment the pH decreased to some extent (ca. 1.5 pH-units), but was still clearly above the level on the control plots. This increase is considerable considering the fact that the lake water was not sprinkled evenly over the whole plot (30 x 30 m). A similar, rather strong decrease in soil acidity has been observed in studies carried out at other sprinkling infiltration plants in Finland, and at these sites the pH has not returned to its original level even 5 years after the cessation of sprinkling infiltration (Helmisaari et al., 2003).

T O P I C 2 Geochemistr y dur ing inf iltration and flow 265

3.5 4.0 4.5 5.0 5.5 6.0 6.5

humus 0 – 5 cm 5 –10 cm 10 – 20 cm 20 – 40 cm

pH control

2001 2002 2003

Figure 1. Mean pH in the humus layer and at different depths in the mineral soil on the control plots, and on the treated plots at the cessation of sprinkling infiltration (2001),

and one year (2002) and two years (2003) after cessation of the treatment.

(22)

C

Ca attiio on n e ex xcch ha an ng ge e cca ap pa acciittyy a an nd d b ba asse e ssa attu urra attiio on n

Cation exchange capacity (CEC) is a measure of the number of negatively charged cation exchange sites in the soil, i.e. the soil’s capacity to bind cations, and hence CEC plays an important role in maintaining site fertility. The cations bound on the cation exchange sites include important plant nutrients such as Mg2+and Ca2+, as well as the acidic cations Al3+ and H+. In forest soils, which are usually relatively coarse textured and have a low content of fine material (e.g. silt and clay) in Finland, the majority of the cation exchange sites are associated with the organic matter in the humus layer and underlying mineral soil. As cation exchange sites are formed through the dis- sociation of so-called functional groups (e.g. carboxyls) in the organic matter, CEC in forest soils is strongly dependent on soil pH: the higher the pH, the higher the proportion of dissociated functional groups, and the greater the CEC value. Sprinkling infiltration considerably increased the CEC of the humus layer, and it remained at approximately the same level for two years after the cessation of the treatment (Fig. 2). Although the increase in pH (Fig. 1) undoubtedly explains part of the increase in CEC, changes will most probably also have occurred in the quality of the organic matter in the humus layer. The increase in available nitrogen (data not shown) and the high concentrations of both Ca and Mg, have probably stimulated microbial degradation of the relatively acidic, slowly decomposing coniferous litter layer characteristic of forest soils. The CEC in the mineral soil layers appeared to have been relatively unaffected by the treatment..

Base saturation is a measure used to depict the proportion of cation exchange sites occupied by so-called base cations (Ca2+, Mg2+, K+, Na+). The maximum value of BS, i.e. 100%, means that all the cation exchange sites are occupied by base cations. Sprinkling infiltration strongly increased BS in the humus layer to over 95% during the treatment, and there was no decrease following cessation of the treatment (Fig. 3). There was also an extremely strong increase in the mineral soil, and BS continued to increase following cessation of the treatment. This is undoubtedly due to the gradual movement of e.g. Ca and Mg cations down the soil profile, following their release in the overlying humus layer. The base cations originate from the lake water and, although the base cation con- centrations in the lake water were relatively low, the total amount of Ca and Mg added in the lake water during the two-year treatment is extremely large due to the large amounts of infiltrated water. These base cations have obviously displaced the other cations (e.g. H+and Al3+, data not shown) from cation exchange sites in the humus layer and mineral soil, thus resulting in an increase in soil pH (Fig. 1).

0 200 400 600

humus 0–5 cm 5–10 cm 10–20 cm 20–40 cm

CEC, meq/kg control

2001 2002 2003

Figure 2. Mean cation exchange capacity (meq/kg dm) in the humus layer and at different depths in the mineral soil on the control plots, and on the treated plots at the cessation of sprinkling infiltration (2001), and one year (2002)

and two years (2003) after cessation of the treatment. Note: owing to the large difference in the bulk density of the humus layer and mineral soil, the CEC values in the humus and mineral soil are not directly comparable.

(23)

E

Ex xcch ha an ng ge ea ab blle e cca allcciiu um m,, m ma ag gn ne essiiu um m a an nd d p po otta assssiiu um m cco on ncce en nttrra attiio on nss

The concentrations of exchangeable Ca and Mg in the humus layer increased 3-fold as a result of the sprinkling infiltration treatment (Table 1). As well as playing an important role in acid buffering processes in the humus layer, Ca and Mg are important macronutrients. It is clear that the increase in exchangeable Ca and Mg, together with the increased availability of nitrogen (data not shown), represents an improvement in the fertility of the site. Despite the fact that the soil is relatively coarse-textured, the increase in site fertility is likely to persist for a considerable length of time because the increase in CEC, and the migration of organic matter down into the underlying mineral soil (data not shown), will reduce the losses of Ca and Mg through leaching.

Despite the input of K in the lake water, the treatment has not significantly affected the exchangeable K concentr- ations in the humus layer (Table 1). This is not a surprising result because K+has a relatively low affinity for cation exchange sites (it is a monovalent cation) and it is not able to compete with the elevated Ca2+and Mg2+concentr- ations (Bohn et al., 1985).

Sprinkling infiltration has strongly reduced the concentrations of extractable P in the humus layer. Despite the fact that there was also an input of P in the lake water, it would appear that the relatively high pH values (above pH 7) in the humus layer have resulted in the immobilization of P (Hartikainen, 1979).

T O P I C 2 Geochemistr y dur ing inf iltration and flow 267

Control

Year

2001 2002 2003

Ca, mg/kg 2516 7313 8033 7605

Mg, mg/kg 308 1003 1030 2198

K, mg/kg 694 638 625 506

P, mg/kg 143 ND 53 55

Table 1. Mean exchangeable Ca, Mg and K and extractable P concentrations in the humus layer on the control plots, and on the treated plots at the cessation of sprinkling infiltration (2001), and one year (2002) and two years (2003) after cessation of the treatment

ND = not determined.

40 60 80 100

humus 0–5 cm 5–10 cm 10–20 cm 20–40 cm BS, %

control 2001 2002 2003

Figure 3. Mean base saturation (%) in the humus layer and at different depths in the mineral soil on the control plots, and on the treated plots at the cessation of sprinkling infiltration (2001),

and one year (2002) and two years (2003) after cessation of the treatment.

(24)

C O N C L U S I O N S

Sprinkling infiltration had a relatively strong effect on the acidity and fertility status of the humus and uppermost mineral soil layers, and it appears that there will not be a very rapid return to the soil conditions prevailing before the treatment. However, the treatment considerably increased the capacity of the soil to withstand acid loads, and increased the concentrations of the important macronutrients Ca and Mg. The only negative effect of sprinkling infiltration appears to be the decrease in the important plant nutrient P, the availability of which decreased owing to the increase in pH. Whether or not this represents a threat to site fertility requires further study.

R E F E R E N C E S

Bohn, H.L., McNeal, B.L. and O’Connor, G.A. (1985). Important ions. In: Soil chemistry (2nd edition), John Wiley &

Sons, New York, pp. 290–328.

Cajander, A.K. (1949). Forest types and their significance. Acta Forestalia Fennica, 56, 1–69.

Hartikainen, H. (1979). Phosphorus and its reactions in terrestrial soils and lake sediments. J. Sci. Agric. Soc. Fin., 51, 537–624.

Helmisaari H-S., Illmer K., Hatva T., Lindroos A-J., Miettinen I., Pääkkönen J. and Reijonen R. (eds) (2003).

Tekopohjaveden muodostaminen: imeytystekniikka, maaperäprosessit ja veden laatu. Metsäntutkimuslaitoksen tiedo- nantoja,902.

Lindroos A-J., Paavolainen L., Smolander A., Derome J. and Helmisaari H-S. (1998). Changes in nitrogen trans- formations in forest soil as a result of sprinkling infiltration. Environ. Pollut., 102, 421–426.

Lindroos A-J., Derome J., Paavolainen L. and Helmisaari H-S. (2001). The effect of lake-water infiltration on the acidity and base cation status of forest soil. Water, Air and Soil Pollut., 131(1–4), 153–167.

Paavolainen L., Smolander A., Lindroos A-J., Derome J. and Helmisaari H-S. (2000). Nitrogen transformations and losses in forest soil subjected to sprinkling infiltration. J. Environ. Qual., 29, 1069–1074.

Tamminen P. (2000). Soil factors. In: Forest Condition in a Changing Environment –The Finnish Case, E. Mälkönen (ed.), Kluwer Academic Publishers, Forestry Sciences, 65, 72– 86.

(25)

Abstract

Nitrogen removal has been sustained in numerous Soil Aquifer Treatment (SAT) systems where ammonia was the primary form of nitrogen in the reclaimed water. During SAT, groundwater is recharged using percolation basins and treatment may occur during percolation in the vadose zone and subsequent transport under saturated con- ditions in the aquifer. Since there is insufficient organic carbon to support heterotrophic denitrification, autotrophic nitrogen removal mechanisms capable of sustaining nitrogen removal were examined. In particular, anaerobic ammonia oxidation (ANAMMOX) was evaluated using column studies and batch studies. Nitrogen removal was sustained in column studies for over 600 days without the addition of supplemental carbon. When nitrate alone was fed to the columns, ammonia pre-adsorbed on the soil was bioavailable and sustained nitrogen removal. When a combination of nitrate and ammonia were fed, nitrate removal rates were similar to the soil column with pre-adsorbed ammonia. When nitrite and ammonia were fed to the soil columns, nitrogen removal rates increased by an order of magnitude as compared to a mixture of nitrate and ammonia. This suggests that the conversion of nitrate to nitrite is the rate-limiting step and soil systems have the capability of converting nitrate to nitrite.

Keywords

Anaerobic ammonia oxidation, nitrate, nitrite.

I N T R O D U C T I O N

Soil Aquifer Treatment has been used extensively for the recharge of groundwater using reclaimed water in arid areas. During Soil Aquifer Treatment (SAT), reclaimed water is percolated through the vadose zone and biological removal mechanisms remove contaminants of concern including organic carbon and nitrogen. This study focuses on how nitrogen removal may be sustained in SAT systems. Nitrogen removal efficiencies of greater than 70% have been sustained using secondary effluent with ammonia nitrogen concentrations in excess of 20 mg-N/L. There is insufficient organic carbon in secondary effluent to sustain the observed nitrogen removal efficiencies. Furthermore, the majority of organic carbon is removed near the soil surface where aerobic conditions exist and conditions are inappropriate for denitrification. During SAT, ammonia is primarily removed by adsorption when water is applied since there is insufficient oxygen to support nitrification. When water is not applied, air enters the soil column and nitrification of adsorbed ammonia will occur. During subsequent additions of reclaimed water, ammonia and nitrate are present in the soil column under anoxic conditions thus creating conditions appropriate for ANAMMOX.

Anaerobic ammonium oxidation may be the oxidation of ammonium with nitrate in the absence of oxygen and organic carbon (Mulder et al. 1995).

+ (1)

+ + 3NO 4N + 9H O + 2H

5NH4 3 2 2

Anaerobic ammonia oxidation during sub-surface transport

Peter Fox and Shivani Shah

VV

Références

Documents relatifs

The model max- imizes the total demand served, locates several facilities with multiple servers in each one, uses a practical demand curve, and measures congestion by counting

Existe-t-il un triangle ABC dont le périmètre est égal à 134 centimètres, le rayon du cercle inscrit est égal à 12 centimètres et le rayon du cercle circonscrit est égal à

[r]

On va utiliser différentes expressions de l’aire S du triangle pour extraire 3 grandeurs a, b, c susceptibles d’être les côtés

Une condition necessaire et suffisante pour existence d'une triangle avec semipérimètre s, le rayon du cercle inscrit r et le rayon du cercle circonscrit R est. On a

New Brunswick is increasing its training of medi- cal students with the opening of the Saint John Campus of Dalhousie University’s medical school.. At the same time,

To  honour  those  Canadian  family  physicians  who  contribute  so  much  to  improve  global  health  and  to  acknowledge  this  special  issue, 

2 Until a refrigerator-stable vaccine becomes available, however, varicella vac- cine will not be incorporated into the recommend- ed immunization schedule in Canada, as most