> Environmental studies 09 07 > Air > Critical Loads of Acidity for Alpine Lakes A weathering rate calculation model and the generalized First-order Acidity Balance (FAB) model applied to Alpine lake catchments > Environmental studies > Air > Critical Loads of Acidity for Alpine Lakes A weathering rate calculation model and the generalized First-order Acidity Balance (FAB) model applied to Alpine lake catchments Published by the Federal Office for the Environment FOEN Bern, 2007 Impressum Editor Federal Office for the Environment (FOEN) FOEN is an agency of the Federal Department of Environment, Transport, Energy and Communications (DETEC). Authors Maximilian Posch Urs Eggenberger Daniel Kurz Beat Rihm CCE, MNP, Bilthoven/NL Institute of Geological Sciences, University of Bern EKG Geo-Science, Bern METEOTEST, Bern FOEN Consultant Beat Achermann, Air Pollution Control and Non-Ionizing Radiation Division Suggested Form of Citation Posch M., Eggenberger U., Kurz D., Rihm B. 2007: Critical Loads of Acidity for Alpine Lakes. A weathering rate calculation model and the generalized First-order Acidity Balance (FAB) model applied to Alpine lake catchments. Environmental studies no. 0709. Federal Office for the Environment, Berne. 69 S. Layout Dominik Eggli, METEOTEST, Bern Cover Picture Chiara Pradella, Lago di Tomè Downloadable PDF file www.environment-switzerland.ch/uw-0709-e (no printed version available) Code: UW-0709-E © FOEN 2007 3 > Table of Contents Table of Contents Abstracts Preface Summary 5 7 8 1 Background 10 2 2.1 2.2 14 14 2.4 2.4.1 2.4.2 2.4.3 2.4.4 Methods Procedure Overview Generalisation of the First-order Acidity Balance (FAB) model Model derivation Input data requirements The Steady-State Water Chemistry (SSWC) model Calculation of Weathering Rates for Catchments Introduction Calibration of Hydrology Calibration of the Reactive Transport of Ions Transfer Functions for the Regional Application 3 3.1 3.2 3.3 3.4 3.5 Input Data Deposition Rates Runoff Weathering Rates Terrestrial Sinks of Nitrogen and Base Cations In-lake Retention 42 42 45 47 50 51 4 Results and Discussion 52 5 Concluding Remarks 58 2.2.1 2.2.2 2.3 Acknowledgements 18 18 23 24 26 26 26 34 39 59 Annexes A1 List of Lakes A2 FORTRAN subroutine genFAB 60 60 63 Indexes Glossary Figures Tables References 66 66 66 67 68 5 > Abstracts > Abstracts In alpine lakes in Southern Switzerland acid deposition is a problem due to slowweathering bedrocks and thin soils. Earlier assessments of critical loads of acidity for these lakes with the Simple Mass Balance (SMB) model and the Steady-State Water Chemistry (SSWC) model led to different results, due to differences in quantifying the weathering of base cations (BC). In this study, a hydrological model was used to quantify the typical groundwater flow through the prevalent bedrock types. A reactive transport model supplied information for estimating the average weathering rates for five lithological classes of bedrock, which are the dominating source of base cations in these catchments. For calculating the critical loads for sulphur and acidifying nitrogen a generalised version of the First-order Acidity Balance (FAB) model was derived, in which BC leaching is explicitly formulated in terms of sources and sinks in the catchment. The generalised FAB model was applied to 100 catchments. The resulting critical loads were compared with the outcome of the SSWC model, which was applicable to 19 lakes for which water-chemistry measurements are available. Overall, the new methodology for calculating critical loads has the advantage of being more processoriented, differentiating better between catchments, and allowing the comparison with S and N depositions. The atmospheric depositions were modelled for 1980, 1995 and 2010. The percentage of lakes protected (i.e. critical loads are not exceeded) increases from 46 % in 1980 via 57 % in 1995 to 73 % in 2010. Keywords: In alpinen Seen der Südschweiz stellen saure Niederschläge wegen des langsam verwitternden Muttergesteins und der dünnen Böden ein Problem dar. Frühere Berechnungen kritischer Eintragsgrenzen (critical loads, CL) für Säure mit der Simple Mass Balance (SMB) und dem Steady-State Water Chemistry (SSWC) Modell führten zu verschiedenen Ergebnissen, da die Verwitterung basischer Kationen (BC) unterschiedlich behandelt wurde. In dieser Studie wurde ein hydrologisches Modell für die Quantifizierung typischer Grundwasserflüsse im Felsuntergrund eingesetzt. Ein reaktives Transportmodell lieferte Abschätzungen der durchschnittlichen Verwitterungsraten für fünf Lithologieklassen, der Hauptquelle von BC in diesen Seen. Zur Berechnung von CL wurde eine verallgemeinerte Version des First-order Acidity Balance (FAB) Modells entwickelt. Dabei wird die BC-Auswaschung explizit als Funktion von Quellen und Senken im Einzugsgebiet formuliert. Das generalisierte FAB Modell wurde auf 100 Seen angewandt. Die resultierenden CL wurden mit dem Ergebnis des SSWC Modells verglichen, welches in 19 Seen mit Wasserchemie-Messungen eingesetzt werden konnte. Insgesamt ist die neue Methode zur Berechnung von CL mehr prozessorientiert, differenziert besser zwischen den Einzugsgebieten und erlaubt den Vergleich mit S- und N-Einträgen. Die atmosphärischen Einträge wurden für die Jahre 1980, 1995 und 2010 modelliert. Der Anteil der geschützten Seen, in welchen die CL nicht überschritten werden, stieg von 46% (1980) auf 57% (1995) und auf 73% (2010). Stichwörter: alpine lakes acid deposition critical loads of acidity lithology base cations weathering FAB model alpine Bergseen saure Niederschläge kritische Eintragsraten Lithologie basische Kationen Verwitterung FAB Modell 6 Critical Loads of Acidity for Alpine Lakes FOEN 2007 Les pluies acides constituent un problème dans les lacs alpins du sud de la Suisse, parce qu’elles altèrent lentement le substratum rocheux et que les sols y ont une faible épaisseur. Les charges critiques d’acidité (critical loads, CL) calculées autrefois en appliquant un bilan massique simple (simple mass balance / SMB) et le modèle de chimie des eaux à l’état d’équilibre (steady-state water chemistry / SSWC) divergeaient car l’altération des cations basiques était traitée différemment. La présente étude met en œuvre un modèle hydrologique servant à quantifier des écoulements typiques à l’intérieur du soubassement rocheux. Un modèle de transport réactif a permis d’estimer les taux moyens d’altération pour les cinq classes lithologiques fournissant le plus de cations basiques aux lacs étudiés. Une version généralisée du modèle du bilan d'acidité du premier ordre (first-order acidity balance / FAB) a été développée pour calculer les charges critiques. Le lessivage des cations basiques y est exprimé explicitement en termes de sources et de puits présents dans le bassin versant. Le modèle généralisé a été mis en œuvre sur 100 captages. Les charges critiques en résultant ont été comparées avec les résultats du modèle SSWC, qui a pu être appliqué à 19 lacs pour lesquels on disposait de mesures hydrochimiques. Dans l’ensemble, la nouvelle méthode de calcul des charges critiques est axée davantage sur les processus, différencie mieux les bassins versants et permet d’opérer des comparaisons avec les dépôts soufrés et azotés. Les dépôts atmosphériques ont été modélisés pour les années 1980, 1995 et 2010. La proportion de lacs protégés, où les charges critiques n’ont pas été dépassées, est passée de 46 % en 1980 à 57 % en 1995, pour atteindre 73 % en 2010. Nei laghi alpini della Svizzera meridionale le piogge acide costituiscono un problema, a causa della roccia madre degradabile solo lentamente e del sottile strato di suolo. In passato, il calcolo dei carichi critici di acidità (critical loads, CL) con il metodo SMB (Simple Mass Bilance) e il modello SSWC (Steady-State Water Chemistry) ha prodotto risultati discordanti, riconducibili a differenze nella quantificazione del rilascio di cationi basici dalla roccia madre durante il suo degrado. In questo studio, è stato utilizzato un modello idrologico per quantificare i flussi tipici delle acque sotterranee attraverso i principali substrati rocciosi. Grazie ad un modello di trasporto reattivo, è stato possibile stimare un tasso di degrado medio per cinque classi litologiche, le quali rappresentano la principale fonte di cationi basici nel bacino imbrifero. Per calcolare i carichi critici di zolfo e di azoto acidifiante è stata sviluppata una versione generalizzata del modello FAB (First-order Acidity Balance), in cui il dilavamento di cationi basici viene esplicitamente formulato come una funzione della loro produzione ed eliminazione. I risultati ottenuti dall’applicazione di tale modello su 100 laghi sono stati confrontati con quelli derivanti dal modello SSWC, quest’ultimo applicato a 19 laghi per i quali erano disponibili misurazioni idrochimiche. Nel complesso, il nuovo metodo di calcolo dei carichi critici a il vantaggio di essere più incentrato sui processi differenziando meglio i singoli bacini imbriferi e permettendo un confronto con gli apporti di zolfo e di azoto. Gli apporti di inquinanti atmosferici sono stati modelizzati per gli anni 1980, 1995 e 2010. La percentuale dei laghi protetti, in cui i carichi critici non vengono superati, aumenta dal 46 per cento (1980) al 57 per cento (1995) e al 73 per cento (2010). Mots-clés : lacs alpins pluies acides charges critiques d’acidité lithologie cations basiques altération modèle FAB Parole chiave: laghi alpini piogge acide carichi critici di acidità litologia cationi basici alterazione modello FAB > Preface > Preface Critical loads play an important role within the Convention on Long-range Transboundary Air Pollution (UNECE) and its Protocols on Further Reduction of Sulphur Emissions (Oslo Protocol, 1994) and to Abate Acidification, Eutrophication and Ground-level Ozone (Gothenburg Protocol, 1999). Critical loads, defined as “a quantitative estimate of an exposure to one or more pollutants below which significant harmful effects on specified sensitive elements of the environment do not occur according to present knowledge”, are the scientific rationale for the development of effects-based air pollution control strategies. For the preparation of the above mentioned Oslo and Gothenburg Protocols, Switzerland applied the so-called “Steady-state (Simple) Mass Balance” approach to calculate critical loads of acidity for alpine lakes and forest ecosystems as proposed at that time in the Convention’s Manual on “Methodologies and Criteria for Mapping Critical Levels/Loads and Geographical Areas where they are Exceeded”. Since the Modelling and Mapping Programme under the Convention now requires the development of critical load functions of acidifying sulphur and nitrogen, the method of choice is the “First-order Acidity Balance” (FAB) model. In cooperation with the Coordination Centre for Effects (CCE) a generalized version of the FAB model was developed, allowing the calculation of critical load functions by taking entire lake catchment properties into account. In addition, substantial improvements could be made in quantifying the weathering of base cations in catchments. Overall, the new methodology for calculating critical loads has the advantage of being more process-oriented, differentiating better between catchments and allowing the comparison with sulphur and nitrogen deposition. The results of the application of the improved weathering rate determination and the generalized FAB model to 100 lake catchments are now part of the Swiss critical loads data set in use for the review of the Gothenburg Protocol. We express our warm thanks to the Coordination Centre for Effects for its support in deriving the generalized FAB model and to all scientists and engineers being involved in the revision of the Swiss data set on critical loads of acidity for alpine lakes. Martin Schiess Head of the Air Pollution Control and Non-Ionizing Radiation Division Federal Office for the Environment (FOEN) 7 Critical Loads of Acidity for Alpine Lakes FOEN 2007 > Summary Acidification of surface waters as a result of deposition of acidifying air pollutants, mainly sulphur and nitrogen, has primarily been witnessed in northern Europe and North America. The release of base cations (BC = Ca2+ + Mg2+ + K+ + Na+) from minerals in the soil and bedrock due to weathering is the major long-term acidity buffering process. Surface waters in central Europe are often buffered by ubiquitous carbonate bedrock. Among the exceptions are high-alpine lakes and rivers of the Lago Maggiore catchment area, which lies half in Southern Switzerland. The bedrock of this region consists of slow-weathering crystalline basement nappes. Sensitive ecosystems regarding acidification became important when acidification was perceived as a consequence of transboundary air pollution within the scope of the UNECE Convention on Long-range Transboundary Air Pollution (LRTAP). Effect and pollutant deposition were linked via the critical load which was defined as «a quantitative estimate of an exposure to one or more pollutants below which significant harmful effects on specified sensitive elements of the environment do not occur according to present knowledge”. Earlier assessments of critical loads of acidity (CL(A)) for alpine lakes in southern Switzerland with the Simple Mass Balance (SMB) model and the Steady-State Water Chemistry (SSWC) model led to substantially different results. The discrepancies were clearly related to differences in the weathering rates (BCw), independently derived from soil and geological information with the SMB and back-calculated from present-day water chemistry with the SSWC. Therefore, this study aims at (1) improving the derivation of weathering rates from catchment properties, (2) developing a more appropriate calculation model instead of the SMB model and (3) applying that model on a regional scale in order to provide more reliable critical loads for Swiss alpine lakes: (1) A detailed study of two lakes involving models for hydrology (groundwater flow) and reactive transport (rock-water interaction) supplies the quantitative information to develop so-called transfer functions. The transfer functions can be used to calculate the average BC weathering of a catchment based on maps and data, which are available on a regional scale, i.e. geological maps, terrain and precipitation data. The units of the geological maps are aggregated to five lithological classes (carbonate bearing rocks, amphibolite, melanocratic granite/gneiss, leucocratic granite/gneiss and quaternary cover). Terrain data are used for calculating hydraulic gradients of the assumed flow paths within the bedrock. The weathering contributed from the marginal soils found in these steep alpine catchments was assumed to be negligible in comparison with the release of base cations to the groundwater percolating the bedrock. (2) Since the Mapping Programme under the LRTAP Convention requests critical loads for sulphur, CLmax(S) and acidifying nitrogen, (CLmax(N)), the model of choice is the First-order Acidity Balance (FAB) model. For this study, a generalised version of 8 > Summary the FAB model has been derived. It differs from the previously published version of FAB in the following points: Base cation leaching is explicitly formulated in terms of sources and sinks in the catchment (deposition, weathering and removal due to uptake by vegetation etc.), instead of plugging in the SSWC model. The number of sub-areas (e.g., land cover classes) in the catchment, for which different fluxes can be specified, is now unrestricted. Individual depositions to different sub-areas of the catchment can be taken into account. (3) In the regional application in Southern Switzerland input data for the transfer functions and for the generalised FAB model are compiled and critical loads are calculated for 100 lake catchments. The lakes are situated between 1650 and 2700 m altitude. The mean lake area is 4 ha and the mean catchment area is 87 ha. The lithology is dominated by granite/gneiss. Carbonate bearing rocks occur in only 20 % of the catchments. Precipitation amounts are in the range of 1.6–2.4 m a-1. The dominating landuse type is bare land (rocks, gravel, glaciers). Grassland covers 35 % of the catchments on the average, and only 2 % is covered by forests. In the 100 lakes, the median value of the resulting weathering rates (BCw) is 596 eq haa-1. The median values for CLmax(S) and CLmax(N) are 570 and 800 eq ha-1 a-1, respectively. 1 The atmospheric deposition of base cations, N and S is calculated with a generalised combined approach for the year 1995. The approach is based on measurements of wet deposition and on emission inventories combined with statistical dispersion models to calculate dry deposition. In the considered catchments, dry deposition contributes only a small part to the total deposition (< 20 %). Using EMEP-results, the 1995 deposition values were rescaled to the years 1980 and 2010. In this period, the emissions and depositions, especially of sulphur, changed substantially, and consequently also the exceedances of the critical loads. The percentage of lakes protected (i.e. non-exceeded) increases from 46 % in 1980 via 57 % in 1995 to 73 % in 2010. For 19 lakes with available water chemistry measurements, also the SSWC model is applied. A comparison of the CL(A) values resulting from the SSWC model with the values for CLmax(S) obtained with the FAB model shows that the SSWC critical loads are generally higher. The divergence must be explainable by the difference in base cation inputs to the catchment (deposition, weathering, uptake), which is used in the generalised FAB model, and the observed base cation flux leaving the lake, which determines the results of the SSWC model. The measured flux of base cations leaving the lake is for about half of the catchments considerably larger than the modelled net input. Also in the balances of input-output fluxes for S and N biases can be observed. Although possible reasons for the differences were identified, these discrepancies warrant further investigations. Overall, the new methodology for calculating critical loads for lakes, the generalised FAB model, has the advantage of (a) being more process-oriented and thus easier to modify or improve, (b) differentiating better between catchments, and (c) allowing the comparison with S and N depositions. 9 Critical Loads of Acidity for Alpine Lakes FOEN 2007 1 > Background 1.1 Introduction Wide-spread acidification of surface waters as a result of deposition of acidifying air pollutants has primarily been witnessed in northern Europe and North America. The main reasons for the susceptibility to acidification of these areas are the climate and the bedrock. Low average annual air temperatures and the predominance of acid silicate mineralogy impair soil formation, and consequently lead to low weathering rates. The release of base cations from minerals in the soil due to weathering is the major long-term acidity buffering process. Unlike in northern Europe, surface waters in central Europe are often buffered by ubiquitous carbonate bedrock. Among the exceptions are high-alpine lakes and rivers of the Lago Maggiore catchment area, which lies half in north-western Italy and half in Southern Switzerland. The bedrock of this area consists of crystalline basement nappes with varied ortho- and paragneiss dominating. This particular geological and climatic environment implies that the sensitivity of headwater lakes and streams above the timberline should be comparable to that of Nordic lake catchments. Sensitive ecosystems regarding acidification became important when acidification was perceived as a consequence of transboundary air pollution and effect related abatement of transboundary air pollution was adopted within the scope of the 1979 Convention on Long-range Transboundary Air Pollution (LRTAP) involving the territory of the United Nations Economic Commission for Europe (UNECE). Effect and pollutant exposure were linked via the critical load which was defined as «a quantitative estimate of an exposure to one or more pollutants below which significant harmful effects on specified sensitive elements of the environment do not occur according to present knowledge» (Nilsson and Grennfelt 1988). The concept was used since the early 1990’s to produce European maps of critical sulphur deposition, which became the basis in the development of sulphur emission reduction scenarios used in the negotiations of the Second Sulphur Protocol (UNECE 1994). 1.2 Earlier Assessments Switzerland participated in this work by compiling critical loads of acidity and sulphur for Swiss terrestrial (forest soils) and aquatic (Alpine lakes) ecosystems (FOEFL 1994). At that time, with respect to the data available in Switzerland and in agreement with procedures recommended in the mapping manual, critical loads were calculated with a simplified steady state mass balance method (SMB, Hettelingh et al. 1991). In Switzerland, also surface water critical loads were assessed with the SMB model, which allowed 10 1 11 > Background considering all potentially sensitive lake catchments covering a total catchment area of 600 km². The critical load of actual acidity, CL(A), was calculated as: (1.1) CL( A) = BC w,C − ANC le ,crit where the average base cation weathering rate of the catchment BCw,C (in eq m-2 a-1) was derived from: (1.2) BC w,C = BC w,class ⋅ d C ⋅ 10 ⎛1 1 ⎞ − 3800 ⎜ − ⎟ ⎝ T 283 ⎠ and the critical leaching of acid neutralizing capacity ANCle,crit (in eq m–2 a–1) was obtained by multiplying the annual average runoff rate Q (in m a–1) with the critical ANC concentration in the runoff water (set at 0.02 eq m–3). BCw,class (in eq m–3 a–1) in the above equation refers to soil type specific weathering rate classes taken from Hettelingh and De Vries (1991). Class averages were allocated to the soil types found in the 1:500,000 soil map of Switzerland (swisstopo 1984). dC (in m) refers to the annual average hydrologically active soil depth in the catchment (set to 1 m), and T (in K) is the annual mean soil temperature at a depth of 0.2 m. Due to inherent problems in directly assessing catchment weathering rates, the generally used method to map critical loads of acidity for surface waters was at that time, however, the Steady-State Water Chemistry (SSWC) model (Henriksen et al. 1990; Brakke et al. 1990; see also Chapter 2.3). The SSWC model was developed and widely applied in the Nordic countries, where the required input was available. With the SSWC model, the critical load of acidity can be derived from annual mean present-day water chemistry, assuming all sulphate in the runoff to originate from sea-salt spray and anthropogenic deposition. Basically, the acid load should not exceed the pristine, non-marine, nonanthropogenic base cation runoff (flux) from the catchment minus a buffer ([ANC]limit) to protect selected biota from being damaged: (1.3) CL( A) = Q⋅ ([ BC * ]0 − [ ANC ]limit ) It is assumed that the catchments were in steady-state regarding deposition inputs during pre-industrial times. The difference between present ([BC*]t) and pristine ([BC*]0) concentration of base cations (BC=Ca+Mg+K+Na) in the surface water is related to the longterm changes in the concentration of strong acid anions by the so called F-factor: (1.4) [BC * ]0 = [BC * ]t − F ⋅ ([SO4* ]t − [SO4* ]0 + [NO3* ]t − [NO3* ]0 ) The F-factor and the historic non-marine sulphate concentration ([SO4*]0 ) are approximated with two empirical functions, both having originally been calibrated in Norway: Critical Loads of Acidity for Alpine Lakes FOEN 2007 (1.5) ⎧sin π [BC * ] [S]) if [BC * ] < [S] t t F = ⎨ (2 1 else ⎩ and (1.6) [ SO4* ]0 = min{[ SO4* ]t ,0.015 + 0.16 ⋅ [ BC * ]t } where [S] = 0.4 eq m–3, assuming all concentrations in the above equations to be in this unit. Non-marine pristine inorganic nitrate concentration (if nitrate was considered in the calculation, see Figure 1) was set to zero. The water chemistry of a series of lakes in the Southern Swiss Alps has irregularly been surveyed since the early 1980s. This data allowed De Jong (1996) to calculate critical loads of acidity for Swiss high-alpine catchments using the SSWC method. The basic findings of this study, i.e. substantial differences between critical load estimates generated with the SMB method and the SSWC method, respectively, could be substantiated (EKG 1997). The discrepancies were found to be clearly related to differences in the weathering rates, independently derived from soil and geological information with the SMB and back-calculated from present-day water chemistry with the SSWC. SMB weathering rates, and consequently critical loads, were found to be without exception much lower than SSWC estimates (Figure 1). Additionally, in the set of lakes considered, there was practically no variation, i.e. SMB weathering rates were all around 360 eq ha–1 a–1. This was considered to be primarily a result of the insufficient resolution of the data source (soil map of Switzerland 1:500,000) used to derive the parent material units, from which catchment weathering rates were calculated. In view of other studies, e.g. Zobrist and Drewer (1990), SSWC estimates appeared to be quite high with around 45 to 55 % of the weathering rates and around 30 to 40 % of the critical acid loads above 1500 eq ha–1 a–1. Correction for an acidification-induced increase of cation exchange was found to reduce critical loads of acidity by up to 24 % (difference between the H90 and the B90 approach) and by up to 53 % (difference between the H93 and the B90 approach). Nonetheless, there was also some basic concern about a default application of the (original) SSWC method to high-alpine catchments. Among the reasons were (1) the importance of current lake water chemistry in the model, which is known to vary annually and seasonally (A. Barbieri, pers. comm.) as well as with lake depth (LSA 1999), (2) the assumption that the catchments are at steady-state with respect to current sulphur deposition, and (3) the assumption that the empirical functions derived in Norway can be applied to alpine catchments. 12 1 > Background 13 Figure 1 > Differences in the distribution of catchment weathering rates and critical loads of acidity of 45 high-alpine lakes in the Ticino area (EKG, 1997). Water chemistry data adopted from De Jong (1996). SMB results adopted from FOEFL (1994); H90: SSWC after Henriksen et al. (1990) considering only sulphate as strong acid anion); B90: simplified SSWC after Brakke et al. (1990) with F=0 (i.e. using present instead of pristine base cation concentration); H93: SSWC after Henriksen et al. (1993) considering both sulphate and nitrate as strong acid anion (see Equation 1.4). 1.3 Aims of this Study Considering these findings and the continual request for updated critical loads, it became prudent to revise the methodology and database used to estimate surface waters critical loads for Switzerland. Since regionalization is a crucial aspect in critical load calculations, and water chemistry is not available for all catchments of interest in the Ticino area, focus was put on improving the methodology to directly estimate catchment weathering rates from mapped catchment properties. Chapter 2.4 of this report describes the conceptual framework and the modelling approach used to calibrate mapped catchment units. Since critical loads are now requested for sulphur and acidifying nitrogen, the model of choice is the First-order Acidity Balance (FAB) model. Chapter 2.2 of the report explains the changes needed in the FAB model to allow the use of independently estimated weathering rates. Chapters 3 and 4 discuss the input data used to do the critical loads calculations and the results obtained. Critical load functions for sulphur and nitrogen were calculated for 100 sensitive lakes in the Ticino area, and the results were submitted to the Coordination Centre for Effects (CCE) of the International Cooperative Programme on Modelling & Mapping in response of the 2004 Call for Data. 14 Critical Loads of Acidity for Alpine Lakes FOEN 2007 2 > Methods 2.1 Procedure Overview Under the Convention on LRTAP, all calculations and mapping of critical loads follow a basic formal procedure (Table 1), which was outlined by Sverdrup et al. (1990). More detailed and updated information about the single steps are compiled in the so called UNECE Mapping Manual (UBA, 2004). Table 1 also summarises the main selections and methodical decisions made in this study as well as references to the detailed descriptions for each step. Table 1 > Workflow for calculating and mapping critical loads. formal procedure selected procedure or item in this study references 1. select receptor type alpine lakes sensitive to acidification chapter 1 2. quantify receptor distribution and abundance 100 lake catchments in Southern Switzerland chapter 2.1 3. determine biological indicator (indicator organisms) fish, invertebrate UBA 2004 4. determine critical chemical value that does not damage the selected biota Acid Neutralising Capacity (ANC) in lake water: [ANC]limit = 20 meq m–3 UBA 2004 5. select/develop computational approaches critical loads: generalized FAB (First-Order Acidity Balance) and SSWC (Steady-state Water Chemistry) models; weathering rates: simplified hydrology model and reactive transport model chapter 2.2 and 2.3 chapter 2.4 6. collect required input data for each receptor deposition (N, S, BC) in 1980, 1995 and 2010, runoff, land use, terrestrial BC and N sinks, catchment properties (land use, lithological units, topographic parameters), water chemistry measurements chapter 3 7. conduct calculation of critical loads and exceedance FAB model for 100 lakes. SSWC model for 19 lakes where water chemistry is available 8. produce maps/statistics critical load functions, cumulative frequency distributions chapter 4 9. check assumptions and quality comparison of the results from FAB and SSWC, N and S budgets chapter 4 After Sverdrup et al. 1990 The occurrence and abundance of potentially sensitive lakes is determined in several steps. First, all water surfaces were extracted from topographic maps within the region shown in Figure 2. According to former studies (see Chapter 1), slow weathering bedrocks can be expected in this region, which covers the Northern part of the canton Ticino and neighbouring areas of other cantons. Second, the Alpine lakes used in this study were selected from the surface waters by meeting the following criteria: 2 15 > Methods – – – – altitude is over 1500 m a.s.l. lake area is greater than 0.5 ha no storage lakes large scale geological map (1:25’000) is available for the catchment Finally, 100 lakes in Southern Switzerland were available for calculating critical loads with the FAB-model (Figure 2, Table 17 in the Annex). They are located at altitudes between 1650 and 2700 m (average 2200 m a.s.l.). Figure 2 > Map of the 100 lakes modelled with genFAB. The 19 lakes with water chemistry measurements during 2000–2003 are shaded in blue color. 2 3 5 11 10 16 19 18 21 24 2526 23 31 205 2830 215 33 20 22 36 3537 38 220 44 217 46 204 201 47 48 49 51 203 54 55 57 218 214 216 60 202 64 65 219 67 707363 68 66 71 76 77 78 79 80 81 222 221 207 84 22388 224 92 206 9 12 17 27 15 4143 50 225 53 72 74 85 86 89 93 9495 20898 99 212 100 102 211 109210 110 209 106 101 213104 K606-01©2004 swisstopo FAB-sites, ID Monitoring 2000-03 EMEP 50x50 km, cell 70/38 0 5 10 km rev. 14.9.2005 Critical Loads of Acidity for Alpine Lakes FOEN 2007 The indicator organisms (fish, invertebrate) and the critical chemical value (ANC concentration greater than 20 meq m–3) are selected following the recommendations in the Mapping Manual (UBA 2004). Following a proposal of ICP-Waters (2000, p. 66) for high altitude lakes, also a [ANC]limit of 30 meq m–3 was tested. However, the results were less plausible, as critical loads became negative for several lakes. Figure 3 shows a flowchart of steps 5 to 7 of Table 1: A detailed study in two catchments involved the calibration of models for hydrology (groundwater flow) and reactive transport (rock-water interaction). The rationale of this procedure is that soils in these catchments are very thin or even absent and therefore the main contribution of base cation weathering does not come from soils but from groundwater percolating the bedrock. As a main result, this ‘calibration’ study provided simplified transfer functions for calculating the average weathering rate of a catchment depending on five classes of lithology and corresponding hydraulic gradients. The transfer functions are used in the regional application to calculate the average weathering rates (BCw,C) in 100 catchments. Within the catchments the lithology classes were derived from geological maps and the hydraulic gradients were calculated from digital terrain data. BCw,C along with input maps related to land-use, deposition, terrestrial sinks and runoff are compiled to apply the generalized FAB model. For 19 lakes, water chemistry measurements are available (Barbieri 2004, see Table 17 in the annex), enabling the SSWC method to be applied as an alternative approach for calculating critical loads. 16 2 17 > Methods Figure 3 > Flowchart of the main procedural steps in this study. Calibration Sites (2 lakes) calibration of hydrology: travel times in different lithologies Measurements of Water Chemistry (19 lakes) Regional Application (100 lakes) MODFLOW model calibration of reactive transport: composition of rockwater MPATH model simplifications for the regional application: transfer functions land-use, deposition, terrestrial sinks input maps terrain, lithology precipitation, runoff weathering rates for catchments generalized FAB model deposition model SSWC model exceedances of critical loads The annual atmospheric depositions and resulting exceedances of the critical loads are calculated for three years: 1980, 1995 and 2010. In this period, the emissions and deposition of sulphur and to a much lesser extent of nitrogen changed substantially, and consequently also the exceedances of the critical loads. Critical Loads of Acidity for Alpine Lakes FOEN 2007 2.2 Generalisation of the First-order Acidity Balance (FAB) model The original version of the FAB model has been developed and applied to lakes in Finland, Norway and Sweden in Henriksen et al. (1993) and is also described in Posch et al. (1997). A modified version was first reported in Hindar et al. (2000) and is fully described in Henriksen and Posch (2001) as well as in Chapter 5.4 of the Mapping Manual (UBA 2004). Here, a generalised version of the First-order Acidity Balance (FAB) model for calculating critical loads of sulphur (S) and nitrogen (N) for a lake is derived. It differs from the previously published version of FAB in the following points: – Base cation leaching is explicitly formulated in terms of its sources and sinks in the catchment, i.e. deposition, weathering and removal due to uptake etc., instead of plugging in the SSWC model, based on water chemistry data. This renders the FAB model completely equivalent to the Simple Mass Balance (SMB) model for a soil profile. – The number of sub-areas (e.g., land cover classes) in the catchment, for which different fluxes can be specified, is now unrestricted. – Individual depositions to different sub-areas of the catchment can be taken into account. In addition, explicit formulae for computing the nodes of the critical load function are given. 2.2.1 Model derivation The lake and its catchment are assumed small enough to be properly characterised by average soil and lake properties. The total catchment area (lake + terrestrial catchment) A consists of the lake area Al =A0 and m different sub-areas Aj (j=1,…,m), comprising the terrestrial catchment: (2.1) m m j =1 j =0 A = Al + ∑ A j = ∑ A j E.g., A1 could be the forested area, A2 the area covered with grass or heathland, A3 the area of bare rocks, etc. Also a subdivision along soil types could be useful. Starting point for the derivation of the FAB model is the charge balance (‘acidity balance’) in the lake water running off the catchment: (2.2) S runoff + N runoff = ∑Y Yrunoff − ANCrunoff 18 2 19 > Methods where ΣY stands for the sum of base cations minus chloride (Ca+Mg+K+Na–Cl), and ANC is the acid neutralising capacity. In the above equation we assume that the quantities are total amounts per time (e.g. eq a–1). To derive critical loads we have to link the ions in the lake water to their depositions, taking into account their sources and sinks in the terrestrial catchment and in the lake. Mass balances in the lake are given by: (2.3) X runoff = X in − X ret , X = S , N , Ca, Mg , K , Na, Cl where Xin is the total amount of ion X entering the lake and Xret the amount of X retained in the lake. The in-lake retention is assumed to be proportional to the input of the respective ion into the lake: (2.4) X ret = ρ X ⋅ X in , X = S , N , Ca, Mg , K , Na, Cl where 0 ≤ ρX ≤ 1 is a dimensionless retention factor. The mass balances then become: (2.5) X runoff = (1 − ρ X ) ⋅ X in , X = S , N , Ca, Mg , K , Na, Cl The total amount of sulphur entering the lake is given by: m (2.6) Sin = ∑ Aj ⋅ S dep , j j =0 where Sdep,j is the total deposition of S per unit area onto land area j. Immobilisation, reduction and uptake of sulphate in the terrestrial catchment are assumed negligible, and sulphate ad/desorption need not be considered since we model steady-state processes only. Equation 2.6 states that all sulphur deposited onto the catchment enters the lake, and no sources or sinks are considered in the terrestrial catchment. For nitrogen we assume that net uptake (= net removal), net immobilisation and denitrification can occur on all sub-areas, possibly at different rates. Thus the amount of N entering the lake is: m (2.7) N in = ∑ A j ⋅ ( N dep , j − N i , j − N u , j − N de, j ) + j =0 where Ndep,j is the total N deposition, Ni,j is the long-term net immobilisation of N (which may include other long-term steady-state sources and sinks), Nu,j the net growth uptake of N and Nde,j is N lost by denitrification, all per unit area for land area j. The symbol (x)+ or x+ is a short-hand notation for max{x, 0}, i.e., x+= x for x > 0 and x+ = 0 for x ≤ 0. The effects of nutrient cycling are ignored and the leaching of ammonium is considered negligible, implying its complete uptake and/or nitrification in the terrestrial catchment. 20 Critical Loads of Acidity for Alpine Lakes FOEN 2007 Note that some of the terms in Equation 2.7 (other than deposition!) can be zero for certain indices; especially for j=0, i.e. the lake itself, one mostly assumes Ni,0+Nu,0 = 0. While immobilisation and net growth uptake are assumed independent of the N deposition, denitrification is modelled as fraction of the available N: (2.8) N de , j = f de, j ⋅ ( N dep , j − N i , j − N u , j ) + for j = 0,..., m where 0 ≤ fde,j < 1 is the (soil-dependent) denitrification fraction for area j. The above equation is based on the assumption that denitrification is a slower process than immobilisation and growth uptake. Inserting Equation 2.8 into Equation 2.7 one obtains: m (2.9) N in = ∑ A j ⋅ (1 − f de, j ) ⋅ ( N dep , j − N i , j − N u , j ) + j =0 For base cations and chloride the amount entering the lake is given by: m (2.10) Yin = ∑ A j ⋅ (Ydep , j + Yw, j − Yu , j ) + , Y = Ca, Mg , K , Na, Cl j =0 where Yw,j is the area weathering flux of ion Y for land cover class j. Equation 2.10 highlights the conceptual difference to the standard FAB model (Henriksen and Posch 2001): Base cation and chloride fluxes are not estimated from the lake water chemistry, as in the SSWC model, but from individual (terrestrial) catchment fluxes. Such an approach has also been tried by Rapp and Bishop (2003), using the PROFILE model for estimating soil weathering rates in catchments. To obtain an equation for critical loads, a link has to be established between a chemical variable and effects on aquatic biota. The most commonly used criterion is the so-called ANC-limit (see above), i.e. a minimum concentration of ANC derived to avoid ‘harmful effects’ on fish: ANCrunoff,crit = A•Q•[ANC]limit, where Q is the catchment runoff. Other criteria, e.g. a critical pH or Al concentration can be considered by calculating the critical ANC concentration from it, as is done in the SMB model (see Chapter 5.3 of the Mapping Manual). Inserting Equations 2.6, 2.9 and 2.10 into Equation 2.5 and Equation 2.2 and dividing by A yields the following equation to be fulfilled by critical depositions (loads) of S and N: (2.11) m m j =0 j =0 (1 − ρ S ) ⋅ ∑ c j ⋅ S dep , j + (1 − ρ N ) ⋅ ∑ c j ⋅ (1 − f de , j ) ⋅ ( N dep , j − N i , j − N u , j ) + = Lcrit where we have defined the sub-area fractions cj: 2 21 > Methods (2.12) c j = Aj A ⇒ m ∑c j =0 j =1 where c0 = r is the lake:catchment ratio. Furthermore we introduced: m (2.13) Lcrit = ∑Y (1 − ρY ) ⋅ ∑ c j ⋅ (Ydep , j + Yw, j − Yu , j ) + − Q ⋅ [ ANC ]limit j =0 where the first sum is over the four base cations minus chloride (Ca+Mg+K+Na–Cl). Note that in the standard FAB model (Henriksen and Posch 2001) Equation 2.13 reads Lcrit=CL(A)=Q•([BC*]0–[ANC]limit). The depositions to the various sub-areas can be written as: (2.14) S dep , j = s j ⋅ S dep and N dep , j = n j ⋅ N dep , j = 0,..., m where Sdep and Ndep are catchment average depositions, and sj and nj dimensionless factors describing the enhanced (or reduced) deposition onto sub-area j. Inserting them into Equation 2.11 yields: (2.15) m ⎛ N + Nu, j aS ⋅ S dep + (1 − ρ N ) ⋅ ∑ c j ⋅ (1 − f de , j ) ⋅ n j ⋅ ⎜ N dep − i , j ⎜ nj j =0 ⎝ ⎞ ⎟ = Lcrit ⎟ ⎠+ with the dimensionless parameter: m (2.16) a S = (1 − ρ S ) ⋅ ∑ c j ⋅ s j j =0 Equation 2.15 defines a function in the (Ndep, Sdep)-plane, the so-called critical load function (see Figure 4) and in the following we look at this function in more detail. We assume that the sub-areas are enumerated in such a way that (2.17) N j −1 = ( N i , j −1 + N u , j −1 ) n j −1 ≤ ( N i , j + N u , j ) n j = N j for j = 1,..., m Between two successive values of Nj the critical load function is linear, but at Nj it changes the slope (another of the large brackets in Equation 2.15 becomes non-zero). The resulting piecewise linear function has (at most) m+2 segments, and every segment is of the form: (2.18) aS ⋅ S dep + aN , k ⋅ N dep = LN , k + Lcrit for N k −1 ≤ N dep ≤ N k , k = 0,..., m + 1 22 Critical Loads of Acidity for Alpine Lakes FOEN 2007 with (by definition) N 1=0 and Nm+1=∞. In Equation 2.18 we introduced the dimensionless parameters: k −1 (2.19) a N ,0 = 0 , a N , k = (1 − ρ N ) ⋅ ∑ c j ⋅ (1 − f de, j ) ⋅ n j , k = 1,..., m + 1 j =0 and the terms: k −1 (2.20) LN ,0 = 0 , LN , k = (1 − ρ N ) ⋅ ∑ c j ⋅ (1 − f de, j ) ⋅ ( N i , j + N u , j ) , k = 1,..., m + 1 j =0 The maximum critical load of sulphur is obtained by setting Ndep = 0 in Equation 2.15: (2.21) CLmax ( S ) = Lcrit aS To compute the maximum critical load of nitrogen one has to find the segment of the critical load function, which intersects the horizontal axis. The first segment is horizontal (since aN,0=0), and this segment extends till Ndep=N0=(Ni,0+Nu,0)/n0 (see Equation 2.17). Each of the following (at most) m+1 straight lines defined in Equation 2.18 intersects the horizontal axis at (setting Sdep = 0): (2.22) N 0,k = ( LN ,k + Lcrit ) a N ,k , k = 1,..., m + 1 And the N0,k which lies between the limits defined in Equation 2.17 defines the maximum critical load for nitrogen. Denoting this specific index with K (1≤K≤m+1), we have: (2.23) CLmax ( N ) = N 0, K = ( LN ,K + Lcrit ) a N , K where N K −1 < N 0, K ≤ N K The first node of the critical load function is (0,CLmax(S)), the second one (N0,CLmax(S)). Note that in most applications uptake and immobilisation in the lake is assumed zero, i.e. N0=0, and thus the second node coincides with the first. The next (maximum) K–1 nodes of this piecewise linear function are given by (Nk,Sk), where Nk is defined in Equation 2.17 and the Sk are obtained as: (2.24) S k = a N ,k ⋅ ( N 0,k − N k ) aS , k = 1,..., K − 1 And the last, at most (K+2)-nd, node is given by (CLmax(N),0). A Fortran subroutine to calculate the nodes of the critical load function for given catchment characteristics can be found in Annex A.2. 2 > Methods Figure 4 > Piece-wise linear critical load function of S and acidifying N for a lake as defined by catchment properties. Here shown for two land use classes characterised by (N1,S1) and (N2,S2) (see eqs.17 and 24). The grey area below the CL function denotes deposition pairs resulting in an ANC leaching greater than Q•[ANC]limit (non-exceedance of critical loads). The critical load exceedance is calculated by adding the N and S deposition reductions needed to reach the critical load function via the shortest path (E→Z): Ex = ΔS+ΔN. 2.2.2 Input data requirements The generalised FAB model needs information on (a) the runoff, (b) the area of lake, catchment and different sub-areas, (c) terrestrial base cation and nitrogen sources and sinks, and (d) parameters for in-lake retention of N, S and base cations. a) Runoff: The runoff Q is the amount of water leaving the catchment at the lake outlet, expressed in m a-1. It is derived from measurements or can be calculated as the difference between precipitation and actual evapotranspiration, averaged over the catchment area, if there are no net losses to the groundwater. A long-term climatic mean annual value should be taken. b) Lake and catchment characteristics: The area parameters Aj (j=0,…, m) can generally be derived from (digital) maps. c) Terrestrial sources and sinks of BC and N: These parameters can be derived the same way as for the SMB model. 23 Critical Loads of Acidity for Alpine Lakes FOEN 2007 The net uptake of base cations and N can be computed from the annual average amount of N in the harvested biomass. If there is no removal of trees or by grazing animals, Nu = 0. Ni is the long-term annual immobilisation (accumulation) rate of N for sustainable soil formation in the catchment. Note that at present, immobilisation may be substantially higher due to elevated N deposition. The denitrification fraction fde depends on the soil type and its moisture status. d) In-lake retention: Concerning in-lake processes, the retention factor for nitrogen ρN (see Equation 2.4) is modelled by a kinetic equation (Kelly et al. 1987) included in the FAB model: (2.25) ρN = sN sN = sN + z /τ sN + Q / r where z is the mean lake depth, τ is the lake’s residence time, r=c0 is the lake:catchment ratio (see Equation 2.12) and sN is the net mass transfer coefficient. There is a lack of observational data for the mass transfer coefficients, especially from European catchments, but Dillon and Molot (1990) give a range of 2–8 m a–1 for sN. Values for Canadian and Norwegian catchments are given in Kaste and Dillon (2003). Alternative methods for calculating the in-lake retention of nitrogen might be evaluated on the basis of monitoring data compiled by Steingruber (2001, page 77), which imply that independent of depth and residence time the nitrogen retention in lake is mainly determined by its areal nitrogen load. An equation analogous Equation 2.25 for ρS, with a mass transfer coefficient sS, is used to model the in-lake retention of sulphur. Baker and Brezonik (1988) give a range of 0.2–0.8 m a–1 for sS. For ρBC no data is available. 2.3 The Steady-State Water Chemistry (SSWC) model The ‘classic’ model for calculation the critical load of acidity for a lake or stream is the SSWC (or ‘Henriksen’) model, which uses (estimated) annual mean values of present-day water chemistry. A derivation of the SSWC model, including many of its variants, and references to the original literature can be found in Chapter 5.4 of the Mapping Manual (UBA 2004). Here we simply summarise the model equations used in this study. In the SSWC model the critical load of acidity, CL(A), is calculated from the principle that the acid load should not exceed the non-marine, non-anthropogenic base cation input 24 2 25 > Methods and sources and sinks in the catchment minus a buffer to protect selected biota from being damaged. This critical load is given by: (2.26) CL( A) = Q ⋅ ([ BC * ]0 − [ ANC ]limit ) where Q is the catchment runoff (in m a-1), [BC*]0 (BC=Ca+Mg+K+Na) is the preacidification concentration of base cations, and [ANC]limit the lowest ANC-concentration that does not damage the selected biota. The star indicates sea salt correction; however, no such correction has been applied to the data for the Swiss lakes. The pre-acidification base cation concentration is calculated with the help of the so-called F-factor from: (2.27) ( [ BC ∗ ]0 = [ BC ∗ ] − F ⋅ [ SO4* ] − [ SO4* ]0 + [ NO3 ] − [ NO3 ]0 ) where [SO4] and [NO3] are the present-day concentrations of sulphate and nitrate and the subscript zero indicates their pre-acidification values. The pre-acidification nitrate concentration is generally assumed zero. Viewing Equation 2.27 as a definition for the F-factor, it shows that it is the rate of change in non-marine base cation concentrations due to changes in strong acid anion concentrations. If F=1, all incoming protons are neutralised in the catchment (only soil acidification), at F=0 none of the incoming protons are neutralised in the catchment (only water acidification). The F-factor was estimated empirically to be in the range 0.2–0.4, based on the analysis of historical data from Norway, Sweden, U.S.A. and Canada (Henriksen 1984). Brakke et al. (1990) later suggested that the F-factor should be a function of the base cation concentration: (2.28) ( ) F = sin π2 [ BC * ] [ S ] where [S] is the base cation concentration at which F=1; and for [BC*]>[S] F is set to 1. The traditional value of [S]=400 meq m–3 (ca. 8 mg Ca L–1) is used here. The pre-acidification sulphate concentration in lakes, [SO4*]0, is assumed to consist of a constant atmospheric contribution and a geologic contribution proportional to the concentration of base cations: (2.29) [ SO4* ]0 = a + b ⋅ [ BC * ] Following Henriksen and Posch (2001), a=0.008 eq m–3 and b=0.17 are used here, as well as a critical ANC-limit of 20 meq m–3. Critical Loads of Acidity for Alpine Lakes FOEN 2007 2.4 Calculation of Weathering Rates for Catchments 2.4.1 Introduction The in-soil weathering rates are very low in those catchments because of very thin soils, low temperature and in many cases slow weathering minerals. Therefore, the weathering rates of base cations (BCw ; BC = Ca2+ + Mg2+ + K+ + Na+) for each catchment were estimated by quantifying rock-water interaction processes also through groundwater recharge using a simplified hydrological model. The geology of the catchments was digitized using regional geological maps and was simplified by classifying it into 5 lithological units: quaternary cover, leucocratic granite/gneiss, melanocratic granite/gneiss, amphibolite, and carbonate bearing rocks (example in Figure 16). Digital elevation maps were used to estimate surface runoff, average linear velocities and the resulting travelling time of the infiltrating water for each individual lithological area in a catchment. Dissolved BC’s of the infiltrating water were estimated using a reactive transport model, where transfer functions for the dependence of «travelling-time» and «mineral dissolution» were calculated for each lithology. Travelling time is essential, since longer reaction time of the water with the bedrock lithology contributes significantly to the overall catchment weathering rate (BCw,C). The contribution from bedrock to BCw,C is restricted to saturated groundwater and infiltrates into the lake mainly at deeper levels. At low porosity, e.g. 2 %, a recharge of 400 mm a–1 has been estimated. The remaining surface runoff is the dominant H+ source, which enters the lake directly (surface runoff), or after relatively short travelling time if the surface is covered by quaternary deposits. 2.4.2 Calibration of Hydrology Before the reactive transport in the subsurface of a basin can be studied in space and time, the hydrological situation needs to be evaluated. First, a fluid flow mass balance calculation for the basin has to be performed, and then particle travel times have to be calculated. For mass balance calculations and particle tracking in the crystalline basement of the two catchments used for calibration, a simplified 3-D groundwater model was derived. Fluid flow and mass transport simulations are conducted with the computer program Visual MODFLOW (Waterloo Hydrogeologic, Inc.). Particle travel times in the overlying gravel aquifer and parameter sensitivity analyses are evaluated using 2-D flow nets and Darcy’s law. 26 2 27 > Methods a) Introduction: For evaluating the hydraulic parameters, the mountain basin of the Lake Superiore (ID=64) and Lake Inferiore (ID=63) was used (see Figure 2 and Annex 1). Three distinct lithological units with different hydrological characteristics are separated. The top gravel layer has an approximate thickness of 0.1 to 1.0m, and is underlain by the crystalline basement, which is composed dominantly of gneiss and amphibolite dykes. Very little is known about the hydrological parameters of the unconfined aquifers. Therefore, hydraulic conductivity (horizontal Kx, vertical Ky), porosity, and specific yield data are taken from the typical range of comparable lithologies (Freeze and Cherry, 1992; Spitz and Moreno, 1996) as shown in Table 2. Specific yield is a property of the aquifer measuring the ability to release groundwater from storage, due to decline in hydraulic head. Hydraulic head is defined as the sum of pressure and elevation heads. In an unconfined aquifer (which is in contact with the atmosphere) it corresponds to the water table. Table 2 > Hydrological characteristics of three distinct lithology units. Lithology Gravel Porosity [%] Ky [m/s] Kx [m/s] 10 Specific Yield [ – ] 15 10 Feldpar-Gneiss 2 10–8 10–7 0.01 Amphibolite 4 10–7 10–6 0.02 –4 –3 0.15 The hydraulic conductivity of the crystalline basement depends largely on the degree of fracturing. The crystalline is in nature a dual porosity aquifer with fractures and a porous matrix. For the top glacial sediment layer a high uncertainty lays in the porosity value, which could be as high as 40 %. The specific yield of 0.15 dominates release and storage of groundwater in this aquifer. An unsaturated zone in the crystalline rocks might separate the horizontal interflow in the gravel layer from the basement aquifer. Thin permeable sediment layers with interflow are common when a layer with low vertical hydraulic conductivity occurs beneath (Fetter, 1994). This interflow may be substantial in the Lake Superiore/Lake Inferiore drainage basin and may contribute significantly to total discharge into the lakes. In the hydrological model only annual mean values are used. Thus, seasonal fluctuations are neglected, e.g. the spring melt, where large amounts of surface flow are observed, as the water can not infiltrate the frozen soil. Furthermore, a constant water level is assumed in the lakes. This water surface is used as a prescribed head boundary (upper boundary of the bedrock aquifer). Critical Loads of Acidity for Alpine Lakes FOEN 2007 b) Conceptual Model: Computer simulations with Visual MODFLOW require defining the model boundaries, defining the model dimensions, and simplifying the aquifer system. a) The mountain range to the South of the lakes (see Figure 5) represents a natural deep reaching groundwater divide and serves as no-flow boundary, whereas the other mountain ranges to the Southeast, West and North are just no-flow boundaries for local-flow but not for regional (deeper reaching) groundwater flow. The Northeast boundary is a natural flux boundary where a stream and partially the crystalline bedrock aquifer discharge. As no head distributions (ground water levels) and flux data are available it must be accepted that all surrounding boundaries are modelled as no-flow boundaries. Other natural discharge areas are the surfaces of Lake Superiore and Lake Inferiore, which represent constant head boundaries. b) The size of the model domain is selected based on the natural boundaries (mountain ranges), and in the Northeast by the extension of Lake Inferiore. c) Due to the steep topography, and subsequently the high vertical hydraulic gradient, it is not possible in the finite difference program Visual MODFLOW to simulate fluid flow through the gravel layer and bedrock aquifer together. As a first approximation the aquifer system is seen as one continuous hydrogeological unit and only groundwater flow through the crystalline basement is simulated. c) Selecting Model Input Data: The 3-D finite difference program Visual MODFLOW requires information about vertical and horizontal hydraulic conductivity, porosity, specific yield (in the case of unconfined aquifers) and the coordinates of all layers. These input data are taken from literature (Freeze and Cherry, 1992; Spitz and Moreno, 1996; see Table 2). The following assumptions lead to further simplifications: d) The aquifer can be represented by a homogeneous porous medium. e) Recharge due to infiltrating rainfall and melting snow, and evapotranspiration are constant over the whole domain. f) The crystalline bedrock aquifer is completely saturated, therefore the water table lies at the surface. 28 2 29 > Methods d) Defining the Model Domain and Discretization: In order to evaluate the water mass balance in the Lake Superiore and Lake Inferiore basin a three-dimensional domain has to be considered. A finite difference model is used, and the domain is discretized with 15 horizontal layers, 75 rows and 118 columns yielding a total of 132750 elements. In the high discharge areas around the lakes a finer grid is chosen. e) Performing Model Simulations and Calibration: All input data are assumed to be constant in time. Therefore, a steady state situation is simulated. The model is calibrated using the prescribed heads of Lake Superiore and Lake Inferiore. No other groundwater heads are known in the basin. In the calibration procedure it is assumed that the water table in the crystalline bedrock aquifer reaches the surface. Model calibration yields a maximum possible recharge into the bedrock aquifer of 400 mm a–1. Evapotranspiration from the saturated groundwater regime is assumed to be 160 mm a–1. f) Results from Model Simulations: In the groundwater mass balance budget the total inflow into the crystalline aquifer system is 1701.46 m³ d–1 (400 mm a–1), and the total discharge is 1712.11 m³ d–1 through the lakes and due to evapotranspiration. For a one-layer system at 400 mm recharge the infiltrations are estimated according to Table 3. Table 3 > Calculation of infiltrations for a one-layer system. Units are m³ d–1. IN OUT Constant head = 2.0 Constant head = 1122.5 Recharge = 1699.4 Evapotranspiration = 589.6 Total in = 1701.5 Total out = 1712.1 IN – OUT = –10.6 % Discrepancy = –0.62 % From calibration simulations it can be concluded that a large amount of rainwater infiltrates into the upper gravel layer (1600 mm a–1 – 400 mm a–1) and reaches the lake system as surface-parallel interflow through the gravel. The amount of interflow is calculated with 2547.0 m³ d–1. In order to evaluate the outflow of the basin, which discharges through the lakes and ultimately through the stream, an annual stream hydrograph is Critical Loads of Acidity for Alpine Lakes FOEN 2007 required. From such hydrograph data it is possible to separate baseflow from interflow plus overland flow. From the total recharge into the crystalline basement about 1/5 (or 339.9 m³/d) infiltrates through amphibolite dikes. For a three-layer system at 1600mm recharge the infiltrations are estimated as shown in Table 4. Table 4 > Estimated infiltrations for a three-layer system. Units are m³ d–1. Layers IN Gravel 2547.0 Amphibolite Gneiss 339.9 1359.5 g) Head Distribution (Groundwater Levels): Calibrating the model with the prescribed heads of 2128m for Lake Superiore and 2074m for Lake Inferiore, and defining the water table at the crystalline surface yields a consistent head distribution with fluid flow towards the two lake system (Figure 5). Due to the head distribution in the basin, velocity vectors are pointing towards the defined discharge areas. Average linear velocities depend on the hydraulic gradient, the vertical and horizontal hydraulic conductivity, and the porosity. 30 2 > Methods Figure 5 > Head distribution isolines and three particle paths in the Lago Superiore (left) and Lago Inferiore (right) area. The coloured areas represent lithology classes: red = gneiss, blue = amphibolite, green = quarterny cover. h) Particle Tracking in the Crystalline Bedrock Aquifer: The pathway of three particles, which have been released in areas with different gradients and distances to the lake system, are shown in Figure 5. Particle 1, which is released at the steep southern mountain slope, reaches Lake Superiore in 9300 days (~25.5 years). The travel distance for particle 1 is about 660m, and the hydraulic gradient 0.48 (vertical distance divided by horizontal distance, dh/dl). Particle 2 is released on the northern ridge, which is relatively close to Lake Superiore and just about 100m higher in elevation. Subsequently, the travel time is much shorter with 2000 days (~5.5 years). Particle 3 is furthest away from the discharge areas, and has a travel time of 20900 days (~57.3 years) before reaching Lake Inferiore. 31 Critical Loads of Acidity for Alpine Lakes FOEN 2007 i) Travel Times: Using Darcy’s law, average linear velocities and resulting travel times are calculated for gravel, amphibolite, and gneiss layer. A sensitivity study with respect to hydraulic gradient (dh/dl), hydraulic conductivity and porosity is performed. 1. Gravel Layer: The travel time for a dissolved conservative particle in a gravel layer with 15 % effective porosity, a hydraulic conductivity of 10 – 4 m s–1, and a gradient of 0.48 is 23.9 days for a distance of 660 m (Table 5). A decreasing gradient to 0.1 results in an increase of the travel time to 114.6 days. If the porosity increases at a constant gradient, travel times will also increase (Table 6). As the porosity changes from 15 to 30 % at a hydraulic gradient of 0.48, the particle travel time changes from 23.9 to 47.7 days. Increasing the hydraulic conductivity by a factor of 10 will decrease the travel time also by a factor of 10, if all other parameters stay constant. 2. Amphibolite Layer: The largest outcrop for amphibolite lies in the south-eastern quadrangle of the study area (Figure 5). Here the average hydraulic gradient through amphibolite is estimated with 0.24. Particles released in the south-eastern area might travel about 210 m through amphibolite, which has an estimated hydraulic conductivity of 10 – 6.5 m s–1. Assuming an effective porosity of 4 %, particles will take about 1292 days (~3.5 years) to travel the distance of 210 m on their way towards Lake Inferiore. Decreasing the hydraulic conductivity to 10 – 7.5m/s will result in a 10-fold travel time of about 35 years. 3. Gneiss Layer: Travel times for particles that are released near the southern border (see particle 1, Figure 5) reach Lake Superiore in about 10,000 days, assuming that the hydraulic gradient is 0.48 (Table 7). Such gradient can only be reached, when the water table lies directly beneath the surface. A lower water table will result in a decreasing hydraulic gradient and subsequently yield longer travel times. If the effective porosity of 2 % is doubled the travel time is doubled as well, assuming all other parameters stay constant (Table 8). 32 2 33 > Methods Table 5 > Maximum total daily flow through gravel layer for different gradients. hydraulic gradient effective porosity hydraulic conductivity specific discharge average linear velocity total daily flow per m width thickness of gravel layer average linear velocity horizontal distance time dh/dl ne K q v Q b v d t [-] [-] [m/d] [m/d] [m/d] [m³/d] [m] [m/d] [m] [d] 0.10 0.15 8.64 0.9 5.8 0.9 1 5.8 660 114.6 0.20 0.15 8.64 1.7 11.5 1.7 1 11.5 660 57.3 0.30 0.15 8.64 2.6 17.3 2.6 1 17.3 660 38.2 0.40 0.15 8.64 3.5 23.0 3.5 1 23.0 660 28.6 0.48 0.15 8.64 4.1 27.6 4.1 1 27.6 660 23.9 average linear velocity total daily flow per m width thickness of gravel layer average linear velocity horizontal distance time Table 6 > Travel times through gravel layer for different porosities. hydraulic gradient effective porosity hydraulic conductivity specific discharge dh/dl ne K q v Q b v d t [-] [-] [m/d] [m/d] [m/d] [m³/d] [m] [m/d] [m] [d] 0.10 0.10 8.64 4.1 41.5 4.1 1 41.5 660 15.9 0.20 0.15 8.64 4.1 27.6 4.1 1 27.6 660 23.9 0.30 0.20 8.64 4.1 20.7 4.1 1 20.7 660 31.8 0.40 0.25 8.64 4.1 16.6 4.1 1 16.6 660 39.8 0.48 0.30 8.64 4.1 13.8 4.1 1 13.8 660 47.7 thickness of gravel layer average linear velocity horizontal distance Time Table 7 > Maximum total daily flow and travel times through gneiss layer for different gradients. hydraulic gradient effective porosity hydraulic conductivity specific discharge average linear velocity total daily flow per m width dh/dl ne K q v Q b v d T [-] [-] [m/d] [m/d] [m/d] [m³/d] [m] [m/d] [m] [d] 0.10 0.02 0.00273 0.0003 0.0137 0.0003 1 0.0137 660 48351.6 0.20 0.02 0.00273 0.0005 0.0273 0.0005 1 0.0273 660 24175.8 0.30 0.02 0.00273 0.0008 0.0410 0.0008 1 0.0410 660 16117.2 0.40 0.02 0.00273 0.0011 0.0546 0.0011 1 0.0546 660 12087.9 0.48 0.02 0.00273 0.0013 0.0655 0.0013 1 0.0655 660 10073.3 34 Critical Loads of Acidity for Alpine Lakes FOEN 2007 Table 8 > Travel times through gneiss layer for different porosities. hydraulic gradient 2.4.3 effective porosity hydraulic conductivity specific discharge average linear velocity total daily flow per m width thickness of gravel layer average linear velocity horizontal distance time dh/dl ne K q v Q b v d t [-] [-] [m/d] [m/d] [m/d] [m³/d] [m] [m/d] [m] [d] 0.10 0.01 0.00273 0.0013 0.1310 0.0013 1 0.1310 660 5036.6 0.20 0.02 0.00273 0.0013 0.0655 0.0013 1 0.0655 660 10073.3 0.30 0.05 0.00273 0.0013 0.0262 0.0013 1 0.0262 660 25183.2 0.40 0.07 0.00273 0.0013 0.0187 0.0013 1 0.0187 660 35256.4 0.48 0.10 0.00273 0.0013 0.0131 0.0013 1 0.0131 660 50366.3 Calibration of the Reactive Transport of Ions a) The used Model and Modelling Conditions: In the used approach the lithologies were treated as porous homogeneous media, neglecting preferential flow paths. The averaged water composition was then calculated from the contribution of each lithology and expressed as field weathering rate BCw.C for each of the 100 catchments. The Model MPATH (Lichtner, 1985, 1988,1992) describes fluid transport in a time-space continuum, assuming pure advection. MPATH calculates the changes in the composition of an infiltrating fluid packet as well as the associated mineralogical changes of the parent rock (Figure 6). Rock-water interaction reactions occur until a quasi stationary state is reached. The next fluid package infiltrates in the now slightly altered parent rock, and so on. The reaction history of each fluid packet therefore behaves differently, producing reaction fronts depending on the reaction kinetics, the composition and velocity of the fluid. Figure 6 > General scheme of the rock-water interaction model MPATH. Water composition Infiltation Minerals Surface Rate K Water composition Mineralogy 2 35 > Methods Reversible homogeneous reactions are considered for aqueous components (speciation, complexation, redox reactions) and non reversible heterogeneous reactions for mineral phases (dissolution and precipitation). Aqueous complexing reactions are assumed to be in local chemical equilibrium, including all redox couples. Sposito (1989) confirmed that in soils the kinetics of complexation is fast enough to assume instant chemical equilibrium. Mineral reactions are described using kinetic rate laws, activity coefficient corrections are based on an extended Debye-Hückel algorithm (The code was developed by Peter Lichtner and is described in detail in Lichtner, 1985 and Lichtner, 1988). At the initial state of the column, mineral compositions and physical properties of a parent rock were used. Rain water composition measured by Barbieri (pers. comm., Barbieri & Pozzi 2001) was used as input solution, representing the water composition, which infiltrates the parent rock. The model has been run over a time span of 100 years. The initial state of development is mainly dependent on the porosity and the velocity of the water front which in the model is assumed to be saturated. The velocity of the percolating front was calculated by ModFlow to be 400 mm per year. The input parameters used are listed in Table 9. Table 9 > Input parameters for the Model MPATH. Lithology specific Yield min VDarcy max VDarcy Discharge Temp m/a m/a mm/a °C Gravel 0.15 2’000 10’000 > pecipitaion 8 Gneiss (L&M) 0.01 5 24 400 8 Amphibolite 0.02 10 48 400 8 Carbonate 0.02 10 48 400 8 Modelling conditions used are: – – Mode Species – – Activity model Area – – Flow Temperature – Constraint types – Output open system (Fluid packet model) all species contained in database are read in consistent with set of primary species extended Debye-Hückel approximation variable surface area according to the two-thirds power of the mineral volume fraction Darcy fluid velocity, VDarcy (m a-1) temperature in °C. Allowed values are in the range 0–300 °C for the EQ3/6 (Wolery et al. 1990) mass balance / mineral constraint / gas constraint / concentration constraint buffer along flow path / concentration (activity of H+ ) constraint / pH constraint or charge balance aqueous concentrations / mineral reaction rates / mineral saturation indices / mineral volume fractions Critical Loads of Acidity for Alpine Lakes FOEN 2007 Figure 7 > Ion concentrations in rock-water for Leucocratic Crystalline Rocks (top) and Melanocratic Crystalline Rock Types (bottom). Y is the best-fit function for BC (dotted line). 36 2 > Methods Figure 8 > Ion concentrations in rock-water for Amphibole Bearing Rocks (top) and Carbonate Rock Types (bottom). Y is the best-fit function for BC (dotted line). 37 Critical Loads of Acidity for Alpine Lakes FOEN 2007 Figure 9 > Ion concentrations in rock-water for quarternary cover. y is the best-fit function for BC. b) Calculated Water Composition: Using the modelling conditions described above, the water composition has then been calculated as function of travelling distance for each of the lithologies. For better reading of Figure 7, Figure 8 and Figure 9, the water composition is shown as a function of residence time (travelling time), corresponding to the time a water package needs to migrate through the bedrock. The figures show the concentrations of Na, K, Ca and Mg as well as the resulting total BC-concentrations approximated by a best-fit function (polynomial of 5th order). Using the best-fit functions of the calculated BC-release in combination with additional catchment specific parameters, transfer functions were defined as described in the next section. 38 39 2 > Methods 2.4.4 Transfer Functions for the Regional Application The measured water composition in the catchment lake is a function of the amount and composition of the infiltrating and surface runoff water, and the mineralisation through mineral dissolution reactions during migration through the host rock. For the calculation of the catchment weathering rates (BCw,C) the following calculation steps were made. The hydrological parameters for the lithology-classes (L) such as conductivity (KL) and porosity (see Chapter 2.4.2) are used to calculate the mass flow (QL): (2.30) QL = dh/dl * KL (Qmax = infiltration) Where: dh is the difference in height to the lake surface and dl is the horizontal distance to the lakeshore (see Figure 10). dh/dl is the hydraulic gradient. dh and dl are calculated from digital terrain data. The amount of infiltrating water is limited by the precipitation, and the difference between precipitation and infiltration is assumed to be surface runoff which reaches the lake directly. Therefore it has no direct influence on the weathering rate of the catchment, but has to be taken into account when the measured water composition in the lake is compared to the modelled water composition. For the area, which is covered by glaciers no infiltration in the host rock is calculated. The total amount of precipitation is infiltrating the quaternary cover (gravel), if present, and below the gravel the water balance is handled according to the specific lithology, the same way as when no gravel would be present. In practice, the lithology underneath the quaternary cover is not known, and therefore the area covered by gravel was distributed among the other lithologies in the catchment according to their relative coverage of the catchment by increasing the weathering rate of the specific lithology. For every lithology in a catchment an average flow path length (DISTL) and an average travelling time (TTL) was approximated by (see Figure 10, left): (2.31) DISTL = dlL + 0.5 * dhL (2.32) TTL = DISTL / (QL / porosityL) The mean flow path length for each lithology was calculated as a weighted average based on the surface of the mapped sub-areas of the lithology. E.g. the catchment in Figure 16 contains three sub-areas of lithology No. 4 (leucocratic granite/gneiss) which are mapped as individual polygons. The area-weighted averaged flow path length was calculated operationally in the GIS by rasterizing the catchment with a resolution of 10x10 m, calculating DIST for each raster-cell and averaging all cells belonging to a specific lithology. The total area of a lithology within the catchment is denoted as AREAL. Critical Loads of Acidity for Alpine Lakes FOEN 2007 Figure 10 > Scheme of the simplified flow path calculation (left)as well as the calculated total areas (AREAL) and the mean flow path lengths (DISTL) per lithology-class in catchment No. 64 (Lago Superiore) (right). In Figure 10 (right) an example for catchment No. 64 is given, where the radius of a sector corresponds to the averaged travelling distance and the area of the sector to the relative surface area of the specific lithology. Depending on the travelling time (TTL), which is assumed to be the average time the water is in contact with the host rock, the BC concentration contributed by every lithology (BC(TTL)) was calculated according to the functions shown in Chapter 2.4.3. At the end, the BC concentrations were multiplied by the mass flow (QL * AREAL), and the results of all lithologies were summed up and then expressed as eq ha–1 a–1: (2.33) BCL = QL * AREAL * BC(TTL) (2.34) BCw,C = ∑ BCL1..L4 + BCGravel Table 10 gives an example of the input data compiled to calculate the average weathering rate of a catchment (BCw,C) by Equations 2.30 to 2.34. 40 2 41 > Methods Table 10 > Mean values of area (AREAL), distance of the flow path (DISTL) and hydraulic gradient (dhL/dlL) per lithology-class (L) in the catchment No. 77 (Lago Nero). L Description AREAL DISTL dhL/dlL KL porosityL 1 carbonate bearing rocks [ha] [m] 0 -- [-] [m s–1] [%] -- 3.17*10-8 2 2 amphibolite 0 -- -- 3.17*10-8 4 7.3 582 0.68 3.17*10 -8 3 melanocratic granite/gneiss 2 4 leucocratic granite/gneiss 32.1 235 0.71 3.17*10-8 2 5 quaternary cover (gravel) 20.2 272 0.57 3.17*10-5 15 8 glacier 0 -- -- -- -- 9 lake surface 12.7 -- -- -- -- Total 72.3 42 Critical Loads of Acidity for Alpine Lakes FOEN 2007 3 > Input Data 3.1 Deposition Rates The atmospheric deposition of base cations, N and S was calculated with a generalised combined approach (FOEFL 1994 and 1996, Rihm & Kurz 2001, Thimonier et al. 2004). Wet deposition is calculated by combining a digital precipitation map (see chapter 3.2) with the concentration fields of the compounds in precipitation water. In Southern Switzerland, a detailed study on wet deposition patterns was carried out (SAEFL, 2001). It presents a regression model for calculating the concentrations in precipitation as a function of altitude, longitude and latitude as shown in Table 11. The regression model is based on 13 monitoring stations in the canton Ticino and Italy. Fortunately, at least one station is situated in the altitude range of the investigated lakes: Robei, 1890 m a.s.l (near lake No. 78). Table 11 > Linear regression models for calculating ion concentrations in precipitation in the canton Ticino, based on measurements from 1993 to 1998 (SAEFL 2001). compound Regression coefficients Sulphate, SO42Nitrate, NO Intercept longitude latitude meq m–3 meq m–3 km–1 meq m–3 km–1 altitude meq m–3 km–1 54.58 0.12 –0.24 –9.18 44.49 0.057 –0.20 –7.29 Ammonium, NH4+ 51.90 0.090 –0.30 –7.98 Base cations, Bc 3- 27.78 0.21 –0.033 –6.92 Sodium, Na+ 9.14 0.045 –0.047 –1.62 Chloride, Cl 9.5 0.053 –0.039 –2.54 - Longitude, latitude and altitude are given in a local coordinate system in km. They can be obtained from Swiss national coordinates x and y (in m) with: longitude = x/1000–668, latitude = y/1000–70. Altitude is the average catchment altitude above sea level. It was available in digital form with a resolution of 25x25 m from the Federal Office of Topography. Bc is defined as the sum of Ca2+, Mg2+ and K+. For all compounds, there is a clear decrease of concentrations with altitude and latitude. Resistance analogue models are used for assessing the dry deposition of NH3 and NO2 gases as well as aerosols (PM10). For these compounds, the concentration fields were 3 43 > Input Data calculated from emission inventories with a resolution of 200m (NH3 100m) by applying statistical dispersion models. The models are described in Thoeni et al. (2004) for NH3, in SAEFL (2004) for NO2 and in SAEFL (2003) for PM10. For HNO3, the concentration field is calculated as a function of altitude. For SO2, the concentration field is determined by geo-statistical interpolation of monitoring results. The concentration fields are multiplied by deposition velocities, which depend on the reactivity of the pollutant, surface roughness and climatic parameters. Deposition velocity values were taken from literature (FOEFL, 1996, with modifications). The resulting patterns of deposition are supposed to be representative for the year 1995. As an example, Figure 11 shows a map of total nitrogen deposition: the gradients in altitude and latitude are evident. The Alpine lake catchments, mostly located near the watersheds, have much lower depositions than the bottoms of the valleys. Figure 11 > Map of modelled N deposition for the year 1995. 2 3 5 10 11 9 16 19 18 21 24 2526 23 31 205 2830 215 20 223336 3537 38 220 44 217 46 204 201 47 48 49 51 203 54 55 57 218 214 216 60 202 64 65 219 66 67 68 707363 71 76 77 78 79 80 81 222 221 207 84 22388 224 92 206 12 17 27 15 4143 50 225 53 72 74 85 86 89 93 9495 20898 99 212 100 102 211 109210 110 209 106 < 400 400 - 700 700 - 1000 1000 - 1500 > 1500 eq /ha /a 101 213104 K606-01©2004 swisstopo 0 5 10 km 44 Critical Loads of Acidity for Alpine Lakes FOEN 2007 For estimating the deposition in 1980 and 2010, the depositions for each lake are scaled backward and forward from 1995 proportionally to the European deposition time series modelled by Schöpp et al. (2003) on the European 150x150 km EMEP-raster. The 2010 scenario corresponds to the NEC Directive Implementation of the Gothenburg Protocol. From that dataset deposition ratios for the relevant years and EMEP-cell were derived as shown in Table 12. The area of investigation is covered by the EMEP-cell no. 24/13 (150x150 km). The ratios are used to scale the 1995 deposition calculated for each catchment. BC deposition is supposed to be time-independent. Table 12 > Deposition of S, NOy and NHx in 1980, 1995 and 2010 for the EMEP-cell 24/13 (from Schöpp et al. 2003) and derived deposition ratios normalised to 1995. Units 1980 1995 2010 Sulphur g S m–2 a–1 3.739 1.405 0.502 NOy g N m–2 a–1 0.754 0.698 0.418 NHx g N m–2 a–1 1.371 1.243 1.213 -- 2.661 1.000 0.357 NOy -- 1.080 1.000 0.599 NHx -- 1.103 1.000 0.976 Deposition: Ratios: Sulphur: Figure 12 shows the resulting deposition values in the 100 catchments with a decrease in S deposition of roughly 85 % and in N deposition of nearly 30 % between 1980 and 2010. At the considered altitude range, dry deposition contributes only a small part to the total deposition, namely 18 % of total nitrogen and 10 % of total sulphur. 3 45 > Input Data Figure 12 > Cumulative frequency distributions of S and N deposition for the 100 catchments in 1980, 1995 and 2010. Bc is the time-independent deposition of base cations. 100 90 80 N 1980 N 1995 70 60 N 2010 S 1980 S 1995 [%] 50 40 S 2010 Bc 30 20 10 0 0 3.2 200 400 600 [eq/ha/a] 800 1000 1200 Runoff The runoff (Q) is the difference between precipitation (P) and evapotranspiration (ET). Precipitation data are taken from a national precipitation map (FOWG 2000, Figure 13) with a resolution of 2x2 km. It contains long-term annual averages. Actual evapotranspiration is available from the Hydrological Atlas of Switzerland (FOWG 1999) on a 1x1 km raster. These are long-term annual averages taking into account climate, altitude, exposition, land-use and soil-properties. The frequency distribution of mean precipitation, evapotranspiration and the resulting runoff within the 100 catchment areas is shown in Figure 14. 46 Critical Loads of Acidity for Alpine Lakes FOEN 2007 Figure 13 > Map of precipitation rates, average 1961–1999 (FOWG 2000). 2 3 5 11 10 15 16 19 18 21 24 2526 23 28 215 31 30 205 20 223336 3537 38 220 44 217 46 204 201 47 48 49 51 203 225 54 55 57 218 214 216 60 202 64 65 219 67 707363 68 66 71 76 77 78 79 80 81 222 221 207 84 86 22388 89 224 92 93 206 99 20898 212 100 102 211 109210 110 209 < 1.25 1.25 - 1.50 1.50 - 1.75 1.75 - 2.00 2.00 - 2.25 > 2.25 106 9 12 17 27 4143 50 53 72 74 85 9495 101 213104 mt©20040920 0 5 10 km m /a 3 47 > Input Data Figure 14 > Cumulative frequency distributions of mean precipitation (P), evapotranspiration (ET) and runoff (Q) for the 100 catchments. 100 P 90 ET 80 Q 70 60 [%] 50 40 30 20 10 0 0.0 3.3 0.5 1.0 1.5 2.0 2.5 [m/a] Weathering Rates For defining the boundaries of the lithological units used for calculating the overall weathering rates of the catchments, four different sources of geological maps were considered. Where available, the maps of the Geological Atlas of Switzerland (scale 1:25’000) were used (Figure 15 top). Where these maps are not yet available, special regional maps were used: Either special geological maps (Figure 15 middle Spezialkarten der Schweiz), original geological maps (Figure 15 bottom geologische Originalkarten der Schweiz). Otherwise, the lithological maps from Boggero et al. (1996) were employed. The maps were simplified by reducing the complex geological information into the five classes of lithology, for which rock-water interactions were calculated using the model MPATH (see Chapter 2.4): 1. carbonate bearing rocks, 2. amphibolites, 3. melanocratic granite/gneiss, 4. leucocratic granite/gneiss, and 5. quaternary cover. The simplified boundaries of the lithologies within the catchments were transferred from the original geological maps to transparent foils and then digitised and geo-referenced on a tablet (see example in Figure 16). By overlaying digital terrain data with the geologylayer, the average horizontal and vertical distances (see Figure 1) from the lithological units to the lakes were calculated using GIS-tools. The digital terrain model had a resolution of 25x25m (source: DHM25©swisstopo). Critical Loads of Acidity for Alpine Lakes FOEN 2007 Figure 15 > Different types of geological maps (supplied by swisstopo, 2006). Geological Atlas (top), special geological maps (middle) and original geological maps (bottom). 48 3 49 > Input Data Figure 16 > Digitized lithological units in the catchment of lake No 77 (Lago Nero). Legend: melanocratic granite/gneiss (3), leucocratic granite/gneiss (4), quaternary cover (5), and surface water (9). Carbonate bearing rocks (1) and amphibolite (2) do not occur. Lattice 0.5x0.5 km. Table 13 > Mean values of area (AREA), distance of the flow path (DIST) and hydraulic gradient (dh/dl) per lithology-class, averaged over the 100 mapped catchments. No Description AREA DIST [ha] [m] dh/dl [-] 5.5 427 0.44 3.1 379 0.47 31.0 395 0.50 1 carbonate bearing rocks 2 amphibolite 3 melanocratic granite/gneiss 4 leucocratic granite/gneiss 15.2 545 0.46 5 quaternary cover 22.0 301 0.40 8 glacier 6.1 645 0.33 9 lake surface 3.9 0 -- 50 Critical Loads of Acidity for Alpine Lakes FOEN 2007 Table 13 shows a statistical summary of the analysed geological maps, lithological units and terrain data. AREA, DIST and dh/dl are explained in Chapter 2.4.4. The Table represents the ‘average’ of all 100 mapped lake catchments. The average catchment area is 86.8 ha and the lake surface 3.9 ha. The dominant lithology is melanocratic granite/gneiss with 31.0 ha on the average. However, the values vary largely among individual lakes (see also Table 14), e.g. carbonate bearing rocks only occur in 20 % of the catchments. The resulting weathering rates of the catchments are listed in Annex A.1 and summarised in Figure 17. Figure 17 > Cumulative frequency distribution of the average weathering rates (BCw,C) for the 100 catchments. Units: eq ha–1 a–1. 100 90 80 70 60 [%]50 40 30 20 10 0 0 3.4 250 500 750 1000 1250 1500 1750 2000 Terrestrial Sinks of Nitrogen and Base Cations For determining terrestrial sinks as well as for calculating deposition, land-use data are necessary. The land-use within the catchment areas was determined on the basis of a raster data set with a resolution of 100x100 m supplied by the Federal Office for Statistics (Geostat, land use statistics). The original 24 categories were aggregated to 3 main categories (j) for the FAB-application: – forest – grassland (mountain meadows and pastures) – bare land (gravel, rocks, glaciers) The lake surface was taken from the digitised geological maps, not from the land use statistics. 3 51 > Input Data Table 14 > Statistics of the land-use categories occurring in the 100 mapped catchments. statistical parameter area [ha] total catchment lake surface forest grassland bare land Minimum 8.38 0.22 0 0 0.47 Maximum 929.88 38.54 47.63 281.31 810.90 Mean 86.82 3.89 1.40 30.44 51.09 Median 52.31 1.97 0 16.25 29.73 Input values for uptake, immobilization and denitrification depending on land-use type are shown in Table 15. The net uptake of BC and N for forests is consistent with the amounts used for the critical load calculations for forests. For grassland, relatively low uptakes are used which reflect goat and sheep grazing. The denitrification factors (fde) correspond to well drained soils. Also for bare land a small value of fde was used accounting for re-emissions from drying bare rocks. Table 15 > Values for uptake, immobilization and denitrification used in the FAB-model application for alpine lakes. parameter land-use categories (j) forest 3.5 units grassland bare land N uptake (Nu) 56 36 0 eq ha–1 a–1 BC uptake (BCu) 54 18 0 eq ha–1 a–1 N immobilisation (Ni) 357 143 0 eq ha–1 a–1 denitrification factor (fde) (see Equation 2.8) 0.3 0.2 0.1 fraction In-lake Retention In the FAB model the retention of sulphur, nitrogen and base cations in the lake is incorporated (see Equations 2.4 and 2.25 in Chapter 2.2). In addition to runoff and lake: catchment ratio, the retention is characterised by the so-called net mass transfer coefficient sX (X=S,N,BC). For base cations no retention was assumed (sBC=0), whereas sS=0.5 m a–1 and sN=5 m a–1 was used for S and N retention, resp., for all lakes. These are average values from the ranges given in Baker and Brezonik (1988) and Dillon and Molot (1990), respectively. 52 Critical Loads of Acidity for Alpine Lakes FOEN 2007 4 > Results and Discussion Using the input data described in Chapter 3, the critical load functions of acidifying N and S for 100 lakes were computed with the generalised FAB model and the result is displayed in Figure 18. Figure 18 > Critical load functions of acidifying N and S for 100 lakes as computed with the generalised FAB model. Also shown are the respective pairs of N and S depositions (points) in the 100 catchments for the year 1995. 100 lakes Sdep 2000 1000 1000 2000 Ndep (eq/ha/yr) To get a better overview over these critical loads, the cumulative distribution functions (CDFs) of the maximum critical loads of S and N are shown in Figure 19. The figure shows that, e.g., the median values for CLmax(S) and CLmax(N) are about 570 and 800 eq ha-1 a-1, respectively. From Figure 18 one can not determine for which lakes critical loads are exceeded (in 1995) and for which not. One can only infer that some lakes are not exceeded, whereas others are exceeded: Certainly those for which all deposition dots lie above the critical load function. An overview over the exceedances is provided in Figure 20. The figure shows the cumulative distribution functions of the exceedances of the acidity critical loads displayed in Figure 18 computed for the depositions of N and S in 1980 (when they were about peaked), in 1995 and in 2010 (after the implementation of the Gothenburg Protocol and other current legislation). The percentage (or number – since there are exactly 100) of lakes protected (i.e. non-exceeded) increases from 46% in 1980 via 57% 4 53 > Results and Discussion in 1995 to 73% in 2010. As can be seen from the cumulative distribution functions, also the amounts by which critical loads are exceeded decrease dramatically over time. Figure 19 > Cumulative distribution functions of CLmax(S) (left) and CLmax(N) (right) of the 100 lakes. 100 lakes 100 % 80 60 40 20 0 0 400 800 1200 1600 2000 2400 Critical Loads (eq/ha/yr) Figure 20 > Cumulative distribution functions of the exceedances of the acidity critical loads for the 100 lakes shown in Figure 18 for the years 1980 (dashed line), 1995 (solid line) and 2010 (thin solid line). 100 lakes 100 % 80 60 40 20 0 0 200 400 600 800 Exceedance (eq/ha/yr) 1000 1200 54 Critical Loads of Acidity for Alpine Lakes FOEN 2007 The novel way of computing lake critical loads by using estimated catchment weathering rates prompts the question how the result compare with critical loads computed in the ‘traditional’ way, i.e. with the SSWC (or ‘Henriksen’) model using water chemistry data. For 19 of the 100 lakes water chemistry measurements are available for the years 2000– 2003 (Barbieri 2004, Annex A.1). And the volume-weighted average concentrations over those years (6–13 water samples taken at a depth of 0.4 m, see column 2 in Table 16) have been used to calculate the critical loads for acidity with the SSWC model (see Section 2.3). Figure 21 shows a comparison of the resulting CL(A) values with the values for CLmax(S) obtained with the FAB model for those 19 lakes. CLmax(S) was chosen for the comparison as the FAB model does not produce values for CL(A) but critical load functions for S and N (see Figure 4). Figure 21 > Comparison of the acidity critical load values obtained with the SSWC model and CLmax(S) output from the FAB model for 19 sites for which water chemistry is available (see also Table 16). CL(A) (SSWC) (eq/ha/yr) 2000 1600 1200 800 400 0 0 400 800 1200 1600 2000 CLmax(S) (FAB) (eq/ha/yr) Figure 21 shows that the SSWC critical loads are generally higher than the corresponding values for CLmax(S). Since in both models the same runoff and critical ANC limit have been used, the difference must be explainable by the difference in base cation inputs to the catchment (deposition, weathering, uptake), which is used in the generalised FAB model, and the observed base cation flux leaving the lake, which determines the results of the SSWC model. A comparison between the net input flux (area weighted BCnet = BCdep–Cldep+BCw–BCu) and the output flux (average Q•[BC–Cl]) is shown in Table 16 and illustrated in Figure 22. 4 55 > Results and Discussion Figure 22 shows that the measured flux of base cations leaving the lake is for about half of the catchments considerably larger than the (mostly) modelled net input, i.e. deposition plus weathering minus net uptake. There could be several reasons for this: Deposition and/or weathering rates are underestimated (the net uptake flux in these high alpine catchments is certainly small) or (the averages of) the measurements do not represent a true annual average. In any case, this discrepancies warrant further investigations. Table 16 > Critical loads and element net in and output fluxes for 19 lakes for which FAB-results and measurements are available (Barbieri 2004). All fluxes are in eq ha–1 a–1. For lake identification see ‘LakeNr’ and the Annex A.1. Nobs is the number of observations (samples). Sin and Nin are defined by Equations 2.6 and 2.9, respectively. LakeNr Nobs CLmax(S) CL(A) BCnet Q•[BC-Cl]ave Sin 64 13 67 7 76 7 77 7 79 81 Q•[S]ave Nin Q•[S]ave 961.4 740.1 1258.6 1218.0 485.3 1242.4 864.7 2152.2 380.0 530.4 470.8 378.0 440.1 1463.3 574.5 319.7 771.1 929.4 1089.5 262.5 1588.7 601.2 1472.2 429.8 736.1 513.9 279.6 2240.2 339.0 940.7 351.9 225.8 6 361.3 1062.3 7 1229.5 1487.5 691.0 1591.9 340.0 668.4 465.5 288.4 1545.6 2238.2 439.3 1023.5 589.6 331.1 84 7 513.5 667.5 88 7 457.3 1222.0 866.0 1218.4 457.9 660.7 500.3 433.4 736.3 1669.5 375.2 398.1 462.1 421.0 89 7 936.9 1598.5 92 7 730.8 501.2 1253.1 2497.0 497.8 1177.5 652.1 416.6 1003.9 878.4 375.8 338.3 465.3 302.6 93 11 1124.6 98 7 553.0 590.4 1460.8 1150.8 535.9 573.1 747.0 627.6 761.9 879.4 1270.2 438.6 454.7 575.3 102 7 777.7 507.6 771.5 1119.4 1341.0 568.2 599.3 807.9 529.4 104 11 535.3 106 7 727.9 484.1 898.4 1053.4 617.4 665.3 844.3 606.5 1771.8 1056.2 2423.3 579.3 797.0 731.0 367.5 109 7 639.0 208 7 475.8 742.3 947.5 1146.0 508.6 497.6 752.2 155.3 1056.3 818.4 1608.8 408.4 560.9 612.1 457.6 210 7 572.5 222 7 287.4 727.2 874.6 1174.3 515.6 529.6 732.3 281.4 314.0 667.2 809.1 439.3 654.5 504.2 350.7 56 Critical Loads of Acidity for Alpine Lakes FOEN 2007 Figure 22 > Comparison between net base cation input flux and average measured BC output flux for the 19 lakes studied. measured Q*[BC]ave (eq/ha/yr) 2500 BC in- and output fluxes 2000 1500 1000 500 0 0 500 1000 1500 2000 2500 Net BC input (eq/ha/yr) Figure 23 > Net (modelled) input and measured average output fluxes of sulphate and total inorganic nitrogen for the 19 lakes studied (averages 2000–2003). S in- and output fluxes 1600 measured Q*[N]ave (eq/ha/yr) measured Q*[S]ave (eq/ha/yr) 1600 1200 800 400 0 0 400 800 1200 Net S input (eq/ha/yr) 1600 N in- and output fluxes 1200 800 400 0 0 400 800 1200 Net N input (eq/ha/yr) Sulphur and nitrogen deposition are not used in the calculation of critical loads. Nevertheless, it is of interest to compare also input-output fluxes for sulphur and nitrogen compounds. Such a comparison is illustrated in Figure 23 (see also Table 16), for which as S 1600 4 > Results and Discussion and N deposition fluxes the mean of the available 1995 and projected 2010 depositions was used to make them compatible with the 2000–2003 average concentrations. Figure 23 (left panel) shows that the amount of sulphur leaving the catchments is for most of the sites (much) larger than the amount deposited (minus a small fraction modelled to be retained in the lake). This could be due to: a) an underestimate of the deposition at certain sites or b) the computed average concentration in the lake not being representative of the annual mean, c) the additional release of sulphate from geological sources in some of the catchments (e.g. pyrite, FeS, which could release sulphur during weathering) and d) the release of previously adsorbed sulphate following the reductions of S emissions and depositions since 1980 (see Figure 12), and/or e) the delay of water reaching the lake because of retention in the terrestrial catchment, which can be up to several decades (see Chapter 2.4). This would mean that a small part of the high S depositions in the 1980ies would reach the lake nowadays. Points c) and d) are supported by the results of a trend-analysis carried out by Steingruber and Colombo (2006) for the same lakes. They show that the trend in sulphate concentrations between 1980 and 2004 is significantly negative in only 14 out of 20 lakes. In the presence of S sources in the catchments, for both the Henriksen and FAB approach the model formulation should be revised. Figure 23 (right panel) shows that in some catchments there is more input of total inorganic N (nitrate plus ammonium) than is leaving via runoff, but the discrepancies are much less pronounced than for sulphur. In fact is not surprising at all that the N input is be larger than the measured output, since in the modelling of the N sinks only a small long-term sustainable immobilisation is assumed, whereas present N immobilisation could be considerably larger. 57 Critical Loads of Acidity for Alpine Lakes FOEN 2007 5 > Concluding Remarks An attempt has been made to improve the calculation of critical loads for Swiss highalpine lakes by explicitly estimating the weathering of base cations in the terrestrial catchment. This was done by combining a hydrological model, which estimates the flow paths through the catchment, with weathering calculations of individual lithological units, resulting in an overall catchment weathering rate. The First-order Acidity Balance (FAB) model was modified to accept these weathering rates as input, thus making it fully compatible with the widely used Simple Mass Balance (SMB) model for soil profiles. In contrast to the SSWC (‘Henriksen’) model the FAB model fully incorporates nitrogen processes and thus also allows assessing a catchment’s sensitivity to N acidity. The critical loads obtained for the 100 lakes have a median value of about 570 eq ha–1 a–1 for CLmax(S) and of about 800 eq ha–1 a–1 for CLmax(N), with the most sensitive below 100 eq ha–1 a–1. Critical loads of S thus calculated were compared with those from the SSWC model and found to be in the same range, albeit somewhat lower than the SSWC values. In this comparison it has to be kept in mind that the calculations of critical loads with the SSWC model depend on the water quality measurements chosen (date of sampling, averages taken, etc.), whereas the FAB model uses only time-independent data (catchment characteristics) as input. As a quality check, catchment input-output fluxes for S, N and base cations (BC) have also been calculated. The output fluxes (runoff) of base actions and especially sulphur are presently (2000–2003 average) higher the corresponding input fluxes (deposition plus weathering minus sinks such as net uptake). For nitrogen the results were the opposite. Whereas this is not surprising for N, it is less clear in the case of S and BC. Many possible reasons were identified, some of them relating to the lack of input data (e.g., sulphate sources in the terrestrial catchment), others due to the inherent simplicity of the critical load models (e.g., neglecting episodes such as spring snowmelt and runoff peaks). Overall, the new methodology has the advantage of (a) being more process-oriented (and thus easier to modify/improve), (b) differentiating better between catchments, and (c) allowing the comparison with S and N depositions. Nevertheless, further improvements, for example in the details of catchment geology, would be desirable. Also, the modelling could be refined, e.g., by taking into account time delays between deposition and leaching to the lake due to long travelling times of ions in the bedrock. Any improvement, however, will depend on the priority given to such a tool for aiding assessing emission reduction, both on a national and international level. 58 5 > Concluding Remarks > Acknowledgements This study was requested and financed by the Federal Office for the Environment (FOEN). We would like to thank Beat Achermann (FOEN) for his helpful support throughout the project and Dr. Niklaus Waber from the Institute of Geological Science, University of Bern, for assistance in hydrology and geochemical modelling. We would also like to thank Dr. Alberto Barbieri and Dr. Sandra Steingruber (Canton Ticino) for supplying the monitoring data and helpful comments as well as Dr. Rosario Mosello and his team (C.N.R - Institute of Ecosystem Study, Verbania Pallanza, Italy) for the fruitful discussions. 59 Critical Loads of Acidity for Alpine Lakes FOEN 2007 > Annexes > A1 List of Lakes List of the 102 lakes. Two lakes could not be modelled with genFAB: ID=63 (downstream lake) and ID=217 (no weathering rate). X and Y are coordinates in the national system. Z is the altitude above sea level. AL is the area of the lake (water surface) and AC the area of the whole catchment. 20 lakes with water chemistry measurements during 2000–2003 have a OBS_ID. Q is the runoff and BCw,C are the calculated average weathering rates of the catchments in keq ha–1 a–1. 60 61 > Annexes Table 17 > List of lakes. ID NAME X [m] 694475 Y [m] 2 Lai da Tuma 165335 3 Lai Urlaun 695436 5 Lai Verd 702572 9 no-name 723520 10 Lai Blau 702526 11 no-name 694982 12 Selvasee 732206 15 Lago Retico 711270 16 no-name 696399 17 Ampervreilsee 731267 159248 18 Lago d’Orsino 684387 19 no-name 696720 20 Laghi d’Orsirora 683450 21 Lago di Froda 694304 22 no-name 683988 23 no-name 685725 24 Lago di dentro 696350 25 Lago della Valletta sup. 683218 26 Laghi della Valletta 27 Curaletschsee 28 Lago di scuro 30 Lago dello Stabbio 31 Z [m] AL [ha] AC [ha] OBS_ID Q [m] BCw,C 2578 2.54 201.3 1.43 0.505 164634 2333 1.54 37.1 1.83 0.336 162606 2798 4.11 51.2 1.98 0.406 161289 2744 0.93 28.0 1.77 1.224 161006 2559 5.30 64.7 1.72 0.445 161030 2432 2.05 18.1 1.57 0.158 160692 2466 2.49 85.1 1.44 0.857 159452 2483 9.22 93.1 1.49 0.555 159237 2776 1.34 15.0 1.71 1.062 2605 0.73 63.3 1.57 0.714 158877 2437 4.10 100.2 1.59 0.497 158908 2681 2.32 36.9 1.72 1.205 158553 2513 3.91 29.8 1.61 0.219 158619 2589 0.91 33.3 1.35 0.557 158416 2414 1.64 47.2 1.49 0.335 158359 2196 2.15 127.9 1.38 0.147 158359 2545 0.57 14.3 1.74 0.156 158191 2536 0.79 21.3 1.54 0.297 683626 158067 2499 2.40 21.9 1.66 0.288 730256 158056 2650 3.30 208.3 1.65 1.251 696555 157612 2528 7.19 42.6 1.59 0.279 697627 157324 2437 8.11 36.6 1.53 0.618 no-name 690725 157333 2592 0.35 61.1 1.32 0.449 33 Lago della Piazza 686418 156733 2210 3.54 98.3 1.50 0.439 35 Lago di Stabiello 694522 156589 2257 0.94 46.0 1.60 3.493 36 no-name 686888 156462 2213 1.57 120.5 1.63 0.494 37 Lago di Tom 695942 156218 2221 9.32 195.6 1.59 1.528 38 Lago di dentro 699560 156280 2420 6.58 55.3 1.18 0.866 41 no-name 729006 156001 2841 0.62 62.8 1.68 1.137 43 no-name 729872 155945 2840 2.87 39.4 1.75 0.319 44 di Lago 693972 155382 2127 3.37 55.9 1.55 1.451 46 Lago dei Campanit 702522 155105 2467 0.83 23.4 1.92 0.406 47 Lago dei Canali 703424 154277 2383 0.75 266.1 1.81 0.883 48 Lago Pécian 701205 154023 2483 1.35 24.4 1.62 0.814 49 Lago Chiera Grande 701371 153091 2499 9.46 73.4 1.63 0.436 50 no-name 727832 153046 2696 2.30 49.9 1.99 0.472 51 Lago di Cari 705817 152275 2383 1.35 49.4 1.37 0.554 53 Laghetto Moesola 732995 150652 2212 6.80 122.2 1.62 0.571 54 Lago di Ravina 691715 150374 2140 1.84 96.8 1.71 1.135 55 no-name 694101 149816 2202 0.89 118.6 1.34 1.911 62 Critical Loads of Acidity for Alpine Lakes FOEN 2007 ID NAME X [m] Y [m] Z [m] AL [ha] AC [ha] OBS_ID Q [m] BCw,C 57 Lago di Prato 693223 148322 2222 2.73 55.5 1.13 2.014 60 Lago di Val Sabbia 686313 148699 2512 1.40 49.3 1.43 0.862 63 Laghetto Inferiore 688630 147849 2230 5.59 57.2 1107 1.39 0.861 1108 1.59 1.038 1.67 0.844 1.80 0.341 1.93 0.533 64 Laghetto Superiore 688009 147804 2314 8.30 124.5 65 Lago del Naret Piccolo 685894 147552 2522 2.32 118.0 66 Lago Cristallina 685651 146884 2474 0.68 19.9 67 Lago della Capannina Leit 698541 146795 2405 2.72 52.4 68 Lago Sfundau 683405 146255 2584 12.91 196.9 1.65 0.950 70 Lago Laiozz 685767 146299 2566 1.46 111.2 1.65 3.612 71 Lago del Corno 673049 145945 2608 2.54 53.2 1.66 2.031 72 Lagh di Stabi 730155 146025 2502 3.47 70.7 1.38 0.507 73 Lago della Zota 687160 145916 2373 1.04 31.6 1.81 0.401 74 Lagh Doss 735695 145754 1696 1.75 41.9 1.52 0.701 76 Lago di Morghirolo 698192 145205 2462 11.89 166.1 1303 1.67 0.797 77 Lago Nero 684556 144804 2507 12.70 72.3 1116 1.76 0.419 78 Lago dei Matörgn 680115 143823 2654 2.47 92.6 1.71 0.513 79 Lago della Froda 686004 143772 2542 1.98 67.3 1.66 0.453 80 Lago del Zött 681832 143034 2569 14.60 929.9 1.73 1.211 81 Lago di Mognola 696084 142870 2386 5.41 197.3 1106 1.63 1.244 84 Lago Barone 701005 139849 2516 6.62 50.8 1205 1.85 0.576 85 Lagh de Trescolmen 733606 139968 2383 1.95 153.9 1.60 1.132 86 Laghetto 704731 139197 2082 16.00 183.3 1.65 0.744 88 Laghetti d’Antabia 681055 137671 2337 6.85 81.7 1120 1.46 0.527 89 Lago dei Porchieirsc 700433 136895 2370 1.46 43.1 1204 1.63 0.914 92 Laghi della Crosa 680335 136082 2366 16.95 193.8 1121 1.47 0.830 93 Lago di Tomè 696320 135380 2098 5.76 294.3 1104 1.71 1.109 94 Lago 721787 134679 2223 1.52 29.1 1.83 0.290 95 Lago di Cava 722737 134548 2276 0.52 65.7 1.95 0.737 98 Lago d Orsalia 683514 132567 2299 2.63 40.7 1.68 0.631 99 Lago d’Efra 708248 132520 2144 1.84 97.9 1.85 0.810 100 Lago Coca 697367 129325 2083 0.62 18.5 1.47 0.317 101 Lago di Canee (Claro) 723905 128569 2378 2.47 32.9 1.84 0.367 102 Lago da Sascòla 687569 126239 1979 3.19 89.7 1130 1.75 0.801 104 L. del Starlaresc da Sgi 702923 125632 1979 1.07 23.4 1202 1.85 0.496 106 Lago d’Alzasca 688387 124656 1989 10.42 110.3 1131 1.74 0.760 109 Laghi dei Pozzöi 679620 124198 2092 1.11 33.3 1125 1.58 0.677 110 Lago Gelato 678268 123460 2243 0.76 25.2 1.58 0.505 201 Lago di Fieud 686361 154780 2316 0.26 36.6 1.59 0.587 202 Lago Tremorgio 698324 148301 2199 38.54 495.1 1.64 1.950 1304 1115 1123 203 Lago Chiera Piccolo 701658 152770 2381 1.26 11.9 1.55 0.208 204 Lago Cadagno 697688 156123 2118 26.22 209.8 1.44 1.568 205 Laghetti di Taneda 696382 156996 2364 0.22 8.4 1.59 0.489 63 > Annexes ID > NAME X [m] Y [m] Z [m] AL [ha] AC [ha] OBS_ID Q [m] BCw,C 206 Lago di Formazzöö 680094 133423 2483 2.58 70.1 1.66 0.663 207 Lago del Piatto 696509 141769 2346 0.48 31.2 1.42 0.528 208 Schwarzsee (Melo/Poma) 681966 132180 2442 0.32 23.6 1.72 0.568 209 Lago della Cavegna 680642 122505 2059 0.54 20.4 1.79 0.477 210 Lago di Sfille 681508 124206 2079 2.82 62.9 1.55 0.599 211 Lago del Pèzz 682686 124272 2126 1.30 31.0 1.82 0.596 212 Lago di Spluga 694948 130757 2083 0.36 27.9 1.84 0.532 213 Laghetto Pianca 701532 125847 2009 0.71 22.3 1.52 0.389 214 no-name 673821 149063 2710 0.56 50.6 1.80 0.861 215 no-name 675960 157309 2761 1.84 36.8 1.98 0.906 216 no-name 677469 149365 2507 0.58 27.4 1.41 0.670 217 no-name 680624 155144 2711 0.70 246.8 1.92 218 Lago del Forna 687075 149091 2468 0.71 52.3 1.47 1.402 1124 1128 219 Lago Scuro 687522 148094 2306 3.11 18.5 1.81 0.371 220 no-name 689678 155930 2396 0.41 9.7 1.51 1.413 221 no-name 700965 141831 2582 0.76 59.1 2.04 0.677 222 Laghetto Gardiscio 701275 142678 2629 1.14 12.1 1.93 0.336 223 no-name 680728 138042 2238 0.78 35.8 1.48 0.851 224 no-name 681050 136350 2369 9.43 209.2 1.47 1.252 225 Laghi di Motella 708790 151128 2339 0.93 35.3 1.79 0.320 1302 A2 FORTRAN subroutine genFAB In Chapter 2.2 a critical load function has been derived. The computation of the distinct nodes of that critical load function with the generalised FAB model for a catchment with given characteristics can be carried out with the aid of following the FORTRAN subroutine: subroutine genFAB (numLC,LCfrc,sfac,nfac,rhoS,rhoN,Niv,Nuv,fdev, & Lcrit,nCL,CLNv,CLSv) ! ! Returns the nCL distinct(!) nodes of the piecewise linear CL function ! in (CLNv(n),CLSv(n),n=1,nCL) computed with the generalised FAB model. ! Note: CLNv(1)=0,CLSv(1)=CLmaxS and CLNv(nCL)=CLmaxN,CLSv(nCL)=0. ! ! Input: ! numLC … number of land cover (LC) classes (incl. lake!) ! LCfrc(j) … fraction of LC j (LCfrc(1)+..+LCfrc(numLC)=1) ! sfac(j) … S-deposition factor on LC j; sfac(j)*Sdep=Sdep,j ! nfac(j) … N-deposition factor on LC j; nfac(j)*Ndep=Ndep,j ! rhoS … S in-lake retention fraction (0<=rhoS<=1) ! rhoN … N in-lake retention fraction (0<=rhoN<=1) ! Niv(j) … net N immobilisation on LC j ! Nuv(j) … net N uptake (N removal) on LC j ! fdev(j) … denitrification fraction on LC j Critical Loads of Acidity for Alpine Lakes FOEN 2007 ! ! ! ! ! ! ! Lcrit … ‘critical’ term (containing BC fluxes and ANClim) Output: nCL … number of distinct nodes (<=numCL+2) CLNv(l) … N-values of nodes (l=1(1)nCL) CLSv(l) … S-values of nodes (l=1(1)nCL) implicit none ! integer, intent(in) :: numLC real, intent(in) :: LCfrc(*), sfac(*), nfac(*), Niv(*), Nuv(*) real, intent(in) :: fdev(*), rhoS, rhoN, Lcrit integer, intent(out) :: nCL real, intent(out) :: CLNv(*), CLSv(*) ! integer :: i, k real :: x0, Nold, Nk, aS integer, allocatable :: indx(:) real, allocatable :: NiNuv(:), aNv(:), LNv(:) ! allocate (indx(numLC),NiNuv(numLC)) do k = 1,numLC NiNuv(k) = (Niv(k)+Nuv(k))/nfac(k) end do call indexr (numLC,Ninuv,indx) ! allocate (aNv(0:numLC),LNv(0:numLC)) aS = 0.; aNv(0) = 0.; LNv(0) = 0. do k = 1,numLC aS = aS+LCfrc(k)*sfac(k) i = indx(k) aNv(k) = aNv(k–1)+LCfrc(i)*(1.–fdev(i))*nfac(i) LNv(k) = LNv(k–1)+LCfrc(i)*(1.–fdev(i))*(Niv(i)+Nuv(i)) end do aS = (1.–rhoS)*aS aNv = (1.–rhoN)*aNv LNv = (1.–rhoN)*LNv ! CLNv(1) = 0. CLSv(1) = Lcrit/aS ! = CLmaxS nCL = 1 Nold = NiNuv(indx(1)) if (Nold > 0.) then ! lake Nu+Ni> 0! CLNv(2) = Nold CLSv(2) = CLSv(1) nCL = 2 end if do k = 1,numLC if (k < numLC) then Nk = NiNuv(indx(k+1)) else Nk = 1.e30 ! infinity end if if (Nk <= Nold) cycle ! skip identical node x0 = (LNv(k)+Lcrit)/aNv(k) nCL = nCL+1 if (x0 <= Nk) then CLNv(nCL) = x0 ! = CLmaxN CLSv(nCL) = 0. exit else CLNv(nCL) = Nk CLSv(nCL) = (x0-Nk)*aNv(k)/aS end if Nold = Nk end do deallocate (indx,NiNuv,aNv,LNv) return end subroutine genFAB The above routine calls the subroutine indexr, which indexes a vector vec(1:n), i.e. returns the vector indx(1:n) such that vec(indx(j)) is in ascending order for j=1,…,n. 64 > Annexes This routine is identical to the subroutine indexx in Press et al. (1992). If the inputs are such that (Ni,j+Nu,j)/nj (see Equation 2.17 in Chapter 2.2) is in ascending order for j=1,…, numLC, then the routine indexr is not needed and indx(j) can be replaced by j everywhere in the above subroutine. 65 66 Critical Loads of Acidity for Alpine Lakes FOEN 2007 > Indexes Glossary ANC acid neutralising capacity (= sum of base cation minus strong acid anions) Bc (sum of) base cations taken up by vegetation (= Ca2+ + Mg2+ + K+) BC (sum of) base cations 2+ 2+ + + (= Bc + Na+ = Ca + Mg + K + Na ) release of BC by weathering of minerals in the soil or BCw bedrock average BCw of a catchment BCw,C CCE Coordination Centre for Effects CL(A)critical loads of acidity CLmax(S) maximum critical load of sulphur CLmax(N) maximum critical load of nitrogen eq equivalents = moles of charges (=molc) EMEP Co-operative Programme for Monitoring and Evaluation of the Long-range Transmission of Air Pollutants in Europe FAB First-order Acidity Balance model LRTAP (Convention on) Long-range Transboundary Air Pollution SMB Simple Mass Balance (model) SSWC Steady-State Water Chemistry (model) UNECE United Nations Economic Commission for Europe Figures Figure 1 Differences in the distribution of catchment weathering rates and critical loads of acidity of 45 high-alpine lakes in the Ticino area (EKG, 1997). Figure 2 Map of the 100 lakes modelled with genFAB. The 19 lakes with water chemistry measurements during 2000–2003 are shaded in blue color. 13 15 Figure 3 Flowchart of the main procedural steps in this study. 17 Figure 4 Piece-wise linear critical load function of S and acidifying N for a lake as defined by catchment properties. 23 Figure 5 Head distribution isolines and three particle paths in the Lago Superiore (left) and Lago Inferiore (right) area. The coloured areas represent lithology classes: red = gneiss, blue = amphibolite, green = quarterny cover. 31 Figure 6 General scheme of the rock-water interaction model MPATH. 34 Figure 7 Ion concentrations in rock-water for Leucocratic Crystalline Rocks (top) and Melanocratic Crystalline Rock Types (bottom). Y is the best-fit function for BC (dotted line). 36 Figure 8 Ion concentrations in rock-water for Amphibole Bearing Rocks (top) and Carbonate Rock Types (bottom). Y is the best-fit function for BC (dotted line). 37 Figure 9 Ion concentrations in rock-water for quarternary cover. y is the best-fit function for BC. 38 Figure 10 Scheme of the simplified flow path calculation (left)as well as the calculated total areas (AREAL) and the mean flow path lengths (DISTL) per lithology-class in catchment No. 64 (Lago Superiore) (right). 40 Figure 11 Map of modelled N deposition for the year 1995. 43 Figure 12 Cumulative frequency distributions of S and N deposition for the 100 catchments in 1980, 1995 and 2010. Bc is the timeindependent deposition of base cations. 45 Figure 13 Map of precipitation rates, average 1961–1999 (FOWG 2000). 46 Figure 14 Cumulative frequency distributions of mean precipitation (P), evapotranspiration (ET) and runoff (Q) for the 100 catchments. 47 Figure 15 Different types of geological maps (supplied by swisstopo, 2006). Geological Atlas (top), special geological maps (middle) and original geological maps (bottom). 48 Figure 16 Digitized lithological units in the catchment of lake No 77 (Lago Nero). 49 67 > Verzeichnisse Figure 17 Cumulative frequency distribution of the average weathering rates (BCw,C) for the 100 catchments. Units: eq ha–1 a–1. Figure 18 Critical load functions of acidifying N and S for 100 lakes as computed with the generalised FAB model. Also shown are the respective pairs of N and S depositions (points) in the 100 catchments for the year 1995. Figure 19 Cumulative distribution functions of CLmax(S) (left) and CLmax(N) (right) of the 100 lakes. Figure 20 Cumulative distribution functions of the exceedances of the acidity critical loads for the 100 lakes shown in Figure 18 for the years 1980 (dashed line), 1995 (solid line) and 2010 (thin solid line). 33 Table 7 Maximum total daily flow and travel times through gneiss layer for different gradients. 33 Table 8 Travel times through gneiss layer for different porosities. 34 Table 9 Input parameters for the Model MPATH. 35 Table 10 Mean values of area (AREAL), distance of the flow path (DISTL) and hydraulic gradient (dhL/dlL) per lithology-class (L) in the catchment No. 77 (Lago Nero). 41 Table 11 Linear regression models for calculating ion concentrations in precipitation in the canton Ticino, based on measurements from 1993 to 1998 (SAEFL 2001). 42 Table 12 Deposition of S, NOy and NHx in 1980, 1995 and 2010 for the EMEP-cell 24/13 (from Schöpp et al. 2003) and derived deposition ratios normalised to 1995. 44 Table 13 Mean values of area (AREA), distance of the flow path (DIST) and hydraulic gradient (dh/dl) per lithology-class, averaged over the 100 mapped catchments. 49 Table 14 Statistics of the land-use categories occurring in the 100 mapped catchments. 51 Table 15 Values for uptake, immobilization and denitrification used in the FAB-model application for alpine lakes. 51 Table 16 Critical loads and element net in and output fluxes for 19 lakes for which FAB-results and measurements are available (Barbieri 2004). 55 Table 17 List of lakes. 61 50 52 53 53 Figure 21 Comparison of the acidity critical load values obtained with the SSWC model and CLmax(S) output from the FAB model for 19 sites for which water chemistry is available (see also Table 16). 54 Figure 22 Comparison between net base cation input flux and average measured BC output flux for the 19 lakes studied. 56 Figure 23 Net (modelled) input and measured average output fluxes of sulphate and total inorganic nitrogen for the 19 lakes studied (averages 2000–2003). Table 6 Travel times through gravel layer for different porosities. 56 Tables Table 1 Workflow for calculating and mapping critical loads. 14 Table 2 Hydrological characteristics of three distinct lithology units. 27 Table 3 Calculation of infiltrations for a one-layer system. Units are m³ d–1. 29 Table 4 Estimated infiltrations for a three-layer system. Units are m³ d–1. 30 Table 5 Maximum total daily flow through gravel layer for different gradients. 33 68 Critical Loads of Acidity for Alpine Lakes FOEN 2007 References Baker LA, Brezonik PL (1988) Dynamic model of in-lake alkalinity generation. Water Resources Research 24: 65-74 Barbieri A, and Pozzi S (2001) Acidifying deposition. Southern Switzerland, Environmental Documentation - Air, Vol. 134, Swiss Agency for the Environment, Forests and Landscape (SAEFL), Bern, 113 pp Barbieri A (2004) Results of water chemistry measurements for 20 alpine lakes in the canton of Ticino, pers. comm, 29.1.2004 Baumgartner A, Reichel E, Weber G (1983) Der Wasserhaushalt der Alpen. Oldenbourg Verlag, München, Wien Brakke DF, Henriksen A, Norton SA (1990) A variable F-factor to explain changes in base cation concentrations as a function of strong acid deposition. Verh. Internat. Verein. Limnol. 24: 146-149 Boggero A, Marchetto A, Barbieri A, Sassi A, Conedera M, Tartari GA, Mosello R (1996) Idrochimica dei laghi alpini del canton Ticino (alpi centrali) in relazione con la chimica delle precipitazioni. Volume 57, Documenta dell’ Istituto Italiano di Idrobiologia, Pallanza, Italy De Jong JE (1996) The methodology for the calculation of critical loads of acidity for high altitude lakes – application on lakes in Canton Ticino (Switzerland). Technical Report: 44, Istituto Italiano di Idrobiologia, Pallanza, Italy Dillon PJ, Molot LA (1990) The role of ammonium and nitrate retention in the acidification of lakes and forested catchments. Biogeochemistry 11: 23-43 EKG (1997) Reevaluation kritischer Säureeinträge für Oberflächengewässer – Konzeptstudie. Eggenberger & Kurz GeoScience, Interim Report, Bern Fetter CW (1994) Applied Hydrogeology. Macmillan, New York FOEFL (ed) (1994) Critical loads of acidity for forest soils and alpine lakes – steady-state mass balance method. Federal Office of Environment, Forests and Landscape Berne, Switzerland, Environmental Series Air, 238: 68pp FOWG (1999) Mean Annual Actual Evaporation 1973–1992. Hydrological Atlas of Switzerland, Federal Office for Water and Geology (FOWG), Bern FOWG (2000) Precipitation Maps 1961–1999. Hydrological Atlas of Switzerland, Federal Office for Water and Geology (FOWG), Bern, based on monitoring data from MeteoSwiss, Zürich Freeze RA, Cherry JA (1979) Groundwater. Englewood Cliffs: PrenticeHall Henriksen A (1984) Changes in base cation concentrations due to freshwater acidification. Verh. Internat. Verein. Limnol. 22: 692-698 Henriksen A, Lien L, Traaen TS (1990) Critical loads for surface waters – chemical criteria for inputs of strong acids. Norwegian Institute for Water Research Oslo, Norway, Acid Rain Research Report 22: 45pp Henriksen A, Forsius M, Kämäri J, Posch M, Wilander A (1993) Exceedance of critical loads for lakes in Finland, Norway and Sweden: Reduction requirements for nitrogen and sulfur deposition. Norwegian Institute for Water Research Oslo, Norway, Acid Rain Research Report 32: 46 pp Henriksen A, Posch M (2001) Steady-state models for calculating critical loads of acidity for surface waters. Water, Air and Soil Pollution: Focus 1: 375-398 Hettelingh J-P, De Vries W (1991) Mapping Vademecum. RIVM Report No. 259101002, Coordination Center for Effects, National Institute of Public Health and Environmental Protection, Bilthoven, The Netherlands Hettelingh J-P, Downing RJ, De Smet, PAM (eds) (1991) Mapping Critical Loads for Europe. CCE Technical Report No. 1, RIVM Report No. 259101001, Coordination Center for Effects, National Institute of Public Health and Environmental Protection, Bilthoven The Netherlands Hindar A, Posch M, Henriksen A, Gunn J, Snucins E (2000) Development and application of the FAB model to calculate critical loads of S and N for lakes in the Killarney Provincial Park (Ontario, Canada). Report SNO 4202-2000, Norwegian Institute for Water Research, Oslo, Norway, 40 pp ICP-Waters (ed) (2000) The 12-year Report: Acidification of Surface Water in Europe and North America; Trends biological recovery and heavy metals. Norwegian Institute for Water Research (NIVA), Oslo, ICP Waters Report 52/2000, 115pp Kaste Ø, Dillon PJ (2003) Inorganic nitrogen retention in acid-sensitive lakes in southern Norway and southern Ontario, Canada – a comparison of mass balance data with and empirical N retention model. Hydrological Processes 17: 2393-2407 Kelly CA, Rudd JWM, Hesslein RH, Schindler DW, Dillon PJ, Driscoll CT, Gherini SA, Hecky RE (1987) Prediction of biological acid neutralization in acid-sensitive lakes. Biogeochemistry 3: 129-140 Lichtner PC (1985) Continuum model for simultaneous chemical reactions and mass transport in hydrothermal systems. Geochim. Cosmochim. Acta 49: 779-800 Lichtner PC (1988) The quasi-stationary state approximation to coupled mass transport and fluid-rock interaction in a porous medium. Geochim. Cosmochim. Acta 52: 143-165 69 > Verzeichnisse Lichtner PC (1992) Time–space continuum description of fluid/rock interaction in permeable media. J. Geophys. Res. 28, 3135-3155 LSA (1999) COLumn Alpine Lake Chemistry - COL-ALC 1997 – Chemistry, physics and hydrobiology. Laboratorio Studi Ambientali (LSA), Lugano/Bellinzona, Switzerland and Federal Office of Environment, Forests and Landscape (FOEFL), Bern, Switzerland, Interim Report, February 1999 Nilsson J, Grennfelt P (1988) Critical Loads for Sulfur and Nitrogen. Nordic Council of Ministers, Miljørapport 1988:15, 418 pp Posch M, Kämäri J, Forsius M, Henriksen A, Wilander A (1997) Exceedance of critical loads for lakes in Finland, Norway and Sweden: Reduction requirements for acidifying nitrogen and sulfur deposition. Environmental Management 21(2): 291-304 Press WH, Teukolsky SA, Vetterling WT, Flannery BP (1992) Numerical Recipes in FORTRAN (2nd ed.) Cambridge University Press, 963 pp Rapp L, Bishop K (2003) Modeling surface water critical loads with PROFILE: Possibilities and challenges. Journal of Environmental Quality 32: 2290-2300 Rihm B, Kurz D (2001) Deposition and Critical Loads of Nitrogen in Switzerland. Water, Air and Soil Pollution 130, 1223-1228 Steingruber S, Colombo L (2006) Impacts of Air Pollution on Alpine Lakes and Rivers – Monitoring Results from 1980 to 2004. Ed: Federal Office for the Environment (FOEN), State of the environment No. …, Berne, in print. Sverdrup H, De Vries W, Henriksen A (1990) Mapping Critical Loads: A guidance manual to criteria, calculations, data collection and mapping. In: UN ECE Annex to chapter 4 of the Mapping Manual. Also published as Miljörapport 1990: Nordic Council of Ministers, Copenhagen swisstopo (ed) (1984) Atlas of Switzerland 1:500’000 (map number 7A: soils, overview). Federal Office of Topography, Wabern-Bern, Switzerland, 2nd Edition swisstopo (2006) Geological Atlas of Switzerland, Special Geological Maps and Original Geological maps. Maps supplied by the Federal Office of Topography (swisstopo) Thimonier A, Schmitt M, Waldner P, Rihm B (2005) Atmospheric deposition on Swiss Long-term Forest Ecosystem Research (LWF) plots. Environmental Monitoring and Assessment 104: 81-118 Thöni L, Brang P, Braun S, Seitler E, Rihm B (2004) Ammonia monitoring in Switzerland with passive samplers: patterns, determinants and comparison with modelled concentrations. Environmental Monitoring and Assessment 98: 95-107 Schöpp W, Posch M, Mylona S, Johansson M (2003) Long-term development of acid deposition (1880-2030) in sensitive freshwater regions in Europe. Hydrology and Earth System Sciences 7(4): 436-446 UBA (ed) ( 2004) Manual on methodologies and criteria for Modelling and Mapping Critical Loads & Levels and Air Pollution Effects, Risks and Trends. Convention on Long-range Transboundary Air Pollution. Umweltbundesamt Berlin, UBA Texte 52/04, distributed at SAEFL (ed) (2001) Acidifying Deposition – Southern Switzerland. Environmental Documentation No. 134. Swiss Agency for the Environment, Forests and Landscape (SAEFL), Berne, 113 pp http://www.icpmapping.org SAEFL (ed) (2003) Modelling of PM10 and PM2.5 ambient concentrations in Switzerland 2000 and 2010. Environmental Documentation No. 169, Swiss Agency for the Environment, Forests and Landscape (SAEFL), Berne SAEFL (ed) (2004) Modelling of NO2 and benzene ambient concentrations in Switzerland 2000 to 2020. Environmental Documentation No. 188, Swiss Agency for the Environment, Forests and Landscape (SAEFL), Berne Spitz K, Moreno J (1996) A Practical Guide to Groundwater and Solute Transport Modeling. John Willey and Sons, New York Sposito G (1989) The chemistry of soils. New York, Oxford University. 277p. Steingruber S (2001) Nutrient transformations in a wetland pond. DISS ETH No. 13939, Zürich. http://e-collection.ethbib.ethz.ch/ecol-pool/diss/ fulltext/eth13939.pdf UNECE (1994) Protocol to the 1979 Convention on Long-range Transboundary Air Pollution on Further Reduction of Sulphur Emissions. United Nations Economic Commission for Europe, Geneva, Switzerland, ECE/EB.AIR/40 Wolery TJ, Jackson KJ, Bourcier WL, Bruton CJ, Viani BE, Knauss KG, Delany JN (1990) Current status of the EQ3/6 software package for geochemical modelling. In: DC Melchior and RL Bassett (eds): Chemical Modeling of Aqueous Systems II: Washington D. C., American Chemical Society Symposium Series 416, pp. 104-116. Zobrist J, Drewer JJ (1990) Weathering processes in Alpine watersheds sensitive to acidification. In: M Johannessen, R Mosello and H Barth (eds): Acidification Processes in Remote Mountain Lakes. Proc. EEC Workshop, Pallanza, Italy, 20-22 June 1989, Air Pollution Research Report 20: 149-161
© Copyright 2024 Paperzz