• No results found

Phosphorus in the sediment of L. Hällerstadsjön: spatial distribution, fractions and release to the water volume

N/A
N/A
Protected

Academic year: 2021

Share "Phosphorus in the sediment of L. Hällerstadsjön: spatial distribution, fractions and release to the water volume"

Copied!
28
0
0

Loading.... (view fulltext now)

Full text

(1)

Linköping University | Department of Physics, Chemistry and Biology Master´s thesis, 60 ECTS | Educational Program: Physics, Chemistry and Biology Spring semester 2016 | LITH-IFM-x-EX—16/3231--SE

Phosphorus in the sediment of L. Hällerstadsjön:

spatial distribution, fractions and release to the

water volume

Nana Osei-Asibey Osafo

Examinator: Karin Tonderski, Linköping University

Tutor: Anders Hargeby

(2)

Avdelning, institution

Division, Department

Department of Physics, Chemistry and Biology Linköping University

URL för elektronisk version

ISBN

ISRN: LITH-IFM-x-EX--16/3231--SE

_________________________________________________________________ Serietitel och serienummer ISSN

Title of series, numbering ______________________________

Språk Language Svenska/Swedish Engelska/English _English _______________ Rapporttyp Report category Licentiatavhandling Examensarbete C-uppsats D-uppsats Övrig rapport _____________ Titel

Title:

Phosphorus in the sediment of L. Hällerstadsjön: spatial distribution, fractions

and release to the water volume

Författare

Author: Nana Osei-Asibey Osafo

Nyckelord

Keyword

Bioavailable phosphorus, Eutrophication, Phosphorus fractionation, Phosphorus flux, Sediments.

Sammanfattning

Abstract: In freshwater systems phosphorus (P) is the limiting element in the cause of eutrophication. In many Swedish lakes, causes of eutrophication have been attributed to more of internal loading than external since the external loading has been fairly well managed. Internal loading is linked to the mobility of sediment P, which are known to be Bioavailable P (BAP). Sediments from Lake Hällerstadsjön in Sweden was studied to know the BAP concentration and its possible release into the water column under reduced conditions. Sediments were sampled at two different depths, 0-5 cm and 5-10cm. BAP was determined by a phosphorus fractionation scheme. Sediments were incubated under oxic and anoxic conditions in the laboratory to evaluate sediment P release. Spatial variation in the distribution of P forms across the lake was also studied, in order to examine possible local patterns, particularly along a transect from the main inlet to the outlet. Fractionation analyses showed a trend of; Residual-P > NaOH-P > HCl-P > BD-P > Loosely bound P. The fractions constituting the BAP was higher at the 0-5 cm sediment depth than 5-10 cm. Sediment P flux was recorded for anoxic but not oxic sediment. BAP correlated significantly with sediment P flux (P= 0.01). Spatially, the P distribution varied both at depths and along a latitudinal transect, from the main inlet to the outlet.

Dredging of the surface sediments with high BAP content would possibly be an effective means of preventing eutrophication of the lake.

.

Datum

(3)

Contents

1

Abstract ... 1

2

Introduction ... 2

2.1

Phosphorus Fractionation ... 2

2.2

Sediment P release ... 3

3

Materials & methods ... 4

3.1

Description of site ... 4

3.2

Sampling ... 4

3.3

Phosphorus Fractionation and Spatial Distribution ... 5

3.4

Sediment Incubation ... 7

3.5

Sediment P Release and Flux ... 8

3.6

Statistical analysis ... 8

4

Results ... 8

4.1

Phosphorus fractionation-BAP... 8

4.2

Spatial Distribution of P ... 11

4.3

Sediment P Release and Flux ... 12

5

Discussion ... 14

5.1

Phosphorus Fractionation-BAP ... 14

5.2

Spatial Distribution of P ... 15

5.3

Sediment P Release and Flux ... 15

6

Conclusion ... 16

7

Ethical and Social Issues ... 16

8

Acknowledgement ... 16

9

Reference ... 17

(4)

1 1 Abstract

In freshwater systems phosphorus (P) is the limiting element in the cause of eutrophication. In many Swedish lakes, causes of eutrophication have been attributed to more of internal loading than external since the external loading has been fairly well managed. Internal

loading is linked to the mobility of sediment P, which are known to be Bioavailable P (BAP). Sediments from Lake Hällerstadsjön in Sweden was studied to know the BAP concentration and its possible release into the water column under reduced conditions. Sediments were sampled at two different depths, 0-5 cm and 5-10cm. BAP was determined by a phosphorus fractionation scheme. Sediments were incubated under oxic and anoxic conditions in the laboratory to evaluate sediment P release. Spatial variation in the distribution of P forms across the lake was also studied, in order to examine possible local patterns, particularly along a transect from the main inlet to the outlet. Fractionation analyses showed a trend of; Residual-P > NaOH-P > HCl-P > BD-P > Loosely bound P. The fractions constituting the BAP was higher at the 0-5 cm sediment depth than 5-10 cm. Sediment P flux was recorded for anoxic but not oxic sediment. BAP correlated significantly with sediment P flux (P= 0.01). Spatially, the P distribution varied both at depths and along a latitudinal transect, from the main inlet to the outlet.

Dredging of the surface sediments with high BAP content would possibly be an effective means of preventing eutrophication of the lake.

(5)

2 2 Introduction

Eutrophication is one of the main environmental problems in aquatic systems. Phosphorus (P) is an element that plays a key role in eutrophication of inland waters. This is due to the fact that P is a limiting nutrient for plant production in most freshwater systems. Sources of P in the waterbodies are mainly characterized under external and internal loadings. The external loadings include run-offs from agricultural farmlands and urban areas, as well as point sources like sewage treatment plants and industries. Internal loading is mainly a result of release of P from the sediments.

Internal loading occurs under conditions like high pH, high temperatures (summer seasons in temperate areas), depletion of dissolved oxygen (DO) etc. (Bostrom, 1988). Studies have shown that external loading have been reduced or managed well during the past years and the major cause of eutrophication in some lakes is the release of P from sediments. (Abrams and Jarell, 1995; Xie et al., 2003). Sediments have been seen to play an essential role for the concentration in the water phase, the transport and eventually the fate of phosphorus in aquatic ecosystems. This is due to the sediment acting both as a site for accumulation and for subsequent release of phosphorus (Søndergaard et al., 1996; Kleeberg et al., 1997).

Orthophosphate anion (HPO42-) is the main P dissolved ion, which could be adsorbed to

sediments as well as be dissolved in water. Orthophosphate is found to be bound to Fe and Al in hydrated oxides, also the mineral lattices with a specific ligand is seen to be a suitable surface with which this bond can occur. Water or hydroxide in this chemical reaction is replaced by the PO43--P anion (Hingston et al. 1972, 1974). Fine clay particles are known to

have a high content of Al and Fe oxides, also with a large surface area for P-binding reactions. This coupled with the slowly settling nature of clay in the water column makes it very efficient in binding dissolved P and gradually transforming it to sediment bound P. Sediment is of interest due to its ability to adsorb nutrients, pesticides and other pollutants which end up in water bodies (Dorioz and Ferhi, 1994; Svendsen et al., 1993). Due to different forms of P that adsorb to sediments, there is need to assess the forms of P and the potential of its release under certain environmental conditions. Phosphorus fractionation could be used in such assessments. This is used because total phosphorus (TP) only gives the total concentration and that in its sense does not give room for predicting potential ecological danger (Psenner et al., 1984). Phosphorus fractionation, which is a sequential extraction of P, has been noted as a useful technique in characterizing various P compounds (De Groot, 1990; Pardo, 1998; Psenner et al., 1998; Zhou et al., 2001). The fractionation scheme is based on differences in reactivity of solid phases to different extractant solutions (Hieltjes et al., 1980). 2.1 Phosphorus Fractionation

This is a scheme mainly to ascertain the various fractions or P forms found in the sediments. Psenner et al. (1984) proposed the very first fractionation scheme, which was then modified by subsequent scientist notable is Jensen and Thamdrup (1993). Different fractions in the scheme were defined as: labile P or loosely bound P, redox sensitive iron (Fe)-bound P (BD-P), hydrated ions of Al-bound P (NaOH-(BD-P), calcium (Ca)-bound P (HCl-(BD-P), and residual-P which consists mainly of organic P and apatite-P.

(6)

3

Loosely-P, Fe-P and Al-P (the fractions that would also be extracted in a single extraction with NaOH) are known to constitute the Bioavailable P (BAP). Particularly the Fe-P could be released in the event of depletion of oxygen and other conditions. Ca-P is only released under acidic conditions. As for the residual-P, Steinberg and Muenster (1985), showed that P in organic matter could either be part of the organic molecule itself or be bound to the organic matter via metallic cations like Al3+ and Fe3+. This P could be released due to both microbial decomposition of organic matter or depletion of oxygen, which may cause a reduction of the Fe3+ and a subsequent release of P from organic-P.

2.2 Sediment P release

Dissolved oxygen (DO) is known to be one of the factors that really influences the release of mobile P from sediments. P is often released when O2 is being depleted and P form bound to

Fe is mainly the one released (Einsele 1938; Mortimer 1941, 1942). As time went on studies showed that even in non-depleted O2 conditions P could be released. This was mostly the

case in very shallow eutrophic lakes. (Nriagu and Dell 1974; Boström et al. 1982; Hupfer et al. 2004). The increase in saturation of P binding sites in sediments or the degradation of organic materials by microbes were plausible reasons for the release of P in oxic conditions. Another instance where P release could be possible in oxic conditions, is when in algal bloom with very high pH surface water mixes with the fluffy surface of sediments of low pH. This elevates the concentration of hydroxyl ion (OH-) which tends to induce the release of P (Drake and Heaney, 1987). Studies have shown how changes in pH could result from photosynthetic activity in which carbon dioxide (CO2) is removed by algae or plants. The

removal of CO2 tends to elevate the pH even to a pH of 10 or more (Boers1991, Frodge

1991). The release of P at high pH is attributed to the desorption of P from ferric hydroxide, where P is replaced with hydroxide (Drake and Heaney, 1987; Boers, 1991; Jensen et al., 1992).

P release mostly is the largest source of internal P loading in a stratified lake with anoxic hypolimnion (Nürnberg 1984, 1987). Fluctuations in localized oxygen also contributes to P release (Penn et al., 2000). Studies have revealed the role sulphur plays in P release under anoxic conditions (Caraco et al., 1993). At lower redox potentials, sulphates and iron are reduced, iron sulphide (FeS) precipitates and iron is removed from the anoxic water by sulphides. In the absence of reoxygenation, P that was bound to iron oxide would then be released (Caraco et al., 1993).

The reduction of external loading has led to the development of several restoration techniques and subsequent management practices. Biomanipulation, i.e. manipulation of the fish

community, the addition of chemicals to bind the phosphorus tightly to the sediment surface and dredging of the upper layer sediments are some techniques have been used in the

restoration plan (Ripl, 1976; Klein, 1997; Søndergaard et al., 2007; Witcheler et al, 2007) The county of Östergötland has been recently awarded a grant to restore lakes by removing the upper sediment layer. Understanding the characteristics of the sediment in L.

Hällerstadsjön, would give an idea of how much of the sediment P is bioavailable and the possibility of the sediment P being released. Studying the sediment in two different depths

(7)

4

would give an idea of how the concentration of P tends to change with depth thus could be a way of knowing if pollution is recent or not. Sampling across the lake would also give an idea of how P is spatially distributed on the lake and that would be vital in planning for the

dredging. The study, which is in connection with the county administration, was done to estimate the amount of BAP in the P forms elucidated by the fractionation procedure and how P is spatially distributed in the lake. Also, the possibility of sediment P being released from BAP was studied in anoxic and oxic conditions.

The hypothesis of the thesis was therefore,

 TP in the sediment would decrease from the inlet to the outlet and

 Bioavailable P would be released in anoxic conditions but not in oxic conditions. 3 Materials and Methods

3.1 Description of L. Hällerstadsjön

Lake Hällerstadsjön is a eutrophic lake located between Östra Ryd and Västra Husby, in the county of Östergotland in Sweden. The lake is found on latitude 58o 27’ (58.45 o) north and longitude 16o 8l (16.1333o) east. The average elevation is 78 meters (256 feet). The lake has an average depth of about 4 m and an area of 1.47 Km2 (SMHI, 2012). It is surrounded by

farmland, forestry, animal husbandry, wetlands and houses. Sediment from Lake Hällerstadsjön serves as a receiving end of drainage water from arable fields and other sources of P.

3.2 Sampling

Sediment samples were collected from the lake with the aid of a gravity core sampler. The undisturbed sediment core was forced upward with a piston inserted from below. A slicer unit was used to remove 2.5 cm high sections of the top of the sediment core each time. This was to enhance the attainment of two different depths, that was 0-5 cm and 5-10 cm. Different depth was necessary to understand the vertical profile of the lakes sediments. Also, this was done to give an idea of how changes in the concentrations of TP content and fractions in the two depths could be related to time. Sampling sites included 17 randomly selected sites. In addition, three sites were selected in the eastern part of the lake, to include a semi-isolated bay (Figure 1). The samples were stored in darkness at 5 oC until the day of analysis.

(8)

5

Figure 1. Map of Lake Hällerstadsjön with sites of sediment sampling on 10th of July 2015 are

indicated by dots and the corresponding site number in the map. 3.3 Phosphorus Fractionation and Spatial Distribution

Samples were analyzed for organic matter (loss on ignition, LOI), TP, and fractions of P. The fractionation scheme used was as proposed by Paludan and Jensen (1995; Figure 2), except that the procedures for separating humic acid bound P (HA-P) and organic P were not performed. Each respective fraction was acidified by a known volume of sulphuric acid and stored in darkness at 5ºC until the day of analysis.

Both total phosphorus (TP) and the dissolved reactive phosphorus (DRP) in the extracts from the fractionation process were determined by the ascorbic acid method according to the Swedish Standards Institute standard (SIS, 1996). The Residual-P was obtained by subtracting all inorganic fractions from the TP. Residual-P constitutes both organic and refractory-P. Loosely bound P, BD-P, NaOH-P and HCl-P constitutes the inorganic P forms. A subsample of each sediment sample was dried to constant mass with the corresponding moisture content recorded. Dried sediment was ground and sieved with a 2 mm mesh size.

(9)

6

0.5 g of the sieved sediment was put in a furnace at a temperature of 550 oC for 2 hours. This was done to estimate the loss on ignition (LOI, organic matter). Total phosphorus (TP) was determined by digestion of the burnt samples with 1 M HCl at 120 ºC and 222 KPa according to Andersen (1976) and Svendsen et al. (1995).

To assess the spatial distribution and the variation of TP and the P forms across the lake area, a regression analysis was conducted for the respective P form as against the latitude (North coordinates). This was chosen, because the outlet of the lake is situated north from the inlet. Thus, increasing latitude could be used to describe the inlet to outlet transect.

The sediment samples were stored at 5oC for 28 weeks, after which the P fractionation scheme was repeated for the sediment samples from the 0-5 cm depth. This was done to check for consistency in the P fractions also after storage and to get background data for the sediment P flux experiment. A paired t-test was conducted using R 0.98.1103 to quantify the significance of any difference due to the storage period.

Data from the repeated P fractionation scheme (only samples from 0-5 cm depth) showed significantly lower levels of Ca-P after storage (p = 0.017, Table 1). Other inorganic

fractions, loosely bound P, Fe-P and Al-P, showed no significant effects of storage (p-values of 0.053, 0.292 and 0.284, respectively). Al-P constituted the largest fraction at both

fractionation occasions with loosely bound P being the smallest fraction (Table 1). The data paved way for the sediments to be used for the incubation experiment.

Table 1. Results of fractionation before (first) and after (second) storing the sediment from 0-5 cm depth for 4 months at 5 ºC. Means and standard deviations for each P form in the two fractionation schemes with the corresponding difference and p values.

Fractions First Fractionation (g P Kg-1 DM) Second Fractionation (g P Kg-1 DM) Difference P-value Loosely-P 0.013 ± 0.007 0.017 ± 0.007 0.004 0.053 Fe-P 0.020 ± 0.013 0.021 ± 0.015 0.001 0.292 Al-P 0.115 ± 0.038 0.105 ± 0.038 -0.01 0.284 Ca-P 0.082 ± 0.021 0.068 ± 0.012 -0.014 0.017*

(10)

7

Figure 2. P fractionation scheme proposed by Paludan and Jensen (1995), but excluding the HA-P and organic-P steps. BD is a strong reducing reagent composed of 0.11 M NaHCO3 and 0.11M

Na2S2O4 in a 1:1 ratio. BD-P, NaOH-P and HCl-P corresponds to Fe-P, Al-P and Ca-P respectively.

3.4 Sediment Incubation

Samples from the 0-5 cm depth were used in a subsequent experiment to determine the release rate of P during oxic and anoxic conditions, this was done by incubating the sediment in both anoxic (N2 gas treatment) and oxic (aeration treatment) conditions. Mass of not less

than 27 g was weighed from each wet sediment into two different jars for respective treatment. A known volume of pre-filtered water from the lake was used to fill each jar containing a pre weighed sediment. A 100 µm mesh size net was used for the filtration to prevent the introduction of zooplankton into the setup. All samples were aerated for 5 days 2 g of wet

sediment, centrifuge at 5000 rpm for 20 minutes

50 ml of deionised water and shake for 1 hr. 50 ml of deionized water for washing + 2.0 ml of H2SO4

Residue

50 ml of BD solution and shake for 1 hour + 50 ml of BD for washing+ 50 ml deionised water for washing + 9.4 ml 1 M H2SO4

BD-P

Residue

Loosely-P

50 ml of 0.1 M NaOH and shake for 1 hour + 50 ml of 0.1 M NaOH for washing + 50 ml of deionised water for washing + 6 ml of 1 M H2SO4.

Residue

NaOH-P

50 ml of 0.5 M HCl and shake for 1

(11)

8

after which half of the samples were treated with N2 gas (anoxic treatment) whiles the other

half continued the oxic treatment. Air or nitrogen gas were bubbled through the upper 2 cm of the water column, using a 1 mm inner diameter silicon tubing. Bubbling speed was adjusted to not disturb the sediment. Incubation time began on this day and continued to the 20th day. The experiment was carried out in a dark chamber at room temperature (approx. 20 ºC).

3.5 Sediment P Release and flux

An aliquot of 25 ml water was drawn from all 20 jars in both conditions of treatment and was filtered with a 0.45 µm pore size. This portion was used for the Soluble Reactive Phosphorus (SRP) determination. SRP was determined using the ascorbic acid method as proposed by the Swedish Standards Institute (SIS, 1996). The P concentrations were determined by means of colorimetry using a Sciencetec spectrophotometer at a wavelength of 880nm. The water used for each analysis was replaced immediately with the filtered lake water after samples were taken, thus 2-3 days before the next analysis. Analysis was carried out in concurrent intervals of days. The pH, DO and Turbidity were measured on the last day of incubation. pH and turbidity were measured with portable Satorius PT-15 and HANNA instruments respectively; whiles DO was measured with Yellow Spring Instruments YSI pro 2030 DO probe.

The resultant average sediment P flux for each sediment sample was calculated by plotting the mass of P (µg) in the water at the time of sampling, against day number. All the

concentrations used are from the analysis of SRP. The rate (µg d-1) was divided by the area of the glass jar (m2) to be able to calculate the release per m2.

3.6 Statistical analysis

Data was statistically treated using R studios 0.98.1103. Paired t-test and linear regression were studied. A p value < 0.05 was considered as significant. A linear regression model (lm) was used to test for relationships between the dependent variable (TP and P forms) and the independent variables latitude and longitude. The relationship between the sediment P flux and the different P fractions in the sediment was also investigated using linear regression. This was done by using the lm code in R 0.98.1103. A paired t-test was used to compare the means of P concentrations in the 0-5 cm depth and 5-10 cm depth. Paired T-Test was also used for comparing sediment P flux in both anoxic and oxic treatments. R 0.98.1103 was used for this analysis.

4 Results

4.1 Phosphorus Fractionation BAP

Five main P forms were recorded from the fractionation scheme, these are, Loosely-P, Fe-P, Al-P, Ca-P and Residual-P. The scheme showed majority of the P in sediment being in the form of Residual-P in both depths studied (Figs 3 and 4). Amongst the inorganic P forms, the majority of P was observed to be bound to Al in both depths as well. It was also observed that all P forms were higher in the 0-5 cm depth. An average 14.5%, which translates to 0.15g P kg-1 DM-1, of the TP was estimated to be the BAP in the 0-5 cm depth layer (Fig. 3). The

(12)

9

Figure 3. Proportion of P forms recorded in the fractionation scheme for the 0-5 cm sediment depth in Lake Hällerstadjön. Concentrations of P forms are expressed in g P Kg-1 DM. N = 20.

Figure 4. Proportion of P forms recorded in the fractionation scheme for the 5-10 cm sediment depth in Lake Hällerstadsjön. Concentrations of P forms are expressed in g P Kg-1 DM. N =17.

It was observed that BAP in the inorganic P forms in 0-5 cm and 5-10 cm ranged between 37% -71% approximately and 31-71% respectively (Figs 5 and 6)

0.01 0.02

0.11

0.08

0.79

L-P Fe-P Al-P Ca-P Res-P

0.006 0.010 0.066

0.059

0.701

(13)

10

Figure 5. Relative contribution of inorganic P forms in the 0-5 cm sediment depth in Lake Hällerstadsjön. BAP forms ranged between 37%-71%. N =20.

Figure 6. Relative contribution of inorganic P forms in the 5-10 cm sediment depth in Lake Hällerstadsjön. BAP forms ranged between 31%-71%. N =17.

There was no DRP observed in the Ca-P extracts of sediment from the 0-5 cm layer, but substantial amounts of DRP in the other extracts. The mean concentrations of DRP in the inorganic fractions corresponded to 0.004, 0.001, and 0.002 g P Kg-1 DM in the L-P, Fe-P, and Al-P, respectively.

0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100%

1A 2A 3A 4A 5A 6A 7A 8A 9A 10A 11A 12A 13A 14A 15A 16A 17A 18A 19A 20A

L-P Fe-P Al-P Ca-P

0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100% 1B 2B 4B 5B 6B 7B 10B 11B 12B 13B 14B 15B 16B 17B 18B 19B 20B

(14)

11

Figure 7. Relative contribution of DRP in inorganic P forms in the 0-5 cm depth in Lake Hällerstadsjön. BAP ranged between 5%-75%. N =20.

4.2 Spatial distribution of P

The spatial distribution of P was observed to vary along the latitude transects and as well as depth. Samples within the 0-5 cm were significantly higher in P concentration as compared to those in 5-10 cm (Loosely-P, Fe-P, Al-P, Ca-P and TP had p values of 0.0001,0.0036, 5.9 x10-5, 0.00054 and 0.006 respectively, paired t-test).

There was however differences amongst P forms and TP within the same depths as against the latitudinal transect. There was a significant increase in Ca-P from inlet to outlet, both in the 0-5 cm and the 5-10 cm sediment layers (Fig. 8). By contrast, in both depths the TP decreased. In 5-10 cm depth, other inorganic P forms showed the same trend as TP but it was the opposite in the 0-5 cm depth. In the 5-10cm layer, statistically significant relations with the latitude was also observed for TP and Fe-P.

-40% -20% 0% 20% 40% 60% 80% 100% 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

(15)

12

Figure 8. Spatial distributions of P forms that were significantly correlated with latitude. I. Ca-P in the 0-5 cm sediment depth, II-IV are P forms in the 5-10 cm depth which are TP, Fe-P and Ca-P

respectively. The R2 = 0.41, 0.25, 0.31 and 0.34 for I-IV respectively. P < 0.05. N=20 (in panel I),

N=17 (panel II-IV).

4.3 Sediment P Release and Flux

There was generally no net release in samples incubated under oxic conditions on each of the days, the average sediment P release under this treatment was –1 µg L-1 (Fig. 9). However, some individual sites recorded a release of sediment P, and the values ranged between -7 and +9 µg L-1 for the experimental period. The corresponding sediment P flux was on average of

-6 µg d-1 m-2 and ranged between -61 µg d-1 m-2 and +28µg d-1 m-2.

In the anoxic treatment, an average of 11µg L-1 was recorded for the sediment P release. The values ranged between 0 and 184 µg L-1. The corresponding sediment P flux in the anoxic

treatment was on average 120 µg d-1m-2 and ranged between 15µg d-1 m-2 and 899 µg d-1 m-2. The sediment P released in anoxic conditions showed an increase from day 2 to day 6 but dipped in day 9. The increase in median release levelled off in days 16 and 20. However, in samples from sites 1, 5-7 there was an increasing sediment P release during the entire experiment except in day 9 (Fig 9). A significant correlation was observed between the release of sediment P in anoxic conditions and the BAP; the p-value was 0.01 with an R2 value of 0.311 (Fig 11). The difference in sediment P flux under the two treatment conditions was significant (P = 0.007, Fig 10).

(16)

13

Figure 9. P concentration in oxic (O) and anoxic (A) during the incubation experiment in µg L-1 for

both anoxic and oxic treatments with respect to days. Treat, refers to the treatment used, O and A represents Oxic and Anoxic respectively. In day 20, a huge outlier of 184 µg L-1 was recorded

Figure 10.P flux for anoxic and oxic treatment of sediment from Lake Hällerstadsjön. One site for anoxic condition recorded an out of range value of 899 µg d-1 m-2 and was not included in this plot

Boxes are medians with 25 and 75 percentiles x=means, N=20.

Days: 2 6 9 13 16 20 Treat: O A O A O A O A O A O A

(17)

14

Figure 11. Relationship between sediment P flux and BAP concentration. Simple linear regression. R2

= 0.311, p =0.01*. N = 20. 5. Discussion

5.1 Phosphorus fractionation - BAP

Bioavailable P is the part of the P that is readily available and released during certain environmental conditions, like low redox potential and changes in pH. The forms of P that constitute BAP are the Loosely-P, the Fe-P and the Al-P. Previous studies have shown that the total amount of P extracted in a single step with NaOH is the most available P for algal consumption (Dorich et al., 1984; Zhou et al., 2001). In this study, Ca-P was not bioavailable as there was no mean DRP concentration recorded. BAP was comparatively lower across the lake, it is however worth mentioning that, these BAP stands a chance of being released

during certain environmental conditions and when released, would have dire consequences on the ecosystem (Hesse, 1971). The concentrations recorded for BAP was similar to that of Zhou et al, (2005). The upper range, however was higher than that recorded in this study. This could be due to the excessive use of inorganic fertilizers in China than it is in Sweden. Phosphorus fractionation scheme results after storage in sediments from the 0-5 cm depth recorded differences in the means of P concentrations. The respective inorganic fractions were slightly higher after storage of the sediments, with the exception of Al-P and Ca-P. There was however no statistical significant difference recorded in the means, except that of Ca-P. This was similar to the results in a study by Lukkari et al., (2007). In that study, long storage caused an increase in both Loosely-P and Fe-P, whiles Al-P and Ca-P were reduced. This was observed in sediments which were not shielded by N2 during sampling and storage.

It was noted that the oxic conditions causes a change in redox-sensitive P fractions like the Fe-P. Storage with N2 is therefore a requirement to ensure sediment integrity when engaging

in long term storage. Sediments used in the release experiment for this study did not deviate significantly from the initial concentrations, hence data recorded in the release experiment is trustworthy. -200 0 200 400 600 800 1000 0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 S ed im en t P f lux BAP concentration

(18)

15 5.2 Spatial distribution of P

The general observation from the study showed that the concentration of P decreased with depth. Decomposition of organic matter, microbial activity and reduction of iron oxides play a vital role in influencing the vertical distribution of P, thus the difference in depths (Sundy et al, 1992; Reddy et al., 1996). The variation observed in the spatial distribution of P in both depth was expected for TP as it could be expected that the concentration of particles from surrounding land areas would be higher in the inlet than the outlet part of the lake. A study by Braskerud et al. (2000) showed that there was higher sedimentation at the inlet than the outlet in constructed wetlands and this could also be the situation in Lake Hällerstadsjön. The spatial distribution could be due to both the distribution of inflowing particles and to the biogeochemical cycles occurring in the lake. Furthermore, effluents from the surrounding sites are not only channeled through a common inlet and this may also vary in composition, which could also be a reason for the variation observed. However, the different trend recorded in most of the fractions, could give an idea of how much time it takes for P to be buried within the sediments. The high concentrations of P forms in the 0-5 cm sediment depth could mean that the P pollution is a bit recent.

5.3 Sediment P Release and flux

In the anoxic incubation, there was an exponential increase in P in the water from the start which started leveling off from day 16. However, samples from site 1, 5-7 were turbid and the sediment P release kept on increasing exponentially. The release of P during anaerobic conditions is likely due to growth of anaerobes which tend to mineralize the organic matter and reduce Fe3+ (Miot et al., 2016).

In the oxic treatment, the concentration change was negative in most days, even though this general trend did not apply to some of the sampling sites, as some tended to fluctuate between negative and positive values. Hence, no release of sediment P was generally

recorded contrary to that of the anoxic treatment. Those results were somewhat similar to the observation made by Penn et al. (2000), who attributed the removal of P from the overlying water to the formation of a microlayer on the sediment surface (Penn et al., 2000). However, though the general trend in the oxic treatment was a negative concentration change, it is worth noting that in some cases there were sediment P release which is contrary to traditional knowledge. For example, in the sample from site 1 a net release (positive values) was

observed throughout the study. Such a positive sediment P flux (desorption) is in line with some other studies (Nriagu and Dell 1974; Boström et al. 1982; Hupfer et al. 2004), and was attributed to either an increased saturation of P binding sites in the sediment or microbial degradation of organic material. It has also been postulated that shallow lakes with large surface areas can experience the desorption of P even in oxic conditions. This tends to show the importance of studying P release in oxic conditions especially in shallow lakes.

The results from L Hällerstadsjön, however, indicate that here the depletion of DO causes the major release of P from sediments. The high flux in samples from sites1 and 5-7 could also be a case of a localised phenomenon. It would therefore be interesting if the study is repeated and special attention is paid to those sites and perhaps adjacent areas. Also, such a study could be included into the dredging plan of the county administration. The significant

(19)

16

correlation between BAP and sediment P release also affirms that BAP could be used to predict the release of P in anoxic conditions.

6 Conclusion

Four different fractions were analysed for inorganic P and these were loosely-P, Fe-P, Al-P and Ca-P. The first three represented the biologically available P (BAP), which represented only 14.5% and 10% of the total P in the 0-5 and 5-10 cm layers, respectively. This study showed that in Hällerstadsjön the BAP could be released and this was influenced by anoxic conditions (low redox potential). The variation in the spatial P distribution and TP decreasing along the latitudinal transect (Inlet to Outlet) probably illustrated the distribution of inflowing particles in runoff into the lake. Dredging of the upper sediment layer would reduce the concentrations of BAP and likely reduce the release of P into the water column.

It is therefore recommended that the concentrations of Fe and other metal oxides and redox sensitive elements could be analysed and studied in another work. This might give a clearer picture on the biogeochemistry of the lake

7 Ethical and Social Issues

There are no real serious ethical nor societal issues with regards to this particular study. However, the dredging plan by the county would require some ethical and societal clearance which would border on how sediments would be dredged and transported from the lake. The accompanying noise, smell and so on would be duly considered by the county board

administration.

8 Acknowledgement

Special thanks goes to Anders Hargeby and Karin Tonderski for their immense support and advise during this whole period. I also thank Tove Bjerg for preparing molybdate solutions as well as setting up the labs to enable me work effectively. I would want to express my

warmest appreciation to the Ghana Education Trust Fund (GETFund) for sponsoring my masters education here in Linköping University.

(20)

17 9 References

Abrams, M.M. and Jarrell, W.M. (1995). Soil-phosphorus as a potential non-point source for elevated stream phosphorus levels. J. Environ. Qual. 24: 132-13

Andersen, J. M (1976). An ignition method for determination of total phosphorus in lake sediments. Water Res. 10: 329-331

Berner, R. A., Ruttenberg, K. C., Ingall, E. D and Rao, J.-L. (1993). The nature of

phosphorus burial in modern marine sediments. In R. Wollast, F. T. Mackenzie, and L. Chou [eds.], Interactions of C, N, P and S Biogeochemical Cycles and Global Change. Springer Verlag, p. 365-378.

Boers, P. (1991). The influence on pH on phosphate release from lake sediments. Water Resources. 25 (3):309-311

Boström, B., Jansson, M., and Forsberg, C. (1982). Phosphorus release from lake sediments. Arch. Hydrobiol. Beih. Ergebn. Limnol. 18:5-59.

Boström. B. (1988). Relations between chemistry. microbial biomass and activity in sediments of a polluted vs. a non-polluted eutrophic lake. Verhandlungen International Verein Limnol 23: 451-459.

Braskerud, B.C., Lundekvam, H. and Krogstad, T., (2000). The impact of hydraulic load and aggregation on sedimentation of soil particles in small constructed wetlands. J. Environ. Qual. 29:2013–2020.

Caraco, N. F., Coles, J. J, and Likens, G.E. (1993). Sulfate control of phosphorus availability in lakes. Hydobiol. 253: 275-280

De Groot, C. and Golterman, H.: (1990). Sequential fractionation of sediment phosphate. Hydrobiol 192:143–148.

Dorich, R., Nelson, D. and Sommers, L. (1985). Estimating phosphorus in suspended sediments by chemical extraction. J. Environ. Qual. 14:400–405.

Dorich, R. A., Nelson D. W. and Sommers, L. E. (1984). Availability of phosphorus to algae from eroded soil fraction[J]. Agric. Ecosys. Environ. 11:253-264.

Dorioz, J. M. and Ferhi, A. (1994). Non-point pollution and management of agricultural areas: phosphorus and nitrogen transfer in an agricultural watershed. Wat. Res., 28 (2): 395-410.

Drake, J. C. and Heaney, S. I. (1987). Occurrence of phosphorus and its potential

remobilization in the littoral sediments of a productive English lake. Freshw. Biol. 17:513-523.

Einsele, W. (1938). Über cheμmische und kolloidchemische Vorgänge in Eisen-Phosphat-Systemen unter limnochemischenund limnogeologischen Gesichtspunkten. Arch.

Hydrobiol. Bd. 33:361-387.

Frodge, J. D., Thomas, G. L. and Pauley, G. B. (1991). Sediment phosphorus loading beneath dense canopies of aquatic macrophytes. Lake Reserv. Mgt. 7(1): 61-71

Froelich, P. N., Bender, M. L., Luedtke, N. A., Heath, G. R., and DeVries, T. (1982). The marine phosphorus cycle. Am. J. Sci. 282:474-511.

Hesse, P. R., (1971). A textbook of soil chemical analysis[M]. London: John Murray, 255-300

(21)

18

Hieltjes, A. and Lijklema, L. (1980). Fractionation of inorganic phosphates in calcareous sediments. J. Environ. Qual. 9(3): 405–407

Hingston, F. J., Posner, A. M., and Quirk, J. P. (1972). Anion adsorption by goethite and gibbsite. I. The role of the proton in determining adsorption envelopes. J. Soil Sci. 23: 177-192.

Hingston, F. J., Posner, A. M., and Quirk, J. P. (1974). Anion adsorption by goethite and gibbsite II. Desorption of anions from hydrous oxide surfaces. J. Soil Sci. 25:16-26. Hupfer, M., Rübe, B. and Schmieder, P. (2004). Origin and diagenesis of polyphosphate in lake sediments: A 31P-NMR study. Limnol. Oceanogr. 49(1):1-10.

Jensen, H. S. and Thamdrup, B. (1993). Iron-bound phosphorus in marine sediments as measured by bicarbonate-dithionite extraction. Hydrobiol 253: 47-59.

Jensen, H.S., Kristensen, P., Jeppesen, E. and Skytthe, A. (1992). Iron: Phosphorus ratio in surface sediment as an indicator of phosphate release form aerobic sediments in shallow lakes. Hydobiol. 235/236, 731-743

Kleeberg, A. and Kozenski, H. (1997). Phosphorus release in a lake and its implications for lake restoration. Hydrobiol 342: 9–26.

Klein Jeremy (1997). Sediment Dredging and Macrophyte Harvest as Lake Restoration Techniques. Student On-line journal Vol 2 No 2.

Lijklema, L., Koelmans, A. A., Portielje, R. (1993). Water quality impacts of sediment pollution and the role of early diagenesis. Water Sci. Tech. 28: 1-12.

Lukkari, K., Leivuori, M. and Hartikainen, H. (2007). Fractionation of sediment phosphorus revisited: II. Changes in phosphorus fractions during sampling and storing in the presence or absence of oxygen. Limnol. Oceanogr.: Meth. 5:445-456.

Miot, J., Jézéquel, D., Benzerara, K., Cordier, L., Rivas-Lamelo, S., Skouri-Panet, F., … Duprat, E. (2016). Mineralogical Diversity in Lake Pavin: Connections with Water Column Chemistry and Biomineralization Processes. Minerals. 6(2): 24.

http://doi.org/10.3390/min6020024

Mortimer, C. H. (1941). The exchange of dissolved substances between mud and water in lakes. J. Ecol. 29:280-329.

Mortimer, C. H. (1942). The exchange of dissolved substances between mud and water in lakes. J. Ecol. 30:147-201.

Nriagu, J. O., and Dell, C. I. (1974). Diagenetic formation of iron phosphates in recent lake sediments. Am. Mineral. 59: 934-946.

Nürnberg, G.K. (1984). The prediction of internal phosphorus loads in lake with hypolimnetic anoxia. Limnol. Oceanogr. 29 (1): 111-124

Nürnberg, G.K. (1987). The comparison of internal phosphorus loads in lake with

hypolimnetic anoxia: laboratory incubation in situ hypolimnetic phosphorus accumulation. Limnol. Oceanogr. 32 (5): 1160-1164

Paludan, C. and Jensen, H. S. (1995). Sequential extraction of phosphorus in freshwater wetland and lake sediment: significance of humic acids. Wetlands 15(4):365-373.

(22)

19

Pardo, P. López-Sánchez, J. and Rauret, G. (1998). Characterisation. validation and comparison of three methods for the extraction of phosphate from sediments. Anal. Chim. Acta 376: 183–195.

Penn, M.R., Auer, M. T., Doerr, S. M., Driscoll, C. T., Brooks, C. M. and Effler, S. W. (2000). Seasonality in phosphorus release rates from sediments in hypereutrophic lake under of matrix of pH and redox conditions. Can. J. Fish. Aquat. Sci. 75: 1033-1041

Pettersson, K., Boström, B and Jacobsen, O.-S. (1988). Phosphorus in sediments: speciation and analysis. Hydrobiol 170:91-101.

Psenner, R., Puesko, R. and Sager, M. (1984). Die Fractionierung Organischer and

Anorganischer Phosphorverbindungen von Sedimenten Versuch einer Definition Okologisch Wichtiger Fractionen. Arch. Hydrobiol. 10: 115–155.

Psenner, R., Boström, B. Dinka, M., Petterson, K., Pucsko, R. and Sage, M. (1998). Fractionation of phosphorus in suspended matter and sediment. Arch. Hydrobiol. Beih. 30: 98–110.

Reddy, K. R., Fisherm, M. and Ivanoff, D. (1996). Resuspension and diffusive flux of nitrogen and phosphorus in a hypereutrophic lake. J. Environ Qual. 25: 363, 1996. Ripl, W. (1976). Biochemical Oxidation of Polluted Lake Sediment with Nitrate: A New Lake Restoration Method. Ambio. 5(3):132-135

Ruttenberg, K. C. (1992). Development of a sequential extraction method for different forms of phosphorus in marine sediments. Limnol. Oceanogr. 37:1460-1482

SIS (1997). Svensk standard SS-EN 1189. Swedish Standards Institute, Stockholm. SMHI Water Web. 2012. http://vattenwebb.smhi.se/ (Retrieved 2015-02)

Søndergaard, M. Windolf, J. and Jeppesen, E. (1996). Phosphorus fractions and profiles in the sediment of shallow Danish lakes as related to phosphorus load. sediment composition and lake. Chem. Water Res. 30(4): 992–1002

Søndergaard, M., Erik, J., Torben, L. L., Christian, S., Egbert, H., Van, N., Rudi, R., Eddy, L. and Rob, P. (2007). Lake restoration: successes, failures and long-term effects. J. App. Ecol. 44:1095–1105

Steinberg, C. and Muenster, U. (1985). Geochemistry and ecological role of humic

substances in lakewater. In G. R. Aiken, D. M. McKnight, R. L. Wershaw, and P. MacCarthy [eds.], Humic Substances in Soil, Sediment, and Water: Geochemistry, Isolation, and

Characterization. John Wiley & Sons, p. 105-145

Sundy, B., Cobeil, C., Silverberg, N. and Mucci, A. (1992). The phosphorus cycle in coastal marine sediments. J. Limnol Oceanogr. 37: 1129-1992.

Svendsen, L. M., Rebsdorf, A. and Nørnberg, P. (1993) Comparison of methods for analysis of organic and inorganic phosphorus in river sediment. Water Res 27: 77-83

Svendsen, L. M., Kronvang, B., Kristensen, P., and Grñsbùll, P. (1995). Dynamics of phosphorus compounds in a lowland river system: Importance of retention and non-point sources. Hydrol. Process. 9: 119-142.

Witcheler, J. V., Steven, D., Koenraad, M., Ivan, H., Vanessa, G., Jochen, V. K., Erik, M., Niels, D. P., Maurice, H., Luc, D. M. and Wim, V. (2007). The Importance of drawdown and

(23)

20

Sediment Removal for the Restoration of Eutrophied Shallow Lake, Kraenepoel (Belgium). Hydrobiol 584:291-303.

Xie, L.Q., Xie, P. and Tang, H. J. (2003). Enhancement of dissolved phosphorus release from sediment to lake water by Microcystis blooms-an enclosure experiment in a hyper-eutrophic. subtropical Chinese lake. Environ. Pol. 122: 391-399

Zhou, Q., Gibson, C. and Zhu, Y. (2001). Evaluation of phosphorus bioavailability in sediments of three contrasting lakes in China and the U.K. Chemos. 42: 221–225.

Zhou, A. M, Wang, D. S, Tang, H. X. (2005). Phosphorus fractionation and bio-availability in Taihu Lake (China) sediments. J. Environ. Sci. 17(3): 384–388

(24)

21 10 Appendix

Appendix 1. Characteristics of Sampled Sediment

ID L-P (g P/Kg DM) Fe-P (g P/Kg DM) Al-P (g P/Kg DM) Ca-P (g P/Kg DM) Residual (g P/Kg DM) TP (g P/Kg DM) Organic matter (%) 1A 0.015 0.041 0.135 0.071 0.671 0.933 17.01 1B 0.005 0.017 0.066 0.044 0.853 0.985 15.67 2A 0.012 0.016 0.130 0.089 0.894 1.141 16.64 2B 0.008 0.013 0.094 0.042 0.921 1.079 18.25 3 0.016 0.012 0.106 0.059 0.694 0.887 12.29 4A 0.005 0.014 0.082 0.048 1.009 1.157 16.64 4B 0.005 0.014 0.058 0.036 0.763 0.875 17.36 5A 0.008 0.024 0.078 0.055 1.007 1.172 17.14 5B 0.004 0.012 0.061 0.033 0.921 1.032 17.17 6A 0.021 0.021 0.127 0.117 1.067 1.353 17.95 6B 0.008 0.020 0.090 0.039 0.984 1.141 16.50 7A 0.015 0.013 0.104 0.045 0.870 1.047 15.98 7B 0.005 0.010 0.078 0.056 0.757 0.907 13.07 8 0.003 0.009 0.074 0.064 0.678 0.828 13.09 9 0.006 0.024 0.097 0.085 0.799 1.011 14.20 10A 0.006 0.008 0.092 0.066 0.720 0.891 14.53 10B 0.005 0.005 0.061 0.052 0.847 0.969 13.41 11A 0.013 0.060 0.194 0.102 1.188 1.555 17.11 11B 0.007 0.006 0.066 0.052 0.807 0.938 15.25 12A 0.005 0.009 0.042 0.096 0.735 0.887 9.48 12B 0.004 0.006 0.043 0.094 0.338 0.485 8.25 13A 0.004 0.013 0.101 0.112 0.520 0.750 13.29 13B 0.003 0.004 0.038 0.059 0.709 0.813 12.87 14A 0.018 0.014 0.118 0.105 0.744 1.000 16.70 14B 0.002 0.006 0.063 0.085 0.719 0.875 13.55 15A 0.021 0.022 0.105 0.068 0.748 0.964 18.03 15B 0.010 0.018 0.102 0.070 0.675 0.875 15.34 16A 0.021 0.012 0.170 0.092 0.361 0.657 17.09 16B 0.004 0.008 0.064 0.044 0.724 0.844 14.97 17A 0.024 0.024 0.184 0.084 0.826 1.141 17.42 17B 0.006 0.007 0.063 0.061 0.160 0.297 13.70 18A 0.020 0.041 0.137 0.094 0.786 1.079 16.60 18B 0.004 0.008 0.082 0.068 0.647 0.809 12.74 19A 0.011 0.011 0.142 0.099 0.904 1.167 13.17 19B 0.011 0.005 0.035 0.066 0.744 0.860 10.61 20A 0.019 0.020 0.081 0.084 0.578 0.782 12.43 20B 0.006 0.007 0.055 0.096 0.352 0.516 15.25

(25)

22

Appendix 2. Laboratory derived pH, Turbidity and DO on last day of incubation. CN and CO are the controls for anoxic and oxic conditions respectively.

ID pH DO(mg/l) Turb. (FTU)

1O 7.72 9.40 14.78 2O 7.74 9.56 14.53 3O 7.86 9.94 8.84 4O 7.54 9.20 9.31 5O 7.95 10.04 14.86 6O 7.96 10.42 18.60 7O 7.92 10.28 20.00 8O 7.86 10.30 10.71 9O 7.93 10.14 24.94 10O 7.82 10.50 11.88 11O 7.98 10.70 46.19 12O 7.95 10.54 9.76 13O 7.90 10.23 15.27 14O 7.98 10.30 27.67 15O 7.94 10.00 40.60 16O 8.01 10.12 27.57 17O 8.02 10.25 81.79 18O 7.93 10.31 31.81 19O 7.88 10.50 17.01 20O 7.83 10.48 11.90 1N 8.50 0.29 6.45 2N 8.78 0.39 4.30 3N 8.18 0.35 7.97 4N 7.88 0.23 16.50 5N 8.16 0.10 62.0 6N 8.61 0.23 7.94 7N 8.47 0.25 11.26 8N 8.09 0.17 11.71 9N 8.50 0.28 1.86 10N 8.42 0.19 2.19 11N 8.96 0.18 1.67 12N 8.69 0.25 1.45 13N 8.10 0.36 2.72 14N 8.72 0.21 2.30 15N 8.86 0.24 0.69 16N 8.88 0.29 1.81 17N 8.70 0.40 2.05 18N 8.59 0.39 1.25 19N 8.75 0.30 2.61

(26)

23

20N 8.27 0.27 1.43

CO 8.08 10.55 6.81

CN 2.58 0.54 1.72

Appendix 3. North, East coordinates for each Sampling site

SITE NORTH EAST

1 6479580 1519982 2 6479506 1519614 3 6479270 1519406 4 6479059 1519401 5 6479661 1519705 6 6479691 1519725 7 6479696 1519695 8 6479821 1519587 9 6479799 1520032 10 6480251 1519890 11 6480840 1519858 12 6481116 1519844 13 6481052 1519622 14 6480888 1519722 15 6480921 1519853 16 6480836 1519747 17 6480657 1519725 18 6480265 1520142 19 6480278 1520247 20 6480375 1520364

(27)

24

Appendix 4. Soluble Reactive Phosphorus concentrations (g P/Kg DM) in extracts from the 0-5 cm sediment depth

ID L-P Fe-P Al-P Ca-P 1 0.00643 0.008145 0.031794 -0.00271 2 0.007439 0.008526 0.021985 -0.00281 3 0.004491 0.004528 0.017096 -0.00226 4 0.003047 0.008711 0.013876 -0.00229 5 0.00563 0.011739 0.01961 -0.00204 6 0.008717 0.013342 0.034148 0.000485 7 0.003101 0.004691 0.013612 -6.7E-05 8 0.003214 0.005415 0.013027 -0.00042 9 0.002502 0.00753 0.015858 -0.00154 10 0.001625 0.001896 0.010296 -0.00038 11 0.006972 0.03124 0.025084 -0.00012 12 0.005285 0.007006 0.00969 -6.9E-05 13 0.003304 0.004978 0.011923 -0.00048 14 0.005394 0.012769 0.018663 -0.00057 15 0.002482 0.008152 0.011275 -0.00088 16 0.004656 0.006244 0.021224 -0.0006 17 0.003285 0.011442 0.016364 -0.001 18 0.004816 0.005817 0.009964 -0.00103 19 0.000862 0.003739 0.012044 -0.00089 20 0.003801 0.012287 0.010931 -0.01238

(28)

References

Related documents

The increasing availability of data and attention to services has increased the understanding of the contribution of services to innovation and productivity in

Generella styrmedel kan ha varit mindre verksamma än man har trott De generella styrmedlen, till skillnad från de specifika styrmedlen, har kommit att användas i större

Parallellmarknader innebär dock inte en drivkraft för en grön omställning Ökad andel direktförsäljning räddar många lokala producenter och kan tyckas utgöra en drivkraft

Närmare 90 procent av de statliga medlen (intäkter och utgifter) för näringslivets klimatomställning går till generella styrmedel, det vill säga styrmedel som påverkar

I dag uppgår denna del av befolkningen till knappt 4 200 personer och år 2030 beräknas det finnas drygt 4 800 personer i Gällivare kommun som är 65 år eller äldre i

Den förbättrade tillgängligheten berör framför allt boende i områden med en mycket hög eller hög tillgänglighet till tätorter, men även antalet personer med längre än

På många små orter i gles- och landsbygder, där varken några nya apotek eller försälj- ningsställen för receptfria läkemedel har tillkommit, är nätet av

Detta projekt utvecklar policymixen för strategin Smart industri (Näringsdepartementet, 2016a). En av anledningarna till en stark avgränsning är att analysen bygger på djupa