ArticlePDF Available

Abstract and Figures

Submerged macrophytes play a key role in north temperate shallow lakes by stabilizing clear-water conditions. Eutrophication has resulted in macrophyte loss and shifts to turbid conditions in many lakes. Considerable efforts have been devoted to shallow lake restoration in many countries, but long-term success depends on a stable recovery of submerged macrophytes. However, recovery patterns vary widely and remain to be fully understood. We hypothesize that reduced external nutrient loading leads to an intermediate recovery state with clear spring and turbid summer conditions similar to the pattern described for eutrophication. In contrast, lake internal restoration measures can result in transient clear-water conditions both in spring and summer and reversals to turbid conditions. Furthermore, we hypothesize that these contrasting restoration measures result in different macrophyte species composition, with added implications for seasonal dynamics due to differences in plant traits. To test these hypotheses, we analyzed data on water quality and submerged macrophytes from 49 north temperate shallow lakes that were in a turbid state and subjected to restoration measures. To study the dynamics of macrophytes during nutrient load reduction, we adapted the ecosystem model PCLake. Our survey and model simulations revealed the existence of an intermediate recovery state upon reduced external nutrient loading, characterized by spring clear-water phases and turbid summers, whereas internal lake restoration measures often resulted in clear-water conditions in spring and summer with returns to turbid conditions after some years. External and internal lake restoration measures resulted in different macrophyte communities. The intermediate recovery state following reduced nutrient loading is characterized by a few macrophyte species (mainly pondweeds) that can resist wave action allowing survival in shallow areas, germinate early in spring, have energy-rich vegetative propagules facilitating rapid initial growth and that can complete their life cycle by early summer. Later in the growing season these plants are, according to our simulations, outcompeted by periphyton, leading to late-summer phytoplankton blooms. Internal lake restoration measures often coincide with a rapid but transient colonization by hornworts, waterweeds or charophytes. Stable clear-water conditions and a diverse macrophyte flora only occurred decades after external nutrient load reduction or when measures were combined.
Content may be subject to copyright.
ORIGINAL RESEARCH
published: 19 February 2018
doi: 10.3389/fpls.2018.00194
Frontiers in Plant Science | www.frontiersin.org 1February 2018 | Volume 9 | Article 194
Edited by:
Janne Alahuhta,
University of Oulu, Finland
Reviewed by:
Rebecca Lester,
Deakin University, Australia
Ludwig Triest,
Vrije Universiteit Brussel, Belgium
*Correspondence:
Sabine Hilt
hilt@igb-berlin.de
Specialty section:
This article was submitted to
Functional Plant Ecology,
a section of the journal
Frontiers in Plant Science
Received: 28 September 2017
Accepted: 01 February 2018
Published: 19 February 2018
Citation:
Hilt S, Alirangues Nuñez MM,
Bakker ES, Blindow I, Davidson TA,
Gillefalk M, Hansson L-A, Janse JH,
Janssen ABG, Jeppesen E, Kabus T,
Kelly A, Köhler J, Lauridsen TL,
Mooij WM, Noordhuis R, Phillips G,
Rücker J, Schuster H-H,
Søndergaard M, Teurlincx S,
van de Weyer K, van Donk E,
Waterstraat A, Willby N and Sayer CD
(2018) Response of Submerged
Macrophyte Communities to External
and Internal Restoration Measures in
North Temperate Shallow Lakes.
Front. Plant Sci. 9:194.
doi: 10.3389/fpls.2018.00194
Response of Submerged Macrophyte
Communities to External and Internal
Restoration Measures in North
Temperate Shallow Lakes
Sabine Hilt 1
*, Marta M. Alirangues Nuñez 1, Elisabeth S. Bakker 2, Irmgard Blindow 3,
Thomas A. Davidson 4, Mikael Gillefalk 1, Lars-Anders Hansson 5, Jan H. Janse 2,6 ,
Annette B. G. Janssen 2,7 , Erik Jeppesen 4,8 , Timm Kabus 9, Andrea Kelly 10, Jan Köhler 1,
Torben L. Lauridsen 4,8, Wolf M. Mooij 2, 11, Ruurd Noordhuis 12 , Geoff Phillips 13 ,
Jacqueline Rücker 14, Hans-Heinrich Schuster 15 , Martin Søndergaard 4,8 , Sven Teurlincx2,
Klaus van de Weyer 16, Ellen van Donk 2, Arno Waterstraat 17, Nigel Willby 13 and
Carl D. Sayer 18
1Department of Ecosystem Research, Leibniz-Institute of Freshwater Ecology and Inland Fisheries, Berlin, Germany,
2Departmnet of Aquatic Ecology, Netherlands Institute of Ecology (NIOO-KNAW), Wageningen, Netherlands, 3Biological
Station of Hiddensee, University of Greifswald, Greifswald, Germany, 4Department of Bioscience, Aarhus University,
Silkeborg, Denmark, 5Department of Biology, Lund University, Lund, Sweden, 6Netherlands Environmental Assessment
Agency (PBL), Den Haag, Netherlands, 7Water Systems and Global Change Group, Wageningen University and Research,
Wageningen, Netherlands, 8Sino-Danish Centre for Education and Research, University of Chinese Academy of Sciences,
Beijing, China, 9Institute of Applied Freshwater Ecology, Seddiner See, Germany, 10 Broads Authority, Norwich,
United Kingdom, 11 Department of Aquatic Ecology and Water Quality Management, Wageningen University and Research,
Wageningen, Netherlands, 12 Deltares, Delft, Netherlands, 13 Biological and Environmental Sciences, University of Stirling,
Stirling, United Kingdom, 14 Department of Freshwater Conservation, Brandenburg University of Technology
Cottbus-Senftenberg, Senftenberg, Germany, 15 Niedersächsischer Landesbetrieb für Wasserwirtschaft, Küsten- und
Naturschutz, Sulingen, Germany, 16 Lanaplan, Nettetal, Germany, 17 Gesellschaft für Naturschutz und Landschaftsökologie,
Kratzeburg, Germany, 18 Department of Geography, Environmental Change Research Centre, University College London,
London, United Kingdom
Submerged macrophytes play a key role in north temperate shallow lakes by stabilizing
clear-water conditions. Eutrophication has resulted in macrophyte loss and shifts to
turbid conditions in many lakes. Considerable efforts have been devoted to shallow lake
restoration in many countries, but long-term success depends on a stable recovery
of submerged macrophytes. However, recovery patterns vary widely and remain to
be fully understood. We hypothesize that reduced external nutrient loading leads
to an intermediate recovery state with clear spring and turbid summer conditions
similar to the pattern described for eutrophication. In contrast, lake internal restoration
measures can result in transient clear-water conditions both in spring and summer
and reversals to turbid conditions. Furthermore, we hypothesize that these contrasting
restoration measures result in different macrophyte species composition, with added
implications for seasonal dynamics due to differences in plant traits. To test these
hypotheses, we analyzed data on water quality and submerged macrophytes from 49
north temperate shallow lakes that were in a turbid state and subjected to restoration
measures. To study the dynamics of macrophytes during nutrient load reduction, we
adapted the ecosystem model PCLake. Our survey and model simulations revealed
the existence of an intermediate recovery state upon reduced external nutrient loading,
characterized by spring clear-water phases and turbid summers, whereas internal lake
Hilt et al. Response of Macrophytes to Restoration
restoration measures often resulted in clear-water conditions in spring and summer
with returns to turbid conditions after some years. External and internal lake restoration
measures resulted in different macrophyte communities. The intermediate recovery state
following reduced nutrient loading is characterized by a few macrophyte species (mainly
pondweeds) that can resist wave action allowing survival in shallow areas, germinate early
in spring, have energy-rich vegetative propagules facilitating rapid initial growth and that
can complete their life cycle by early summer. Later in the growing season these plants
are, according to our simulations, outcompeted by periphyton, leading to late-summer
phytoplankton blooms. Internal lake restoration measures often coincide with a rapid
but transient colonization by hornworts, waterweeds or charophytes. Stable clear-water
conditions and a diverse macrophyte flora only occurred decades after external nutrient
load reduction or when measures were combined.
Keywords: aquatic plants, biomanipulation, eutrophication, lake restoration. nutrient load reduction, PCLake, plant
traits, regime shift
INTRODUCTION
Shallow lakes are the most abundant freshwater ecosystems on
earth (Verpoorter et al., 2014). In their pristine state, they are
often characterized by abundant submerged vegetation which can
stabilize clear-water conditions (Scheffer et al., 1993) and plays a
key role in the functioning of the ecosystem (Hilt et al., 2017).
Several mechanisms contribute to a positive feedback between
macrophytes and clear water conditions. As a consequence,
shallow lakes are resistant to increasing nutrient loading up
to a critical threshold, above which their macrophytes collapse
and the lakes shift into a turbid, phytoplankton-dominated state
(Scheffer et al., 1993). In recent centuries, excessive nutrient
loading has resulted in a loss of macrophytes and shift to this
turbid state in many temperate shallow lakes (e.g., Körner, 2002;
Phillips et al., 2016).
Sayer et al. (2010a) suggested a typical pattern of lake
macrophyte loss, defining a so-called “crashing” state lying
between the stable clear-water state featuring a diverse plant
community and the final turbid state lacking in macrophytes.
This crashing state is characterized by the occurrence of only a
few macrophyte species that can complete their life cycle during
clear-water conditions in spring and early summer while later
in summer, cyanobacteria blooms often occur. Eventually, the
remaining macrophyte stands are also lost and give way to year-
round phytoplankton dominance (Sayer et al., 2010a,b, 2016).
Under these conditions, several ecosystem functions and services
deteriorate, including biodiversity support, nutrient retention,
provision of water of drinking or swimming quality (Hilt et al.,
2017).
Hence, considerable efforts and financial resources have been
devoted to the restoration of shallow lakes in many countries
in recent decades (Jeppesen et al., 2005). The success of lake
restoration in the long-term depends critically on the stable
recovery of submerged macrophytes (Hilt et al., 2006). However,
the turbid state is stabilized by feedback mechanisms that can
prevent macrophyte re-colonization even at reduced nutrient
loading. In theory, only the reduction of nutrient levels below a
critical threshold or a significant reduction in the abundance of
planktivorous and benthivorous fish (e.g., by biomanipulation or
natural fish kills) will lead to a recovery of clear-water conditions
and a return of macrophytes (Scheffer et al., 1993). In practice,
reductions in the external nutrient load to shallow lakes often fail
to deliver macrophyte recovery (Jeppesen et al., 2005). Similarly,
biomanipulation of the fish community in turbid shallow lakes
has produced variable effects on macrophytes in shallow lakes
(Hansson et al., 1998; Bergman et al., 1999; Søndergaard et al.,
2008; Jeppesen et al., 2012; Bernes et al., 2015; Sayer et al., 2016).
Overall, the response of macrophyte communities to different
types of lake restoration measures remains to be fully understood
(Jeppesen et al., 2005; Bakker et al., 2013).
We hypothesize that (1) external lake restoration measures
leading to nutrient load reduction in turbid temperate shallow
lakes result in macrophyte re-establishment in a reversed
sequence to the one described by Sayer et al. (2010a,b) for
advancing eutrophication. An intermediate recovery state should
occur where the water is clear in spring but dominated by
phytoplankton and thus turbid in late summer, until, eventually,
seasonally stable conditions characterized by high water clarity in
both spring and summer would dominate (Figure 1). In contrast,
lake internal measures such as biomanipulation or phosphorus
precipitation are expected to result in transient clear-water
conditions in spring and summer if either zooplankton is
sufficiently released from fish predation or internal phosphorus
loading from sediments is reduced enough to control summer
phytoplankton. Such conditions are supposed to occur only
temporarily in the absence of additional external nutrient load
reduction (Figure 1). We hypothesize that (2) these contrasting
types of restoration measures result in different macrophyte
community composition and seasonal patterns in plant
abundance. Specific macrophyte communities with short growth
seasons should dominate during the intermediate recovery
state following external nutrient loading reduction, while
species with longer growing season requirements are predicted
to temporarily establish following lake-internal measures
(Figure 1). The establishment of stable clear conditions with a
Frontiers in Plant Science | www.frontiersin.org 2February 2018 | Volume 9 | Article 194
Hilt et al. Response of Macrophytes to Restoration
FIGURE 1 | Response patterns of turbid north temperate shallow lakes to different restoration measures: (1) External restoration measures (reduction of external
nutrient loading) are expected to lead to an intermediate recovery state with clear-water conditions in spring and turbid water in summer and specific macrophyte
communities with short growth seasons and eventually stable clear conditions with a diverse macrophyte flora if nutrient loading is reduced sufficiently or additional
internal measures are applied (reversed order as suggested for eutrophication by Sayer et al., 2010a,b). Thresholds in phosphorus (P) loading are based on
simulations using PCLake (see Figure 5). (2) Lake-internal measures (biomanipulation, sediment suction dredging) leading to unstable clear-water conditions with
specific macrophyte communities that may collapse resulting in a shift back to turbid conditions unless nutrient loading is reduced, or (3) a combination of external
and internal restoration leading to stable clear-water conditions with an abundant and diverse macrophyte community.
diverse macrophyte community is thus assumed to require both,
external and internal measures (Figure 1).
To test these hypotheses, we analyse existing data on the water
quality and submerged macrophytes of 49 temperate shallow
lakes that had deteriorated to a turbid state and subsequently
were subject to either external or internal restoration measures
or both. In addition, we use an adapted version of the ecosystem
model PCLake (Janse et al., 2008) to simulate the response of
water clarity and macrophyte biomass to external nutrient load
reduction and to detect any thresholds in nutrient loading for
macrophyte recovery. Traits of the typical macrophyte species
found after external and internal lake restoration measures
are compared to provide a mechanistic understanding of the
observed re-colonization patterns.
MATERIALS AND METHODS
Literature and Data Search on Macrophyte
Species Recovery
We started our literature review with the 22 shallow lakes
described in detail in the study by Jeppesen et al. (2005) on the
response of lakes to reduced external nutrient loading. However,
only two of these 22 lakes were turbid before the nutrient
load reduction (Müggelsee, Veluwemeer) and both showed an
increase in macrophyte coverage following the intervention. The
rest showed no change, macrophyte declines or lacked suitable
data (Jeppesen et al., 2005). Therefore, we searched for more
examples in published and unpublished studies on lakes in
Germany, The Netherlands, Denmark, southern Sweden and
UK where shallow lakes are abundant and experience similar
eutrophication problems and climatic conditions. We selected
lakes that had lost most of their submerged macrophytes during
a turbid phase and subsequently had been subjected to either
external restoration via nutrient load reduction (summarized
in Table 1), or internal restoration using biomanipulation or
sediment dredging (Table 2). We did not carry out a full
systematic review of available data, but instead focussed on
known lakes within the research network of the authors where
at least partial recovery through restoration measures was
evident.
For all lakes, we retrieved information from the turbid
period and its macrophyte assemblage, macrophyte composition
after external/internal restoration, total phosphorus (TP)
concentrations in the water and Secchi depth in spring (April-
June) and summer (July–September) using published studies or
questionnaire responses provided by co-authors. As is commonly
the case, data on TP concentrations and Secchi depth were very
diverse, ranging from multi-year weekly measurements to
single values. The data were merged into a single value for
each season and lake by using means (Figures 2A,B) and
raw data are available as a supplement. We also analyzed
the occurrence of dominant macrophyte species during the
recovery period. To visualize potentially typical recovery
patterns, the long-term changes in nutrient concentrations,
water transparency and macrophyte occurrence are shown
in more detail for three lakes restored only by reductions in
external nutrient loading (Müggelsee, Veluwemeer, Eemmeer,
for details see no. 1, 19 and 20 in Table 1) and for three lakes
Frontiers in Plant Science | www.frontiersin.org 3February 2018 | Volume 9 | Article 194
Hilt et al. Response of Macrophytes to Restoration
TABLE 1 | The response of north temperate shallow lakes in Germany (DE), United Kingdom (UK), The Netherlands (NL), and Denmark (DK) to external nutrient load reduction (x: data on Secchi disk transparency and
total phosphorus concentrations were available and used in Figure 2A).
No. Lake Country Size
(ha)
Depth
(max/mean)
(m)
Measures of nutrient load
reduction
Period of turbid
conditions, major
remaining macrophyte
species and coverage (%
lake area)
Period of intermediate recovery
conditions, major macrophyte
species and coverage (% lake
area)
Period of stable
clear conditions
and macrophyte
species
Data
Figure 2A
References
1 Großer
Müggelsee
DE 750 8/4.9 Since 1989: improved
wastewater treatment in
catchment
1970–1989
Sparse stands of
P. pectinatus
1990–2013
P. pectinatus, P. perfoliatus,
P. crispus, Z. palustris
ca. 5% (Figure 4)
2014-now
P. pectinatus,
P. perfoliatus,
Najas marina,
E. nuttallii
ca. 25% (Figure 4)
xHilt et al., 2013, S. Hilt
unpubl. data
2 Großer
Wannsee
282 9.8/5.5 Nutrient load reduction in
catchment
?-ongoing
P. pectinatus, P. crispus
<5%
not yet reached Hilt and Grünert, 2008;
Van de Weyer, 2011
3 Galen-becker
See
590 1.8/0.8 Since 1972: reduction of P
loading by treatment in
upstream reservoir, since
2007 by regulation of whole
water supply to lake
1995–2002 2003–2006
P. pectinatus
2008–now
Chara contraria,
C. tomentosa,
C. globularis,
C. intermedia,
Nitellopsis obtusa,
N. marina f.
intermedia,
P. pectinatus,
Z. palustris
xWaterstraat, 2008,
unpubl. data
4 Dümmer 1350 4/1.1 Reduction of P load by
wastewater treatment plants
by 95%
1960–2011
P. pectinatus (rare)
2011–ongoing
P. pectinatus, P. crispus, Z. palustris,
P. pusillus
25-50%
not yet reached x Blüml et al., 2008;
Schuster, unpubl. data
5 Schwielow-
see
786 9.1/2.8 Nutrient load reduction in
catchment area (River Havel)
?2005 2006–now
P. pectinatus
not yet reached x Kabus et al., 2007
6 Wuster-witzer
See
172 9.2/3.4 Nutrient load reduction in
catchment area
?2005 2006–ongoing
P. pectinatus
not yet reached x Kabus et al., 2007
7 Grimnitz-see 777 10.3/4.5 Since 1994: sewage
treatment plant in operation
1970–1990
P. pectinatus
1991–?
P. pectinatus
?–now
2008: C. contraria,
N. obtusa,
P. pectinatus,
P. perfoliatus,
C. demersum, L.
trisulca
2001: C. contraria,
C. globularis,
N. obtusa,
P. pectinatus
xMauersberger and
Mauersberger, 1996;
Gervais et al., 1999;
Kabus and
Mauersberger, 2011
(Continued)
Frontiers in Plant Science | www.frontiersin.org 4February 2018 | Volume 9 | Article 194
Hilt et al. Response of Macrophytes to Restoration
TABLE 1 | Continued
No. Lake Country Size
(ha)
Depth
(max/mean)
(m)
Measures of nutrient load
reduction
Period of turbid
conditions, major
remaining macrophyte
species and coverage (%
lake area)
Period of intermediate recovery
conditions, major macrophyte
species and coverage (% lake
area)
Period of stable
clear conditions
and macrophyte
species
Data
Figure 2A
References
8 Wardersee 357 10.8/3.7 Wastewater treatment
plants
? 1996–2006
P. pectinatus, P. perfoliatus,
P. crispus, Z. palustris
? x Landesamt für Natur
und Umwelt des
Landes
Schleswig-Holstein,
1997; Heinzel and
Martin, 2006
9 Hemmels-
dorfer
See
450 6 (northern
part)
1998: P load reduction ?-1978-? ?-2006-?
P. pectinatus, Z.palustris,
P. perfoliatus, R. circinatus
?Heinzel and Martin,
2006
10 Großer
Varchen-tiner
See
182 1.7/? Lowered nutrient input from
agricultural catchment since
1990
?–? ?-2012-?
P. pectinatus,
M. spicatum
90%
2016
M. spicatum,
Najas marina ssp.
intermedia
x Kabus unpubl. data
11 Großer
Dambecker
See
94 2.1/0.8 Lowered nutrient input from
agricultural catchment since
1990
?- 2007
few P. pectinatus
2010-ongoing
P. pectinatus,
C. globularis
70%
Not yet reached x Kabus unpubl. data
12 Langer See 130 3.8/2.2 Since 1990: catchment
restoration, improved
wastewater treatment
?-1997-2001-? ?-2011-2015-ongoing
M. spicatum, C. demersum,
N. marina
not yet reached x Rücker et al., 2015,
unpubl. data
13 Nonnensee 76 2.2/? Re-flooded area (mid
1990’s)
?- 2012
very few
P. pusillus
2012, 2016
P. pectinatus,
M. spicatum, Ceratophyllum spp.
not yet reached Kabus unpubl. data
14 De Wittsee 24 2.1/1.4 Improvement of wastewater
treatment
1970-2009
Nuphar lutea
2009-ongoing
E. nuttallii, Lemna
not yet reached x Van de Weyer, unpubl.
data
15 Steinhuder
Meer
3000 2.9/1.35 Improvement of wastewater
treatment
1960-1998 none 1999-2001:
2002-03: E. nuttallii
2003-08: P. perfoliatus, P. crispus
2009-?: shift back to turbid
not yet reached Hussner et al., 2014
16 Felbrigg Lake UK 2.7 1.3/0.9 Creation of pre-lake wetland
in 2012 resulting in N-limited
conditions. Cormorant
predation on rudd
1960-2013
P. pectinatus, P. pusillus,
P. crispus, Z. palustris
2014-ongoing
C. demersum, Chara spp.
P. pectinatus, P. crispus, P. pusillus
unclear whether
already reached
Sayer et al., 2010a,b;
Sayer et al. unpubl.
data
17 Barton Broad 75 2/1 Progressive increase in
number of effluents with P
removal
1974-1990
none
1990-2000
C. demersum, P. crispus
<1%
reached after
sediment removal
1996 (Table 2, no.
41)
xPhillips et al., 2005,
2015
(Continued)
Frontiers in Plant Science | www.frontiersin.org 5February 2018 | Volume 9 | Article 194
Hilt et al. Response of Macrophytes to Restoration
TABLE 1 | Continued
No. Lake Country Size
(ha)
Depth
(max/mean)
(m)
Measures of nutrient load
reduction
Period of turbid
conditions, major
remaining macrophyte
species and coverage (%
lake area)
Period of intermediate recovery
conditions, major macrophyte
species and coverage (% lake
area)
Period of stable
clear conditions
and macrophyte
species
Data
Figure 2A
References
18 Wolderwijd NL 2650 2.5/1.5 1982-89: Flushing with
nutrient-poor water
1969-1975
P. pectinatus, P. perfoliatus
(Figure 5)
1976-1995
P. pectinatus,
P. perfoliatus, P. pusillus (Figure 5)
reached after
biomanipulation
carried out since
1990 (Table 2, no.
25, Figure 5)
xScheffer et al., 1992;
Noordhuis et al., 2016
19 Veluwe-meer 3400 5/1.55 1982-89: Flushing with
nutrient-poor water
1975-76
P. pectinatus
0-5%
(Figure 4)
1977-1995
P. pectinatus, P. perfoliatus,
P. pusillus, Characeae
10-15% (Figure 4)
1996-now
P. pectinatus,
C. aspera,
C. contraria,
N. obtusa
30-85% (Figure 4)
xScheffer et al., 1992;
Van den Berg et al.,
1998, 1999; Noordhuis
et al., 2016
20 Eemmeer 1520 ?/2.1 Since 1995: improved
sewage treatment and
closure of treatment plant
1970-1999 P. pectinatus
1–5% (Figure 4)
2000-ongoing P. crispus,
P. pectinatus, P. pusillus 10-40%
(Figure 4)
not yet reached,
but first Characeae
visible since 2010
(Figure 4)
xNoordhuis et al., 2016
21 Arresø DK 3987 5.6/3.1 Improved sewage
treatment, artificial lakes on
the main inlet stream,
reduced
catchment.fertilization
1989-1996
none
?- 2011: C. globularis, C. vulgaris,
M. spicatum, P. perfoliatus,
P. berchtoldii, P. cripus, P. pectinatus
unclear whether
already reached
xJeppesen et al.,
2007a,b Søndergaard,
unpubl. data
Frontiers in Plant Science | www.frontiersin.org 6February 2018 | Volume 9 | Article 194
Hilt et al. Response of Macrophytes to Restoration
TABLE 2 | The response of north temperate shallow lakes in The Netherlands (NL), Sweden (SE), Denmark (DK), United Kingdom (UK) ,and Germany (DE) to biomanipulation, natural fish kills or other lake-internal
measures (x: data on Secchi disk transparency and total phosphorus concentrations were available and used in Figure 2B).
No. Lake Country Size
(ha)
Depth
(max/mean)
(m)
Measures applied Period of turbid
conditions and major
remaining macrophyte
species
Periods of clear conditions,
macrophyte species and
coverage (% lake area)
Period of return
to turbid
conditions
Data
Figure 2B
References
22 Duiniger-meer NL 30 ?/1 1992,1993,1994: Fish
removal
?-1992 1992-ongoing
Chara globularis, C. vulgaris
Nitellopsis obtusa (2000-04)
Varying
macrophyte cover,
but no shift back
to turbid
xVan Berkum et al.,
1995; Meijer et al.,
1999; Riegman, 2007;
Verhofstad et al., 2017;
Van Donk unpubl. data
23 Ijzeren Man 11 2.3/2.3 1989: Complete removal of
fish biomass by pumping
dry, restocking with pike
fingerlings, roach, rudd, ide
and tench, sediments
removed
1960s1989 Within 2 months of fish removal,
macrophytes covered 50% of lake,
Characeae
1995-?,
varying
macrophyte cover
xMeijer et al., 1999; Van
Donk unpubl. data
24 Noorddiep 4.5 ?/1.5 1988: Biomanipulation ?-1989 1989-?
E. canadensis,
C. demersum
clear for at least 8
years despite TP
250 µg L1
xMeijer et al., 1999; Van
Donk unpubl. data
25 Wolderwijd 2650 5/1.5 1990: Biomanipulation See Table 1 (lake no. 18) 1992-
C. contraria, C. vulgaris
Figure 5
none x Meijer and Hosper,
1997; Noordhuis et al.,
2016
26 Zwemlust 1.5 2.5/1.5 1987: Lake drained empty,
fish completely removed,
restocked with pike and
rudd
1999: Temporary lowering
of water level, fish removal
?-1987 1988–1996
1988–1989: E. nuttallii,
1990–1991: C. demersum
1992–1994: P. berchtoldii
1995–1996: E. Nuttallii
1997–1999 x Van de Bund and Van
Donk, 2002;
Verhofstad et al., 2017
27 Terra Nova 85 ?/1.4 2003: (Removal of
planktivorous and
benthivorous fish)
1987–2003
1994: sparse stands of
C. demersum, P. lucens,
P. obtusifolius, P. pectinatus,
M. spicatum
2004-?
C. demersum, E. nuttallii,
P. obtusifolius, Nitella mucronata,
N. marina, Utricularia vulgaris
Van de Haterd and Ter
Heerdt, 2007
28 Galgje 3.1 ?/1.1 1987: Removal of all
planktivorous and 85% of
benthivorous fish in 1987
?-1988 1988: Within 2 months of fish removal
macrophyte covered lake, Characeae
2014: C. demersum, L. minor,
P. crispus, S. polyrhiza
Meijer et al., 1999;
Verhofstad et al., 2017
29 Loender-
veense
Plas
270 ?/2.7 2004/05: Removal of 95%
of fish stock
1980s2004
P. perfoliatus
2005-present
E. nuttallii, N. marina, C. globularis,
C. connivens
none Pot and Ter Heerdt,
2014; Verhofstad et al.,
2017
30 Naarder-meer 1,042 ?/1.0 1993–1996: Sediment
dredging
1980–1989
Z. palustris, P. pectinatus
1990–1995-?
Najas marina, M. spicatum,
C. globularis, R. circinatus,
C. demersum, N. obtusa
Bootsma et al., 1999
(Continued)
Frontiers in Plant Science | www.frontiersin.org 7February 2018 | Volume 9 | Article 194
Hilt et al. Response of Macrophytes to Restoration
TABLE 2 | Continued
No. Lake Country Size
(ha)
Depth
(max/mean)
(m)
Measures applied Period of turbid
conditions and major
remaining macrophyte
species
Periods of clear conditions,
macrophyte species and
coverage (% lake area)
Period of return
to turbid
conditions
Data
Figure 2B
References
31 Finjasjön SE 1,100 12/2.7 since 1970: nutrient load
reduction, 1987:
suction-dredging of
sediments, 1992-2014.
removal of cyprinids
(Abramis brama, Rutilus
rutilus)
?-1994
M. spicatum, E. canadensis,
P. perfoliatus
1995-?
E. canadensis (disappeared 1997),
M. spicatum
P. perfoliatus
?Annadotter et al., 1999;
Strand and Weisner,
2001; Lage et al., 2015
32 Ringsjön
(Western Bay)
1,480 5.4/3.1 1992: Removal of about
50% of cyprinid fish
?-1992 P. crispus, P. perfoliatus,
P. pectinatus,
M. spicatum,
P. lucens, E. canadensis
2000–2005 Strand, 1999; Hansson
unpubl. data
33 Vasatorp-
dammen
2.1 1.4/1.1 1992: Fish removal by
rotenone
1989–1992 none 1993–1996
P. natans, P. obtusifolius,
C. demersum, C. globularis
?Blindow et al., 2000
34 Væng DK 15 1.9/1.2 1986-88 and 2007-09: Fish
removal
?-1986 1989–1996 and
2010-now
E. canadensis
1997–2009 x Jeppesen et al., 1991;
Søndergaard et al.,
2017
35 Arreskov 317 3.7/1.9 1991: Fish removal ?-1991 1992–1998
Characeae,
P. pectinatus, P. pusillus, P. crispus,
Z. palustris,C. demersum,
Z. pedunculata
xLauridsen et al., 2003b;
Søndergaard et al.,
unpubl. data
36 Alderfen UK 5.2 1.2/1 1979: Isolation from inflow,
1990: natural fish kill,
1992–1993: sediment
removal, 1995, 2000: fish
removal
Several turbid periods C. demersum
20–65%
several phases of
macrophyte
decline (e.g.,
1994, 1999–2000)
xMoss et al., 1986,
1990; Perrow et al.,
1997; Hoare et al.,
2008; Phillips et al.,
2015
37 Cockshoot
Broad
5.5 1.2/1 1992: Isolation from river,
sediment removal;
1989/90, 1996-2002,
2004-08:
fish removal
1970–1980
none
1990-2012
C. demersum, N. marina,
Z. palustris
40-58%
none x Moss et al., 1986;
Hoare et al., 2008;
Phillips et al., 2015
38 Hoveton Little
Broad Pound
End
15.5 1.5/1.0 1990: Suction dredging,
1990-1999: several fish
removals from isolated bay
(Pound End)
1970s1991 1995-2006
C. demersum, N. marina
Low macrophyte
abundance since
2001
xHoare et al., 2008;
Phillips et al., 2015
39 Ormesby
Great Broad
40 1.5/0.9 1995: Fish removal 1970–1989
Chara, Z. palustris,
P. pectinatus, C. demersum,
M. spicatum, P. pusillus,
P. crispus
1995–2010
C demersum, E. canadensis,
Z. palustris, P. pectinatus, Chara,
P. friesii
67%
none x Phillips et al., 2015
(Continued)
Frontiers in Plant Science | www.frontiersin.org 8February 2018 | Volume 9 | Article 194
Hilt et al. Response of Macrophytes to Restoration
TABLE 2 | Continued
No. Lake Country Size
(ha)
Depth
(max/mean)
(m)
Measures applied Period of turbid
conditions and major
remaining macrophyte
species
Periods of clear conditions,
macrophyte species and
coverage (% lake area)
Period of return
to turbid
conditions
Data
Figure 2B
References
40 Cromes 4.3 1.2/1 1988: Sediment removal,
1992: Barrier to isolate from
river,1999: natural fish kill,
2004: Sediment removal
? 1995-ongoing
C. demersum
80%
xPerrow et al., 1997;
Phillips et al., 2015
41 Barton Broad 75 2/1 1980: P reduction from
effluents upstream 1996:
sediment removal
see Table 1 (lake no. 17) 2000-2012
C. demersum, P. pectinatus,
E. canadensis
12%
xPhillips et al., 2005,
2015
42 Schollener
See
DE 95 1/<1 2002: Natural fish kill during
summer flood
1980 to 2003 none 2004
C. demersum, N. marina,
P. berchtoldii, P. crispus
2005-ongoing? x Knösche, 2008
43 Rangsdorfer
See
272 2.5/1.5 2009/10: Natural winter fish
kill
?-2009
(probably several decades)
2010–2011
C. demersum, M. spicatum,
P. crispus
2012- ongoing Hussner et al., 2014
44 Schwandter
See
16.5 2.5/1.6 2002: P-precipitation with
aluminum sulfate
2009/10: Natural winter fish
kill
?-1995-2002
sparse stands of
C. demersum, P. crispus
2003-2010
C. demersum, P. crsipus
100%
2011–2015-?:
shift to turbid state
after carp stocking
xMathes, 2007; Nixdorf
et al., 2013; Hussner
et al., 2014
5 Ivenacker See 73.3 1.9/1.1 2009/10: Winter fish kill
2012/13: sediment dredging
?-2009 2010-?
Characeae
? x Nixdorf et al., 2013
46 Schloßsee
Buggen-
hagen
9.8 2.9/0.8 1990-96:
Improved wastewater
treatment,
1997: Sediment dredging
?-1997 1998-?, 2006-?
C. demersum, C. submersum.
C. hispida
2003 x Mathes, 2007
47 Möllener See 18.7 2.2/2 2006: P-precipitation with
aluminum sulfate
? 2007–2009-?
Characeae (occurred 3 years after
treatment)
?Hussner et al., 2014
48 Bachtel-
weiher
4.8 2/1.6 2002: Lake drained empty,
fish completely removed,
restocked with pike
?-2002 2003–2006-?
M. spicatum, R. trichophyllus,
C. globularis, N. mucronata
50-90%
?Hussner et al., 2014
49 Herren-wieser
Weiher
6.7 4.7/1.8 2001: Lake drained empty,
partial sediment removal,
fish completely removed,
restocked with pikeperch
?-2001 2002
Mainly R. circinatus, P. berchtoldii,
P. crispus, P. lucens, P. pectinatus,
C. contraria, C. globularis
75%
2003–2005-?
(after illegal carp
stocking)
Hussner et al., 2014
Frontiers in Plant Science | www.frontiersin.org 9February 2018 | Volume 9 | Article 194
Hilt et al. Response of Macrophytes to Restoration
FIGURE 2 | Total phosphorus (TP) concentrations and Secchi depth in spring
(April–June) and summer (July–September) of different north temperate
shallow lakes (A) before and after external nutrient load reductions during the
turbid, the intermediate recovery and the clear-water state (for details see
Table 1) and (B) before (turbid) and after (clear) biomanipulation or other
lake-internal measures (for details see Table 2).
restored through biomanipulation of the fish community
(Noorddiep, Wolderwijd, Zwemlust, for details see no. 24–26 in
Table 2).
Mann-Whitney U-tests were performed to compare lake
size, maximum and mean depths between lakes with different
restoration measures (external nutrient load reduction vs.
internal measures). Total phosphorus concentrations and Secchi
disk transparency were compared between the different states
(turbid, intermediate, clear) in lakes with external nutrient load
reduction) using Kruskal-Wallis tests and subsequent posthoc
comparisons (separately for different seasons). The same was
done for lakes with internal measures using Mann-Whitney U
tests. All statistical tests were run in SPSS.
PCLake Simulations
Simulating the response of water clarity and macrophyte biomass
to external nutrient load reduction and detecting thresholds of
nutrient loading for either intermediate recovery or clear state
required adaptation of the established ecosystem model PCLake.
This model has previously been used to estimate threshold
responses of shallow lakes to nutrient loading (Janse et al.,
2008; Janssen et al., 2017), and to simulate the response of
temperate shallow lakes to climate warming (Mooij et al., 2007),
to mowing of macrophytes (Kuiper et al., 2017) and—in a variant
of the model with three plant species—to biomanipulation and
herbivory (Janse et al., 1998). PCLake consists of a number
of coupled ordinary differential equations that describe the
most important biotic (submerged macrophytes, phytoplankton,
detritivorous macrozoobenthos, zooplankton, zooplanktivorous
fish, benthivorous fish, and piscivorous fish) and abiotic (detritus,
inorganic material, dissolved phosphorus, ammonium, and
nitrate) components of both the water column and the top-
layer of the sediment in a non-stratifying shallow lake (Janse,
1997, 2005). All organic components (apart from predatory
fish) are modeled in terms of dry weight (DW), nitrogen
(N), and phosphorus (P), and hence the nutrient-to-dry-weight
ratios of the organic components are variable. Internal fluxes
of nutrients between the sediment layer and the pelagic zone,
including internal loading, are accounted for and modeled
dynamically.
For our simulations we used the default settings of a lake in
PCLake. This default lake represents a relatively shallow lake with
an average depth of 2 m and is relatively small with a maximum
fetch of 1,000 m, an areal hydraulic loading of 20 mm days1
(=7.2 m year1), no infiltration or seepage, no surrounding
wetland zone, and a lightly clayish sediment (30% dry matter, of
which 10% organic matter, and 10% fine mineral material) (Janse
2005). Due to small size and shallowness, the lake is dominated
by macrophytes when nutrient loads are sufficiently low, but as
nutrient loads increase the lake switches to a turbid state. This
switch occurs rather suddenly due to the positive feedbacks in the
model (Janse 2005) that lead to a critical transition (e.g., Scheffer
and Carpenter, 2003). A common method to determine critical
transitions is bifurcation analysis. In this approach the model is
run to equilibrium several times, each with a different nutrient
load. For each run, the yearly average phytoplankton chlorophyll-
a concentration and macrophyte biomass are calculated. To
assess the presence of hysteresis this procedure is repeated twice
for each level of nutrient load, the first starting from a clear
lake and the second from a turbid lake. Where the equilibrium
outcomes of these two runs with identical nutrient load differ,
hysteresis is inferred. Here we ran the model for nutrient loads
ranging from 0.1 to 2.5 mg m2days1(0.4–9.0 kg ha1year1)
to cover a wide range of the eutrophication axis. The output of
the bifurcation analysis is a load-response curve (or bifurcation
plot) showing the effect of nutrient load on the biomass of
primary producers. The point of a sudden switch marks the
critical transition(s).
To simulate the influence of temperature and light on
the response of different macrophyte species to nutrient load
reduction we had to make two adjustments to the original
formulations of PCLake, while maintaining the modeling of
macrophytes as one functional group. First, the original power
function for temperature limitation of macrophytes was replaced
Frontiers in Plant Science | www.frontiersin.org 10 February 2018 | Volume 9 | Article 194
Hilt et al. Response of Macrophytes to Restoration
by a temperature optimum curve LT(-):
LT=e0.5
σ2h(TTopt)2(Tref Topt )i
Here, σ(C) is the temperature constant based on a Gaussian
curve, T(C) is the water temperature, Topt (C) is the optimum
temperature for macrophytes and Tref (C) is the reference
temperature used to normalize the limitation function to 1
(Janse, 2005). With this function the model is flexible to simulate
macrophytes with different temperature optima. The second
adjustment is the timing of root allocation which occurs in
the autumn when macrophytes store energy to overwinter, for
instance, in propagules. In the original PCLake, this timing was
linked to a specific day in the year while submerged plants
are known to respond to physiological and environmental cues,
such as light availability, to determine timing of root allocation
(Madsen, 1991). Hence, we decided to link the timing of root
allocation to a minimum daily light availability for macrophytes,
following Madsen (1991),Van Dijk and Van Vierssen (1991), and
Van Dijk and Janse (1993). The available light for macrophytes is
based on the light availability over the depth of the water column
corrected for periphyton shading which is estimated by:
PAR =PAR0D
LN PAR0
PARbot
D1εperip hyton
In which PAR (W m2) is an approximation of the
average Photosynthetic Available Radiation (PAR) for plant
photosynthesis, PAR0(W m2) is the PAR available at the top of
the macrophyte layer, PARbot (W m2) is the PAR available at the
bottom of the macrophyte layer, D (m) is the depth and εperiphyton
is the shading by periphyton. In order to restrict complexity,
we refrained from adding periphyton as an extra compartment
to the model, but instead used the empirical relationship of
Vadeboncoeur et al. (2006) to estimate periphyton chlorophyll-a
biomass:
LOG10 Chlaperip hyton=c1LOG10 (TP)+c2
where c1=1.79 and c2=0.85 and TP is the in-lake
total phosphorus concentration (mg m3). The periphyton
chlorophyll-a biomass is then used to estimate the shading effect:
εperi phyton =Chlaperip hyton Sf(L)f(T)f(α)
where Sis the specific light attenuation by periphyton (m g1)
and correction factors for light limitation f(L), temperature
limitation f(T) and available plant surface area for periphyton
growth f(α). The shading effect of periphyton affects the timing
of root allocation through light availability to macrophytes as well
as the light limitation of macrophyte shoots.
PCLake has previously been calibrated following a Bayesian
approach to parameter estimation and uncertainty analysis using
data from nearly 40 temperate shallow lakes (Janse et al., 2010).
Although this calibration did not account for the specific effect
of periphyton, the data used for model calibration most likely
integrate this effect indirectly (Kuiper, 2016). By adding the effect
of periphyton to PCLake implicitly, we thus have to technically
recalibrate the model. This implies adjusting the parameter
settings of the model, such that given the same boundary
conditions, the model produces the same output. Therefore we
have calibrated the adjusted model manually by lowering the
half saturation light constant of vegetation and decreasing the
parameter for dark respiration of vegetation (see Table 3 for new
and calibrated parameter settings). PCLake is implemented in
DATM (Mooij et al., 2014). For the full overview of parameter
settings and model formulations, please see the DATM-file in the
Supplementary Material.
RESULTS
Lake Water Quality following External and
Internal Restoration
Our literature review provided information on water quality
and macrophyte development in 21 turbid lakes that were
subject to external nutrient loading reduction without
additional in-lake measures (Table 1) and 28 lakes with in-
lake restorative measures. Some of these measures were preceded
or accompanied by external nutrient loading reduction (Table 2).
Lakes with internal measures were on average smaller
than lakes with external nutrient load reduction alone, while
maximum depths were higher but mean depths were similar
(Table 4). Turbid conditions lasted from 1 year (Lake Veluwe)
to 51 years (Dümmer). Often, however, exact timing and
duration of the turbid period are unknown (Tables 1,2). During
the turbid phase, spring and summer TP concentrations were
high (>0.15 mg L1), while Secchi disk transparencies
were low (0.4 m), with considerable differences between lakes
(Figures 2A,B,Table 4).
TABLE 3 | Parameter settings for PCLake.
Parameter Description Unit ValueaSource
σTemperature constant
based on a Gaussian curve
C 20 -
Topt Optimum temperature for
macrophytes
C 20 -
Tref Reference temperature C 20
c1Slope of logistic curve
periphyton
1.79 Vadeboncoeur
et al., 2006
c2Intercept logistic curve
periphyton
0.85 Vadeboncoeur
et al., 2006
S Light attenuation by
periphyton
m g10.03 Van Dijk, 1993
Lmin Minimal light availability cue
needed for plants to initiate
increased root allocation
W m291.2 Calibrated
hLveg Half saturation light constant
of vegetation at 20C
W m212 (17) Calibrated
kDresp Dark respiration rate of
vegetation
day10.015 (0.02) Calibrated
aValues between brackets are the original value.
Frontiers in Plant Science | www.frontiersin.org 11 February 2018 | Volume 9 | Article 194
Hilt et al. Response of Macrophytes to Restoration
TABLE 4 | Size, maximum and mean depths, total phosphorus concentrations and Secchi disk transparency in lakes after external nutrient load reduction or implementation of internal measures.
Lakes State Size (ha) Maximum depth (m) Mean depth (m) Total phosphorus concentrations (µg L1) Secchi disk transparency (m)
spring summer spring summer
mean median mean median mean median mean median mean median mean median mean median
External nutrient
load reduction
Turbid 984 450 5.0 3.9 2.4 1.8 281 163 A 317 244A 0.66 0.45 A 0.46 0.4 A
Intermediate 97 94 A 206 206 A 1.23 1.11 AB 0.75 0.6 AB
Clear 53 51 B 59 49 B 1.74 2.17 B 1.17 1.12 B
P(Kruskal-Wallis test) 0.02 0.007 0.042 0.02
Internal measures Turbid 273 18 2.8 2.1 1.5 1.3 229 156 a 256 150 a 0.62 0.45 a 0.45 0.40 a
Clear 168 81 a 170 103 b 1.12 0.96 b 1.10 0.99 b
P(Mann-Whitney U test) <0.001 0.016 0.076 0.051 0.011 0.026 <0.001
Size and depths include all lakes, while TP and transparency data were only available for selected lakes (see Figure 2,Tables 1,2). Different letters indicate significant differences between values of different states (external nutrient load
reduction: Kruskal-Wallis test, capital letters; internal measures: Mann-Whitney U-test, small letters).
External nutrient loads were usually reduced following
improved sewage treatment in the catchment. Lake Veluwe and
Wolderwijd were flushed with nutrient-poor water (Table 1). The
most commonly applied internal measure was biomanipulation
in the form of removal of benthivorous and planktivorous fish
(in some cases by draining/pumping the lake dry). In nine
lakes, sediment removal was applied as an additional or the sole
measure and in five lakes natural fish kills occurred during severe
winters or as a result of a summer flood (Table 2), emulating
the effects of planned biomanipulation. During the first years
after nutrient load reduction, TP concentrations were lower than
during the turbid period, but still about twice as high in summer
as in spring. Spring water transparency was higher than during
the turbid period, while summer values were still low (Table 4,
Figure 2A). This phase has been found to last up to 20 years
(e.g., Müggelsee), but often its start has not been recorded and/or
lakes have not yet reached stable clear conditions (Table 1).
Lakes Müggelsee, Veluwemeer, and Eemmeer show a similar
intermediate recovery state with spring water transparencies
being higher than during turbid summer conditions which lasted
for about 20 years (Figure 3). A switch back from intermediate
to turbid conditions has only been observed in Lake Steinhuder
Meer, in this case about 10 years after macrophytes returned
(Table 1).
After implementation of in-lake measures, TP concentrations
were at the same level as before, in both spring and
summer, whereas water transparency in spring and summer was
significantly higher than before restoration (Table 4,Figure 2B).
These unstable clear conditions often only lasted for a few
years and many lakes shifted back to turbid conditions (e.g.,
no. 23, 26, 32, 34, 36, 38, 42, 43, 44, 46, 49 in Table 2). Fish
stock reductions in Wolderwijd, Zwemlust and Noorddiep led
to clear conditions both in spring and summer (Figure 4). Lake
Zwemlust shifted back to turbid conditions after 9 years, while
Noorddiep stayed clear for at least 8 years with no further
information on subsequent periods.
Only six of the lakes with external nutrient load reduction
(25%, no. 1, 3, 7, 10, 18, and 19 in Table 1) reached stable
clear-water conditions with lower ambient TP concentrations
and higher Secchi depths than during the intermediate recovery
state both in spring and summer (Figure 2A,Table 4). In
four of the 28 lakes (14%, no. 25, 29, 37, 39 in Table 2,
Lake Wolderwijd in Figure 4), stable, longer-term clear-water
conditions with a diverse macrophyte flora were obtained after
the application of in-lake measures. For several lakes, their
longer-term development is not known (Table 2).
Model Simulations on Lake Response to
External Nutrient Load Reduction
The results of the simulations using the adjusted PCLake model
revealed three stages for lakes that undergo external nutrient
load reduction: a turbid state, an intermediate recovery state and
a clear state (Figure 5A). For the default lake in PCLake, the
turbid state occurs if the P load exceeds 1.3 mg P m2days1.
The intermediate recovery state occurs between the two critical
transitions that appear at a P load of 1.06 and 1.3 mg P m2
Frontiers in Plant Science | www.frontiersin.org 12 February 2018 | Volume 9 | Article 194
Hilt et al. Response of Macrophytes to Restoration
FIGURE 3 | Total phosphorus (TP) concentrations and Secchi depth in spring (April–June) and summer (July–September) and macrophyte coverage in Lake
Müggelsee; Lake Veluwe and Lake Eem during the turbid (green), the intermediate recovery (brown) and the clear-water (blue) state (for lake details see Table 1).
days1. If the P load drops below 1.06 mg P m2days1the lake
turns into a clear state. The critical P loads for the intermediate
recovery state are smaller in the adjusted model including
periphyton than with the original formulation of PCLake (critical
transition between approximately 1 and 2 mg P m2days1,
Janse et al., 2008). The periphyton effect thus reduces the
threshold for intermediate recovery state of the default lake by
about a quarter of what it would be without periphyton. The
reason for this is that in case of hysteresis, the position of the
highest critical nutrient load is mainly determined by macrophyte
characteristics, while the position of the lowest critical nutrient
load is mainly determined by phytoplankton characteristics. This
is in agreement with the results of the sensitivity analysis of
PCLake (Janse et al., 2010). Since periphyton negatively affects
the performance of macrophytes, the highest critical nutrient
load was reduced, while phytoplankton characteristics were
unaltered and the lowest critical nutrient load (1 mg P m2
days1) thus did not change.
The turbid state was characterized by lack of macrophytes
and enhanced phytoplankton biomass with increasing nutrient
loading (Figure 5A). The intermediate state was characterized by
a changed phenology of the primary producers. With increasing
nutrient loading, the phytoplankton summer peak shifted to
earlier dates of the year and this advancement of phytoplankton
was mirrored by an abbreviated macrophyte growing season
(Figure 5A). The shorter growing season was a direct effect
of the inclusion of periphyton shading in our adapted version
of the model. The shading of macrophytes by periphyton was
most severe at the peak of summer when the light input and
water temperature are high. As a result, macrophyte growth was
limited to the period just after the clear water phase in spring
until the start of the summer phytoplankton bloom. The clear
state was characterized by an increase of macrophyte biomass
with increasing nutrient loading (Figure 5A). Phytoplankton
production was restricted to the spring bloom peak and there was
no summer bloom.
The bifurcation plot (Figure 5B) shows a less sudden
transition of macrophytes and phytoplankton compared to the
abrupt transition between the phytoplankton-dominated turbid
state and the macrophyte-dominated clear state, often seen in
bifurcation plots in literature (e.g., Janse et al., 2008). The
gradual course of the bifurcation plot is due to the inclusion
of periphyton shading of macrophytes. This shading permits
high biomass of both macrophytes and phytoplankton within
Frontiers in Plant Science | www.frontiersin.org 13 February 2018 | Volume 9 | Article 194
Hilt et al. Response of Macrophytes to Restoration
FIGURE 4 | Total phosphorus (TP) concentrations and Secchi depth in spring (April–June) and summer (July–September) and macrophyte coverage in Lake
Wolderwijd, Lake Zwemlust and Lake Noorddiep before (green and brown) and after (blue) biomanipulation (for lake details see Table 2). In Lake Zwemlust,
P. berchtoldii occurred instead of P. pectinatus, and the coverage in Lake Noorddiep was only estimated based on the information that it was higher than 25% (Gulati
and Van Donk, 2002).
the same year during the crashing or intermediate recovery
phase. Furthermore, the region of hysteresis is tilted, leading
to a less abrupt and thus more realistic critical transition from
phytoplankton dominance to macrophyte dominance and back
again.
Macrophyte Species Recovery following
External and Internal Restoration
During the turbid phase, three lakes with subsequent external
nutrient load reductions were reported to lack macrophyte
stands altogether and six and one lakes had sparse stands of
Potamogeton pectinatus (also known as Stuckenia pectinata) and
P. pusillus, respectively, while no information was available
for the remaining lakes (Table 1). For lakes with internal
measures, information on macrophyte species present during the
turbid phase is available for 10 lakes. Apart from P. pectinatus
and P. perfoliatus, plants such as Ceratophyllum demersum,
M. spicatum or E. canadensis are mentioned, if indeed plant
stands were present at all (Table 2).
During the intermediate recovery state P. pectinatus was the
dominant macrophyte species in two thirds of the analyzed lakes
with reduced external nutrient loading. Other pondweed species
such as P. perfoliatus, P. crispus and P. pusillus or Zannichellia
palustris were also found in several lakes during the intermediate
recovery state, while other groups such as Characeae or Elodea
species were much less common (Table 1,Figure 6). Lakes
Müggelsee, Veluwemeer and Eemmeer were all dominated by
P. pectinatus during the intermediate recovery state which lasted
for about 20 years (Figure 3). Müggelsee and Veluwemeer seem
to have entered a stable clear state with more diverse submerged
vegetation (including Characeae in Veluwemeer) in 2011 and
1996, respectively, while Eemmeer has not yet reached that
phase despite the recent detection of Characeae (Figure 3). Three
other lakes also reached a stable clear state and were colonized
by different species of Characeae and/or Najas marina, Elodea
species and C. demersum (Table 1).
The dominant macrophytes occurring after lake internal
measures were Characeae, C. demersum, Elodea species or
N. marina (Table 2,Figure 6). Often, lakes had only one or two
dominant species, and in at least 10 cases, lakes switched back
to turbid conditions and lost their macrophytes again (Table 2,
Figure 4). The response of macrophytes occurred gradually in
Frontiers in Plant Science | www.frontiersin.org 14 February 2018 | Volume 9 | Article 194
Hilt et al. Response of Macrophytes to Restoration
FIGURE 5 | (A) Time series of simulation using PCLake of chlorophyll-aconcentrations and macrophyte shoot biomass for different phosphorus (P) loadings within
the clear, intermediate recovery and turbid states. Darker colors are associated with higher P loading simulations. (B) Hysteresis plots showing yearly mean simulated
values of chlorophyll-aconcentrations and macrophyte shoot biomass for different phosphorus (P) loadings within the clear, intermediate recovery and turbid states.
Arrows denote the directions of the hysteresis effects.
Wolderwijd, showing an increase in Characeae coverage, while
in the much smaller Zwemlust and Noorddiep, macrophytes
immediately covered large areas, E. nuttallii, E. canadensis and
C. demersum dominating (Figure 4,Table 2). In four lakes, stable
clear-water conditions were obtained after in-lake measures, all
being characterized by Characeae dominance (Table 2,Figure 4).
DISCUSSION
Our analysis of long-term data from 49 temperate shallow
lakes during their recovery from a turbid phase reveals
that both a reduction of the external nutrient loading and
implementation of lake-internal measures often result in the
occurrence of an intermediate state (Figure 1) that can last
for several decades. External nutrient load reductions are often
followed by the re-occurrence of a spring clear-water phase
that opens a “window of opportunity” for macrophyte re-
colonization, but with only a short growth period due to turbid
conditions during summer. This pattern was confirmed by our
model simulations (Figure 5). As hypothesized, macrophyte re-
establishment following nutrient load reduction occurs in a
reversed sequence to the one described for eutrophication by
Sayer et al. (2010a,b). In contrast, lake internal measures such as
fish or sediment removal often result in clear-water conditions
during spring and summer. This clear-water state is, however,
often only temporary and lakes frequently shift back to turbid
conditions one or a few years after the restoration. Likely, the
duration of the clear-water conditions is related to nutrient
loading and the intensity of the restoration effort (Hansson et al.,
1998). Only in a few examples have longer lasting clear-water
conditions been observed. These required spring and summer TP
concentrations below 0.05 mg L1.
We also have evidence for our second hypothesis, namely
that different types of restoration measures influence the
macrophyte community composition. P. pectinatus and a few
other pondweeds most often recolonise temperate shallow
lakes with reduced external nutrient loading and dominate
during the intermediate recovery state. The implementation of
Frontiers in Plant Science | www.frontiersin.org 15 February 2018 | Volume 9 | Article 194
Hilt et al. Response of Macrophytes to Restoration
FIGURE 6 | The 10 most common macrophyte taxa in north temperate
shallow lakes during re-colonization after reduction of external nutrient loading
(light gray) or implementation of internal measures (dark gray; for more details
see Tables 1,2).
internal restoration measures results in the establishment of
a different community, often consisting of a small selection
from either hornwort (C. demersum), charophytes, water weeds
(E. canadensis, E. nuttallii) or naiad (N. marina). Only in a
few cases have lakes reached a state of clear-water conditions
during spring and summer and with a more diverse macrophyte
community.
Submerged Macrophyte Survival during
the Turbid Phase
Whether and which macrophyte stands or propagules survive
during the turbid phase depends on the occurrence of
macrophyte species before the shift to turbid conditions and on
the length and severity of the turbid phase (e.g., Vari and Toth
2017). Seed banks in shallow lake sediments have often been
assumed to be insufficient for recovery of submerged vegetation
by germination, due to low numbers of viable seedlings, strong
seed dormancy, strict germination cues and the reliance of many
species upon vegetative reproduction (Haag, 1983; Kautsky, 1990;
Rodrigo et al., 2013; Baldridge and Lodge, 2014). In contrast,
De Winton et al. (2000) and Verhofstad et al. (2017) have
shown that seed banks from even the most degraded lakes are
capable of an emergence response and thus offer a potential
means to restore vegetation. In our survey, the duration of
complete macrophyte loss was often unrecorded, but periods of
several decades are common (Tables 1,2). If macrophyte stands
survived during this period and were recorded, these were often
sparse stands of P. pectinatus, a species commonly associated
with the crashing phase during eutrophication (Sayer et al.,
2010a). This species survives in very shallow water even under
phytoplankton dominance (Hilt et al., 2013) and its strongly
apical growth form may allow it to survive at greater depths
in turbid water. The shallow littoral, especially in larger lakes,
is strongly disturbed by wave action and only species with
high anchorage and breaking strength can survive under these
conditions (Schutten et al., 2005). Potamogeton pectinatus has
a high breaking strength (Brewer and Parker, 1990) and its
phenotypic plasticity allows it to form short plants in shallower
water (Idestam-Almquist and Kautsky, 1995). In contrast, other
common species in eutrophic lakes such as E. canadensis or
M. spicatum have been described as deep water species with lower
tensile strength (Brewer and Parker, 1990), while C. demersum
has no roots for anchorage. These species are thus less likely
to persist through severely turbid states in very shallow littoral
areas and to serve as remnant populations for re-colonization, at
least in larger lakes. Nevertheless, they have been reported during
turbid phases in four lakes included in our survey, most probably
in wind-protected areas or bays.
Knowledge of the survival of propagules in sediments during
turbid phases is limited. In general, charophyte oospores and
macrophyte seeds have been found to survive up to 150 years
(Kaplan and Muer, 1990; De Winton et al., 2000; Alderton et al.,
2017). Germination tests with sediments have been suggested
before implementing lake restoration measures to forecast the
potential for macrophyte recovery from internal sources (Hilt
et al., 2006), however, these are not routinely applied. Re-
colonizing macrophyte clones of several species that originated
from periods before eutrophication have been found (Sand-
Jensen et al., 2008), but knowledge about the origin of re-
colonizing macrophytes post-restoration remains scarce (Bakker
et al., 2013).
Response of Macrophytes to Nutrient Load
Reductions
Shallow temperate lakes often show a relatively rapid response
to reductions in external phosphorus loading, characterized by
a reduction in phytoplankton biomass during spring and early
summer (Jeppesen et al., 2005). During late summer, however,
the response is delayed because of sustained remobilisation of
phosphorus from the sediment (Søndergaard et al., 2013). As a
consequence, high phytoplankton and cyanobacterial abundance
are often reasserted in summer (Sommer et al., 2012) resulting in
turbid water and preventing macrophyte growth. In our survey,
such conditions occurred at spring TP concentrations of around
0.1 mg L1, while summer concentrations were twice as high.
The increased water transparency in spring and early summer
seems to be exploited by certain macrophyte species, in our
survey mainly P. pectinatus along with P. perfoliatus, P. crispus
and Z. palustris. These macrophyte species are characterized by
specific traits that may explain their prevalence. Firstly, they can
compress their whole life cycle into the short clear-water period
in spring and early summer due to early germination from tubers,
turions or seeds, shortening time to peak biomass and allowing
early formation of overwintering tubers and seeds (Table 5).
Secondly, they can have short growth forms that can establish
in very shallow habitats (e.g., Van Vierssen, 1982a). Thus, they
are often the species that survive in turbid conditions in shallow
Frontiers in Plant Science | www.frontiersin.org 16 February 2018 | Volume 9 | Article 194
Hilt et al. Response of Macrophytes to Restoration
TABLE 5 | Characteristics of macrophyte species/groups typically recolonising temperate shallow lakes after external nutrient load reduction or following implementation of lake-internal measures such as
biomanipulation.
Restoration
measure
Macrophyte
species
Re-production
mode
Germination Colonization
mode
Timing of peak
biomass
Specific features Epiphyton
density
Allelopathic
activity
Susceptibility to
herbivory
External nutrient
load reduction
Potamogeton
pectinatus
Mainly by tubers,
although seeds
are formed1
Between
10-15C7
Rhizomatic
growth11
June19 Concentration of
biomass below
water surface1
Tubers rich in
carbohydrates32
Low13 Low2High3,14,28
P. perfoliatus Turions29, seeds Rhizomatic
growth11
High20
P. crispus Mainly by
turions26
Turions can sprout
in late summer, but
shoot elongation
after winter above
10C26
Turions Late spring to
early summer25
Low2
P. pusillus Turions4Rhizomatic
growth4,12
Low13 Low2High14
Zannichellia
palustris
Seeds5Mainly
temperature-
dependent, above
12-16C5,27
Seeds12 Short, but
numerous shoots
allow development
in very shallow
water27
Tolerant to
disturbance by
wave action27
Low13 Low2
Internal measures Elodea canadensis Vegetatively6May-June22, can
be evergreen
Fragments,
vegetative growth
(peripheral
propagation)4,12
Sudden biomass
collapses6,8,9
High13 Medium2Medium3
E. nuttallii Vegetatively17 Can be evergreen Fragments,
vegetative
growth12
Aug.-Oct.21 Extremely high
reproductive
capacity16,17
Sudden biomass
collapses
High13
Low15
Medium2Medium18
Ceratophyllum
demersum
Vegetatively
(dormant apices)
April22, can be
evergreen
Fragments,
vegetative
growth12
August31 Sudden biomass
collapses10
Low15 High2Low
Najas marina Annual, seeds Late in (temperate)
season (>20C)30
Seeds11 Mid August11 Medium2Low18
Characeae Oospores Can be evergreen,
overwintering by
shoot apices
Oospores Late summer23 Low15 Medium2Medium3
1Van Wijk (1989),2Hilt and Gross (2008),3Dorenbosch and Bakker (2011),4Barrat-Segretain and Bornette (2000),5Bytnerowicz and Carruthers (2014),6Simberloff and Gibbons (2004),7Madsen and Adams (1988),8Rørslett et al.
(1985),9Strand and Weisner (2001),10Van de Bund and Van Donk (2002),11Vari and Toth (2017),12Capers (2003),13Lalonde and Downing (1991),14 Hidding et al. (2010),15Grutters et al. (2017),16 Simpson (1990),17 Josefsson
(2011),18Pot and Ter Heerdt (2014),19Körner (2001),20 Choi et al. (2002),21Best and Dassen (1987),22 Best (1977),23Talling and Parker (2002),24Lillie et al. (1997),25 Woolf and Madsen (2003),26Tobiessen and Snow (1984),27Van
Vierssen (1982b),28Weisner et al. (1997),29 Wolfer and Straile (2004),30Van Vierssen (1982c),31Best (1982),32 Spencer (1986).
Frontiers in Plant Science | www.frontiersin.org 17 February 2018 | Volume 9 | Article 194
Hilt et al. Response of Macrophytes to Restoration
margins (see Submerged Macrophyte Survival during the Turbid
Phase) and then expand into deeper water with improvements in
clarity during nutrient load reduction. Rhizomatic growth from
remaining P. pectinatus stands has been shown, via microsatellite
analyses, to be the dominant re-colonization mode in Lake
Müggelsee; more recently established P. pectinatus stands had
lower genotype diversity and were comprised of only a small
subset of genotypes from shallower areas (Hilt et al., 2013).
Thirdly, energy reserves in vegetative propagules such as tubers
of P. pectinatus allow early onset of growth independent of
light availability (Spencer, 1986). In addition, P. pectinatus can
concentrate large parts of its biomass just under the water surface
and thus survive in relatively turbid water (Van Wijk, 1988). An
initial colonization of formerly turbid lakes with P. pectinatus
during recovery has also been observed in deeper, stratifying
lakes. Thus, in Lake Tegel (Germany), this species dominated for
more than 20 years after the start of phosphorus stripping in the
major inflow (Hilt et al., 2010).
Usually, the maximum colonization depth of macrophytes
during the intermediate recovery phase is low (around 1 m)
and consequently, depending on the lake morphometry, only
small parts of the lake bed might be covered. In contrast,
very shallow lakes may reach over 50% cover (Table 1). This
suggests that in “deeper” shallow lakes, macrophyte coverage
during the intermediate phase may be insufficient to stabilize
clear-water conditions during later summer. Based on the
findings of Søndergaard et al. (2016) submerged macrophyte
coverage on average needs to pass a threshold of 20% of lake
area to markedly lower phytoplankton densities. In principle,
small stands can be sufficient as a refuge for phytoplankton-
grazing zooplankton against fish predation (Lauridsen et al.,
1996; Portielje and Van der Molen, 1999). However, abundant
colonial and filamentous cyanobacteria which often dominate
the summer phytoplankton communities during lake recovery
cannot be effectively controlled by zooplankton grazers (Wang
et al., 2010 and references therein). Bottom-up stabilizing
mechanisms of macrophytes on water clarity such as nutrient
competition, increased sedimentation within stands and reduced
sediment resuspension will be inefficient at low plant coverage
(Blindow et al., 2014). Low coverage is, however, not the
only reason why macrophytes in the intermediate recovery
phase cannot stabilize clear-water conditions in late summer,
as shown in the case of the very shallow Lake Dümmer, where
cyanobacteria blooms still occurred in summer despite high
macrophyte coverage.
Our model simulations suggest that high periphyton shading
triggers macrophyte disappearance in summer. Periphyton
shading, often accompanied by herbivory (Hidding et al., 2016),
has been shown to impair macrophyte development in empirical
studies (e.g., Jones et al., 2002; Jones and Sayer, 2003; Roberts
et al., 2003) and is argued to be a major factor in the failure
of macrophytes to establish even decades after the start of
nutrient loading reduction, despite suitable water clarity for
plant re-establishment in spring (Phillips et al., 2005). In our
adapted PCLake model, periphyton biomass was dependent on
TP concentrations in the water, based on the positive correlation
between chlorophyll content of periphyton on hard substrata
and TP in the water column (Vadeboncoeur et al., 2006). In
eutrophic shallow, temperate lakes, periphyton is often top-
down controlled by a cascading effect from omnivorous fish
that feed on periphyton grazers such as snails and chironomid
larvae (Jones and Sayer, 2003). Thus, nutrient load reductions
will only reduce periphyton shading after the fish biomass built
up during the turbid period has also been reduced, which
may take 10–15 years (Jeppesen et al., 2005). Furthermore,
the observed dominant macrophyte species in the intermediate
recovery phase after nutrient load reduction (Table 1) show little
or no allelopathic activity that might hamper periphyton growth
(Table 5), thus making them more susceptible to shading by
periphyton.
Cyanobacteria have been shown to potentially inhibit
submerged macrophyte growth via allelopathy (Zheng
et al., 2013), but whether this mechanism contributes to
the disappearance of macrophytes during the recovery phase in
summer is unknown. Most of the dominant macrophyte species
during intermediate recovery after nutrient load reductions are
also highly susceptible to herbivory due to their low content of
polyphenols, low carbon to nitrogen ratio and low dry matter
content (Elger and Willby, 2003; Dorenbosch and Bakker,
2011,Table 5). Periphyton shading may further increase the
sensitivity of macrophytes to herbivory (Hidding et al., 2016).
Finally, fine-leaved species such as P. pectinatus,P. pusillus and
Z. palustris also suffer from leaf plucking by omnivorous fish
during periods of low zooplankton abundance when those fish
switch to macroinvertebrate prey found in the periphyton of
macrophytes (Körner and Dugdale, 2003). Such leaf plucking by
fish can lead to a considerable leakage of nutrients from injured
macrophyte tissue, thereby further stimulating phytoplankton
growth (Hansson et al., 1987). Overall, while being well-suited
for survival during turbid phases and for exploiting the clear-
water conditions in spring for re-colonization, other traits of
macrophyte species typical of the intermediate recovery phase
following nutrient load reduction prevent their survival during
later summer (Table 5).
Stable clear-water conditions in spring and summer with more
diverse macrophyte vegetation were observed when both spring
and summer TP concentrations reached about 0.05 mg L1. This
value corresponds well with a threshold for low cyanobacterial
abundance in shallow lakes found by Jeppesen et al. (2005) and
Triest et al. (2016) and the average critical loading for shifts
from turbid to clear conditions estimated for Dutch shallow
lakes by Janse (2005) and in eastern England by Phillips et al.
(2015). Whether external nutrient load reductions alone were
responsible for the observed low in-lake TP concentrations
in the lakes in our survey that reached stable clear-water
conditions, however, remains questionable. It seems that in
most cases additional changes in either the fish community
(Lake Veluwe, Galenbecker See) and/or exotic mussel invasions
(Lake Müggelsee, Eemmeer) contributed to the observed trend.
In Lake Veluwe, several severe winters, an increase in bream
(Abramis brama) fisheries between 1993 and 1997 and the
increase in zebra mussel densities are all thought to have
contributed to a break in the dominance of cyanobacteria, thus
allowing for the prevalence of stable clear-water conditions
Frontiers in Plant Science | www.frontiersin.org 18 February 2018 | Volume 9 | Article 194
Hilt et al. Response of Macrophytes to Restoration
with charophyte dominance since 1996 (Noordhuis et al., 2016,
Figure 4). Once established, these dense charophyte beds provide
more efficient stabilizing mechanisms for clear-water conditions
than rooted angiosperms (Blindow et al., 2014). Characeae also
successfully replaced P. pectinatus in Lake Wolderwijd after
biomanipulation (Figure 4), while in Swedish Lake Krankesjön
a similar development has been observed, the reasons for which
are unknown (Blindow, 1992; Hargeby et al., 1994; Hansson et al.,
2010). In Lake Müggelsee and Eemmeer, the additional influence
of a sudden invasion of the quagga mussel (Dreissena rostriformis
bugensis) in around 2013 might have contributed to a decline in
TP concentrations and increased water transparencies (Figure 4,
S. Hilt, unpublished, Noordhuis et al., 2016). This species can
colonize soft substrates and thus cover much larger areas than
those previously occupied by the zebra mussel (D. polymorpha)
(Karatayev et al., 2015). In Lake Eemmeer, quagga mussels
filtered the lake volume about five times a day in 2013 (Noordhuis
et al., 2016).
Macrophyte recovery in Steinhuder Meer and Langer See
deviates from the suggested pattern in that both are dominated
by species more typical of lakes having undergone restoration
with internal measures (Table 1). In Steinhuder Meer, a
strong reduction of the fish population has been observed
which was attributed to cormorant activities (Niedersächsischer
Landesbetrieb für Wasserwirtschaft, Küsten- und Naturschutz,
2011). Cormorant effects on fish populations are also suggested
for Felbrigg Lake (C. Sayer, unpublished). Through a natural
increase in cormorants, the lake food web configuration
was likely affected in ways comparable to those of lakes
undergoing biomanipulation. This makes Steinhuder Meer and
Felbrigg Lake cases of external load reduction with added
unintentional internal measures (natural biomanipulation)
possibly accelerating recovery. Therefore, these cases show closer
correspondence to macrophyte community patterns of lakes
having undergone internal measures. They also illustrate that
parallel biological processes may be at play in recovering lakes
that need to be considered in unison to understand the speed and
trajectory of macrophyte recovery.
Response of Macrophytes to
Biomanipulation in Shallow Temperate
Lakes
In contrast to lakes undergoing only reduced external nutrient
loading, submerged macrophytes often respond very quickly
in shallow lakes subjected to biomanipulation by fish removal
(Hansson et al., 1998; Bakker et al., 2013), even at rather
high nutrient concentrations (Figure 2). Macrophytes colonizing
these lakes are often “pioneer” species, such as Elodea or
Ceratophyllum, re-colonizing from either seeds, oospores or
fragments and characterized by high growth rates (Tables 2,4).
Similar species have also been recorded in lakes following
natural fish kills (Sayer et al., 2016) or implementation of
other in-lake restoration measures such as sediment dredging
and phosphorus precipitation (Table 2). Fish removal may
indirectly (due to more periphyton grazing invertebrates) reduce
periphyton shading in summer, a major mechanism preventing
macrophyte survival after nutrient load reduction (see Response
of Macrophytes to Nutrient Load Reductions). The relevance
of this process for macrophyte recovery after fish removal
has not yet been directly tested, although mesocosm trials
in which the density of periphyton grazers are manipulated
produce predictable outcomes in terms of periphyton biomass
and macrophyte composition (Elger et al., 2009). Excretion of
allelopathic substances, which has been detected for many of the
typical species that colonize after biomanipulation (Table 5), may
also contribute to lower periphyton densities.
In general, macrophyte species typically occurring after
introduction of in-lake restoration measures allow for a longer
period of high macrophyte cover and dampen seasonal changes
in phytoplankton abundance as described for “stable” lakes prior
to major eutrophication (Sayer et al., 2010b). Both, Elodea and
charophyte species can remain evergreen in temperate lakes (e.g.,
Søndergaard et al., 2017), thus extending their positive influence
on water quality to seasons outside the influence of annual
species.
In many cases, however, mass developments of monocultures
occur. Monocultures of Elodea or Ceratophyllum species are often
unstable in terms of interannual persistence (Table 5) and can
collapse leading to a shift back to turbid conditions as in, for
instance, Lakes Zwemlust, Væng, and Alderfen Broad (Table 2).
Characeae seem less often involved in sudden collapses, although,
exceptions are known, for example Schlosssee Buggenhagen
(Table 2) or Lake Botshol (Rip et al., 2007). If lakes remain
clear for several consecutive years, which is usually only the case
at lower nutrient concentrations, a more diverse macrophyte
community develops (Table 2,Lauridsen et al., 2003a). Lauridsen
et al. (2003b) assumed that differences in the success of
biomanipulation in Danish and Dutch shallow lakes might be
attributable to variation in pioneer macrophyte species; thus,
Elodea and Potamogeton species, typical for Danish lakes, were
preferred over charophytes by macrophyte-grazing waterfowl
(Weisner et al., 1997). Indeed, increasing top-down control
of periphyton-grazing invertebrates by omnivorous fish, which
increase in abundance in the period after a biomanipulation,
may render macrophytes more susceptible to herbivory (Hidding
et al., 2016).
Conclusions and Implications for Lake
Management
Our analyses suggest that the composition of the macrophyte
community and their seasonal abundance in shallow lakes
during recovery from turbid, highly eutrophic conditions often
depends on remnant macrophyte stands, the specific restoration
measure applied and additional stochastic influences on water
clarity such as winter fish kills, cormorant predation on fish
or introduction of invasive filter-feeding mussel populations. In
turn, the prevailing macrophyte community can influence lake
water quality.
Reductions in external nutrient loading often result in
the re-occurrence of spring clear-water phases exploitable by
a few macrophyte species (mainly pondweeds) with specific
traits. Resistance to wave action permits survival during the
Frontiers in Plant Science | www.frontiersin.org 19 February 2018 | Volume 9 | Article 194
Hilt et al. Response of Macrophytes to Restoration
turbid phase in very shallow areas, in particular in larger
lakes. During recovery these plants germinate early in spring
from energy-rich vegetative propagules and complete their
life cycle in early summer, when phytoplankton takes over.
This intermediate recovery phase may, in some cases, last for
several decades before a more diverse and abundant submerged
macrophyte community develops that stabilizes clear-water
conditions during the entire potential growing season (Figure 1,
Table 1). Our model simulations suggest that, if the premature
termination of macrophyte growth can be prevented, the summer
phytoplankton peak responsible for turbid water and potentially
harmful algae blooms will also be reduced. Simulations also
revealed that at high periphyton shading, the intermediate
recovery phase is shifted to lower nutrient loads compared with a
scenario with lower periphyton shading. Therefore, if periphyton
shading can be reduced external restoration measures could
potentially be effective at a higher nutrient load. Macrophyte
recovery during the intermediate recovery state might be
facilitated by establishing exclosures to protect certain areas from
herbivory by birds and/or predation of periphyton grazers by
omnivorous fish, a lake-wide biomanipulation of fish, or internal
measures, such as TP precipitation to lower water column
TP concentrations in summer. Additional, usually unintended
internal changes, such as reductions in fish abundance by
commercial fisheries, natural fish kills or exotic mussel invasions
can facilitate a shift to clearer conditions in summer and
further aid the establishment of a more diverse macrophyte
community.
In contrast, fish stock reductions, natural fish kills and
sediment removal via suction-dredging can, by themselves,
temporarily restore clear-water conditions in spring and summer,
even at high nutrient concentrations and then allow rapid
colonization by pioneer macrophyte species from in situ seeds,
oospores or vegetative fragments. Fish stock reductions might
thus be a suitable short-term management strategy, but fish
removal needs to be frequently repeated.
Lasting macrophyte recovery can only be achieved in
combination with reduced nutrient loading (Figure 1), a need
that is further accentuated under future climate change scenarios,
where cyanobacterial shading of macrophytes will likely be more
severe (Kosten et al., 2012). Although global, political measures
against ongoing climate warming are slow, local restoration
efforts may reduce the combined stress from e.g. eutrophication
and climate warming (Moss et al., 2011; Scheffer et al., 2015),
and thereby serve as a buffer against further deterioration of
macrophyte beds and the ecosystem services that derive from
lakes and reservoirs (Urrutia-Cordero et al., 2016).
AUTHOR CONTRIBUTIONS
SH conceived the presented idea, wrote the manuscript and
performed the literature research. MA, EB, IB, TD, L-AH, EJ, TK,
AK, JK, TL, RN, GP, JR, H-HS, MS, KvdW, EvD, AW, NW, and
CS provided lake data. MG, JJ, AJ, WM, and ST performed the
modeling. All authors contributed to discussions and the writing
of different parts of the text.
FUNDING
MA and MG were supported by the German Research
Foundation (DFG, grant no. SU 623/1-1 and GRK 2032/1,
respectively). AJ is supported by the Netherlands Environmental
Assessment Agency (PBL) and ST by STOWA (grant no.
443.269). L-AH was supported by the BiodivERsA ERA-
net LIMNOTIP. EJ, MS, TD, and TL were supported by
MARS (Managing Aquatic ecosystems and water Resources
under multiple Stress) funded under the 7th EU Framework
Programme (Contract No.: 603378). JR was supported by
the German Ministry of Education and Research (project
NITROLIMIT, grant no. 033L041 A).
ACKNOWLEDGMENTS
We thank all technicians of IGB Berlin responsible for the
long-term measuring program in Lake Müggelsee. Antje Barsch,
Nadine Baadke (Landesumweltamt Brandenburg) and Antje
Köhler (Senat Berlin) provided data for selected German lakes.
The Environment Agency provided water quality data and the
Broads Authority macrophyte data for the lakes in the Broads
National Park (UK). We acknowledge linguistic improvements
by Anne Mette Poulsen. We thank two reviewers for their helpful
comments.
SUPPLEMENTARY MATERIAL
The Supplementary Material for this article can be found
online at: https://www.frontiersin.org/articles/10.3389/fpls.2018.
00194/full#supplementary-material
REFERENCES
Alderton, E., Sayer, C. D., Davies, R., Lambert, S. J., and Axmacher, J. C. (2017).
Buried alive: Aquatic plants survive in “ghost ponds” under agricultural fields.
Biol. Conserv. 212, 105–110. doi: 10.1016/j.biocon.2017.06.004
Annadotter, H., Cronberg, G., Aagren, R., Lundstedt, B., Nilsson, P. A., and
Ströbäck, S. (1999). Multiple techniques for lake restoration. Hydrobiologia 395,
77–85. doi: 10.1023/A:1017011132649
Bakker, E. S., Sarneel, S. M., Gulati, R. D., Liu, Z., and Van Donk, E.
(2013). Restoring macrophyte diversity in shallow temperate lakes: biotic
versus abiotic constraints. Hydrobiologia 710, 23–37. doi: 10.1007/s10750-012-
1142-9
Baldridge, A. K., and Lodge, D. M. (2014). Long-term studies of crayfish-invaded
lakes reveal limited potential for macrophyte recovery from the seed bank.
Freshw. Sci. 33, 788–797. doi: 10.1086/677070
Barrat-Segretain, M. H., and Bornette, G. (2000). Regeneration and colonization
abilities of aquatic plant fragments: effect of disturbance seasonality.
Hydrobiologia 421, 31–39. doi: 10.1023/A:1003980927853
Bergman, E., Hansson, L.-A., Persson, A., Strand, J., Romare, P., Enell, M.,
et al. (1999). Synthesis of theoretical and empirical experiences from nutrient
and cyprinid reductions in Lake Ringsjön. Hydrobiologia 404, 145–156.
doi: 10.1023/A:1003788900521
Bernes, C., Carpenter, S. R., Gårdmark, A., Larsson, P., Persson, L., Skov, C.,
et al. (2015). What is the influence of a reduction of planktivorous and
Frontiers in Plant Science | www.frontiersin.org 20 February 2018 | Volume 9 | Article 194
Hilt et al. Response of Macrophytes to Restoration
benthivorous fish on water quality in temperate eutrophic lakes? A systematic
review. Environ. Evidence 4, 7. doi: 10.1186/s13750-015-0032-9
Best, E. P. H. (1977). Seasonal changes in mineral and organic components
of Ceratophyllum demersum and Elodea canadensis.Aquat. Bot. 3, 337–348.
doi: 10.1016/0304-3770(77)90038-9
Best, E. P. H. (1982). The aquatic macrophytes of Lake Vechten: species
composition, spatial distribution and production. Hydrobiologia 95, 65–77.
doi: 10.1007/BF00044477
Best, E. P. H., and Dassen, J. H. A. (1987). A seasonal study of growth
characteristics and the levels of carbohydrates and proteins in Elodea nuttallii,
Polygonum amphibium and Phragmites australis.Aquat. Bot. 28, 353–372.
doi: 10.1016/0304-3770(87)90011-8
Blindow, I. (1992). Long- and short-term dynamics of submerged
macrophytes in two shallow eutrophic lakes. Freshw. Biol. 28, 15–27.
doi: 10.1111/j.1365-2427.1992.tb00558.x
Blindow, I., Hargeby, A., and Hilt, S. (2014). Facilitation of clear-water
conditions in shallow lakes by macrophytes: differences between
charophyte and angiosperm dominance. Hydrobiologia 737, 99–110.
doi: 10.1007/s10750-013-1687-2
Blindow, I., Hargeby, A., Wagner, B. M. A., and Andersson, G. (2000). How
important is the crustacean plankton for the maintenance of water clarity in
shallow lakes with abundant submerged vegetation? Freshw. Biol. 44, 185–197.
doi: 10.1046/j.1365-2427.2000.00552.x
Blüml, V., Körner, F., Marxmeier, U., Richter, M., and Schönheim, A. (2008).
Entwicklung und aktuelle Situation der Verlandungsvegetation des Dümmers
(Niedersachsen). Osnabr. Naturwiss. Mitteil. 33/34, 19–46.
Bootsma, M. C., Barendregt, A., and Van Alphen, J. C. A. (1999). Effectiveness
of reducing external nutrient load entering a eutrophicated shallow lake
ecosystem in the Naardermeer nature reserve, The Netherlands. Biol. Conserv.
90, 193–201. doi: 10.1016/S0006-3207(99)00027-0
Brewer, C. A., and Parker, M. (1990). Adaptations of macrophytes to life in moving
water: upslope limits and mechanical properties of stems. Hydrobiologia 194,
133–142. doi: 10.1007/BF00028414
Bytnerowicz, T. A., and Carruthers, R. I. (2014). Temperature-dependent models
of Zannichellia palustris seed germination for application in aquatic systems.
Environ. Exp. Bot. 104, 44–53. doi: 10.1016/j.envexpbot.2014.03.006
Capers, R. S. (2003). Macrophyte colonization in a freshwater tidal wetland (Lyme,
CT, USA). Aquat. Bot. 77, 325–338. doi: 10.1016/j.aquabot.2003.08.001
Choi, C., Bareiss, C., Walenciak, O., and Gross, E. M. (2002). Impact of polyphenols
on growth of the aquatic herbivore Acentria ephemerella.J. Chem. Ecol. 28,
2223–2238. doi: 10.1023/A:1021049332410
De Winton, M. D., Clayton, J. S., and Champion, P. D. (2000). Seedling emergence
from seed banks of 15 New Zealand lakes with contrasting vegetation histories.
Aquat. Bot. 66, 181–194. doi: 10.1016/S0304-3770(99)00074-1
Dorenbosch, M., and Bakker, E. S. (2011). Herbivory in omnivorous fishes:
effect of plant secondary metabolites and prey stoichiometry. Freshw. Biol. 56,
1783–1797. doi: 10.1111/j.1365-2427.2011.02618.x
Elger, A., and Willby, N. J. (2003). Leaf dry matter content as an integrative
expression of plant palatability: the case of freshwater macrophytes. Func. Ecol.
17, 58–65. doi: 10.1046/j.1365-2435.2003.00700.x
Elger, A., Willby, N. J., and Cabello-Martinez, M. (2009). Invertebrate
grazing during the regenerative phase affects the ultimate
structure of macrophyte communities. Freshw. Biol. 54, 1246–1255
doi: 10.1111/j.1365-2427.2009.02171.x
Gervais, F., Berger, S., Schönfelder, T., and Rusche, R. (1999). Basic limnological
characteristics of the shallow eutrophic lake Grimnitzsee (Brandenburg,
Germany). Limnologica 29, 105–119. doi: 10.1016/S0075-9511(99)80
058-9
Grutters, B. M. C., Gross, E. M., Van Donk, E., and Bakker, E. S. (2017). Periphyton
density is similar on native and non-native plant species. Freshw. Biol. 62,
906–915. doi: 10.1111/fwb.12911
Gulati, R. D., and Van Donk, E. (2002). Lakes in the Netherlands, their origin,
eutrophication and restoration: state-of-the-art review. Hydrobiologia 478,
73–106. doi: 10.1023/A:1021092427559
Haag, R. W. (1983). Emergence of seedlings of aquatic macrophytes from lake
sediments. Can. J. Bot. 61, 148–156. doi: 10.1139/b83-014
Hansson, L.-A., Annadotter,H., Bergman, E., Hamrin, S. F., Jeppesen, E., Kairesalo,
T., et al. (1998). Biomanipulation as an application of food-chain theory:
constraints, synthesis, and recommendations for temperate lakes. Ecosystems
1, 558–574. doi: 10.1007/s100219900051
Hansson, L.-A., Johansson, L., and Persson, L. (1987). Effects of fish grazing on
nutrient release and succession of primary producers. Limnol. Oceanogr. 32,
723–729. doi: 10.4319/lo.1987.32.3.0723
Hansson, L.-A., Nicolle, A., Brönmark, C., Hargeby, A., Lindström, Å., and
Andersson, G. (2010). Waterfowl, macrophytes, and the clear water state of
shallow lakes. Hydrobiologia 646, 101–109. doi: 10.1007/s10750-010-0169-z
Hargeby, A., Andersson, G., Blindow, I., and Johansson, S. (1994). Trophic
web structure in a shallow eutrophic lake during a dominance shift
from phytoplankton to submerged macrophytes. Hydrobiologia 279, 83–90.
doi: 10.1007/BF00027843
Heinzel, K., and Martin, C. (2006). Monitoring der Qualitätskomponente
Makrophyten für WRRL und FFH-RL in schleswig-holsteinischen Seen;
Gutachten im Auftrag des Landesamt für Natur und Umwelt des Landes
Schleswig-Holstein. Available online at: http://www.umweltdaten.landsh.de/
nuis/wafis/seen/Berichte_Gutachten/Ufer_Unterwasservegetation/Bericht_
Makrophyten_2006_WRRL_Heinzel_GFN.pdf.
Hidding, B., Bakker, E. S., Hootsmans, M. J. M., and Hilt, S. (2016). Synergy
between shading and herbivory triggers plant loss and regime shifts in aquatic
systems. Oikos 125, 1489–1495. doi: 10.1111/oik.03104
Hidding, B., Bakker, E. S., Keuper, F., De Boer, T., De Vries, P. P., and Nolet,
B. A. (2010). Differences in tolerance of pondweeds and charophytes to
vertebrate herbivores in a shallow Baltic estuary. Aquat. Bot. 93, 123–128.
doi: 10.1016/j.aquabot.2010.04.002
Hilt, S., Adrian, R., Köhler, J., Monaghan, M. T., and Sayer, C. (2013). Clear,
crashing, turbid and back long-term changes of macrophyte assemblages in a
shallow lake. Freshw. Biol. 58, 2027–2036. doi: 10.1111/fwb.12188
Hilt, S., Brothers, S., Jeppesen, E., Veraart, A., and Kosten, S. (2017). Translating
regime shifts in shallow lakes into changes in ecosystem functions and services.
Bioscience 67, 928–936. doi: 10.1093/biosci/bix106
Hilt, S., and Gross, E. M. (2008). Can allelopathically active submerged
macrophytes stabilise clear-water states in shallow eutrophic lakes? Bas. Appl.
Ecol. 9, 422–432. doi: 10.1016/j.baae.2007.04.003
Hilt, S., Gross, E. M., Hupfer, M., Morscheid, H., Mählmann, J., Melzer, A.,
et al. (2006). Restoration of submerged vegetation in shallow eutrophic lakes
guideline and state of the art in Germany. Limnologica 36, 155–171.
doi: 10.1016/j.limno.2006.06.001
Hilt, S., and Grünert, U. (2008). Praxistest zur Bewertung von Makrophyten
in Berliner Seen im Rahmen der Umsetzung der EU-Wasserrahmenrichtlinie.
Report for Senate (Berlin), 1–58.
Hilt, S., Van de Weyer, K., Köhler, A., and Chorus, I. (2010). Submerged
macrophyte responses to reduced phosphorus concentrations in two peri-
urban lakes. Rest. Ecol. 18, 452–461. doi: 10.1111/j.1526-100X.2009.00
577.x
Hoare, D., Phillips, G., and Perrow, M. (2008). Broads Lake Restoration Strategy.
Appendix 4. Review of Biomanipulaton. Norwich: Broads Authority, 1–59.
Hussner, A., Gross, E. M., Van de Weyer, K., and Hilt, S. (2014).
Handlungsempfehlung zur Abschätzung der Chancen einer Wiederansiedlung
von Wasserpflanzen bei der Restaurierung von Flachseen Deutschlands.
Arbeitsblatt: Deutsche Gesellschaft für Limnologie (DGL),
Idestam-Almquist, J., and Kautsky, L. (1995). Plastic responses in morphology
of Potamogeton pectinatus L. to sediment and above-sediment conditions
at two sites in the northern Baltic proper. Aquat. Bot. 52, 205–216.
doi: 10.1016/0304-3770(95)00499-8
Janse, J. H. (1997). A model of nutrient dynamics in shallow lakes in relation to
multiple stable states. Hydrobiologia 342, 1–8. doi: 10.1023/A:1017018812215
Janse, J. H. (2005). Model Studies on the Eutrophication of Shallow Lakes and
Ditches, Dissertation, Wageningen University.
Janse, J. H., De Senerpont Domis, L. N., Scheffer, M., Lijklema, L., Van
Liere, L., Klinge, M., et al. (2008). Critical phosphorus loading of
different types of shallow lakes and the consequences for management
estimated with the ecosystem model PCLake. Limnologica 38, 203–219.
doi: 10.1016/j.limno.2008.06.001
Janse, J. H., Scheffer, M., Lijklema, L., Van Liere, L., Sloot, J. S., and Mooij, W.
M. (2010). Estimating the critical phosphorus loading of shallow lakes with the
ecosystem model PCLake: sensitivity, calibration and uncertainty. Ecol. Model.
221, 654–665. doi: 10.1016/j.ecolmodel.2009.07.023
Frontiers in Plant Science | www.frontiersin.org 21 February 2018 | Volume 9 | Article 194
Hilt et al. Response of Macrophytes to Restoration
Janse, J. H., Van Donk, E., and Aldenberg, T. (1998). A model study on the stability
of the macrophyte-dominated state as affected by biological factors. Wat. Res.
32, 2696–2706. doi: 10.1016/S0043-1354(98)00049-9
Janssen, A. B. G., de Jager, V. C. L., Janse, J. H., Kong, X., Liu, S., Ye,
Q., et al. (2017). Spatial identification of critical nutrient loads of large
shallow lakes: implications for Lake Taihu (China). Wat. Res. 119, 276–287.
doi: 10.1016/j.watres.2017.04.045
Jeppesen, E., Kristensen, P., Jensen, J. P., Søndergaard, M., Mortensen, E., and
Lauridsen, T. (1991). Recovery resilience following a reduction in external
phosphorus loading of shallow, eutrophic Danish lakes: duration, regulating
factors and methods for overcoming resilience. Mem. Ist. Ital. Idrobiol. 48,
127–148.
Jeppesen, E., Søndergaard, M., Meerhoff, M., Lauridsen, T. L., and Jensen,
J. P. (2007a). Shallow lake restoration by nutrient loading reduction
some recent findings and challenges ahead. Hydrobiologia 584, 239–252.
doi: 10.1007/s10750-007-0596-7
Jeppesen, E., Meerhoff, M., Jacobsen, B. A., Hansen, R. S., Søndergaard, M.,
Jensen, J. P., et al. (2007b). Restoration of shallow lakes by nutrient control
and biomanipulation - the successful strategy varies with lake size and climate.
Hydrobiologia 581, 269–285. doi: 10.1007/s10750-006-0507-3
Jeppesen, E., Søndergaard, M., Jensen, J. P., Havens, K., Anneville, O., Carvalho,
L., et al. (2005). Lake responses to reduced nutrient loading an analysis of
contemporary data from 35 European and North American long term studies.
Freshw. Biol. 50, 1747–1771. doi: 10.1111/j.1365-2427.2005.01415.x
Jeppesen, E., Søndergaard, M., Lauridsen, T. L., Davidson, T. A., Liu, Z.,
Mazzeo, N., et al. (2012). Biomanipulation as a response tool to combat
eutrophication: recent advances and future challenges. Adv. Ecol. Res. 47,
411–488. doi: 10.1016/B978-0-12-398315-2.00006-5
Jones, J. I., and Sayer, C. D. (2003). Does fish-invertebrate-periphyton
cascade precipitate plant loss in shallow lakes? Ecology 84, 2155–2167.
doi: 10.1890/02-0422
Jones, J. I., Young, J. O., Eaton, J. W., and Moss, B. (2002). The influence of
nutrient loading, dissolved inorganic carbon and higher trophic levels on
the interaction between submerged plants and periphyton. J. Ecol. 90, 12–24.
doi: 10.1046/j.0022-0477.2001.00620.x
Josefsson, M., (2011). NOBANIS - Invasive Species Fact Sheet - Elodea canadensis,
Elodea nuttallii and Elodea callitrichoides NOBANIS. Available online at:
www.nobanis.org. Date of Access (Accessed Nov 28, 2012).
Kabus, T., Schumann, M., Weiß, G., and Kalhoff, M. (2007). Beitrag zur Flora der
Havel und angrenzender Flächen zwischen Potsdam und Pritzerbe. Verh. Bot.
Ver. Berlin Brandenburg 140, 25–55.
Kabus, T., and Mauersberger, R., (2011). Liste und Rote Liste der
Armleuchteralgengewächse (Characeae) des Landes Brandenburg. Natursch.
Landschaftspfl. Brbg. 20, 1–32.
Kaplan, K., and Muer, T. (1990). Beobachtungen zum diasporenreservoir im
Bereich ehemaliger Heideweiher. Florist. Rundbriefe 24, 38–45.
Karatayev, A. Y., Burlakova, L. E., and Padilla, D. K. (2015). Zebra versus quagga
mussels: a review of their spread, population dynamics, and ecosystem impacts.
Hydrobiologia 746, 97–112. doi: 10.1007/s10750-014-1901-x
Kautsky, L. (1990). Seed and tuber banks of aquatic macrophytes in
thE Askö area, northern Baltic proper. Holarct. Ecol. 13, 143–148.
doi: 10.1111/j.1600-0587.1990.tb00600.x
Knösche, R. (2008). Wiederfund von Najas marina L. spp. marina im Schollener
See (Elbe-Havel-Winkel, Sachsen-Anhalt). Mitteil. Florist. Kart. Sachsen Anhalt
13, 41–51.
Körner, S. (2001). Development of submerged macrophytes in shallow
Lake Müggelsee (Berlin, Germany) before and after its switch to the
phytoplankton-dominated state. Arch. Hydrobiologia 152, 395–409.
doi: 10.1127/archiv-hydrobiol/152/2001/395
Körner, S. (2002). Loss of submerged macrophytes in shallow lakes in
North-Eastern Germany. Intern. Rev. Hydrobiologia 87, 375–384.
doi: 10.1002/1522-2632(200207)87:4&lt;375::AID-IROH375&gt;3.0.CO;2-7
Körner, S., and Dugdale, T. (2003). Is roach herbivory preventing re-colonization
of a shallow lake with submerged macrophytes? Hydrobiologia 506, 497–501.
doi: 10.1023/B:HYDR.0000008561.67513.ec
Kosten, S., Huszar, V. L. M., Bécares, E., Costa, L., Van Donk, E., Hansson, L.-A.,
et al. (2012). Warmer climates boost cyanobacterial dominance in shallow lakes.
Glob. Chang. Biol. 18, 118–126. doi: 10.1111/j.1365-2486.2011.02488.x
Kuiper, J. J. (2016). Making Eco Logic and Models work: An Integrative Approach to
Lake Ecosystem Modelling, Dissertation, Wageningen University.
Kuiper, J. J., Verhofstad, M. J., Louwers, E. L., Bakker, E. S., Brederveld, R. J., Van
Gerven, L. P. A., et al. (2017). Mowing submerged macrophytes in shallow
lakes with alternative stable states: battling the good guys? Environ. Manage.
59, 619–634. doi: 10.1007/s00267-016-0811-2
Lage, S., Annadotter, H., Rasmussen, U., and Rydberg, S. (2015). Biotransfer of
β-N-methylamino-L-alanine (BMAA) in a eutrophicated freshwater lake. Mar.
Drugs 13, 1185–1201 doi: 10.3390/md13031185
Lalonde, S., and Downing, J. A. (1991). Epiphyton biomass is related to lake
trophic status, depth, and macrophyte architecture. Can. J. Fish. Aquat. Sci. 48,
2285–2291. doi: 10.1139/f91-268
Landesamt für Natur und Umwelt des Landes Schleswig-Holstein (1997). Zustand
und Belastungsquellen des Warder Sees. Flintbek: Landesamt für Natur und
Umwelt des Landes Schleswig-Holstein.
Lauridsen, T. L., Jensen, J. P., Jeppesen, E., and Sondergaard, M. (2003a).
Response of submerged macrophytes in Danish lakes to nutrient
loading reductions and biomanipulation. Hydrobiologia 506, 641–649.
doi: 10.1023/B:HYDR.0000008633.17385.70
Lauridsen, T. L., Sandsten, H., and Hald Møller, P. (2003b). Restoration
of a shallow lake by introducing Potamogeton spp.: impact of waterfowl
grazing. Lakes Reserv. Res. Manage. 8, 177–187. doi: 10.1111/j.1440-1770.2003.
00224.x
Lauridsen, T., Pedersen, L. J., Jeppesen, E., and Søndergaard, M. (1996).
The importance of macrophyte bed size for composition and horizontal
migration of cladocerans in a shallow lake. J. Plankt. Res. 18, 2283–2294.
doi: 10.1093/plankt/18.12.2283
Lillie, R. A., Budd, J., and Rasmussen, P. W. (1997). Spatial and temporal variability
in biomass density of Myriophyllum spicatum L. in a northern temperate lake.
Hydrobiologia 347, 69–74. doi: 10.1023/A:1003059000277
Madsen, J. D. (1991). Resource allocation at the individual plant level. Aquat. Bot.
41, 67–86. doi: 10.1016/0304-3770(91)90039-8
Madsen, J. D., and Adams, M. S. (1988). The germination of Potamogeton
pectinatus tubers: environmental control by temperature and light. Can. J. Bot.
66, 2523–2526. doi: 10.1139/b88-343
Mathes, J. (2007). Erfahrungen mit Seentherapien in Mecklenburg-Vorpommern.
Available online at: http://www.lake-jewel.net/download/Dateien/
Seesanierung%20Mecklenburg-Vorpommern.pdf
Mauersberger, H., and Mauersberger, R. (1996). Die Seen im Biosphärenreservat
Schorfheide-Chorin: eine ökologische Studie. Dissertation, Universität
Greifswald.
Meijer, M. L., De Boois, I., Scheffer, M., Portielje, R., and Hosper, H. (1999).
Biomanipulation in shallow lakes in The Netherlands: an evaluation of 18 case
studies. Hydrobiologia 408, 13–30. doi: 10.1023/A:1017045518813
Meijer, M. L., and Hosper, H. (1997). Effects of biomanipulation in the large
and shallow Lake Wolderwijd, The Netherlands. Hydrobiologia 342, 335–349.
doi: 10.1023/A:1017088701621
Mooij, W. M., Brederveld, R. J., De Klein, J. J. M., DeAngelis, D. L., Downing,
A. S., Faber, M., et al. (2014). Serving many at once: how a database approach
can create unity in dynamical ecosystem modelling. Environ. Mod. Softw. 61,
266–273. doi: 10.1016/j.envsoft.2014.04.004
Mooij, W. M., Janse, J. H., De Senerpont Domis, L. N., Hülsmann, S., and
Ibelings, B. W. (2007). Predicting the effect of climate change on temperate
shallow lakes with the ecosystem model PCLake. Hydrobiologia 584, 443–454.
doi: 10.1007/s10750-007-0600-2
Moss, B., Balls, H., Irvine, K., and Stansfield, J. (1986). Restoration of two lowland
lakes by isolation from nutrient-rich water sources with and without removal
of sediment. J. Appl. Ecol. 23, 391–414. doi: 10.2307/2404025
Moss, B., Kosten, S., Meerhoff, M., Battarbee, R. W., Jeppesen, E., Mazzeo, N.,
et al. (2011). Allied attack: climate change and nutrient pollution. Inland Wat.
1, 101–105. doi: 10.5268/IW-1.2.359
Moss, B., Stansfield, J., and Irvine, K. (1990). Problems in the restoration of a
hypertrophic lake by diversion of a nutrient-rich inflow. Verh. Intern. Ver.
Limnol. 24, 568–572. doi: 10.1080/03680770.1989.11898801
Nixdorf, B., Rücker, J., Dolman, A. M., Wiedner, C., Hilt, S., Kasprzak, P., et al.
(2013). Prozessverständnis als Grundlage für die Gewässerbewirtschaftung
Fallbeispiele für Limitation, Konkurrenz, Gewässerstruktur und
Nahrungsnetzsteuerung. Korr. Wasserwirtschaft 60, 693–701.
Frontiers in Plant Science | www.frontiersin.org 22 February 2018 | Volume 9 | Article 194
Hilt et al. Response of Macrophytes to Restoration
Niedersächsischer Landesbetrieb für Wasserwirtschaft, Küsten- und Naturschutz,
NLWKN. (2011). Leitfaden Maßnahmenplanung Oberflächengewässer.
Wasserrahmenrichtlinie Band 3 Steinhuder Meer. 1–33.
Noordhuis, R., Van Zuidam, B. G., Peeters, E. T. H. M., and Van Geest, G. J.
(2016). Further improvements in water quality of the Dutch Border lakes:
two types of clear states at different nutrient levels. Aquatic Ecol. 50, 521–539.
doi: 10.1007/s10452-015-9521-8
Perrow, M. R., Schutten, J. H., Howes, J. R., Holzer, T., Madgwick, F. J., and
Jowitt, A. J. D. (1997). Interactions between coot (Fulica atra) and submerged
macrophytes: the role of birds in the restoration process. Hydrobiologia 342,
241–255. doi: 10.1023/A:1017007911190
Phillips, G., Bennion, H., Perrow, M., Sayer, C. D., Spears, B., and Willby, N.
(2015). A Review of Lake Restoration Practices and their Performance in the
Broads National Park, 1980-2013. Report for the Broads Authority, Norwich
and Natural England. Available online at: http://www.broads-authority.gov.uk/
news-and-publications/publications-and- reports/conservation- publications-
and-reports/water-conservation-reports
Phillips, G., Kelly, A., Pitt, J. A., Sanderson, R., and Taylor, E. (2005).
The recovery of a very shallow eutrophic lake, 20 years after the
control of effluent derived phosphorus. Freshw. Biol. 50, 1628–1638.
doi: 10.1111/j.1365-2427.2005.01434.x
Phillips, G. L., Willby, N., and Moss, B. (2016). Submerged macrophyte decline in
shallow lakes: what have we learnt in the last forty years? Aquat. Bot. 135, 37–45.
doi: 10.1016/j.aquabot.2016.04.004
Portielje, R., and Van der Molen, D. T. (1999). Relationships between
eutrophication variables: from nutrient loading to transparency. Hydrobiologia
408, 375–387. doi: 10.1023/A:1017090931476
Pot, R., and Ter Heerdt, G. N. J. (2014). Succession dynamics of aquatic
lake vegetation after restoration measures: increased stability after 6 years
of development. Hydrobiologia 737, 333–345. doi: 10.1007/s10750-014-
1835-3
Riegman, R. (2007). Evaluatie Rapport Vijftien jaar Actief Biologisch Beheer (ABB)
in het Duinigermeer (in Dutch). Reest aen Wieden Waterschap.
Rip, W. J., Ouboter, R. L., and Los, H. L. (2007). Impact of climatic fluctuations
on Characeae biomass in a shallow, restored lake in The Netherlands.
Hydrobiologia 584, 415–424. doi: 10.1007/s10750-007-0608-7
Roberts, E., Kroker, J., Körner, S., and Nicklisch, A. (2003). The role of periphyton
during the re-colonization of a shallow lake with submerged macrophytes.
Hydrobiologia 506, 525–530. doi: 10.1023/B:HYDR.0000008560.73832.1c
Rodrigo, M. A., Rojo, C., Alonso-Guillén, J. L., and Vera, P. (2013). Restoration
of two small Mediterranean lagoons: the dynamics of submerged macrophytes
and factors that affect the success of revegetation. Ecol. Eng. 54, 1–15.
doi: 10.1016/j.ecoleng.2013.01.022
Rørslett, B., Berge, D., and Johansen, S. W. (1985). Mass invasion
of Elodea canadensis in a mesotrophic, South Norwegian lake
impact of water quality. Verh. Intern. Ver. Limnol. 22, 2920–2926.
doi: 10.1080/03680770.1983.11897803
Rücker, J., Barsch, A., and Nixdorf, B. (2015). Besser, aber noch nicht gut“
Ökologischer Zustand der Seen in Brandenburg 2014. Wasserwirtschaft 12,
41–47.
Sand-Jensen, K., Pedersen, N. L., Thorsgaard, I., Moeslund, B., Borum, J., and
Brodersen, K. P. (2008). 100 years of vegetation decline and recovery in Lake
Fure, Denmark. J. Ecol. 96, 206–271. doi: 10.1111/j.1365-2745.2007.01339.x
Sayer, C. D., Burgess, A., Kari, K., Davidson, T. A., and Rose, N. (2010a). Long-
term dynamics of submerged macrophytes and algae in a small and shallow,
eutrophic lake: implications for the stability of macrophyte-dominance. Freshw.
Biol. 55, 565–583. doi: 10.1111/j.1365-2427.2009.02353.x
Sayer, C. D., Davidson, T. A., and Jones, J. I. (2010b). Seasonal dynamics
of macrophytes and phytoplankton in shallow lakes: a eutrophication-
driven pathway from plants to plankton? Freshw. Biol. 55, 500–513.
doi: 10.1111/j.1365-2427.2009.02365.x
Sayer, C. D., Davidson, T. A., Rawcliffe, R., Langdon, P. G., Leavitt, P. R.,
Cockerton, G., et al. (2016). Consequences of fish kills for long-term trophic
structure in shallow lakes: implications for theory and restoration. Ecosystems
16, 1289–1309. doi: 10.1007/s10021-016-0005-z
Scheffer, M., Barrett, S., Carpenter, S. R., Folke, C., Green, A. J., Holmgren, M.,
et al. (2015). Creating a safe operating space for iconic ecosystems. Science 347,
1317–1319. doi: 10.1126/science.aaa3769
Scheffer, M., and Carpenter, S. R. (2003). Catastrophic regime shifts in
ecosystems: linking theory to observation. Trends Ecol. Evol. 18, 648–656.
doi: 10.1016/j.tree.2003.09.002
Scheffer, M., De Redelijkheid, M. R., and Noppert, F. (1992). Distribution and
dynamics of submerged vegetation in a chain of shallow eutrophic lakes. Aquat.
Bot. 42, 199–216. doi: 10.1016/0304-3770(92)90022-B
Scheffer, M., Hosper, S. H., Meijer, M. L., Moss, B., and Jeppesen, E. (1993).
Alternative equilibria in shallow lakes. Trends Ecol. Evol. 8, 275–279.
doi: 10.1016/0169-5347(93)90254-M
Schutten, J., Dainty, J., and Davy, A. J. (2005). Root anchorage and its
significance for submerged plants in shallow lakes. J. Ecol. 93, 556–571.
doi: 10.1111/j.1365-2745.2005.00980.x
Simberloff, D., and Gibbons, L. (2004). Now you see them, now you don’t!
population crashes of established introduced species. Biol. Invasions 6, 161–172.
doi: 10.1023/B:BINV.0000022133.49752.46
Simpson, D. A. (1990). Displacement of Elodea canadensis Michx. by Elodea
nuttallii (Planch.) St. John in the British Isles. Watsonia 18, 173–177.
Sommer, U., Adrian, R., De Senerpont Domis, L. N., Elser, J. J., Gaedke, U.,
Ibelings, B. W., et al. (2012). Beyond the Plankton Ecology Group (PEG)
Model: mechanisms driving plankton succession. Ann. Rev. Ecol. Evol. Syst. 43,
429–448. doi: 10.1146/annurev-ecolsys-110411-160251
Søndergaard, M., Bjerring, R., and Jeppesen, E. (2013). Persistent internal
phosphorus loading in shallow eutrophic lakes. Hydrobiologia 710, 95–107.
doi: 10.1007/s10750-012-1091-3
Søndergaard, M., Larsen, S. E., Johansson, L. S., Lauridsen, T. L., and Jeppesen,
E. (2016). Ecological classification of lakes: uncertainty and the influence of
year-to-year variability. Ecol. Indic. 61, 248–257. doi: 10.1016/j.ecolind.2015.
09.024
Søndergaard, M., Lauridsen, T. L., Johansson, L. S., and Jeppesen, E. (2017).
Repeated fish removal to restore lakes: case study of Lake Væng, Denmark-
two biomanipulations during 30 years of monitoring. Water 9, 43.
doi: 10.3390/w9010043
Søndergaard, M., Liboriussen, L., Pedersen, A. R., and Jeppesen, E. (2008). Lake
restoration by fish removal: short- and long-term effects in 36 Danish lakes.
Ecosystems 11, 1291–1305. doi: 10.1007/s10021-008-9193-5
Spencer, D. F. (1986). Early growth of Potamogeton pectinatus L. in response
to temperature and irradiance: morphology and pigment composition. Aquat.
Bot. 26, 1–8. doi: 10.1016/0304-3770(86)90002-1
Strand, J. (1999). The development of submerged macrophytes in
Lake Ringsjön after biomanipulation. Hydrobiologia 404, 113–121.
doi: 10.1023/A:1003728730563
Strand, J. A., and Weisner, S. E. B. (2001). Dynamics of submerged macrophyte
populations in response to biomanipulation. Freshw. Biol. 46, 1397–1408.
doi: 10.1046/j.1365-2427.2001.00746.x
Talling, J. F., and Parker, J. E. (2002). Seasonal dynamics of phytoplankton
and phytobenthos, and associated chemical interactions, in a shallow
upland lake (Malham Tarn, northern England). Hydrobiologia 487, 167–181.
doi: 10.1023/A:1022934705359
Tobiessen, P., and Snow, P. D. (1984). Temperature and light effects on the
growth of Potamogeton crispus in Collins Lake, New York State. Can. J. Bot.
62, 2822–2826. doi: 10.1139/b84-376
Triest, L., Stiers, I., and Van Onsem, S. (2016). Biomanipulation as a nature-
based solution to reduce cyanobacterial blooms. Aquat. Ecol. 50, 461–483.
doi: 10.1007/s10452-015-9548-x
Urrutia-Cordero, P., Ekvall, M. K., and Hansson, L. A. (2016). Local food web
management increases resilience and buffers against global change effects on
freshwaters. Sci. Rep. 6:29542. doi: 10.1038/srep29542
Vadeboncoeur, Y., Kalff, J., Christoffersen, K., and Jeppesen, E. (2006).
Substratum as a driver of variation in periphyton chlorophyll and
productivity in lakes. J. N. Am. Benth. Soc. 25, 379–392. doi: 10.1899/0887-
3593(2006)25[379:SAADOV]2.0.CO;2
Van Berkum, J. A., Klinge, M., and Grimm, M. P. (1995). Biomanipulation
in the duinigermeer, first results. Neth. J. Aquat. Ecol. 29, 81–90.
doi: 10.1007/BF02061791
Van de Bund, W. J., and Van Donk, E. (2002). Short-term and long-term
effects of zooplanktivorous fish removal in a shallow lake: a synthesis
of 15 years of data from Lake Zwemlust. Freshw. Biol. 47, 2380–2387.
doi: 10.1046/j.1365-2427.2002.01006.x
Frontiers in Plant Science | www.frontiersin.org 23 February 2018 | Volume 9 | Article 194
Hilt et al. Response of Macrophytes to Restoration
Van de Haterd, R. J. W., and Ter Heerdt, G. N. J. (2007). Potential for
the development of submerged macrophytes in eutrophicated shallow
peaty lakes after restoration measures. Hydrobiologia 584, 277–290.
doi: 10.1007/s10750-007-0593-x
Van den Berg, M. S., Scheffer, M., Coops, H., and Simons, J. (1998). The role of
characean algae in the management of eutrophic shallow lakes. J. Phycol. 34,
750–756. doi: 10.1046/j.1529-8817.1998.340750.x
Van den Berg, M. S., Scheffer, M., Van Nes, E., and Coops, H. (1999).
Dynamics and stability of Chara sp. and Potamogeton pectinatus in a
shallow lake changing in eutrophication level. Hydrobiologia 408, 335–342.
doi: 10.1023/A:1017074211970
Van de Weyer, K. (2011). Erfassung und Bewertung der Makrophytenvegetation in
der Unteren Havel und im Wannsee im Jahr 2011. Report for Senate (Berlin).
1–111.
Van Dijk, G. M. (1993). Dynamics and attenuation characteristics of periphyton
upon artificial substratum under various light conditions and some additional
observations on periphyton upon Potamogeton pectinatus L. Hydrobiologia 252,
143–161. doi: 10.1007/BF00008152
Van Dijk, G. M., and Janse, J. H. (1993). Modeling resource allocation in
Potamogeton pectinatus L. J. Aquat. Plant Manage. 31, 128–134.
Van Dijk, G. M., and Van Vierssen, W. (1991). Survival of a Potamogeton
pectinatus L.population under various light conditions in a shallow
eutrophic lake (Lake Veluwe) in The Netherlands. Aquat. Bot. 39, 121–129.
doi: 10.1016/0304-3770(91)90027-3
Van Vierssen, W. (1982a). The ecology of communities dominated by
Zannichellia taxa in Western Europe. 1. Characterization and autecology of
the Zannichellia taxa. Aquat. Bot. 12, 103–155. doi: 10.1016/0304-3770(82)90
010-9
Van Vierssen, W. (1982b). The ecology of communities dominated by Zannichellia
taxa in Western Europe. II. Distribution, synecology and productivity
aspects in relation to environmental factors. Aquat. Bot. 13, 385–483.
doi: 10.1016/0304-3770(82)90073-0
Van Vierssen, W. (1982c). Some notes on the germination of seeds of
Najas marina L. Aquat. Bot. 12, 201–203. doi: 10.1016/0304-3770(82)
90015-8
Van Wijk, R. J. (1988). Ecological studies on Potamogeton pectinatus L. I. General
characteristics, biomass production and life cycles under field conditions.
Aquat. Bot. 31, 211–258. doi: 10.1016/0304-3770(88)90015-0
Van Wijk, R. J. (1989). Ecological studies on Potamogeton pectinatus L. 3.
Reproductive strategies and germination ecology. Aquat. Bot. 33, 271–299.
doi: 10.1016/0304-3770(89)90042-9
Vari, A., and Toth, V. R. (2017). Quantifying macrophyte colonisation strategies -
A field experiment in a shallow lake (Lake Balaton, Hungary). Aquat. Bot. 136,
56–60. doi: 10.1016/j.aquabot.2016.09.006
Verhofstad, M. J. J. M., Alirangues Núñez, M. M., Reichman, E. P., van Donk, E.,
Lamers, L. P. M., and Bakker, E. S. (2017). Mass development of monospecific
submerged macrophyte vegetation after the restoration of shallow lakes: roles of
light, sediment nutrient levels, and propagule density. Aquat. Bot. 141, 29–38.
doi: 10.1016/j.aquabot.2017.04.004
Verpoorter, C., Kutser, T., Seekell, D. A., and Tranvik, L. J. (2014). A global
inventory of lakes based on high-resolution satellite imagery. Geophys. Res. Lett.
41, 6396–6402. doi: 10.1002/2014GL060641
Wang, X., Qin, B., Gao, G., and Paerl, H. W. (2010). Nutrient enrichment and
selective predation by zooplankton promote Microcystis (Cyanobacteria) bloom
formation. J. Plankt. Res. 32, 457–470. doi: 10.1093/plankt/fbp143
Waterstraat, A. (2008). Die Fischfauna des Galenbecker Sees und seiner Zu- und
Abflüsse. In: 70 Jahre Naturschutzgebiet Galenbecker See. Staatliches Amt für
Umwelt und Natur Ueckermünde. 27–34.
Weisner, S. E., Strand, J. A., and Sandsten, H. (1997). Mechanisms regulating
abundance of submerged vegetation in shallow eutrophic lakes. Oecologia 109,
592–599. doi: 10.1007/s004420050121
Wolfer, S. R., and Straile, D. (2004). Spatio-temporal dynamics and plasticity
of clonal architecture in Potamogeton perfoliatus.Aquat. Bot. 78, 307–318.
doi: 10.1016/j.aquabot.2003.11.005
Woolf, T. E., and Madsen, J. D. (2003). Seasonal biomass and carbohydrate
allocation patterns in southern Minnesota curlyleaf pondweed populations. J.
Aquat. Plant Manage. 41, 113–118.
Zheng, G., Xu, R., Chang, X., Hilt, S., and Wu, C. (2013). Cyanobacteria
can allelopathically inhibit submerged macrophytes: effects of Microcystis
aeruginosa extracts and exudates on Potamogeton malaianus.Aquat. Bot. 109,
1–7. doi: 10.1016/j.aquabot.2013.02.004
Conflict of Interest Statement: The authors declare that the research was
conducted in the absence of any commercial or financial relationships that could
be construed as a potential conflict of interest.
The handling Editor is currently co-organizing a Research Topic with one of
the authors EB, and confirms the absence of any other collaboration.
Copyright © 2018 Hilt, Alirangues Nuñez, Bakker, Blindow, Davidson, Gillefalk,
Hansson, Janse, Janssen, Jeppesen, Kabus, Kelly, Köhler, Lauridsen, Mooij,
Noordhuis, Phillips, Rücker, Schuster, Søndergaard, Teurlincx, van de Weyer, van
Donk, Waterstraat, Willby and Sayer. This is an open-access article distributed
under the terms of the Creative Commons Attribution License (CC BY). The use,
distribution or reproduction in other forums is permitted, provided the original
author(s) and the copyright owner are credited and that the original publication
in this journal is cited, in accordance with accepted academic practice. No use,
distribution or reproduction is permitted which does not comply with these terms.
Frontiers in Plant Science | www.frontiersin.org 24 February 2018 | Volume 9 | Article 194
... The community composition of epiphyton biofilms is known to change with season and may be specific to their macrophyte substrate, showing mutualistic, commensal, and competitive relationships with their substrate as they interact for resources or engage in chemical communication (Wijewardene et al. 2022). Overall, epiphyton has been found to be decisive for macrophyte decline under increasing nutrient loading (Phillips, Eminson, and Moss 1978;Phillips, Willby, and Moss 2016;Jones et al. 2002;Jones and Sayer 2003), can delay macrophyte recovery even when nutrient loading has been reduced (Roberts et al. 2003;Hilt et al. 2018), and affect macrophyte sensitivity to herbivory (Hidding et al. 2016). ...
... Most submerged macrophyte species in temperate lakes follow an annual life cycle with storage of energy in bulbs and rhizomes and associated loss of leafy parts (the substrate for epiphyton) at the end of the growing season. In spring, macrophyte stems and leaves emerge again and often are immediately colonized by epiphyton (Hilt et al. 2018). Field data of epiphyton were based on weekly, fortnightly, or monthly sampling of epiphyton biomass from artificial substrates (plastic or glass) exposed in the lakes in early spring mimicking the start of the growing season of submerged macrophytes, or from submerged macrophytes or underwater reed stems (Phragmites australis). ...
... As our modeled lake is shallow, we did not invoke the option to model stratification and the hypolimnion. In PCLake+, the macrophyte growth season was initiated through temperature increase in spring (Madsen and Adams 1988;Van Wijk 1989), whereas the onset of winter storage (root allocation and shoot die-off) was defined by an integrated light threshold (Hilt et al. 2018;Janssen et al. 2019). Summer biomass of most temperate region macrophytes carries over into the following growth period in the form of winter storage rather than overwintering leafy biomass (Hilt et al. 2018). ...
Article
Full-text available
Submerged macrophytes are key components in many freshwater and marine ecosystems, contributing to ecosystem functions and services. In temperate shallow lakes, spring epiphyton shading can be decisive for submerged macrophyte development, potentially leading to macrophyte collapse and a shift to undesired, turbid conditions. Global change can alter epiphyton phenology; however, the consequences for submerged macrophytes and their stabilizing effects on clear‐water conditions remain to be elucidated. Based on field data, we propose a general epiphyton shading phenology for submerged macrophytes in temperate shallow lake ecosystems. We express the temporal dynamics of epiphyton shading in terms of onset and relative increase (slope) of epiphyton development as well as epiphyton grazing impacts (onset, duration) using a Boltzmann function. This function is added to the ecosystem model PCLake+ as a customizable, macrophyte‐specific shading factor. We then assess how changes in the epiphyton phenology and the presence of grazing on epiphyton affects submerged macrophyte biomass in a generic temperate shallow model lake under control and warm winter scenarios. The results from the model provide a proof‐of‐concept that epiphyton shading can provoke macrophyte loss and shifts between alternative equilibria. Threshold values for critical shifts depend on epiphyton shading phenology. Earlier onset and longer duration of grazing can maintain macrophytes in nutrient or climate conditions under which they would otherwise collapse. Our results show the pivotal importance of epiphyton phenology in determining lake ecosystem‐wide responses stressing the need for better incorporation of epiphyton into both models and monitoring.
... Submerged macrophytes play a key role in the functioning of shallow lakes (Carpenter and Lodge 1986;Jeppesen et al. 1998;Hilt et al. 2018). They may be abundant and cover a large part of the bottom in oligomesotrophic shallow lakes, whereas their abundance is low in eutrophic-hypertrophic lakes where the growth conditions are poor because of turbid water and poor light conditions (Mitchell 1989;Scheffer et al. 1993;Sayer et al. 2010). ...
... Besides, macrophytes form habitats that provide refuge for invertebrates and vertebrates and substrate for periphytic organisms (Jeppesen et al. 1998;Scheffer 2001;Pełechaty 2004;Messyasz and Kuczynska-Kippen 2006). These functional features of macrophytes can help to maintain a clear-water state of temperate shallow lakes (Moss 1990;Scheffer et al. 1993;Jeppesen et al. 1998;Hilt et al. 2018). Consequently, submerged macrophytes are used to counteract eutrophication after reduction of external nutrient loading, and the long-term success of restoration of eutrophic shallow lakes depends critically on stable recovery of submerged macrophytes (Hilt et al. 2006). ...
... growth rate 0.039 day −1 ; Fu and He 2015). Light intensity is considered the main limiting factor for the growth, biomass, and distribution of submerged macrophytes under eutrophic conditions (Jupp and Spence 1977;Franklin 2008;Hilt et al. 2018). Accordingly, we found that the total biomass of both species was negatively correlated with the biomass of phytoplankton (as Chl a) and positively with transparency (Table 2). ...
Article
Full-text available
Shallow lakes are often rich in submerged macrophytes that play an important role in the functioning of shallow lake ecosystems. Omnivorous fish can intensify eutrophication by enhancing the growth of phytoplankton, leading to higher water turbidity and light attenuation with subsequent negative effects on the growth of submerged macrophytes. However, the effects of different omnivorous fish species on aquatic ecosystem are not necessarily the same. A 70-day outdoor experiment was conducted in mesocosms containing meadow-forming arrowhead (Sagittaria sagittifolia Linnaeus, 1753) and canopy-forming clasping leaf pondweed (Potamogeton perfoliatus Linnaeus, 1753) to test the hypothesis that two different species of omnivorous fish, the native species rudd (Scardinius erythrophthalmus (Linnaeus, 1758) and the invasive Prussian carp (Carassius auratus (Linnaeus, 1758), have different effects on plankton and benthic algae communities, water quality, and submerged macrophyte growth. Both fish species increased the phytoplankton biomass and the concentrations of total and inorganic suspended solids and decreased water transparency, with Prussian carp having a stronger effect than rudd. The fish also changed the phytoplankton composition and decreased the densities of zooplankton, except for rotifers that were most pronouncedly impacted by Prussian carp. Rudd decreased the dry weight of the above-ground parts of arrowhead and pondweed roots, while Prussian carp decreased the biomass of the whole pondweed. Canonical analyses showed that the impact of all studied factors on pondweed growth was 2.6 times greater in the Prussian carp treatments. Our findings thus showed differential effects of the two omnivorous fish species on plankton communities, water quality, and submerged macrophytes, with Prussian carp having a stronger impact than rudd.
... Meanwhile, the reduction of nutrient levels in lakes can also create more opportunities for lake management, such as the implementation of endogenous passivation, vegetation restoration (J. Chen et al., 2023;Huisman et al., 2018;Sondergaard et al., 2017), and vegetation restoration (Hilt et al., 2018). Exogenous load reduction should combine humanities and natural sciences , such as improving public participation and strengthening the construction of pollution treatment facilities (SSP1) (Wei et al., 2023). ...
Article
Full-text available
Assessing future chlorophyll‐a (Chl‐a) changes is crucial for developing effective lake management programs. Future assessments should integrate socio‐economic and climatic factors, including extreme climate impacts. However, such assessments are lacking. This study developed a novel framework and integrated model to analyze Chl‐a response in Chaohu Lake to climate (e.g., precipitation, temperature, and wind) and socio‐economic (e.g., population, fertilizer application, and livestock farming). The results indicated that by around 2050, the Chl‐a concentration would change by −6.5% to −0.1% compared to around 2020. While socio‐economic and climatic factors significantly altered nitrogen and phosphorus loading, exogenous loading had a minimal effect on Chl‐a due to high endogenous releases. The decrease in Chl‐a is mainly due to the increase in precipitation. Chl‐a reduction was primarily driven by increased precipitation (0.6%–12.9%), leading to −9.4% to −4.4% Chl‐a changes. The runoff increase also provided an opportunity to treat the heavily polluted lake area. Air temperature increases of 5.1%–9.2% resulted in Chl‐a increases of 0.9%–3.1%. On annual scales the effects of precipitation and temperature were mainly due to changes in mean values. The effects of variability were significant at seasonal scales, for example, lower spring temperatures favored a decrease in summer Chl‐a. There was uncertainty about the impact of future wind speeds. This study emphasized the importance of comprehensively quantifying the impacts of external pressures in lake water quality assessment and provided a reference and pathway support for the assessment and management of large, shallow lakes.
... To address these challenges, this study focused on Potamogeton crispus, a globally distributed submerged aquatic plant that is critical for nutrient absorption (Cao et al. 2009;Li et al. 2015). Its brief emergence phase in late spring provides a valuable window for remote sensing extraction (Chi 2009;Hilt et al. 2018). High spatial and temporal resolution Sentinel-2 imagery was used to analyze the phenological characteristics of Potamogeton crispus across eight lakes. ...
Article
Full-text available
Submerged aquatic vegetation (SAV), particularly Potamogeton crispus (P.C.), plays a critical role in lake ecosystems by supporting nutrient cycling and sediment stabilization. Accurate and large-scale monitoring of SAVs is essential for effective ecosystem management and conservation. However, traditional classification methods struggle with large-scale applications because of their reliance on extensive sample data and the influence of regional phenological variations, which limits their generalizability. This study introduces the Potamogeton crispus index (PCI), a novel approach developed using time-series Sentinel-2 imagery processed on the Google Earth Engine (GEE) platform. The PCI automatically captures the differences between P.C. and other aquatic vegetation during the emergence and senescence phases by utilizing maximum normalized difference vegetation index (NDVI) composites and the Otsu thresholding method, enabling large-scale extraction. Four representative lakes with significant P.C. presence were selected as study areas. The results of phenological-stage identification revealed that in Baoan Lake, the P.C. completely died at the earliest (DOY = 160), whereas in Baiyang Lake, the P.C. persisted the longest (DOY = 189). The field and high-resolution satellite image validation results demonstrated overall classification accuracies ranging from 89.9% to 95.22%, with kappa coefficients between 0.85 and 0.92. Compared with alternative methods, including the floating-leaf vegetation sensitivity index (FVSI), submerged vegetation sensitivity index (SVSI), modified normalized difference water index (MNDWI), and NDVI, the PCI achieved a 5% to 18% improvement in mapping accuracy. Further application across four additional lakes with diverse geographic and climatic conditions confirmed the robustness and adaptability of the proposed method. These findings underscore the PCI’s reliability and scalability for large-scale monitoring of P.C., demonstrating its potential for mapping other types of SAVs.
... Indeed, McLay (1974) hypothesized that the lowered growth rate of the floating minute duckweed Lemna perpusilla in zones of water with the submerged Stuckenia pectinata was attributable to higher pH in the water; probably as result of photosynthesis by the submerged plants. For hornwort (Ceratophyllum demersum), as one of the most abundant submerged species in small ponds and ditches in Europe (Birk and Willby 2010;Hilt et al. 2018;Szabó et al. 2024), high pH together with nitrogen limitation were found to be the most important growth-limiting factors maintaining its dominance over the free-floating duckweed (Szabó et al. 2022b). In laboratory experiments, hornwort-induced rise in alkalinity to pH 10 reduced the growth and tissue nitrogen and phosphorus content in duckweed (Szabó et al. 2022b). ...
Article
Full-text available
Submerged and free‐floating macrophytes are the two main vegetation types that can alternatively dominate in small shallow ponds and ditches. Submerged macrophytes provide more aquatic ecosystem services and cause fewer problems with anoxia and greenhouse gas emissions than free‐floating plants, but are inferior competitors for light. High pH values have been proposed as a contributor to submerged plant dominance, but threshold values for pH‐induced growth reduction of floating plants by submerged macrophytes are not known. In this study, we conducted laboratory experiments to test whether submerged plants ( Ceratophyllum demersum ) can effectively limit the growth of free‐floating competitors ( Lemna gibba ) by exceeding a critical threshold pH value. We used field data to test if and when such pH values can be achieved within dense submerged macrophyte stands. The pH values that caused 50% reduction in fresh weight‐ and chlorophyll‐based growth of L. gibba were 9.6 and 9.8, respectively, and the growth was negligible above pH 10.0. The photochemical efficiency of photosystem II and chlorophyll content of the floating plants declined rapidly above pH 9 and 9.5, respectively. Similarly, phosphorus concentration decreased and dry matter content increased sharply in L. gibba fronds above pH 9.5, respectively. We thus expect a critical pH threshold for L. gibba at around 9.5. Our mesocosm and field data show that the critical pH threshold for L. gibba growth is regularly exceeded within dense stands of bicarbonate‐using submerged macrophytes in lentic water‐bodies. Such conditions can be prevalent in the upper water layer during most of the day (12 h) in summer. The literature showed that many duckweed species (and other common free‐floating plants) in temperate and tropical water bodies show growth inhibition above pH 8. Therefore, high pH created by submerged macrophytes that use HCO 3 ⁻ in photosynthesis can be a major mechanism by which dominance over free‐floating competitors is maintained in densely vegetated lentic eutrophic water bodies. We propose that alternative stable states occur between free‐floating and submerged macrophytes along a pH gradient. Active management of pH may, thus, help in maintaining the desired state.
Article
Narmada River, one of the largest waterways in India, plays a critical role in supporting aquatic ecosystems. This study examines the relationship between macrophyte diversity and nitrogen levels across fifteen sampling stations (A1-A15) in the central India (Anuppur, Dindori, Mandla, and Jabalpur) during the pre-monsoon seasons. Three macrophyte indices like RMNI (River Macrophyte Nutrient Index), MIR (Macrophyte Index for Rivers), and RMHI (River Macrophyte Hydraulic Index) were employed to assess ecological conditions and water quality. A total of 20 macrophyte species from 14 families were identified, with Lemna minor, Potamogeton pectinatus, Polygonum persicaria, Ceratophyllum demersum, Persicum hydropiper, and Potamogenton natans being the most prevalent in nutrient-rich environments. The principal component analysis is conducted to identify the similarities in the variables and to extract the key factors that exhibit strong internal correlations while remaining distinct from other factors. The analysis revealed significant spatial variability across monitoring stations, with A-5, A-6, A-10, A-11, A-12, A-13, A-14, and A-15 exhibiting elevated orthophosphates concentration. Notably, station A5 showed moderate level of DO, BOD and COD, whereas sulphate was comparatively low. These findings underscore the heterogeneity of water quality conditions along the Narmada River and highlight the influence of localized anthropogenic inputs. The clustering analysis grouped stations on their ecological status, highlights variations in macrophytes indices and water quality parameters. This study highlights the macrophytes' utility as ecological indicators for assessing riverine nutrient dynamics and overall environmental health, and further emphasizing the need for site-specific interventions to mitigate pollution and enhance ecological resilience.
Article
While biomanipulation and chemical treatments have been applied to speed up the recovery of shallow eutrophic lakes through top-down and bottom-up effects, the efficacy of a combined approach has received less attention. We conducted a large-scale (4 ha) restoration experiment in an isolated part of Lake Yanglan, a shallow eutrophic lake in subtropical China. Here, lanthanum-modified bentonite and polyaluminium chloride were applied after fish removal, followed by transplantation of submerged macrophytes. Samples were collected from within the experimentally restored area and the unrestored area of the lake throughout the study period (August 2017 to May 2018), and data were compared for three periods: 1) fish removal period, 2) chemical treatment and macrophyte transplantation period, and 3) after full restoration. Compared to the unrestored area, water clarity (Secchi depth, SD) increased, total suspended solids and total nitrogen (TN) concentrations decreased significantly in the restored area across the whole study period, while total phosphorus (TP), particulate phosphorus, and soluble reactive phosphorus concentrations declined significantly in the second and third period. Phytoplankton biomass (chlorophyll α concentrations, Chl a) decreased significantly during fish removal in period 1 and after full restoration in period 3, but not in the intervening chemical treatment in period 2. After full restoration, mean SD had increased by a factor of 6.2 (from 29 to 181 cm), mean TN had decreased by 26 % (from 1.68 to 1.25 mg/L), TP by 72 % (from 0.18 to 0.05 mg/L), and Chl a by 78 % (from 49 to 11 μg/L) in the restored area compared to the control. The mobile phosphorus content of surface sediments significantly decreased after full restoration. The zooplankton to phytoplankton biomass ratio in the restored area increased after full restoration, peaking in March when Daphnia abundance was high, indicating enhanced grazing control on phytoplankton. However, the ratio was low in the warm months, likely due to fish recruitment that led to stronger predation on zooplankton. Our eight-month experiment showed that a clear-water state can be successfully restored using a combined approach of biomanipulation and chemical in a subtropical shallow lake. Yet, given that external and internal nutrient loading may increase and the zooplankton to phytoplankton biomass ratio decrease in summer due to the region's subtropical monsoon climate, maintaining a stable clear-water state in the long term may require additional measures such as periodic thinning of the fish stock and/or chemical treatment.
Article
Full-text available
Background and aim Sediments serve as the primary reservoir for nitrogen and phosphorus nutrients in lakes, and the release of these nutrients plays a crucial role in contributing to lake eutrophication. The restoration of submerged vegetation has emerged as a promising area of lake ecosystem research and is recognized as an effective method for managing eutrophic lakes. However, the validity of previous findings may be compromised by artificial experimental conditions and the brief duration of the studies, potentially underestimating the long-term effectiveness of submerged plants in restoring eutrophic lakes. Methods We restored a eutrophic lake through the reconstruction of submerged vegetation. Continuous sampling and monitoring of lake water and sediments were conducted throughout the project. By analyzing the spatiotemporal variations and factors driving nitrogen and phosphorus concentrations in the overlying water and sediments, we explored the effects of the restoration of submerged vegetation on nitrogen and phosphorus loading in the sediment‒water system. Results The restoration of submerged vegetation significantly improved the environmental conditions in overlying water in the lake by increasing transparency and dissolved oxygen levels while reducing nitrogen and phosphorus concentrations. Moreover, throughout their growth period, submerged plants could significantly reduce nitrogen and phosphorus loads in surface sediments (0–5 cm), particularly the concentrations of organic nitrogen (ON), ammonium nitrogen (NH4⁺-N), Ca-bound P (Ca-P), and P bound to Fe, Al and Mn oxides and hydroxides (Fe–P). Conclusion Our findings suggest that the restoration of submerged vegetation effectively controlled the nutrient loads of nitrogen and phosphorus in the overlying water and sediments of the lake. To further stabilize and extend the function of submerged plants in controlling internal nutrients and purifying water quality, it is necessary to strengthen the construction of a multi-seasonal and age-structured submerged plant community to promote the benign cycle of the lake ecosystem. Graphical abstract
Article
Full-text available
Aus Sicht des Arten- bzw. Natur- und Gewässerschutzes handelt es sich um eine wichtige Gruppe, da viele Arten der Armleuchter-algeneine hohe Bedeutung als Indikatoren der Gewässergüte haben und sich die Verbreitung mancher Arten auf nährstoffarme und somit allgemein gefährdete Gewässertypen beschränkt. Die von anderen Armleuchteralgen besiedelten Kleingewässer und temporären Gewässer sind im Zuge der Melioration in der Landwirtschaft selten geworden. Für das Land Brandenburg erschien 1993 eine erste Fassung der Roten Liste derArmleuchter- algen, die bereits alle heute für Brandenburg bekannten Arten ent- hielt und einstufte. Aufgrund des Erkenntnisfortschritts, aber auch aufgrund der veränderten Gefährdungssituation mancher Arten, ist inzwischen eine Neufassung der Roten Liste notwendig geworden, die hier vorgelegt wird und zugleich die aktuellen, veränderten Einstufungskriterien für Rote Listen berücksichtigt.
Article
Full-text available
The widespread loss of wetlands due to agricultural intensification has been highlighted as a major threat to aquatic biodiversity. However, all is not lost as we reveal that the propagules of some aquatic species could survive burial under agricultural fields in the sediments of ‘ghost ponds’ - ponds in-filled during agricultural land consolidation. Our experiments showed at least eight aquatic macrophyte species to germinate from seeds and oospores, following 50–150 years of dormancy in the sediments of ghost ponds. This represents a significant proportion of the expected macrophyte diversity for local farmland ponds, which typically support between 6 and 14 macrophyte species. The rapid (< 6 months) re-colonisation of resurrected ghost ponds by a diverse aquatic vegetation similarly suggests a strong seed-bank influence. Ghost ponds represent abundant, dormant time capsules for aquatic species in agricultural landscapes around the globe, affording opportunities for enhancing landscape-scale aquatic biodiversity and connectivity. While reports of biodiversity loss through agricultural intensification dominate conservation narratives, our study offers a rare positive message, demonstrating that aquatic organisms survive prolonged burial under intensively managed agricultural fields. We urge conservationists and policy makers to consider utilizing and restoring these valuable resources in biodiversity conservation schemes and in agri-environmental approaches and policies.
Article
Full-text available
Ongoing eutrophication frequently causes toxic phytoplankton blooms. This induces huge worldwide challenges for drinking water quality, food security and public health. Of crucial importance in avoiding and reducing blooms is to determine the maximum nutrient load ecosystems can absorb, while remaining in a good ecological state. These so called critical nutrient loads for lakes depend on the shape of the load-response curve. Due to spatial variation within lakes, load-response curves and therefore critical nutrient loads could vary throughout the lake. In this study we determine spatial patterns in critical nutrient loads for Lake Taihu (China) with a novel modelling approach called Spatial Ecosystem Bifurcation Analysis (SEBA). SEBA evaluates the impact of the lake's total external nutrient load on the local lake dynamics, resulting in a map of critical nutrient loads for different locations throughout the lake. Our analysis shows that the largest part of Lake Taihu follows a nonlinear load-response curve without hysteresis. The corresponding critical nutrient loads vary within the lake and depend on management goals, i.e. the maximum allowable chlorophyll concentration. According to our model, total nutrient loads need to be more than halved to reach chlorophyll-a concentrations of 30–40 μg L⁻¹ in most sections of the lake. To prevent phytoplankton blooms with 20 μg L⁻¹ chlorophyll-a throughout Lake Taihu, both phosphorus and nitrogen loads need a nearly 90% reduction. We conclude that our approach is of great value to determine critical nutrient loads of lake ecosystems such as Taihu and likely of spatially heterogeneous ecosystems in general.
Article
Full-text available
Non‐native plants increasingly dominate the vegetation in aquatic ecosystems and thrive in eutrophic conditions. In eutrophic conditions, submerged plants risk being overgrown by epiphytic algae; however, if non‐native plants are less susceptible to periphyton than natives, this would contribute to their dominance. Non‐native plants may differ from natives in their susceptibility to periphyton growth due to differences in nutrient release, allelopathy and architecture. Yet, there is mixed evidence for whether plants interact with periphyton growth through nutrient release and allelopathy, or whether plants are neutral so that only their architecture matters for periphyton growth. We hypothesised that (1) non‐native submerged vascular plants support lower periphyton density than native species, (2) native and non‐native species are not neutral substrate for periphyton and interact with periphyton and (3) periphyton density increases with the plant structural complexity of plant species. We conducted an experiment in a controlled climate chamber where we grew 11 aquatic plant species and an artificial plant analogue in monocultures in buckets. These buckets were inoculated with periphyton that was collected locally from plants and hard substrate. Of the 11 living species, seven are native to Europe and four are non‐native. The periphyton density on these plants was quantified after five weeks. We found that the periphyton density did not differ between non‐native and native plants and was not related to plant complexity. Three living plant species supported lower periphyton densities than the artificial plant, one supported a higher periphyton density and the other plants supported similar densities. However, there was a strong negative correlation between plant growth and periphyton density. We conclude that the periphyton density varies greatly among plant species, even when these were grown under similar conditions, but there was no indication that the interaction with periphyton differs between native and non‐native plant species. Hence, non‐native plants do not seem to benefit from reduced periphyton colonisation compared to native species. Instead, certain native and non‐native species tolerate eutrophic conditions well and as a consequence, they seem to host less periphyton than less tolerant species.
Article
Shallow lakes respond to nutrient loading reductions. Major findings in a recent multi-lake comparison of data from lakes with long time series revealed: that a new state of equilibrium was typically reached for phosphorus (P) after 10–15 years and for nitrogen (N) after <5–10 years; that the in-lake Total N:Total P and inorganic N:P ratios increased; that the phytoplankton and fish biomass often decreased; that the percentage of piscivores often increased as did the zooplankton:phytoplankton biomass ratio, the contribution of Daphnia to zooplankto biomass, and cladoceran size. This indicates that enhanced resource and predator control often interact during recovery from eutrophication. So far, focus has been directed at reducing external loading of P. However, one experimental study and cross-system analyses of data from many lakes in north temperate lakes indicate that nitrogen may play a more significant role for abundance and species richness of submerged plants than usually anticipated when total phosphorus is moderate high. According to the alternative states hypothesis we should expect ecological resistance to nutrient loading reduction and P hysteresis. We present results suggesting that the two alternative states are less stable than originally anticipated. How global warming affects the water clarity of shallow lakes is debatable. We suggest that water clarity often will decrease due to either enhanced growth of phytoplankton or, if submerged macrophytes are stimulated, by reduced capacity of these plants to maintain clear-water conditions. The latter is supported by a cross-system comparison of lakes in Florida and Denmark. The proportion of small fish might increase and we might see higher aggregation of fish within the vegetation (leading to loss of zooplankton refuges), more annual fish cohorts, more omnivorous feeding by fish and less specialist piscivory. Moreover, lakes may have prolonged growth seasons with a higher risk of long-lasting algal blooms and at places dense floating plant communities. The effects of global warming need to be taken into consideration by lake managers when setting future targets for critical loading, as these may well have to be adjusted in the future. Finally, we highlight some of the future challenges we see in lake restoration research.
Article
Background In recent decades, many attempts have been made to restore eutrophic lakes through biomanipulation. Reducing the populations of planktivorous and benthivorous fish (either directly or through stocking of piscivorous fish) may induce ecosystem changes that increase water transparency and decrease the risk of algal blooms and fish kills, at least in the short term. However, the generality of biomanipulation effects on water quality across lake types and geographical regions is not known. Therefore, we have undertaken a systematic review of such effects in eutrophic lakes in temperate regions throughout the world. Methods Searches for literature were made using online publication databases, search engines, specialist websites and bibliographies of literature reviews. Search terms were developed in English, Danish, Dutch and Swedish. Identified articles were screened for relevance using inclusion criteria set out in an a priori protocol. To reduce the risk of bias, we then critically appraised the combined evidence found on each biomanipulation. Data were extracted on outcomes such as Secchi depth and chlorophyll a concentration before, during and/or after manipulation, and on effect modifiers such as lake properties and amounts of fish removed or stocked. Results Our searches identified more than 14,500 articles. After screening for relevance, 233 of them remained. After exclusions based on critical appraisal, our evidence base included useful data on 128 biomanipulations in 123 lakes. Of these interventions, 85% had been made in Europe and 15% in North America. Meta-analysis showed that removal of planktivores and benthivores (with or without piscivore stocking) leads to increased Secchi depth and decreased chlorophyll a concentration during intervention and the first three years afterwards. Piscivore stocking alone has no significant effect. The response of chlorophyll a levels to biomanipulation is stronger in lakes where fish removal is intense, and in lakes which are small and/or have high pre-manipulation concentrations of total phosphorus. Conclusions Our review improves on previous reviews of biomanipulation in that we identified a large number of case studies from many parts of the world and used a consistent, repeatable process to screen them for relevance and susceptibility to bias. Our results indicate that removal of planktivorous and benthivorous fish is a useful means of improving water quality in eutrophic lakes. Biomanipulation tends to be particularly successful in relatively small lakes with short retention times and high phosphorus levels. More thorough fish removal increases the efficacy of biomanipulation. Nonetheless successes and failures have occurred across a wide range of conditions.
Article
Shallow lakes, the most prevalent type of freshwater ecosystems, can shift between clear states with macrophyte dominance and turbid, phytoplankton-dominated states. Such transformations, commonly termed regime shifts, have gained increasing attention in recent decades. Of 1084 studies documenting regime shifts, only 28% investigated the consequences for ecosystem functions and services such as habitat (13%), carbon processing (4%), or nutrient retention (4%). Although there is general consensus that a clear macrophyte state supports a higher diversity of aquatic organisms than a turbid one, the effects of shifts on primary production, carbon burial, greenhouse-gas emissions, and nutrient retention remain ambiguous. Shifts between the two states also affect drinking-water quality and the recreational value of lakes, leading to conflicting management measures and potentially deteriorating natural functions. We call for more comprehensive studies on the effects of regime shifts on ecosystem functions in shallow lakes to guide their sustainable management.
Article
After restoration, eutrophicated shallow freshwaters may show mass development of only one or two submerged macrophyte species, lowering biodiversity and hampering recreation. It is unclear which environmental factors govern this high percentage of the volume inhabited (PVI²) by submerged macrophytes, and whether the development of a more diverse, low canopy vegetation is likely to occur if dominant species decline in abundance. We hypothesized that (1) adequate light and high sediment nutrient availability leads to massive development of submerged macrophytes, and (2) that macrophyte species richness is low at high PVI, but that this is not caused by a lack of viable propagules of non-dominant species (especially charophytes). To test these hypotheses, fifteen shallow waters in the Netherlands were studied with respect to submerged vegetation (including propagules), water, and sediment characteristics. The probability of high submerged macrophyte PVI is highest in shallow waters where light availability in the water layer and phosphorus availability in the sediment are abundant. These conditions typically occur upon restoration of eutrophic waterbodies by reducing water nutrient loading or applying biomanipulation. Other factors, as top-down control, can additionally influence realised PVI. Viable propagules of species other than the dominant ones, including charophytes, were found in most of the sediments, indicating that once the dominant species declines, there is local potential for a diverse submerged vegetation to develop. Results can be used to predict when mass development occurs and to tackle the factors causing mass development.