+ All Categories
Home > Documents > Physical and chemical limnology of alpine lakes and pools ...llens/VAN HILDE/Literatuur... ·...

Physical and chemical limnology of alpine lakes and pools ...llens/VAN HILDE/Literatuur... ·...

Date post: 08-Jul-2020
Category:
Upload: others
View: 2 times
Download: 0 times
Share this document with a friend
23
PRIMARY RESEARCH PAPER Physical and chemical limnology of alpine lakes and pools in the Rwenzori Mountains (Uganda–DR Congo) Hilde Eggermont James M. Russell Georg Schettler Kay Van Damme Ilse Bessems Dirk Verschuren Received: 14 November 2006 / Revised: 29 March 2007 / Accepted: 18 April 2007 / Published online: 18 July 2007 ȑ Springer Science+Business Media B.V. 2007 Abstract This study describes the physical and chemical properties of 17 Afroalpine lakes (>2 m deep) and 11 pools (<2 m deep) in the Rwenzori mountains, Uganda-DR Congo, with the aim to establish the baseline conditions against which to evaluate future environmental and biological changes in these unique tropical ecosystems, and to provide the foundation for lake-based paleoenvironmental studies. Most Rwenzori lakes are located above 3,500 m elevation, and dilute (5–52 lS/cm specific conductance at 25ŶC) open systems with surface in- and outflow. Multivariate ordination and pairwise correlations between environmental variables mainly differ- entiate between (1) lakes located near or above 4,000 m (3,890–4,487 m), with at least some direct input of glacial meltwater and surrounded by rocky catchments or alpine vegetation; and (2) lakes located mostly below 4,000 m (2,990– 4,054 m), remote from glaciers and surrounded by Ericaceous vegetation and/or bogs. The former group are mildly acidic to neutral clear- water lakes (surface pH: 5.80–7.82; Secchi depth: 120–280 cm) with often above-average dissolved ion concentrations (18–52 lS/cm). These lakes are (ultra-) oligotrophic to mesotrophic (TP: 3.1– 12.4 lg/l; Chl-a: 0.3–10.9 lg/l) and phosphorus- limited (mass TN/TP: 22.9–81.4). The latter group are mildly to strongly acidic (pH: 4.30– 6.69) waters stained by dissolved organic carbon (DOC: 6.8–13.6 mg/l) and more modest trans- parency (Secchi-disk depth: 60–132 cm). Ratios of particulate carbon, particulate nitrogen and chlorophyll a in these lakes indicate that organic matter in suspension is primarily derived from the lakes’ catchments rather than aquatic pri- mary productivity. Since key features in the Rwenzori lakes’ abiotic environment are strongly tied to temperature and catchment hydrology, these Afroalpine lake ecosystems can be expected to respond sensitively to climate change and glacier melting. Handling editor: S. Declerck Electronic supplementary material The online version of this article (doi: 10.1007/s10750-007-0741-3) contains supplementary material, which is available to authorized users. H. Eggermont (&) Á K. Van Damme Á I. Bessems Á D. Verschuren Limnology Unit, Department of Biology, Ghent University, K. L. Ledeganckstraat 35, 9000 Gent, Belgium e-mail: [email protected] J. M. Russell Department of Geological Sciences, Brown University, BOX 1846, Providence, RI 02912, USA G. Schettler Section Climate Dynamics and Sediments, GeoForschungsZentrum Potsdam, Telegrafenberg C328, Potsdam 14473, Germany 123 Hydrobiologia (2007) 592:151–173 DOI 10.1007/s10750-007-0741-3
Transcript

PRIMARY RESEARCH PAPER

Physical and chemical limnology of alpine lakes and poolsin the Rwenzori Mountains (Uganda–DR Congo)

Hilde Eggermont Æ James M. Russell ÆGeorg Schettler Æ Kay Van Damme ÆIlse Bessems Æ Dirk Verschuren

Received: 14 November 2006 / Revised: 29 March 2007 / Accepted: 18 April 2007 / Published online: 18 July 2007� Springer Science+Business Media B.V. 2007

Abstract This study describes the physical and

chemical properties of 17 Afroalpine lakes (>2 m

deep) and 11 pools (<2 m deep) in the Rwenzori

mountains, Uganda-DR Congo, with the aim to

establish the baseline conditions against which to

evaluate future environmental and biological

changes in these unique tropical ecosystems,

and to provide the foundation for lake-based

paleoenvironmental studies. Most Rwenzori

lakes are located above 3,500 m elevation, and

dilute (5–52 lS/cm specific conductance at 25�C)

open systems with surface in- and outflow.

Multivariate ordination and pairwise correlations

between environmental variables mainly differ-

entiate between (1) lakes located near or above

4,000 m (3,890–4,487 m), with at least some

direct input of glacial meltwater and surrounded

by rocky catchments or alpine vegetation; and (2)

lakes located mostly below 4,000 m (2,990–

4,054 m), remote from glaciers and surrounded

by Ericaceous vegetation and/or bogs. The

former group are mildly acidic to neutral clear-

water lakes (surface pH: 5.80–7.82; Secchi depth:

120–280 cm) with often above-average dissolved

ion concentrations (18–52 lS/cm). These lakes

are (ultra-) oligotrophic to mesotrophic (TP: 3.1–

12.4 lg/l; Chl-a: 0.3–10.9 lg/l) and phosphorus-

limited (mass TN/TP: 22.9–81.4). The latter

group are mildly to strongly acidic (pH: 4.30–

6.69) waters stained by dissolved organic carbon

(DOC: 6.8–13.6 mg/l) and more modest trans-

parency (Secchi-disk depth: 60–132 cm). Ratios

of particulate carbon, particulate nitrogen and

chlorophyll a in these lakes indicate that organic

matter in suspension is primarily derived from

the lakes’ catchments rather than aquatic pri-

mary productivity. Since key features in the

Rwenzori lakes’ abiotic environment are strongly

tied to temperature and catchment hydrology,

these Afroalpine lake ecosystems can be

expected to respond sensitively to climate change

and glacier melting.

Handling editor: S. Declerck

Electronic supplementary material The online version ofthis article (doi: 10.1007/s10750-007-0741-3) containssupplementary material, which is available to authorizedusers.

H. Eggermont (&) � K. Van Damme �I. Bessems � D. VerschurenLimnology Unit, Department of Biology, GhentUniversity, K. L. Ledeganckstraat 35, 9000 Gent,Belgiume-mail: [email protected]

J. M. RussellDepartment of Geological Sciences, BrownUniversity, BOX 1846, Providence, RI 02912, USA

G. SchettlerSection Climate Dynamics and Sediments,GeoForschungsZentrum Potsdam, TelegrafenbergC328, Potsdam 14473, Germany

123

Hydrobiologia (2007) 592:151–173

DOI 10.1007/s10750-007-0741-3

Keywords Alpine lakes � Afro-alpine zone �Climate change � East African mountains �Limnology � Pools � Rwenzori � Water chemistry

Introduction

The Rwenzori Mountains, the legendary ‘Moun-

tains of the Moon’, straddle the equator along the

border between Uganda and the Democratic

Republic of Congo. Though surpassed in height

by both Kilimanjaro (5,996 m) and Mt. Kenya

(5,199 m), the high-mountain range as a whole is

more extensive than either of these (Osmaston,

1989). Unlike the other mountains in equatorial

East Africa, the Rwenzoris are not an extinct

volcano, but comprise an uplifted complex of

Precambrian rock (Ebinger, 1989). This complex

has subsequently been dissected by erosion,

resulting in six separate mountains all rising over

4,500 m: Mts. Stanley (5,109 m), Speke (4,889 m),

Baker (4,842 m), Gessi (4,715 m), Emin

(4,791 m) and Luigi di Savoia (4,626 m). Each

of these consists of several peaks, the highest

being Margharita on Mt. Stanley. All mountains

were glaciated until historic times, but ice caps on

Mts. Gessi, Emin and Luigi di Savoia have now

completely disappeared. The Rwenzoris are dot-

ted with numerous lakes mainly occupying gla-

cially-excavated and moraine-dammed basins

formed after the last glacial period (Osmaston,

1989; Osmaston & Kaser, 2001). Streams

descending from the eastern slopes of the moun-

tain range are generally small (Busulwa & Bailey,

2004), yet their sheer number discharges a large

volume of water to the economically important

Lakes Edward and George, and constitute the

most important headwaters of the White Nile

River north of Lake Victoria (Yeoman, 1989). As

in other tropical high-mountain regions, the

glaciers themselves are also important freshwater

reservoirs that store precipitation during the wet

season and release meltwaters during dry periods,

thus buffering seasonal stream flow (Bradley

et al., 2006). The Rwenzori therefore comprise a

vital water catchment, upon which an estimated

500,000 Ugandans directly depend for their water

supply (Howard, 1991). In addition, the mountain

range holds an outstanding diversity of biota, with

many species endemic to the Albertine Rift

region (Kingdom, 1989). Recognising this impor-

tance, the Rwenzori were gazetted as a UNESCO

World Heritage site in 1994.

Climate change and the associated retreat of

Rwenzori’s glaciers may constitute an immediate

threat to its mountain ecosystems. Observations

of glacial termini confirm rapid glacial regression

from 1906 to present. The retreat is consistent

with warming of the tropical middle troposphere

in recent decades (Hastenrath & Kruss, 1992;

Gaffen et al., 2000; Taylor et al., 2006), though a

longer-term decrease in precipitation and in-

creased air humidity may also have affected

glacier mass balance (Hastenrath, 2001; Kaser

et al., 2004). At the current pace, all remaining

glaciers will disappear within the next two

decades (Kaser & Osmaston, 2002; Taylor et al.,

2006). This in turn will directly affect hydrological

processes and biology in the mountain lakes

immediately downstream from these glaciers

(e.g., Hauer et al., 1997; Koinig et al., 2002).

Other factors that may increasingly affect these

alpine lakes are the long-range atmospheric

deposition of toxic pollutants, acids (e.g., Psenner

& Schmidt, 1992; Mosello et al., 2001; Rogora

et al., 2001) and dust (e.g., Psenner, 1999). Hence,

establishing baseline data for the now still rela-

tively pristine Rwenzori lakes will be critical to

our understanding of the ecological and biodiver-

sity effects of future climate and environmental

changes in the region.

Previous scientific expeditions to the Rwenzori,

primarily conducted during the early 20th century,

focused almost exclusively on terrestrial biodiver-

sity (e.g., Ogilvie-Grant, 1908; Heron, 1909; Old-

field, 1910; Burgeon, 1937; Hedberg, 1951;

Kimmins, 1959; Salt, 1987; Wilson, 1995). Limno-

logical surveys have been sparse and/or incom-

plete. De Heinzelin and colleagues conducted

basic surveys of eight lakes on the Congo side of

the Rwenzori range (de Heinzelin & Mollaret,

1956; Mollaret, 1961). Subsequent studies focused

only on four lakes on the Ugandan side, located

along tourist routes in the Bujuku-Mubuku river

drainage (Loffler, 1964, 1968a, b; Livingstone,

1962, 1967). Cholnoky (1964) visited seven lakes

in the remote Kamusongi and Nyamughasana

152 Hydrobiologia (2007) 592:151–173

123

valleys to inventory their lacustrine diatom flora,

but did not collect morphological, physical or

chemical data on these lakes.

This study aims to create a reference database

on the physical and hydrochemical characteristics

of a representative number of high-elevation

lakes and pools in the Rwenzori Mountains, to

permit monitoring of how climate-driven envi-

ronmental change will affect their unique fauna

and flora, and their ecosystem functioning. As on

Mt. Kenya (Loffler, 1964, 1968a, b) and mountain

ranges elsewhere (e.g., Marchetto et al., 1994;

Kamenik et al., 2001), we expect the physical and

chemical properties of Rwenzori lakes to be

strongly linked to elevation (mean annual tem-

perature), local catchment characteristics (type of

vegetation cover, bedrock composition), and

whether or not a lake is fed directly by glacial

meltwater. If this is indeed the case, Rwenzori

lakes and their aquatic biota will likely undergo

major impacts from current climate warming and

predicted disappearance of the glaciers. In view of

conserving a potentially unique tropical cold-

stenothermic aquatic fauna, we also aim to

appraise whether deepwater benthic environ-

ments in thermally-stratified lakes at lower ele-

vations could substitute for shallow-water benthic

environments in (un-)stratified lakes higher up

the mountain. Finally, the assembled data on how

key environmental variables (bottom tempera-

ture and oxygen, pH, transparency and nutrients)

are structured among lakes within the study

region will support studies relating the distribu-

tion of aquatic biota to local environmental

gradients, and the use of their remains preserved

in lake-sediment records as proxy indicators in

paleoenvironmental reconstruction.

Materials and methods

Description of the study sites

All but one of the lakes surveyed for this study are

located on the Ugandan side of the mountain

range between 0�16¢–0�24¢ N and 29�52–29�59¢ E,

and part of the Batoda, Bujuku, Butawu, Kamu-

songi, Mahoma and Nyamugasani River drainages

(Fig. 1; Table 1). The exception is Lac du Speke,

headwater lake of the Luusilubi River and situ-

ated immediately across the Ugandan border in

DR Congo. With few exceptions the Rwenzori

lakes were formed by glacial activity (Table 1).

During the Last Glacial Maximum (LGM,

21 kyears BP) the local snowline extended down

to 3,000 m (Mahaney, 1989), but since most lakes

are located above 3,700 m they are probably of

Holocene age; at least some of their basins were

formed only following glacier retreat after the

Little Ice Age (de Heinzelin, 1962). Most lakes

(e.g., Batoda, Bigata, and Kachope) were created

after a glacial valley was dammed by terminal or

recessional moraines, but those above 4,200 m

occupy glacially-scoured basins. Some of the latter

are quite young (i.e., ~60 years or less; Osmaston,

2006), as their basin was exposed by recent glacier

regression. Examples include Lake Ruhandika at

the foot of Speke glacier, and Lake Irene below

Stanley glacier. All rock pools above 4,400 m are

also formed by glacial scouring below the head-

wall of former glaciers; those at lower elevations

are marsh or river features. Lake Bujuku appears

to be dammed by a landslide off the North slope of

Mt. Baker rather than a moraine (Livingstone,

1967). Lake Mahoma is located at 2,990 m within

the LGM terminal moraine, and is extraordinary

in that its size, shape, and depth all suggest that

unlike any other lake in the African tropics

(Livingstone, 1967) its basin was formed after

the thawing of a block of ice detached from the

retreating glacier (i.e., a kettle lake; Wetzel, 1983).

In many lakes, shoreline features such as wave

cuts in unconsolidated material and water marks

on rocks and large boulders showed evidence of

seasonal lake-level changes, presumably linked to

river discharge variation between wet and drier

seasons. However, all of the lakes are hydrolog-

ically open, limiting the extent to which their

water level can fluctuate.

The Rwenzori are wetter than other East

African mountains, with annual rainfall varying

with altitude from 2,000 to 3,000 mm, and being

heaviest on the eastern slope, which faces the

prevailing winds. On the Uganda side heavy rain

can occur any time of year, but the most rainy

periods are from mid-March to May and from

September to mid-December (Temple, 1961;

Osmaston, 1965). The equatorial position of the

Hydrobiologia (2007) 592:151–173 153

123

mountain range creates daily air temperature

oscillations between – 5 and 20�C in the Alpine

and Nival zones, an order of magnitude greater

than the seasonal variation in maximum daytime

temperature. Occasional night-time freezing

occurs from ~3,000 m altitude (the present-day

boundary between Bamboo and Ericaceous

zones); at 4,000 m (the Ericaceous-Alpine zone

boundary) freezing occurs on 80–90% of the

nights (Rundel, 1994).

The Rwenzori, like other tropical mountains,

exhibit discrete vegetation belts distributed across

the altitudinal gradient (Fries & Fries, 1948;

Hedberg, 1951). Except for Lake Mahoma, which

is surrounded by a mix of montane and bamboo

forest, all study sites are located in the Ericaceous

(~3,100–3,800 m), Alpine (~3,800–4,300 m) or

Nival zones (>4,300 m). The Ericaceous zone is

characterized by broad-leaved trees (Hypericum

spp., Hagenia abyssinica and Rapanea rhododen-

droides), arborescent heathers (Erica spp.)

draped with lichens (Usnea spp.), flowering

shrubs (typically Helicrysum guilelmii), scattered

tree groundsel (mostly Senecio longeligulatus),

Fig. 1 Map of the Rwenzori Mountains, showing glaciers, river drainages and location of the 17 study lakes (numbers referto their listing in Table 1)

154 Hydrobiologia (2007) 592:151–173

123

Ta

ble

1O

ve

rvie

wo

fth

est

ud

ysi

tes

(17

lak

es

an

d1

1p

oo

ls)

list

ed

pe

rd

rain

ag

e,

wit

hin

dic

ati

on

of

the

iro

rig

in,

mix

ing

reg

ime

(se

ete

xt)

,a

nd

ma

jor

en

vir

on

men

tal

va

ria

ble

s

No

.S

ite

na

me

Sa

mp

led

ate

LA

T(N

)L

ON

G(E

)L

oca

tio

nC

atc

hm

en

tty

pe

Ori

gin

Mix

ing

reg

ime

EL

EV

(ma

sl)

La

kes

1L

ac

du

Sp

ek

e7

/15

/20

06

0�2

4.3

21

¢2

9�5

2.8

69

¢M

t.S

pe

ke

Alp

ine

Gla

cia

lly

-sco

ure

d2

a-b

4,2

35

2M

ah

om

a7

/27

/20

06

0�2

0.7

34

¢2

9�5

8.1

02

¢M

ah

om

aF

ore

st/B

am

bo

oK

ett

le3

2,9

90

3B

uju

ku

7/1

1/2

00

60�2

2.6

88

¢2

9�5

3.5

76

¢M

ub

uk

u-B

uju

ku

Alp

ine

*L

an

dsl

ide

-da

mm

ed

2b

3,8

91

4R

uh

an

dik

a7

/12

/20

06

0�2

3.4

53

¢2

9�5

3.4

33

Mt.

Sp

ek

eN

iva

lG

laci

all

y-s

cou

red

N/A

4,3

41

5Ir

en

e7

/14

/20

06

0�2

2.9

36

¢2

9�5

2.8

57

¢M

t.S

tan

ley

Niv

al

Gla

cia

lly

-sco

ure

dN

/A4

,48

76

Up

pe

rK

ita

nd

ara

7/1

9/2

00

50�2

1.1

71

¢2

9�5

3.2

41

¢B

uta

wu

Alp

ine

Mo

rain

e-d

am

me

d2

a4

,01

87

/17

/20

06

7L

ow

er

Kit

an

da

ra7

/17

/20

06

0�2

0.9

47

¢2

9�5

3.1

94

¢B

uta

wu

Alp

ine

Mo

rain

e-d

am

me

d2

a3

,98

98

Lo

we

rK

ach

op

e1

7–

18

/07

/20

05

0�2

0.0

66

¢2

9�5

2.2

96

¢B

uta

wu

Eri

cace

ou

sM

ora

ine

-da

mm

ed

2a

-b3

,84

17

/19

/20

06

9M

idd

leK

ach

op

e1

7–

18

/07

/20

05

0�2

0.0

40

¢2

9�5

2.4

18

¢B

uta

wu

Eri

cace

ou

s*M

ora

ine

-da

mm

ed

2a

3,8

43

10

Up

pe

rK

ach

op

e7

/16

/20

05

0�1

9.9

32

¢2

9�5

2.5

67

¢B

uta

wu

Eri

cace

ou

sM

ora

ine

-da

mm

ed

2a

3,9

61

7/1

9/2

00

61

1B

ato

da

7/1

4/2

00

50�1

7.9

77

¢2

9�5

2.9

72

¢B

ato

da

Eri

cace

ou

sM

ora

ine

-da

mm

ed

2a

4,0

17

7/2

1/2

00

61

2K

op

ell

o7

/12

/20

05

0�1

8.6

12

¢2

9�5

3.5

04

¢N

ya

mu

ga

san

iA

lpin

e*

Mo

rain

e-d

am

me

d2

b4

,01

71

3B

iga

ta7

/11

/20

05

0�1

8.3

96

¢2

9�5

3.5

40

¢N

ya

mu

ga

san

iA

lpin

e*

Mo

rain

e-d

am

me

d3

3,9

83

14

Afr

ica

7/9

/20

05

0�1

7.6

97

¢2

9�5

3.7

70

¢N

ya

mu

ga

san

iA

lpin

e*

Mo

rain

e-d

am

me

d2

b3

,94

31

5K

an

ga

ny

ika

7/8

/20

05

0�1

7.1

89

¢2

9�5

3.8

47

¢N

ya

mu

ga

san

iE

rica

ceo

us

Mo

rain

e-d

am

me

d2

a3

,82

27

/23

/20

05

16

Ka

tun

da

7/2

3/2

00

60�1

6.7

98

¢2

9�5

3.7

28

¢N

ya

mu

ga

san

iE

rica

ceo

us

Mo

rain

e-d

am

me

d1

3,7

82

17

Nsu

ran

ja7

/6/2

00

50�1

7.5

79

¢2

9�5

4.5

01

¢N

ya

mw

am

ba

Eri

cace

ou

s*M

ora

ine

-da

mm

ed

2a

-b3

,83

4

Po

ols

18

Ka

mu

son

gi’

sp

oo

l7

/20

/20

05

0�2

2.3

40

¢2

9�5

3.0

03

¢M

t.S

tan

ley

Niv

al

Gla

cia

lly

-sco

ure

dN

/A4

,50

91

9M

ba

him

ba

’sp

oo

l7

/20

/20

05

0�2

2.3

71

¢2

9�5

2.9

09

¢M

t.S

tan

ley

Niv

al

Gla

cia

lly

-sco

ure

dN

/A4

,55

52

0M

uh

esi

’sp

oo

l7

/20

/20

05

0�2

2.3

67

¢2

9�5

2.8

88

¢M

t.S

tan

ley

Niv

al

Gla

cia

lly

-sco

ure

dN

/A4

,57

02

1B

ale

ng

ek

an

ia’s

po

ol

7/1

4/2

00

60�2

2.9

79

¢2

9�5

2.7

39

¢M

t.S

tan

ley

Niv

al

Gla

cia

lly

-sco

ure

dN

/A4

,57

32

2S

alo

mo

n’s

po

ol

7/1

4/2

00

60�2

2.9

46

¢2

9�5

2.8

59

¢M

t.S

tan

ley

Niv

al

Gla

cia

lly

-sco

ure

dN

/A4

,48

02

3Z

ap

ha

na

’sp

oo

l7

/14

/20

06

0�2

3.0

04

¢2

9�5

2.7

73

¢M

t.S

tan

ley

Niv

al

Gla

cia

lly

-sco

ure

dN

/A4

,54

32

4T

un

aN

oo

dle

po

ol

7/1

5/2

00

50�2

0.0

59

¢2

9�5

2.3

53

¢B

am

wa

nja

rap

ass

Alp

ine

*M

ars

hp

oo

lN

/A4

,38

02

5Jo

sep

ha

t’s

po

ol

7/1

9/2

00

50�2

1.6

76

¢2

9�5

3.2

52

¢B

uta

wu

Alp

ine

Ma

rsh

po

ol

N/A

4,1

00

26

Ba

gu

ma

’sp

oo

l7

/7/2

00

50�1

7.8

43

¢2

9�5

3.9

86

¢N

ya

mu

ga

san

iE

rica

ceo

us*

Ma

rsh

po

ol

N/A

3,9

94

27

Ze

ph

an

ia’s

po

ol

7/1

3/2

00

50�1

8.3

85

¢2

9�5

3.0

83

¢N

ya

mu

ga

san

iA

lpin

e*

Ma

rsh

po

ol

N/A

4,2

24

28

Mu

tin

da

po

ol

7/5

/20

05

0�1

6.4

65

¢2

9�5

5.6

51

¢K

am

uso

ng

iE

rica

ceo

us*

Riv

er

po

ol

N/A

3,5

07

Hydrobiologia (2007) 592:151–173 155

123

Ta

ble

1co

nti

nu

ed

Sit

en

am

eZ

max

(m)

Lm

ax

(m)

Are

a(h

a)

TE

MP

bo

tto

m(�

C)

CO

ND

bo

tto

m(l

S/c

m)

O2

surf

ace

(mg

/l)

O2

bo

tto

m(m

g/l

)p

Hb

ott

om

(un

its)

SE

CC

HI

(cm

)

La

kes

La

cd

uS

pe

ke

17

.02

75

3.9

12

5.1

51

27

.96

5.0

45

.20

12

0M

ah

om

a2

5.6

32

54

.76

31

2.5

24

15

.36

0.4

55

.10

82

Bu

juk

a1

3.5

40

07

.87

36

.95

50

7.4

00

.24

5.8

81

80

Ru

ha

nd

ika

3.0

30

0.0

09

(4.0

0)

57

.59

4.3

76

.03

(16

6)

Ire

ne

3.5

50

0.0

16

(4.0

0)

77

.59

4.3

75

.70

(18

6)

Up

pe

rK

ita

nd

ara

14

.53

00

3.2

58

4.9

12

89

.40

7.3

97

.00

28

04

.72

32

7.8

67

.82

6.7

52

80

4.8

23

08

.63

7.6

16

.88

28

0L

ow

er

Kit

an

da

ra1

1.0

30

02

.86

05

.93

33

7.4

77

.09

6.4

52

40

Lo

we

rK

ach

op

e8

.32

00

0.8

12

7.3

52

27

.66

5.7

56

.29

13

07

.43

25

6.4

64

.61

5.7

61

10

7.3

92

3.5

7.0

65

.18

6.0

31

20

Mid

dle

Ka

cho

pe

3.3

15

01

.00

77

.33

27

7.6

06

.57

6.5

61

25

Up

pe

rK

ach

op

e1

2.0

35

04

.59

36

.36

19

8.8

56

.78

6.8

61

20

6.3

62

26

.53

5.1

05

.53

11

06

.36

20

.57

.69

5.9

46

.20

11

5B

ato

da

15

.04

50

9.5

30

6.0

91

27

.53

5.9

55

.08

80

15

.56

.10

12

6.7

96

.30

4.8

49

06

.10

12

7.1

66

.13

4.9

68

5K

op

ell

o1

4.5

35

03

.00

75

.51

10

9.5

52

.31

4.8

81

12

Big

ata

18

.02

00

1.4

41

5.1

72

48

.00

0.2

25

.74

13

2A

fric

a4

.72

70

1.0

01

5.4

81

99

.39

3.3

05

.37

90

Ka

ng

an

yik

a2

5.0

75

01

1.2

34

6.9

51

37

.82

5.1

15

.12

85

7.3

11

46

.17

5.8

74

.73

10

07

.13

13

.57

.00

5.2

94

.93

92

.5K

atu

nd

a1

0.5

40

03

.76

57

.52

14

6.3

56

.44

5.2

21

17

Nsu

ran

ja1

2.5

30

02

.15

56

.78

12

7.1

73

.80

4.3

16

0

me

an

(n=

17

)1

2.6

30

03

.60

26

.29

20

.79

7.5

94

.37

5.6

11

35

min

imu

m3

.03

00

.00

92

.89

5.0

05

.36

0.2

24

.31

60

ma

xim

um

25

.67

50

11

.23

41

2.5

25

0.0

09

.55

7.3

97

.00

28

0S

D6

.61

64

3.2

49

1.4

81

1.1

41

.08

2.3

50

.81

65

Po

ols

Ka

mu

son

gi’

sp

oo

l<

2<

10

<0

.00

3N

/AN

/AN

/AN

/AN

/AN

/AM

ba

him

ba

’sp

oo

l<

2<

10

<0

.00

3N

/AN

/AN

/AN

/AN

/AN

/AM

uh

esi

’sp

oo

l<

2<

10

<0

.00

3N

/AN

/AN

/AN

/AN

/AN

/AB

ale

ng

ek

an

ia’s

po

ol

<2

<1

0<

0.0

03

N/A

N/A

N/A

N/A

N/A

N/A

Sa

lom

on

’sp

oo

l<

2<

10

<0

.00

3N

/AN

/AN

/AN

/AN

/AN

/AZ

ap

ha

na

’sp

oo

l<

2<

10

<0

.00

3N

/AN

/AN

/AN

/AN

/AN

/A

156 Hydrobiologia (2007) 592:151–173

123

and giant lobelias (Lobelia stuhlmannii). The

most abundant plants in the Alpine zone are

Carex tussocks, as well as Helichrysum stuhlman-

nii and Alchemilla spp. Various Senecio and

Lobelia spp. (primarily S. adnivalis, S. friesiorum

and L. wollastonii) are common in ravines and

other sheltered or well-watered sites. The Nival

zone is typified by bare rock, moss, liverworts and

lichens. The altitudinal vegetation belts are not

always clearly separated, and transition zones

with mixed vegetation are common throughout

the range.

Sample collection and laboratory analyses

In July of 2005 and 2006, 17 lakes (depth

Zmax > 2 m; surface area at least 90 m2 but

usually >1 ha) and 11 pools (Zmax < 2 m; surface

area < 30 m2) were sampled (Table 1). Because

of their small size, pools lack an official name; we

named them after guides, porters, or people from

the research team. Latitude (LAT), longitude

(LONG) and elevation (ELEV, in meter above

sea level) were recorded by GPS (Garmin eTrex

Vista; mean error 10 m). Water samples were

taken ~30 cm beneath the surface at mid-lake

locations, but near the shore in pools. Lake

bathymetry was determined by GPS-guided echo-

sounding (handheld depth sounder Echotest II),

and bathymetric maps were drawn in Surfer v. 7.0

(Golden Software, Inc.) using kriging as the

gridding method. Lake surface area (Area) and

maximum length (Lmax) were subsequently cal-

culated using ImageJ v. 1.37 software (W. Ras-

band, unpublished software); for pools these

morphometric parameters were visually esti-

mated from the shore.

Temperature (T), pH, oxygen (O2) and con-

ductivity (COND: specific conductance at 25�C,

in lS/cm) of the water column were measured at

1-m intervals with a Hydrolab Quanta CTD

profiler. Transparency (SECCHI) was measured

using a 22-cm diameter Secchi disk. Due to logistic

constraints no T or O2 measurements were done in

the pools, and their pH and conductivity was

measured up to 4 h after sampling. The majority

of lakes and pools were sampled only once, but

repeat sampling of five lakes in 2005 and 2006

indicates quasi-stationary dry-season conditions inTa

ble

1co

nti

nu

ed

Sit

en

am

eZ

max

(m)

Lm

ax

(m)

Are

a(h

a)

TE

MP

bo

tto

m(�

C)

CO

ND

bo

tto

m(l

S/c

m)

O2

surf

ace

(mg

/l)

O2

bo

tto

m(m

g/l

)p

Hb

ott

om

(un

its)

SE

CC

HI

(cm

)

Tu

na

No

od

lep

oo

l<

2<

10

<0

.00

3N

/AN

/AN

/AN

/AN

/AN

/AJo

sep

ha

t’s

po

ol

<2

<1

0<

0.0

03

N/A

N/A

N/A

N/A

N/A

N/A

Ba

gu

ma

’sp

oo

l<

2<

10

<0

.00

3N

/AN

/AN

/AN

/AN

/AN

/AZ

ep

ha

nia

’sp

oo

l<

2<

10

<0

.00

3N

/AN

/AN

/AN

/AN

/AN

/AM

uti

nd

ap

oo

l<

2<

10

<0

.00

3N

/AN

/AN

/AN

/AN

/AN

/A

me

an

(n=

7)

<2

<1

0<

0.0

03

N/A

N/A

N/A

N/A

N/A

N/A

min

imu

m<

2<

10

<0

.00

3N

/AN

/AN

/AN

/AN

/AN

/Am

ax

imu

m<

2<

10

<0

.00

3N

/AN

/AN

/AN

/AN

/AN

/AS

D<

2<

10

<0

.00

3N

/AN

/AN

/AN

/AN

/AN

/A

Ab

bre

via

tio

ns

are

as

de

fin

ed

inth

ete

xt;

N/A

=n

od

ata

av

ail

ab

le.

Va

lue

sin

bo

ldty

pe

face

are

the

me

an

so

ftw

om

ea

sure

me

nts

inco

nse

cuti

ve

ye

ars

;T

an

dS

EC

CH

Iv

alu

es

for

Ire

ne

an

dR

uh

an

dik

a(b

etw

ee

nb

rack

ets

)a

ren

ot

me

asu

red

bu

te

stim

ate

da

se

xp

lain

ed

inth

ete

xt

Hydrobiologia (2007) 592:151–173 157

123

consecutive years. Samples for analysis of cations

(Ca, Mg, Na, K, Li, Ba, Sr, Fe, Mn), anions (F, Cl,

NO3, SO4), dissolved silica (Si), total dissolved

sulfur (TS), dissolved phosphorus (PO4-P), dis-

solved inorganic carbon (DIC) and dissolved

organic carbon (DOC) were filtered through

Sartoriusr cellulose acetate filters of 0.45 lm

mesh. Filtrates for cation, PO4-P and TS analysis

were stabilized with nitric acid (Merck, Suprapur

quality) and stored in 50 ml pre-cleaned polyeth-

ylene bottles. Filtered samples for analysis of

anions, Si and dissolved carbon were stored

untreated in separate bottles. For determination

of total phosphorus (TP) and total nitrogen (TN),

unfiltered samples were fixed by adding concen-

trated sulfuric acid. Samples for analysis of

particulate organic carbon (POC), particulate

organic nitrogen (POC) and chlorophyll a (Chl-

a) were taken by filtering 180–420 ml water

through pre-ashed Whatmanr glass microfibre

filters (0.45 lm). The filters carrying suspended

material were folded in half, placed in plastic Petri

dishes and wrapped in aluminium foil. All water

samples were kept dark and cool (filters were

stored frozen) during the field campaign and until

their shipment to the laboratory.

Cations, Si, PO4-P and TS were analysed by

Inductively Coupled Plasma Atom Emission

Spectrometry (ICP-AES; IRIS, Thermo Elemen-

tal). F, Cl, NO3 and SO4 were analysed by ion-

exchange chromatography (DX100, Dionex). The

highly dilute nature of the studied waters (5–

52 lS/cm specific conductance at 25�C) resulted

in measured values for several chemical species

(K, Na, PO4-P, F, Cl, SO4, NO3) that are close to

or below the detection limit of the analytical

methods employed. Risk of contamination was

minimized by repeated pre-flushing of membrane

filters and autosampler vials with the individual

waters. We also took account of the non-linearity

of IC calibration curves by increasing the density

of calibration points in the low-concentration

range. Summation of all the measured ion con-

centrations yielded a systematic excess of calcu-

lated conductivity relative to conductivity

measured in the field. We tentatively attribute

this to the leaching of cations from microscopic

(<0.45 lm) organic particles remaining in the

filtered water samples, after field addition of nitric

acid. However, no systematic variation in cation

excess occurs in relation to the time elapsed

between sample collection and analysis.

DIC and DOC (excluding volatile organic

carbon components, i.e. the Non-Purgeable

Organic Carbon or NPOC) were measured by

IR-spectrometry (TOC-2000A, Shimadzu). The

NPOC measurement protocol included acidificat-

ion of the sample by addition of hydrochloric acid,

sparging of the sample with high purity air, and

catalytic burning of the sample aliquots at 680�C.

DIC and field pH data were combined to calculate

the in situ HCO3, CO3 and CO2 concentrations,

following Wetzel & Likens (1990). TP was deter-

mined by wet oxidation in an acid persulphate

solution (120�C, 30 min), and TN (as nitrate plus

nitrite) by wet oxidation in an alkaline persul-

phate solution (120�C, 30 min), both following

Grasshoff et al. (1983). Chl-a was determined by

high-pressure liquid chromatography (HPLC),

following the protocol of Wright et al. (1991,

1997). POC and PON were measured by flash

combustion/thermal conductivity in a CE Instru-

ments NC2100 elemental analyzer. The reported

POC and PON values are means of duplicate

measurements, using subsamples of each filter.

Data screening and statistical analyses

The substantial differences in size and depth

ranges of Rwenzori lakes and pools (Table 1)

justifies separate analysis of their thermal and

chemical regimes. This study focuses on the

comparative limnology of the true lakes, with

pools referred to only in passing. The initial data

set thus includes 17 sites and 39 environmental

variables (see Electronic supplementary mate-

rial). In lakes sampled twice, we use the average

values of the two measurement series. Four

variables were removed because their concentra-

tions were below detection limit at more than half

of the study sites (NO3, PO4-P, Li) or were

calculated as negligible (CO3, due to low pH).

The remaining variables were tested for normal-

ity using Shapiro–Wilk tests (Shapiro et al., 1968).

Eight variables were normalized using either log

(Tsurface, Tbottom, CONDsurface) or log

(x + 1) transformation (SO4, TS, TPU, H2CO3,

HCO3). Two variables (F and Mn) had to be

158 Hydrobiologia (2007) 592:151–173

123

excluded because their distribution could not be

normalized.

We calculated a Pearson correlation matrix to

quantify relationships between the 33 environ-

mental variables retained for multivariate analy-

sis. We then used principal components analysis

(PCA) to identify the principal environmental

gradients structuring the data set. Some variables

which are intrinsically strongly correlated with

others were excluded to avoid distortion of the

analysis due to redundancy. Specifically, Lmax was

removed in favour of lake Area. Surface-water

measurements of T, COND and pH were

removed in favour of the corresponding bottom

water measurements, because the latter reflect

long-term conditions while surface-water data are

more subject to daily (i.e. depending on the time

of day) and seasonal fluctuations (i.e. whether the

measurement occurred just before or after a

deep-mixing event). Surface and bottom water

measurements of oxygen, on the other hand, were

both kept in because lack of correlation between

the two suggests that site-specific basin morphom-

etry or primary production affects local bottom

oxygen regimes. Missing Tsurface and SECCHI

data for Lakes Irene and Ruhandika were esti-

mated using the significant regressions between

Tsurface and ELEV and between SECCHI and

DOC at the other 15 sites. The Tbottom for these

two lakes was set at 4�C, because (1) the

regression between Tbottom and ELEV in the

other lakes indicated that their Tbottom is

unlikely to be significantly higher than 4�C; and

(2) estimated Tsurface suggests at least daytime

warming of their water surface.

For PCA analysis we further added a set of

categorical variables representing the dominant

vegetation type in each lake catchment, namely:

bare rocks (Nival), alpine vegetation (Alpine),

alpine vegetation dominated by Carex swamp

(Alpine*), Ericaceous vegetation (Ericaceous),

Ericaceous vegetation dominated by Carex

swamp (Ericaceous*), and a mix of montane

and bamboo forest (Forest/Bamboo). The final

set of 35 variables was centred and standardized

to allow comparison of disparate variables (ter

Braak & Smilauer, 1998). We refrained from

testing the significance of the PCA ordination

axes since the theory that has been developed for

these tests suffers intrinsic methodological flaws

(ter Braak & Smilauer, 2002). Finally, we per-

formed redundancy analysis (RDA) to assess

whether selected external variables significantly

influenced the physical and chemical properties of

Rwenzori lakes. For this we used six predictor

variables (ELEV; Ericaceous; Erica-

ceous* + Alpine*, Alpine, Nival, and Forest/

Bamboo) and 26 response variables (the same

set as for the PCA, but excluding the predictor

variables and morphometric parameters), and

performed a series of RDAs constrained to one

single predictor at a time. Statistical testing of

every predictor variable was done with the

random permutation procedure of CANOCO

(reduced model, 999 permutations; ter Braak &

Smilauer, 2002) with manual variable selection.

Ternary graphs, correlation matrices and nor-

mality tests were generated using the software

package STATISTICA 5.5 (Statsoft, 2000). Multi-

variate statistics were performed using the package

CANOCO v. 4.5 (ter Braak & Smilauer, 2002).

Results

Multivariate statistical analysis

The first two PCA axes together account for 51.9%

of the environmental variance in our Rwenzori

lake data set (k1 = 0.298 and k2 = 0.221; Fig. 2).

PCA axis 1, which explains 29.8% of the total

variation, mainly captures gradients in land cover

(e.g. Nival: presence/absence of bare rocks), ele-

vation, bottom water temperature, major cations,

dissolved Si, and levels of the trace elements Ba, Fe

and Sr. Bottom water temperature and elevation

point in opposite directions, reflecting the strong

inverse relationship between them (see correlation

matrix, see Electronic supplementary material).

All lakes located below 4,000 m plot in the left

quadrants, and with the exception of Batoda all

lakes above 4,000 m plot in the right quadrants.

Lake Africa and the Kachope lakes have Mg, Na,

K, Sr and Si concentrations above the mean,

explaining their clustering with these variables in

(or close to) the upper left quadrant. Lakes in the

lower left quadrant (e.g. Nsuranja, Batoda, Kang-

Hydrobiologia (2007) 592:151–173 159

123

anyika) are typified by enhanced Ba and Fe

concentrations. PCA axis 2, which explains 22.1%

of the total variation, predominantly captures

gradients of Secchi depth, DOC, TN, TP, major

anions, and bottom pH. Here, Secchi depth and

DOC point in opposite directions, reflecting the

strong inverse relationship between them. As

reflected by the narrow angles between variables

pointing towards the lower left quadrant, less

transparant lakes (high DOC content) are more

acidic and hold more nutrients (high TP and TN);

they also tend to be larger and deeper than other

lakes. The opposite is true for lakes clustering in (or

close to) the upper right quadrant. As indicated by

the land-cover centroids, the former type of lakes

are predominantly surrounded by Forest/Bamboo,

Ericaceous vegetation and/or Carex swamp,

whereas the latter type are enclosed by bare rocks

or Alpine vegetation. Some lakes from the same

valley (drainage) but in different vegetation zones

plot in the same quadrant, e.g. Africa, Kanganyika

and Katoda.

RDA analysis indicates that elevation

(F = 3.400, P = 0.003) and Nival (i.e. presence/

absence of bare rocks; F = 3.915, P = 0.007), Alpine

(F = 2.297; P = 0.026) and Ericaceous* + Alpine*

vegetation (i.e. presence/absence of Carex swamps;

F = 2.376, P = 0.035) contribute significantly and

independently to explained variation in within-lake

environmental conditions: RDA axes 1 and 2

together explain 41.1% of the observed variance

(k1 = 0.277 and k2 = 0.134).

Lake morphometry and catchment

characteristics

Lake morphometry is summarized in Table 1 and

illustrated in Figs. 3–5. Zmax has a significant (P <

0.05) relationship with both Lmax (r = 0.57) and

Area (r = 0.60). Lake Kanganyika is by far the

largest and deepest of all lakes (Zmax = 25.0 m;

Lmax = 750 m; Area = 11.2 ha), while Lake Ru-

handika is the smallest and shallowest (Zmax =

3.0 m; Lmax = 30 m; Area = 0.009 ha). Lakes

dammed by moraines or land-slides in steep valleys

(e.g. the Kachope and Kitandara lakes; Figs. 3, 4)

are typically long-drawn between in- and outflow,

whereas lakes in glacially-scoured basins (e.g. Lac

du Speke; Fig. 3) and the Mahoma kettle (Fig. 3)

are more circular and lack a well-defined inlet. Most

moraine-dammed lakes have a single major inflow

(e.g. Upper and Lower Kitandara; Fig. 3), in some it

is supplemented by several smaller inflowing

streams that drain the local catchment (e.g. Batoda

and Bujuku; Figs. 3, 4). None of the lakes is steep-

sided all around, so they all have (at least locally) a

clear littoral, shallow-water zone. Except for lakes

Ruhandika and Irene in the Nival zone whose basin

lack continuous vegetation cover, all surveyed

Rwenzori lakes are surrounded either by bogs with

Carex and Sphagnum (e.g. Africa, Bigata, Middle

Kachope, Nsuranja), or steep rocky slopes with

Ericaceous (e.g., Kanganyika, Katunda, Upper and

Lower Kachope) or Alpine vegetation (e.g., Buj-

uku, Kopello, Kitandara, Speke), as defined above.

This vegetation is still pristine in all drainage basins

except the Kachope lakes (Fig. 4), where deliber-

ately-set recent fires burned much of the vegetation.

Thermal regime

None of the investigated Rwenzori lakes had ice

cover during our visits. Excluding the atypically

warm Lake Mahoma at 2,990 m (13.5�C), mid-

day surface temperatures ranged from 5.5�C in

Fig. 2 Principal Components Analysis (PCA) biplotshowing relationships between the 17 Rwenzori studylakes (full circles) and 35 environmental variables (arrowsand open circles)

160 Hydrobiologia (2007) 592:151–173

123

Irene to 9.1�C in Bigata, with a mean of 7.8�C

(Table 1). Bottom water temperatures ranged

from 4.8�C in Upper Kitandara to 7.5�C in

Katunda, again excluding Mahoma (12.5�C).

The expected inverse correlation between water

temperature and elevation is distinctly stronger

using values of bottom temperature (r = – 0.93,

P < 0.001) than of surface temperature

(r = – 0.82; P < 0.001). When excluding Lake

Mahoma, the relationship of surface temperature

with elevation even lacks statistical significance

(r = – 0.44, P > 0.05; versus r = – 0.84, P <

0.001 for bottom temperature).

Water-column profiles of temperature, dis-

solved O2, conductivity and pH (Fig. 6) indicate

that all surveyed Rwenzori lakes undergo stratifi-

cation, also the shallow lakes Africa (Zmax =

4.7 m; Fig. 6d) and Middle Kachope (3.3 m).

Based on the magnitude of surface-to-bottom

contrast in the above-mentioned parameters

observed at the time of our visit(s), we differentiate

three types of mixing regime (Table 1). Type 1 is

defined for Lake Katunda (Fig. 6a), the only lake

where temperature, oxygen, pH and conductivity

were uniform at all depths, except for modest mid-

day warming in the upper 1 m. As normal wind-

driven mixing in this lake most probably extends

deeper than 1 m, we surmise that the lack of a clear

thermocline in our profile reflects an event of deep

convective mixing shortly before the measurement

was taken. The type 2 mixing regime comprises

lakes with a slightly to strongly clinograde oxygen

curve, but where the bottom water remains well-

oxygenated ([O2] > 1 mg/l). These lakes display a

more or less well-developed thermocline at 3–5 m

depth and (often) reduced pH values in the

hypolimnion, but no significant conductivity gra-

dient between surface and deep water. In the

Fig. 3 Bathymetry and surrounding land cover of Upper and Lower Kitandara lakes (Butawu drainage), Lac du Speke(Mt. Speke), Mahoma (Mahoma drainage) and Bujuku (Mubuku-Bujuku drainage)

Hydrobiologia (2007) 592:151–173 161

123

context of benthic habitat conditions we further

differentiate between lakes with hypolimnetic O2

concentrations close to those at the surface (Type

2a; e.g., Kanganyika, Fig. 6b) and lakes where

hypolimnetic O2 has been markedly depleted

(Type 2b; e.g., Kopello and Africa, Fig. 6c, d).

Type 3 comprises lakes with a strongly clinograde

oxygen curve, and anoxic bottom waters with

significantly higher conductivity than surface

waters (Fig. 6e, d).

Transparency, water chemistry and primary

production

Secchi depth transparency in the 17 study lakes

ranged from 60 to 280 cm (Table 1), and DOC

ranged from 3 to 13.6 mg/l (see Electronic

supplementary material). F. Secchi depth was

significantly correlated with both DOC

(r = – 0.66; P < 0.01) and Fe (r = – 0.61; P <

0.05), which also co-vary among them (r = 0.79;

P < 0.001). Low-transparency lakes were either

tea-coloured (e.g. Nsuranja, Batoda, Africa,

Kanganyika) or green (Mahoma only).

Surface-water conductivity ranged from 5 to

52 lS/cm (mean 17.3 lS/cm) in the lakes, and

from 5 to 18 lS/cm (mean 9.6 lS/cm) in the pools

(Table 1). The principal cation concentration

sequence in the lakes is Ca > K > Na > Mg, with

means of 1.953, 0.511, 0.420 and 0.346 mg/l,

respectively (see Electronic supplementary mate-

rial for individual values). Pools display a similar

cation concentration pattern (Fig. 7a), with

means of 0.867, 0.548, 0.359 and 0.226 mg/l.

Anion concentration patterns exhibit consider-

ably less coherence (Fig. 7b), with the sequence

SO4 > HCO3 > Cl in lakes of the Bujuku and

Butawu drainages (means of 1.608, 0.672 and

0.414 mg/l, respectively) and SO4 > Cl > HCO3

in lakes on Mt. Speke, Mt. Stanley (including the

Fig. 4 Bathymetry and surrounding land cover of lakes Nsuranja (Nyamwamba drainage), Upper, Middle and LowerKachope (Butawu drainage), and Batoda (Batoda drainage)

162 Hydrobiologia (2007) 592:151–173

123

rock pools) and in the Mahoma, Batoda and

Nyamughasani drainages.

Surface-water pH in the lakes ranged from

4.30 to 7.80, and is significantly (P < 0.01)

correlated with DOC (r = – 0.68), Ca

(r = 0.78) and DIC (r = 0.71). Dissolved Si

concentrations in the surface water varied from

0.11 to 1.90 mg/l but with a fairly high mean

value of 1.21 mg/l: in all but four lakes, surface

Si concentrations were near or above 1 mg/l.

The four exceptions are the two lakes in the

Nival zone (Ruhandika: 0.11 mg/l; Irene,

0.14 mg/l), Lac du Speke on the DR Congo side

of the range (0.26 mg/l), and our July 2006

measurement in Upper Kachope (0.22 mg/l; see

Electronic supplementary material).

Surface-water TP concentrations in the sur-

veyed lakes ranged from 3.1 to 68.4 lg/l. TN

ranged from 155.4 to 695.8 lg/l, and was signifi-

cantly (P < 0.01) correlated with DOC (r = 0.76).

TN:TP varied between 9:1 and 81:1, Chl-a between

0.29 and 10.9 mg/l, POC:Chl-a between 85:1 and

Fig. 5 Bathymetry and surrounding land cover of lakes Kopello, Afrika, Bigata, Kanganyika and Katunda (Nyamugasanidrainage)

Hydrobiologia (2007) 592:151–173 163

123

2,471:1, and POC:PON between 11:1 and 14:1 (see

Electronic supplementary material).

In agreement with the ordination of lakes and

environmental variables in PCA (Fig. 2), distinct

patterns among Rwenzori lakes with regard to the

inter-related variables of Secchi depth, DOC, pH,

TP, TN:TP and Chl-a permit the distinction of two

groups of lakes. The first group includes six lakes

(Lac du Speke, Bujuku, Ruhandika, Irene, Upper

and Lower Kitandara) that are located near or

above 4,000 m (3,890–4,487 m) in the Alpine and

Nival zones, and receive at least some direct input

of glacial meltwater. These lakes are mildly acidic

to pH-neutral (surface pH: 5.80–7.82) and have

above-average transparency (Secchi depth: 120–

280 cm); their DOC content is noticeable but

modest (3.0–5.5 mg/l), and epilimnetic NO3 con-

centration is always well above the detection limit

(0.2–2.9 lg/l). These lakes have very low to mid-

range TP (3.1–12.4 lg/l), TN (155–393 lg/l) and

Chl-a (0.3–10.9 lg/l), and their range of TN/TP

values (22.9–81.4 by mass) suggests that primary

production is phosphorus-limited.

The second group of 11 lakes are mostly

located below 4,000 m (2,990–4,054 m), more

remote from the glaciers and surrounded by

Ericaceous vegetation or, when in the Alpine

zone, fringed by a Carex-Sphagnum bog. These

lakes are mildly to strongly acidic (surface pH:

4.30–6.69), and less transparent (Secchi depth: 60–

132 cm) usually because they are stained by high

concentrations of dissolved organic carbon

(DOC: 6.8–13.6 mg/l). The exception is Lake

Mahoma at the Forest/Bamboo transition, which

Fig. 6 Profiles of temperature, oxygen, conductivity andpH in lakes Katunda (a), Kanganyika (b), Kopello (c),Africa (d), Bigata (e) and Mahoma (f). In case of repeated

measurements, full and dotted lines indicate profiles of2005 and 2006, respectively

164 Hydrobiologia (2007) 592:151–173

123

is acidic (pH: 5.20) and modestly transparent

(Secchi depth: 82 cm) but less burdened with

humic acids (DOC: 5.5 mg/l) and has a greenish

colour. In almost all lakes of this second group

including Mahoma, both PO4-P and NO3 concen-

trations are below the detection limit of 0.1 lg/l.

Compared to the first group, these lakes have

higher epilimnetic concentrations of TP (10.2–

68.4 lg/l) and TN (213–696 lg/l), but their range

of Chl-a (1.0–10.6 lg/l) is similar.

All sampled Rwenzori lakes have high to very

high concentrations of suspended organic matter

(POC range: 0.72–4.13 mg/l; PON range: 0.08–

0.34 mg/l; see Electronic supplementary mate-

rial), with no marked difference between the two

groups of lakes defined above. The carbon-to-

nitrogen ratio (C:N) of particulate organic matter

(POC/PON), with ranges from 8.3 to 12.1 with a

mean value of 10.5.

Discussion

Thermal and stratification regimes

The temperature of mountain lakes depends on

elevation (through its effect on air temperature),

on the proximity of a glacier (specifically, whether

or not the lake is directly fed by glacial meltwa-

ter), on the number of hours of direct insolation,

and on their exposure to wind. In stratifying lakes,

bottom temperatures integrate at least the daily

variation in surface-water temperature, and are

therefore more representative of mean local

conditions. The observed strong relationship

between bottom temperature and elevation in

our lake data set (r = – 0.84, P < 0.001) mirrors

the highly significant correlation between mean

annual air temperature and elevation in the

Rwenzori (r = – 0.98, P < 0.001; Eggermont

et al., unpublished temperature logger data),

indicating that surface air temperature is a key

factor in determining the temperature regime of

Rwenzori lakes. The scatter that does occur in the

relationship between lake temperature and ele-

vation is nevertheless significant, and points to

substantial influences of lake hydrology and

exposure on local thermal regimes. For example,

Upper Kitandara Lake is colder than other lakes

at similar elevation, likely because it is fed

directly by glacial meltwater from Mts. Baker

and Stanley, and furthermore located in a narrow

valley tunnelling cold wind from the glaciated

Fig. 7 Ternary diagram showing major cation (a) and anion(b) proportions of individual Rwenzori lakes (full circles)and pools (open circles), with codes as in Table 1. Alsoindicated (as square symbols) are the mean values for theUgandan Rwenzori lakes (R1, n = 17; this study), Rwenzoripools (R3, n = 7; this study), Congolese Rwenzori lakes (R2,

n = 5, derived from de Heinzelin & Mollaret, 1956), andselected freshwater (<1,000 lS/cm) lakes in the Ugandanlowlands (L1, n = 13, derived from Kilham 1971; L2, withn = 9, derived from Kizito et al., 1993; L3, with n = 21,Verschuren et al., unpublished data). Kizito et al. (1993) andde Heinzelin & Mollaret (1956) do not report anion data

Hydrobiologia (2007) 592:151–173 165

123

peaks. In other lakes, deep wind-driven mixing

during periods of incipient surface heating can

raise their bottom temperature above the value

predicted from its relationship with elevation.

Such lakes (e.g., Lower Kachope; Table 1) also

tend to feature modest surface warming and a

weakly pronounced thermocline. Latent heat

losses from the Rwenzori mountain lakes through

evaporation must be generally minimal, due to

the prevailing cool and humid conditions.

Hutchinson and Loffler (1956) and Loffler

(1957, 1960, 1964) described lakes of tropical

high-mountain regions in Africa and South

America as frequently (even daily) stratifying

and mixing at temperatures below 10�C, and

therefore designated all of them as ‘cold poly-

mictic’. Ruttner (1963) adopted their reasoning

when assigning this term to all lakes that are not

ice-covered and circulate daily at temperatures at

or slightly above 4�C. Lewis (1983) noted the

arbitrary nature of the 10�C boundary, and

presented an amended classification of mixing

regimes which resolved this and other problems

with the Hutchinson-Loffler scheme (Wetzel,

2001). In the Lewis (1983) classification, cold-

monomictic lakes stratify because surface tem-

perature drops well below 4�C (often resulting in

seasonal ice cover) and no longer mix with the

heavier, but warmer, bottom water. Depending

on wind stress during the ice-free season, such

lakes have bottom temperatures at or below 4�C

for much of the year. When such lakes also

stratify for short or longer periods during the ice-

free season by surface heating above 4�C, they are

designated as cold polymictic. The Lewis (1983)

classification is still somewhat unsatisfactory for

tropical high-mountain lakes, which can briefly

freeze over at night but during daytime are

usually ice-free year-round (Loffler 1964). In

terms of the ambient temperature regime for

aquatic flora and fauna, the most important

differentiation here is between mono- or poly-

mictic lakes that stratify due to surface heating

(and thus have hypolimnia of 4�C or higher) and

mono- or polymictic lakes that stratify due to

surface cooling (and thus have hypolimnia of 4�C

or cooler). In this respect, all 17 Rwenzori lakes

we surveyed so far are warm stratifying. Most

lakes in the Alpine Zone (Lac du Speke, Upper

Kitandara, Kopello, Bigata, Africa) have bottom

temperatures slightly above 4�C (range 4.8–

5.5�C). Lake Mahoma at 2,990 m has a bottom

temperature of 12.5�C, a clear reflection of the

higher mean annual air temperature at this lower

elevation. The intermediate bottom temperatures

of lakes in the Ericaceous zone (range 6.1–7.5�C;

n = 7) must at least partly also be due to the

influence of ambient air temperature, but can also

result because deep wind-driven mixing continues

into periods of incipient surface heating. Cold-

stratifying mountain lakes, although not repre-

sented in our data set, do occur in tropical Africa:

Lewis Tarn and Curley Pond on Mt. Kenya, and

Lac Gris and Lac Blanc on the drier DR Congo

side of the Rwenzori all have bottom tempera-

tures well below 4�C (data from Loffler (1964)

and de Heinzelin & Mollaret (1956).

Based on our current data set, lakes on the

Uganda side of the Rwenzori broadly classify into

three types of mixing regime. In Type 1 (Lake

Katunda, Fig. 6a), uniform temperature, oxygen,

pH and conductivity throughout the water column

except the very surface suggests that it undergoes

a continuous cycle of weak day-time stratification

and night-time mixing. Given the absence of

seasonal ice cover, this lake classifies as continu-

ous warm polymictic. Type 2 groups lakes char-

acterized by a more or less well-developed

primary thermocline at the depth of wind-driven

mixing (here 3–5 m depth), and modest to strong

oxygen depletion in the hypolimnion, however

without it becoming completely anoxic. This

suggests that these lakes stratify for at least

several weeks or months at a time, alternating

with periods of mixing when fresh oxygen is

supplied to the hypolimnion. These lakes classify

as discontinuous warm polymictic. Evidently, a

more productive Type 2 lake that mixes fre-

quently may better prevent hypolimnetic oxygen

depletion than a less productive lake mixing

rarely, consequently the magnitude of oxygen

depletion observed at any given moment (Fig. 6b,

d) is a poor guide to the frequency of mixing.

Some Type 2 lakes with either modest or strong

oxygen gradients may even circulate only once

each year, and hence classify as warm monomictic.

Accurate determination of mixing frequency in

individual Rwenzori lakes will require year-round

166 Hydrobiologia (2007) 592:151–173

123

water-column monitoring. At this time we differ-

entiate between Type 2a and Type 2b lakes

because the relative magnitude of oxygen deple-

tion has a significant influence on benthic habitat

quality, and thus potentially on benthic inverte-

brate species distribution. Finally, Type 3 lakes

(Bigata, Mahoma; Fig. 6e, f) are characterized by

a strongly clinograde oxygen curve and com-

pletely anoxic bottom waters with significantly

higher conductivity than surface waters. This

suggests that stratification persists throughout

the year, due to a chemical density difference

between bottom and surface waters. These two

lakes are meromictic, i.e. they lack at least one

complete circulation of the water column per

year. The cause of their chemical density gradient

is probably biological, i.e. a gradual accumulation

over many years of chemical substances resulting

from the decomposition of organic matter (Wet-

zel, 1983). In the case of Mahoma, at least, the

onset of meromixis sometime in the past may

have been promoted by wind shelter in its steep-

sided catchment.

Factors influencing transparency

The Secchi-depth transparency of the Rwenzori

lakes (0.60–2.80 m; mean 1.35 m) is on average

distinctly lower than that of 14 alpine lakes on

Mt. Kenya surveyed by Loffler (1968a), which

had a mean Secchi depth of 3.80 m (range 1.0–

10.0 m). Tight coupling between DOC and Fe,

and a significant inverse relationship between

Secchi depth with both DOC and Fe, in the

Rwenzori lakes indicates that high dissolved Fe

concentrations are stabilized by organic com-

plexing substances (humic acids), and that their

presence is the principal regional control on

water-column transparency. As evident from the

PCA, these three variables were all strongly tied

to the type of catchment vegetation. The least

transparent lakes (SECCHI <1.00 m: Nsuranja,

Africa, Batoda and Kanganyika) are all sur-

rounded by lush Ericaceous vegetation and/or

sizable bogs. The most transparent lakes (SEC-

CHI >1.80 m: Upper and Lower Kitandara,

Bujuku) are located in the Alpine zone near

glaciers, lack Carex bogs, and have low DOC

(<4 mg/l) and Fe (<0.1 mg/l) contents.

This differentiation between lakes with differ-

ent types of catchment vegetation also showed

from the RDA. The Nival and Alpine location of

lakes without bogs each contributes significantly

and independently to explained variation in the

lake data set, as do the combined group of Alpine

and Ericaceous locations with bogs. RDA group-

ings of all Ericaceous zone locations with and

without bogs, or all Alpine zone locations with

and without bogs, do not independently contrib-

ute to explained variation in the data set, pointing

to the decisive influence of the Carex-Sphagnum

bogs on the aquatic environment.

DOC is a key controlling factor in the under-

water penetration of solar ultraviolet radiation

(UV) (reviewed by Vincent & Pienitz, 1996; see

also Pienitz and Vincent, 2000). DOC more

strongly absorbs the biologically damaging UV-

A radiation (Scully & Lean, 1994), and is thus a

natural sunscreen for aquatic biota. Very likely,

lack of this UV protection in the more transpar-

ent Rwenzori lakes explains the presence of

Chydoridae with strongly pigmented carapaces

in some of them (Eggermont et al., in press; see

also Hansson, 2004; Tollrian & Heibl, 2004).

In glaciated mountain ranges elsewhere, low

lake-water transparency is often caused by high

loads of suspended inorganic solids in the glacial

meltwater passing through them. In comparison,

Rwenzori glaciers carry very little sediment on

their surface, and the streams flowing out from

beneath the glaciers are clear, not turbid as

typical meltwater streams (Busulwa & Bailey,

2004). Glacially-derived sediment input certainly

does affect some of the Rwenzori lakes (e.g.,

Bujuku with its well-developed delta and silty

shorelines), but overall a lake’s proximity to the

glaciers does not seem to exert a major influence

on water-column transparency.

Factors influencing water chemistry and pH

Rwenzori lake-waters are primarily derived from

surface runoff, and only secondarily from direct

rainfall and glacier melting. Hence, bedrock

composition can be expected to exert dominant

control on lake chemistry. In the adjacent

lowlands of western Uganda, most lakes occur

on volcanic bedrock (Melack, 1978), and receive

Hydrobiologia (2007) 592:151–173 167

123

high inputs of Na, Mg and K relative to Ca

(Fig. 7a). High weathering rates in the sub-humid

tropical climate there produce large amounts of

HCO3 (Fig. 7b), and evaporative concentration of

dissolved solids often leads to biologically

induced precipitation of carbonate, causing fur-

ther loss of Ca. Bedrock in the Rwenzori is more

variable. The highest peaks are composed of

volcanic amphibolite (Osmaston & Pasteur,

1972). Below it occur softer schists, and at lower

elevations the range is primarily composed of

weathered Precambrian gneiss. The somewhat

higher conductivity of lakes in the Bujuku and

Butawu river systems (18–52 lS/cm; n = 7) likely

reflect the weathering of amphibolite and schists

exposed on high peaks nearby. The alpine lakes in

this group (Bujuku, Upper and Lower Kitandara)

also have elevated SO4 concentrations. The lower

conductivity of lakes in the other river systems

(5–21 lS/cm; n = 8) reflects slow weathering of

the underlying gneiss. As expected, the lowest

conductivity values (5–8 lS/cm) occur in lakes

(Irene, Ruhandika) and pools in the Nival zone

that are surrounded by bare rock. Thus, factors

contributing to the extremely dilute character of

Rwenzori lake waters, besides slow-weathering

bedrock and low temperatures, are the steep

mountain slopes which allow percolating water to

pass quickly, and the near-continuous rainy

weather, which dilutes the scarce ions released

by weathering. Given the extremely low conduc-

tivity of most Rwenzori lakes, salts carried by rain

(Na, Cl and SO4) might have a noticeable

influence on their hydrochemistry. But in fact,

the average concentrations of conservative ions in

the lakes (see Electronic supplementary material)

are even lower than values reported for lowland

rainfall (Visser, 1961). This is likely due to the

great distance to the Indian Ocean, and efficient

rain-out of sea-spray salts before reaching Rwenz-

ori’s slopes.

The cation concentration sequence in Rwenzori

lakes and pools surveyed in this study (Ca > -

K > Na > Mg) is strongly biased toward products

of feldspar weathering, as in Congolese Rwenzori

lakes (de Heinzelin & Mollaret, 1956) but deviat-

ing from that in the lowland lakes of western

Uganda where Mg, Na and/or K concentrations

often surpass Ca. Also the anion pattern (the

SO4 > HCO3 > Cl and SO4 > Cl > HCO3 se-

quences occur in about equal proportion) deviate

from the HCO3 > Cl > SO4 sequence found in

most lowland Uganda lakes (Melack, 1978; Kil-

ham, 1971; Kizito et al., 1993; Verschuren et al.,

unpubl. data).

The acidity of many Rwenzori surface waters is

primarily due to the presence of humic substances

leached from organic top soil and the bogs (viz.

the significant inverse relationship between DOC

and pH). The circum-neutral pH of lakes in the

Bujuku and Butawu drainages (6.55–7.81; n = 7),

including DOC-stained lakes surrounded by lush

Ericaceous vegetation and/or Carex-Sphagnum

bogs (Bujuku, Middle Kachope), must primarily

derive from the buffering capacity of their catch-

ment’s rock geochemistry rather than CO2 uptake

by phytoplankton.

Nutrients and primary production

Surface-water TP concentrations in Rwenzori

lakes (3.1–68.4 lg/l) are high on average compared

to mountain lakes in Europe (2–7 lg/l; Psenner,

1989; Marchetto et al., 1994; Camarero et al., 1995;

Mosello et al. 1995; Skjelkvale & Wright, 1998;

Kopacek et al., 2000), North America (2.5–15 lg/l;

Schindler, 2000) and Asia (average 7.3 lg/l, Lacoul

& Freedman, 2005). Lakes in the Nyamughasani

and Nyamwamba valleys, in particular, contained

very high TP levels (>20 lg/l). However, since

PO4-P in all surveyed lakes was below detection

limit, and TP is positively correlated with DOC, it

follows that most of the measured TP must be

biologically unavailable phosphorus bound to

allochthonous organic matter. Lakes with a

POC:Chl-a ratio greater than 100:1 likely derive

most of their POC input from terrestrial or detrital

sources (Eppley et al., 1977). POC:Chl-a ratios in

the Rwenzori lakes (85–2,471; mean 625:1) are

decidedly extreme, again indicative that most

organic matter suspended in the water column is

produced from dead terrestrial vegetation rather

than the result of autochthonous production. This

conclusion is further supported by the carbon-to-

nitrogen ratio (C:N) of particulate organic matter

(POC/PON), with ranges from 8.3 to 12.1. Organic

matter produced by algae tends to have a C:N ratio

of 6–9, while vascular land plants produce organic

168 Hydrobiologia (2007) 592:151–173

123

matter with C:N ratios of 20 and greater (Meyers &

Teranes, 2001). Wetland soils, and Sphagnum-

dominated wetlands in particular, tend to have C:N

values of 9–17, i.e. lower than typical terrestrial

organic matter and only slightly above those of

lacustrine organic matter (Meyers & Teranes,

2001). The C:N values recorded here can be

interpreted to represent organic matter derived

primarily from the lakes’ catchments, particularly

wetlands soils. This is confirmed by both PCA

(Fig. 2) and RDA ordinations, which showed total

nutrient concentrations to be strongly linked to the

type of surrounding land cover, with the lowest

values recorded in lakes surrounded by bare rock

or alpine vegetation without bogs, and the highest

values recorded in lakes surrounded by Ericaceous

vegetation and/or bogs.

Lake productivity classifications on the basis of

epilimnetic TP, TN and Chl-a concentration (Fors-

berg & Ryding, 1980; Downing & McCauley, 1992)

identify the group of six high-alpine, transparent

Rwenzori lakes as ultra-oligotrophic (Irene) to

mesotrophic (Ruhandika) and phosphorus-limited

(TN/TP by mass: 23:1–81:1; PO4-P below detection

limit; NO3: 0.2–1.4 lg/l). In the humic, DOC-

stained lakes at lower elevation, autochthonous

primary production must be limited both by

phosphorus and light penetration. As both PO4-P

and NO3 in these lakes are below detection limit,

nitrogen limitation may also affect algal commu-

nity structure. However the very low TN:TP ratios

in some lakes (9:1 to 12:1 in Kopello, Africa and

Kanganyika) more likely results from the pre-

dominant contribution of dead, allochthonous

organic matter to total nutrient concentrations

(McNeely et al., 1979).

Summary and perspectives

In the context of ecological and biodiversity

research, aquatic habitat in mountain lakes on

the Ugandan side of the Rwenzori range is

structured along two major environmental

gradients. With regard to the limnetic habitat,

lakes differentiate between (1) slightly acidic to

pH-neutral, phosphorus-limited clear-water lakes

near or above 4,000 m elevation, with at least

some direct input of glacial meltwater and

surrounded by alpine vegetation or rocky catch-

ments; and (2) more strongly acidic, light- and

phosphorus-limited humic lakes mostly below

4,000 m elevation, remote from the glaciers and

surrounded by Ericaceous vegetation and/or bogs.

With regard to benthic habitat, the Rwenzori

lakes surveyed to date have mixing regimes

ranging from polymictic to meromictic, resulting

in hypolimnetic (profundal) oxygen regimes rang-

ing from near-saturated over mildly depleted to

completely anoxic, despite generally low primary

productivity. Excluding the mid-elevation Lake

Mahoma, the measured range in surface-water

temperature among Rwenzori lakes (5.5–9.1�C) is

relatively modest, overlapping substantially with

the range of bottom-water temperatures (4.8–

7.5�C). Hence, we do not expect the benthic fauna

of well-oxygenated profundal habitat to be mark-

edly different from that on similar substrate in

shallow-water habitat. We do expect species

turnover in profundal benthic communities of

Rwenzori lakes along the oxygen gradient con-

trolled by mixing frequency.

Alpine lakes in remote and undisturbed regions

are known early-warning systems for more wide-

spread environmental change (e.g. Schmidt &

Psenner, 1992; Sommaruga-Wograth et al., 1997;

Skjelkvale & Wright, 1998; Battarbee et al., 2002;

Rogora et al., 2003). As shown from this study, the

physical limnology, hydrochemistry and nutrient

budget of Rwenzori lakes varies closely with

climate-controlled characteristics of their local

abiotic environment, in particular air temperature,

proximity to glaciers and type of vegetation cover.

It follows that Rwenzori lake ecosystems may

indeed prove to be very sensitive to anthropogenic

climate change and glacier retreat.

In temperate regions, air temperature is rec-

ognized to influence the timing of ice break-up

and growing season in lakes (Livingstone, 1997;

Palecki & Barry, 1986). Climatic warming in the

Rwenzori may enhance the thermal stratification

of lakes and thus reduce their frequency of

complete water-column mixing and aeration.

Combined with high inputs of terrestrial organic

matter, this may result in more widespread

occurrence of low-oxygen conditions in the hypo-

limnion, profoundly changing the biogeochemis-

try and aquatic community structure of these lake

Hydrobiologia (2007) 592:151–173 169

123

systems. Rising regional air temperature will also

likely result in a gradual upward shift of vegeta-

tion belts (Beniston, 2005; Walther et al., 2005),

which will in turn affect the lakes’ nutrient budget

and DOC content, water-column transparency

and UV penetration. The low buffering capacity

of most Rwenzori lakes also makes them vulner-

able to the acidification associated with incom-

plete organic decomposition (see Psenner &

Schmidt, 1992; Schmidt & Psenner, 1992; Wright

& Schindler, 1995; Koinig et al., 1996). Glacier

retreat will cause a larger proportion of the

catchment to be exposed to physical and chemical

weathering (Wright & Schindler, 1995; Skjelkvale

& Wright, 1998) changing nutrient and ion

budgets accordingly. In comparison, the thermal

effects of changes in glacial meltwater input may

be modest, except in lakes now in close proximity

of the remnant glaciers.

The limnological data resulting from this study

may serve as an important quantitative baseline

for long-term environmental monitoring of the

Rwenzori lakes. The exact amplitude and direc-

tion of ecosystem change due to climatic warm-

ing, however, depend on site-specific processes

and are not readily predictable (Psenner &

Schmidt, 1992; Koinig et al., 1998). Use of these

limnological data to create paleoenvironmental

calibration data sets will, when applied to sedi-

mentary records, help to improve insight in the

long-term natural variability of these ecosystems

and in their actual sensitivity to climate-driven

environmental change.

Acknowledgements The fieldwork was conducted underUganda NCST research clearance NS21 and UgandaWildlife Authority permit UWA/TBDP/RES/50, withlogistic support from Rwenzori Mountaineering Services.We greatly thank Halewijn Missiaen and Bob Rumes forfield assistance, Angelica Alcantara for lab assistance,Renaat Dasseville for pigment analyses (at GhentUniversity, Belgium) and Lei Chou for TP and TNanalyses (at the Univerisite Libre de Bruxelles,Belgium). Elie Verleyen is acknowledged forconstructive comments on the statistical analyses, andSteven Declerk and two anomymous referees forsupportive critique which helped to improve themanuscript. This research was sponsored by the Fund forScientific Research of Flanders, the Leopold III-fundBelgium (Belgium), the Stichting Ter Bevordering vanhet Wetenschappelijk Onderzoek in Afrika (Belgium),and the U.S. National Geographic Society (grant 7999-06).

H.E. is a postdoctoral fellow with the Fund for ScientificResearch of Flanders (FWO-Vlaanderen).

References

Battarbee, R. W., R. Thompson, J. Catalan, J. A. Grytnes& H. J. B. Birks, 2002. Climate variability and eco-system dynamics of remote alpine and arctic lakes: theMOLAR project. Journal of Paleolimnology 28: 1–6.

Beniston, M., 2005. Climatic change and its possible im-pacts on the alpine region. Revue de Geographie –Journal of Alpine Research 93(2): 25–32.

Bradley, R. S., M. Vuille, H. F. Diaz & W. Vergara, 2006.Threats to water supplies in the tropical Andes. Sci-ence 312: 1755–1756.

Burgeon, L., 1937. Liste des Coleopteres recoltes au coursde la Mission Belge au Ruwenzori. Bruxelles-InstitutRoyal Colonial Belge, 140 pp.

Busulwa, H. S. & R. G. Bailey, 2004. Aspects of thephysico-chemical environment of the Rwenzori rivers,Uganda. African Journal of Ecology 42: 87–92.

Camarero, L., J. Catalan, A. Boggero, A. Marchetto,R. Mosello & R. Psenner, 1995. Acidification in highmountain lakes in central, southwest and southeastEurope (Alps, Pyrennes, Pirin). Limnologica 25: 141–156.

Cholnoky, B. J., 1964. Die Diatomeenflora einiger Gew-asser des Ruwenzorigebirges in Zentralafrika. NovaHedwiga 8: 55–101.

de Heinzelin, J., 1962. Carte des extensions glaciaires duRwenzori (versant Congolais). Biuletyn Peryglacjalny11: 133–139.

de Heinzelin, J. & H. Mollaret, 1956. Biotopes de hautealtitude. Ruwenzori I. Institut des Parcs Nationaux duCongo belge, Exploration du Parc National Albert,Deuxieme serie, Bruxelles.

Downing, J. A. & E. McCauley, 1992. The nitro-gen:phosphorus relationship in lakes. Limnology andOceanography 37: 936–945.

Ebinger, C. J., 1989. Tectonic development of the westernbranch of the East African Rift system. GeologicalSociety of America Bulletin 101: 885–903.

Eggermont, H., K. Van Damme & J. M. Russell, in press.Rwenzori Mountains (Mountains of the Moon):headwaters of the White Nile. In Dumont, H. J. (ed.),The Nile Basin. Monographiae Biologicae, Springer.

Forsberg, C. & S. Ryding, 1980. Eutrophication parame-ters and trophic state indices in 30 Swedish waste-receiving lakes. Archiv fur Hydrobiologie 89: 189–207.

Fries, R. E. & T. C. E. Fries, 1948. Phytogeographical re-searches on Mt. Kenya and Mt. Aberdara, BritishEast Africa. Kundliga Svenska Vetenskapsakadem-iens Handlingar 25: 1–54.

Gaffen, D. J., M. A. Sargent, R. E. Habermann & J. R.Lanzante, 2000. Sensitivity of tropospheric andstratospheric temperature trends to radiosonde dataquality. Journal of Climatology 13: 1776–1796.

170 Hydrobiologia (2007) 592:151–173

123

Grasshoff, K., M. Ehrhardt & K. Kremling, 1983. Methodsof Seawater Analysis. Verlag Chemie, Weinheim,419 pp.

Hansson, L. A., 2004. Plasticity in pigmentation induced byconflicting threats from predation and UV radiation.Ecology 85: 1005–1016.

Hastenrath, S., 2001. Variations of East African climatedruing the past two centuries. Climatic Change 50:209–217.

Hastenrath, S. & P. Kruss, 1992. Glacial fluctuation andvegetation succession on Tyndall Glacier, Mt Kenya.Mountain Research and Development 25: 68–75.

Hauer, F. R., J. S. Baron, D. H. Campbell, K. D. Fausch,S. W. Hostetler, G. H. Leavesley, P. R. Leavitt, D. M.McKnight & J. A. Stanford, 1997. Assessment of cli-mate change and freshwater ecosystems of the RockyMountains, USA and Canada. Hydrological Processes11: 903–924.

Hedberg, O., 1951. Vegetation belts of the East Africanmountains. Svensk Botanisk Tidskrift 45: 140–202.

Heron, F. A., 1909. Ruwenzori expedition reports. 12.Lepidoptera Rhopalocera. Transactions of the Zoo-logical Society of London. Volume XIX. Part II.Zoological Society of London, London.

Howard, P., 1991. Nature Conservation in Uganda’sTropical Forest Reserves. IUCN, Gland.

Hutchinson, G. E. & H. Loffler, 1956. The thermal clas-sification of lakes. Proceedings of the Natural Aca-demic Sciences USA 42: 84–86.

Kamenik, C., R. Schmidt, G. Kum & R. Psenner, 2001.The influence of catchment characteristics on thewater chemistry of mountain lakes. Arctic, Antarcticand Alpine Research 33: 404–409.

Kaser, G., D. R. Hardy, T. Molg, R. S. Bradley & T. M.Hyera, 2004. Modern glacier retreat on Kilimanjaroas evidence of climate change: observations andfacts. International Journal of Climatology 24(3):229–329.

Kaser, G. & H. Osmaston, 2002. Tropical Glaciers. Cam-bridge University Press, Cambridge, 207 pp.

Kilham, P., 1971. Biochemistry of African lakes and rivers.Unpublished PhD, Duke University, Durham, N.C.,199 pp.

Kimmins, D. E., 1959. Ruwenzori Expedition 1952. Vol-ume II. Number 8–10. 8. Odonata. 9. Trichoptera. 10.Ephemeroptera, Plecoptera and Neuroptera. Trusteesof the British Museum (Nat. Hist), London.

Kingdom, J., 1989. Island Africa. Princeton UniversityPress, Princeton, NJ, 287 pp.

Kizito, Y. S., A. Nauwereck, L. J. Chapman & W. Koste,1993. A limnological survey of some Western UgandaCrater Lakes. Limnologica 23(4): 335–347.

Koinig, K. A., C. Kamenik, R. Schmidt, A. Agusti-Pana-reda, P. Appleby, A. Lami, M. Prazakova, N. Rose,Ø. A. Schnell, R. Tessadri, R. Thompson & R. Psenner,2002. Environmental changes in an alpine lakes(Gossenkollesee, Austria) over the last two centuries– the influence of air temperature on biologicalparameters. Journal of Paleolimnology 28: 147–160.

Koinig, K. A., R. Schmidt & R. Psenner, 1996. Effects ofair temperature changes and acid deposition on the

pH history of three high alpine lakes. In Mayama,S. & I. Koizumi (eds), 14th Diatom Symposium.Koeltz Scientific Books, Koenigstein, 467–478.

Koinig, K. A., R. Schmidt, S. Wograth, R. Tessadri &R. Psenner, 1998. Climate change as the primarycause for pH shifts in a high alpine lake. Water, Airand Soil Pollution 104: 167–180.

Kopacek, J., E. Stuchlık, V. Straskrabova & P. Pseriakova,2000. Factors governing nutrient status of mountainlakes in the Tatra Mountains. Freshwater Biology 43:369–383.

Lacoul, P. & B. Freedman, 2005. Physical and chemicallimnology of 34 lentic waterbodies along a tropical-to-alpine altitudinal gradient in Nepal. InternationalReview of Hydrobiology 90(3): 254–276.

Lewis, W. M., Jr., 1983. A revised classification of lakesbased on mixing. Canadian Journal of Fisheries andAquatic Sciences 40: 1779–1787.

Livingstone, D. A., 1962. Age of deglaciation in theRuwenzori Range, Uganda. Nature 194: 859–860.

Livingstone, D. A., 1967. Postglacial vegetation of theRuwenzori Mountains in Equatorial Africa. Ecologi-cal Monographs 37: 26–51.

Livingstone, D. M., 1997. Break-up dates of alpine lakes asproxy data for local and regional mean surface airtemperatures. Climate Change 37: 407–439.

Loffler, H., 1957. Die klimatischen Typen des holomiktis-chen Sees. Mitteilungen der Geographischen Gesell-schaft 99: 35–44.

Loffler, H., 1960. Limnologische Untersuchungen an chi-lenischen und peruanischen Binnengewassern. ArchivGeophysik 3: 155–254.

Loffler, H., 1964. The limnology of tropical high mountainlakes. Verhandlungen Internationale Vereinigung derLimnologie 15:176–193.

Loffler, H., 1968a. Die Hochgebirgsseen Ostafrikas.Hochgebirgsforschung 1: 1–61.

Loffler, H., 1968b. Tropical high mountain lakes. Collo-quium Geographicum 9: 57–76.

Mahaney, W. C., 1989. Quaternary and Environmental Re-search on East African Mountains. Balkema, Rotterdam.

Marchetto, A., R. Mosello, R. Psenner, A. Barbieri,G. Bendetta, D. Tait & G. A. Tartari, 1994. Evalua-tion of the level of acidification and critical loads forAlpine lakes. Ambio 23: 150–159.

McNeeley, R. N., V. P. Neimanis & L. Dwyer, 1979. WaterQuality Source Book: A Guide to Water QualityBranch. Ministry of Supply and Services, Ottawa,Ontario, Canada, 88 pp.

Melack, J. M., 1978. Morphometrical, physical and chem-ical features of the volcanic crater lakes of westernUganda. Archiv fur Hydrobiologie 84: 430–453.

Meyers, P. A. & Teranes, J. L., 2001. Sediment organicmatter. In Last, W. M. & J. P. Smol (eds), TrackingEnvironmental Change Using Lake SedimentsVolume 2 : Physical and Geochemical Methods.Kluwer Academic Publishers, Dordrecht, 239–269.

Mollaret, H., 1961. Biotopes de haute altitude RuwenzoriII et Virunga. Institut des Parcs Nationaux du Congoet du Ruanda-Urundi, Exploration du Parc NationalAlbert, Deuxieme serie, Bruxelles.

Hydrobiologia (2007) 592:151–173 171

123

Mosello, R., A. Lami, A. Marchetto, M. Rogora, B. Wat-hne, L. Lien, J. Catalan, L. Camarero, M. Ventura,R. Psenner, K. Koinig, H. Thies, S. Sommaruga-Wograth, U. Nickus, D. Tait, B. Thaler, A. Barbeiri &R. Harriman, 2001. Trends in the water chemistry ofhigh-altitude lakes in Europe. Water, Air and SoilPollution, Focus 2: 75–89.

Mosello, R., B. M. Wathne, L. Lien & H. J. B. Birks, 1995.AL:PE projects: water chemistry and critical loads.Water, Air and Soil Pollution 85: 493–498.

Ogilvie-Grant, W. R., 1908. Ruwenzori expedition 1905–1906. 16: Aves. Transactions of the Zoological Societyof London. Volume XIX. Part IV. Zoological Societyof London, London.

Oldfield, T., 1910. Ruwenzori expedition 1905–1906. 17:Mammalia. Transactions of the Zoological Society ofLondon. Volume XIX. Part V. Zoological Society ofLondon, London.

Osmaston, H. A., 1965. The past and present climate andvegetation of Ruwenzori and its neighbourhood. PhDthesis, Oxford, 238 pp.

Osmaston, H. A., 1989. Glaciers, glaciations and equilib-rium line altitudes on the Ruwenzori. In Mahaney, W.C. (eds), Quaternary & Environmental Research onthe East African Mountains. Balkema, Rotterdam,7–30.

Osmaston, H. A., 2006. Guide to the Rwenzori. Mountainsof the Moon. Rwenzori Trust, UK, 288 pp

Osmaston, H. & G. Kaser, 2001. Glaciers and glaciations,Rwenzori Mountains National Parc. 1:100,000 Map.Henry Osmaston, Ulverston.

Osmaston, H. A. & D. Pasteur, 1972. Guide to the Ru-wenzori. The Mountains of the Moon. West ColProductions, Reading, UK.

Palecki, M. A. & R. G. Barry, 1986. Freeze-up and break-upof lakes as an index of temperature changes during thetransition seasons: a case study for Finland. Journal ofClimate and Applied Meteorology 25: 893–902.

Pienitz, R. & W. F. Vincent, 2000. Effect of climate changerelative to ozone depletion on UV exposure in sub-arctic lakes. Nature 404: 484–487.

Psenner, R., 1989. Chemistry of high-mountain lakes insiliceous catchments of the central eastern Alps.Aquatic Sciences 51: 108–125.

Psenner, R., 1999. Living in a dusty world: airborne dust asa key factor for Alpine lakes. Water, Air and Soilpollution 112: 217–227.

Psenner, R. & R. Schmidt, 1992. Climate-driven pH con-trol of remote alpine lakes and effects of acid depo-sition. Nature 356: 781–783.

Rogora, M., A. Marchetto & R. Mosello, 2001. Trends inthe chemistry of atmospheric deposition and surfacewaters in the Lago Maggoriore watershed. Hydrologyand Earth System Science 5: 379–390.

Rogora, M., R. Mosello & S. Arisci, 2003. The effect ofclimate warming on the Hydrochemistry of Alpinelakes. Water, Air and Soil Pollution 148: 347–361.

Rundel, P. W., 1994. Tropical alpine climates. In Rundel,P. W., A. P. Smith & F. C. Meinzer (eds), TropicalAlpine Environments. Plant Form and Function.Cambridge University Press, 21–44.

Ruttner, F., 1963. Fundamental of Limnology. Universityof Toronto Press, Toronto, 295 pp.

Salt, G., 1987. Insects and other invertebrate animals col-lected at high altitudes in the Rwenzori and on MountKenya. African Journal of Ecology 25: 95–106.

Schindler, D. W., 2000. Aquatic problems caused by hu-man activities in Banff National Park, Alberta, Can-ada. Ambio 29: 401–407.

Schmidt, R. & R. Psenner, 1992. Climate changes andanthropogenic impacts as causes for pH fluctuations inremote high alpine lakes. Documenta Istituto ItalianoIdrobiologia 32: 31–57.

Scully, N. M. & D. R. S. Lean, 1994. The attenuation ofultraviolet radiation in temperate lakes. Archiv furHydrobiologie, Beihefte, Ergebnisse der Limnologie43: 135–144.

Shapiro, S. S., M. B. Wilk & H. J. Chen, 1968. A com-parative study of various tests of normality. Journal ofthe American Statistical Association 63: 1343–1372.

Skjelkvale, B. L. & R. F. Wright, 1998. Sensitivity to aciddeposition and global climate change. Ambio 27(4):280–286.

Statsoft Inc., 2000. Statistica for Windows. ComputerProgram Manual, Tulsa.

Sommaruga-Wograth, S., K. A. Koinig, R. Schmidt,R. Sommaruga, R. Tessadri & R. Psenner, 1997.Temperature effects on the acidity of remote alpinelakes. Nature 386: 64–67.

Taylor, R. G., L. Mileham, C. Tindimugaya, A. Majugu,A. Muwanga & B. Nakileza, 2006. Recent glacialrecession in the Rwenzori Mountains of East Africadue to rising air temperature. Geophysical ResearchLetters 33: L10402, doi:10.1029/2006GL025962.

Temple, P. H., 1961. A final report of the Makerere Col-lege I.G.Y. expeditions to Ruwenzori. Department ofGeography, Makerere.

ter Braak, C. J. F. & P. Smilauer, 2002. CANOCO Ref-erence Manual and Canodraw for Windows User’sGuide: Software for Canonical Community Ordina-tion (version 4.5). Microcomputer Power, Ithaca, NY,USA, 500 pp.

Tollrian, R. & C. Heibl, 2004. Phenotypic plasticity in pig-mentation in Daphnia induced by UV radiation and fishkairomones. Functional Ecology 18(4): 497–502.

Vincent, W. F. & R. Pienitz, 1996. Sensitivity of high-altitude freshwater ecosystems to global change:temperature and solar ultraviolet radiation. Geosci-ence Canada 23: 231–236.

Visser, S., 1961. Chemical composition of rain water inKampala, Uganda, and its relation to meterologicaland topographical conditions. Journal of GeophysicalResearch 66: 3759–3765.

Walther, G. R., S. Beissner & C. A. Burga, 2005. Trends inthe upward shift of alpine plants. Journal of Vegeta-tion Science 16(5): 541–548.

Wetzel, R. G., 1983. Limnology (2nd edn.). SaundersPublishing, Philadelphia, PA, 767 pp.

Wetzel, R. G., 2001. Limnology: Lake and River Ecosys-tems. Academic Press, San Diego, 1006 pp

Wetzel, R. G. & E. Likens, 1990. Limnological Analysis.Springer-Verlag, London, 391 pp

172 Hydrobiologia (2007) 592:151–173

123

Wilson, S., 1995. Bird and Mammal Checklists for Ugan-da’s National Parks. Institute of Environment andNatural Resources, Makerere University, Kampala,Uganda.

Wright, S. W. & J. W. Jeffrey, 1997. High resolution sys-tem for chlorophylls and carotenoids of marine phy-toplankton. In Jeffrey, S. W., R. F. C. Mantoura &S. W. Wright (eds), Phytoplankton Pigments inOceanography: A Guide to Advanced Methods.SCOR-UNESCO, Paris, 327–341.

Wright, S. W., S. W. Jeffrey, R. F. C. Mantoura, C. A.Llewellyn, T. Bjørnland, D. Repeta & N. Welsch-

meyer, 1991. Improved HPLC method for the analysisof chlorophylls and carotenoids from marine phyto-plankton. Marine Ecology Progress Series 77:183–196.

Wright, R. F. & D. W. Schindler, 1995. Interactions of acidrain and global changes: effects on terrestrial andaquatic ecosystems. Water, Air and Soil Pollution 85:359–364.

Yeoman, G., 1989. Africa’s Mountains of the Moon.Savitri Books, London, UK.

Hydrobiologia (2007) 592:151–173 173

123


Recommended