ArticlePDF Available

Impacts of Tamarix-mediated soil changes on restoration plant growth

Authors:

Abstract and Figures

QuestionDo soils impacted by Tamarix spp. affect the growth of plants used for restoration through altered soil chemistry and/or plant-soil feedbacks? LocationThe Bighorn River, the Yellowstone River and the Fort Peck Reservoir, Montana, western USA. Methods Soil was collected from paired subsites where Tamarix was either present or absent along three water bodies. To evaluate chemical and biological soil effects on plant growth, eight plant species (Achnatherum hymenoides, Astragalus cicer, Dalea candida, Elymus lanceolatus, Leymus cinereus, Pascopyrum smithii, Ratibida columnifera and Trifolium pratense) commonly used in restoration projects at Tamarix-invaded sites were grown in the collected soil. Plant-soil feedbacks were evaluated by growing two species (D. candida and P. smithii) in greenhouse soils inoculated with small amounts of the field soils. Germination, emergence and growth characteristics were compared between Tamarix-invaded and un-invaded subsites and across water bodies. ResultsSeedling emergence and plant relative growth rate, total biomass production and allocation of resources to roots and shoots were not negatively affected in field soils or in greenhouse soil inoculated with soil from areas where Tamarix was present. In fact, overall, plants emerged earlier and produced more biomass in soils affected by Tamarix than in soils from where Tamarix was not present. These results indicate that for sites in the northern range of Tamarix, restoration would not be inhibited by Tamarix-induced soil changes. Conclusions Tamarix is a relatively new invader in the northern USA, and little is known about its impacts in this area or the potential implications for restoration. However, our results indicate that neither altered soil chemistry nor plant-soil feedbacks negatively impact native plant growth, and restoration efforts would not be hindered by Tamarix-induced changes to soil chemistry or microbiota.
Content may be subject to copyright.
Applied Vegetation Science 16 (2013) 438–447
Impacts of Tamarix-mediated soil changes on
restoration plant growth
Erik A. Lehnhoff & Fabian D. Menalled
Keywords
Exotic plants; Impacts; Invasive species;
Non-indigenous species; Plant-soil feedbacks;
Restoration; Riparian; Saltcedar;Tamarisk
Abbreviations
AMF = arbuscular mycorrhizal fungi; NIS =
non-indigenous plant species; PGC = plant
growth centre; PSF = plant-soil feedb acks; S:R
= ratio of shoots to roots; RGR = relative
growth rate
Nomenclature
USDA, NRCS (2012). The PLANTS Database
(http://plants.usda.gov, 5 November 2012)
National Plant Data Team, Greensboro, NC,
USA
Received 16 April 2012
Accepted 11 October 2012
Co-ordinating Editor: Amy Symstad
Lehnhoff, E.A. (corresponding author,
erik.lehnhoff@montana.edu) & Menalled, F.D.
(menalled@montana.edu): Department of Land
Resources and Environmental Sciences,
Montana StateUniversity, Bozeman, Montana,
USA
Abstract
Question: Do soils impacted by Tamarix spp. affect the growth of plants used for
restoration through altered soil chemistry and/or plant-soil feedbacks?
Location: The Bighorn River, the Yellowstone River and the Fort Peck Reser-
voir, Montana, western USA.
Methods: Soil was collected from paired subsites where Tamarix was either
present or absent along three water bodies. To evaluate chemical and biological
soil effects on plant growth, eight plant species (Achnatherum hymenoides,Astraga-
lus cicer,Dalea candida,Elymus lanceolatus,Leymus cinereus,Pascopyrum smithii,
Ratibida columnifera and Trifolium pratense) commonly used in restoration pro-
jects at Tamarix-invaded sites were grown in the collected soil. Plant-soil feed-
backs were evaluated by growing two species (D. candida and P. smithii)in
greenhouse soils inoculated with small amounts of the field soils. Germination,
emergence and growth characteristics were compared between Tamarix-invaded
and un-invaded subsites and across water bodies.
Results: Seedling emergence and plant relative growth rate, total biomass pro-
duction and allocation of resources to roots and shoots were not negatively
affected in field soils or in greenhouse soil inoculated with soil from areas where
Tamarix was present. In fact, overall, plants emerged earlier and produced more
biomass in soils affected by Tamarix than in soils from where Tamarix was not
present. These results indicate that for sites in the northern range of Tamarix,res-
toration would not be inhibited by Tamarix-induced soil changes.
Conclusions: Tamarix is a relatively new invader in the northern USA, and little
is known about its impacts in this area or the potential implications for restoration.
However, our results indicate that neither altered soil chemistry nor plant-soil
feedbacks negatively impact native plant growth, and restoration efforts would
not be hindered by Tamarix-induced changes to soil chemistry or microbiota.
Introduction
Restoration of plant communities after the removal of
non-indigenous plant species (NIS) is complicated by
many factors, including potential changes in soil properties
(Weidenhamer & Callaway 2010), depletion of the native
species seed bank (French et al. 2011) and increased
resource availability that promotes colonization by other
NIS (Davis et al. 2000). Also, microbe-mediated plant-soil
feedbacks (PSF) may condition the soil in a manner that
alters its suitability for the same and other plant species
(Bever et al. 1997). Plant-soil feedbacks can determine
subsequent plant growth, abundance, persistence and
competitive success through changes in decomposition
rates (Hobbie 1992; Knops et al. 2002), resource availabil-
ity (Vinton & Burke 1995; Ehrenfeld et al. 2001; Clark
et al. 2005) and/or pathogenic and mutualistic interactions
(Hamel et al. 2005; Wolfe et al. 2005).
Legacy effects of PSF are important in invaded plant
communities as they could either enhance or decrease the
performance of plants (Callaway et al. 2004; Kulmatiski &
Beard 2011). Positive PSF occur when plants accumulate
beneficial microbes in their rhizosphere, including mycor-
rhizal fungi and nitrogen-fixing bacteria, which enhance
the growth and competitive ability of conspecifics relative
to other plant species. On the other hand, negative PSF
enhance species turnover rates through the accumulation
of pathogenic microbes, parasites and herbivores in the
Applied Vegetation Science
438 Doi:10.1111/avsc.12011 ©2012 International Association for Vegetation Science
rhizosphere. In nature, positive and negative PSF seldom
occur in isolation, and there is ample evidence that the net
effect of these co-occurring events plays a vital role in eco-
system organization, functioning and dynamics (Harrison
& Bardgett 2010).
The control of Tamarix chinensis and T. ramosissima,and
hybrids of the two (Gaskin & Schaal 2002), hereafter
referred to as Tamarix, and restoration of Tamarix-invaded
sites has recently become the focus of many western land
managers (Shafroth & Briggs 2008). Tamarix are shrubs or
small trees native to Asia that have invaded riparian areas
in the southwestern and more recently the northern and
northwestern USA (Pearce & Smith 2007; Lehnhoff et al.
2011). Once established in riparian areas, Tamarix can limit
recruitment of native Populus and Salix species (Shafroth
et al. 1995) and become the dominant tree (Friedman
et al. 2005). Tamarix can alter soil chemistry by increasing
soil salinity and nutrient content (Bagstad et al. 2006;
Lehnhoff et al. 2012), reducing soil pH (Ladenburger et al.
2006) and decreasing the abundance of arbuscular mycor-
rhizal fungi (Beauchamp et al. 2005; Lehnhoff et al.
2012). While the effects of Tamarix-altered soil chemistry
on native plant establishment have been studied, the
impacts of PSF on the success of restoration plants at
Tamarix-occupied sites have not.
Restoration of Tamarix-impacted sites is difficult, and
may be unsuccessful because of interacting many factors.
Tamarix invasion is often closely related to anthropogenic
alterations of river hydrology (Stromberg et al. 2007).
Because of altered hydrology, simply removing Tamarix via
burning, herbicide or mechanical means may not lead to
the desired changes in plant communities (McDaniel &
Taylor 2003; Harms & Hiebert 2006). Invasion by other
NIS after Tamarix removal is also a concern that affects res-
toration success. In a review of restoration projects at 28
Tamarix sites in the southwestern USA, Bay & Sher (2008)
found that 19% of plant species colonizing the sites were
NIS. Finally, while it is difficult to decouple Tamarix-
induced changes to soil properties from the effects of flow
alteration, soil chemistry and PSF may have effects on the
establishment and growth of restoration plants. Indeed,
ecosystems may be altered, either directly by anthropo-
genic changes or indirectly by NIS through modified soil
chemistry or PSF, to the point where they can be consid-
ered ‘novel’ (Seastedt et al. 2008), thereby further compli-
cating restoration by making restoration targets
ambiguous.
To maximize the probability of successful restoration
after NIS management, it is crucial to first understand the
implications of such potentially novel ecosystems, includ-
ing altered soil chemistry and PSF, and second, to select
plant species adapted to the conditions present, including
both altered soil and hydrology. This study focused on
evaluating plant performance in soils altered by the pres-
ence of Tamarix. The specific objectives of this work were
to (1) assess the suitability of eight plant species (seven
indigenous and one not indigenous to Montana) for
growth at Tamarix-invaded sites on three water bodies
with different hydrology, and (2) investigate the existence
of Tamarix-induced PSF. For the first objective, we hypoth-
esized that plant species performance would vary between
soils from different water bodies because of differences in
soil biological and chemical characteristics. While there is
evidence that Tamarix increases soil nutrients (Ladenburg-
er et al. 2006; Lehnhoff et al. 2012; Meinhardt & Gehring
2012), which could be beneficial for plant growth, we
tested the common assumption that plants would perform
worse in soils from where Tamarix was present because of
increased salinity. For the second objective, no a priori
hypotheses were developed due to the lack of pre-existing
information on the potential impacts of Tamarix on PSF.
However, previous research has shown a variety of impacts
of Tamarix on soil biota that could potentially lead to nega-
tive PSF. For example, Meinhardt & Gehring (2012)
showed that the presence of Tamarix reduced the coloniza-
tion of neighbouring Populus trees by beneficial arbuscular
mycorrhizal fungi (AMF), and Moseman et al. (2009)
found altered diversity and activity of nitrogen-fixing
bacteria in Tamarix-invaded wetlands.
Methods
Site descriptions and soil collection
Three replicate sites were selected along each of three
water bodies with contrasting hydrology in Montana: the
dam-controlled Bighorn River, which experiences annual
flooding; the unregulated Yellowstone River, which also
floods annually but with higher flow than the Bighorn
River; and the Fort Peck Reservoir, which has fluctuating
water levels (Fig. 1). The Bighorn River sites included the
public access points of Arapooish, General Custer and
Grant Marsh. The Yellowstone River sites included the
public access points of Bundy Bridge, Duck Creek Bridge
and Isaac Homestead. Finally, the Fort Peck Reservoir sites
included the Dam, Dry Arm and Sand Arroyo. The oldest
living Tamarix trees at the Yellowstone, Bighorn River and
Fort Peck Reservoir sites were 23, 37 and 15 yr respectively
(Lehnhoff et al. 2011). At each of these sites we selected
two subsites one with Tamarix present and one without
Tamarix. Vegetation at the adjacent subsites was similar,
with invaded subsites with Tamarix simply adding to the
species richness but not otherwise changing species com-
position or diversity (Lehnhoff et al. 2012). At each sub-
site, 20 aliquots of soil, including the overlying plant litter,
were collected with a shovel at randomly located positions
to a depth of 15 cm and placed into two 18.9-L plastic
Applied Vegetation Science
Doi: 10.1111/avsc.12011©2012 International Association for Vegetation Science 439
E.A. Lehnhoff & F.D. Menalled Tamarix-affected soil
buckets. At the subsites with Tamarix present, soil was col-
lected from directly under Tamarix trees, where there was
generally no other vegetation present. Composite soil sam-
ples from each subsite were collected from the buckets and
analysed for organic matter (OM), nitrate (NO
3), phos-
phorus (P) and potassium (K
+
) concentrations. In a previ-
ous study, samples were also collected and analysed for
electrical conductivity (EC), pH, calcium (Ca
2+
), K
+
,
sodium (Na
+
)andmagnesium(Mg
2+
), and the sodium
adsorption ratio (SAR) was calculated. These data were
previously reported in Lehnhoff et al. (2012; Table 1).
Buckets of soil were taken to the Montana State University
Plant Growth Center (PGC) and kept in cold storage
(12.8 °C) until the experiments began ca. 4 wk later.
Plant growth in field-collected soils
The species for this study were all mycorrhizal and
included (1) four grasses Leymus cinereus (Scribn. & Merr.)
A. Love (basin wildrye), Achnatherum hymenoides Roem. &
Schult. (Indian ricegrass), Elymus lanceolatus (Scribn. & J.G.
Sm.) Gould subsp. Lanceolatus (thickspike wheatgrass) and
Pascopyrum smithii (Rydb.) A. Lo
¨ve (western wheatgrass),
and (2) four forbs Ratibida columnifera (Nutt.) Woot. &
Standl. (stillwater prairie coneflower), Astragalus cicer (Cic-
er milkvetch), Trifolium pratense L. (red clover) and Dalea
candida (Michx.) ex Willd. (antelope prairie clover). Except
for T. pratense, all species are indigenous to Montana and
all of them are commonly used for restoration. The four
grasses are specifically recommended by the Montana Nat-
ural Resources Conservation Service for re-vegetation of
Tamarix-invaded areas in Montana (USDA 2010), and the
forbs are commonly used in restoration of prairie sites that
are similar to areas above the typical river high water levels
and in the reservoir drawdown area. Seeds for all species
were obtained from the Bridger Plant Materials Center in
Bridger, Montana.
To evaluate the germinability of seeds, 20 seeds of each
species were placed on germination paper (regular weight
blue paper; Anchor Paper Co., St. Paul, MN, USA) in 11 9
11 93.5 cm germination trays, with three replicates for
each species. The trays were covered with lids, kept in the
greenhouse in ambient light and at temperatures of 21 °C
(day, 16 h) and 16.5 °C (night, 8 h), and the germination
paper was moistened daily with distilled water. The total
number of germinants in each tray was recorded at 11 d.
The growth study evaluated the joint biologically and
chemically mediated impacts of Tamarix on plant seedling
emergence and growth. Eight plant species were grown in
soils collected fromthe nine field sites (18subsites) as a com-
pleteblock,randomizeddesigntoassessthespecies’potential
for restoration planting at Tamarix-invaded areas. For these
experiments, subsite soils were individually placed into 3.8-cm
diameter by 21-cm deep pots (Ray Leach ‘Cone-tainer’,
model SC10,hereafter ‘conetainer’), with four replicates for
each species-subsite combination. Five seeds were planted
per conetainer, daily seedling emergence was recorded, and
seedlings were subsequently thinned to one per pot. Plants
were grown for 10 wk in the PGC with ambient light and
temperatures of 21 °C (day, 16 h) and 16.5 °C(night,8 h),
and watered threetimes daily with anautomatic mistingsys-
tem. Plants were then removed from the conetainers,
washed over a screen to remove soil from the roots, and the
above-and below-groundportions of theplants placedsepa-
rately into envelopes for drying. Plant material was dried at
40 °Cfor1 wk and weighedtothenearest0.001 g.
Seed germinability was calculated as the mean number
of seeds that had germinated at the end of 11 d. For seeds
in conetainers, the mean time (days) to >50% emergence
was calculated as in Menalled et al. (2005) to provide a
relative assessment of species emergence times in different
soils, with the assumption that earlier emerging species
would have a competitive advantage over undesirable
weedy species at restorations sites. Effects of the different
soils on seedling emergence were evaluated with a nested
ANOVA with subsite nested within water body. The ratio
of above- to below-ground biomass, i.e. shoot to root ratio
(S:R), was calculated as the shoot biomass divided by root
Fig. 1. Soil collection locations on the Fort Peck Reservoir, Bighorn River and Yellowstone River sites in Montana, USA.
Applied Vegetation Science
440 Doi:10.1111/avsc.12011 ©2012 International Association for Vegetation Science
Tamarix-affected soil E.A. Lehnhoff & F.D. Menalled
biomass. Nested ANOVA models were also used to evalu-
ate plant biomass and S:R, and data were log transformed
prior to analysis to meet the assumptions of normality. All
data analyses were performed in R (R version 2.12.1; R
Foundation for Statistical Computing, Vienna, AT).
Plant-soil feedback
The PSF study addressed the Tamarix biologically mediated
effects on plant growth. The study was conducted as a
complete block, randomized design with soil from the 18
subsites, two soil treatments, two plant species, two plant
harvest times and four replicates. One half of the soil from
each subsite was steam-pasteurized (Lindig soil treatment
system, 1 h at 70 °C) and the other half was untreated.
Pots (10-cm base, 16-cm top, 41-cm high; I.E.M. Plastics,
Reidsville, NC, US) were filled with a 1:1:1 mix of mineral
soil, Canadian sphagnum peat moss and washed sand, all
pasteurized. All pots were then inoculated with the field
soil, with half of them randomly receiving steamed soil
and the other half untreated soil. Soil inoculation was con-
ducted by mixing 69 cm
3
(ca. 1% of the pot volume) of the
field-collected soil within the top 2 cm of the greenhouse
soil mix and then watering. The small amount of inoculum
was used to avoid altering the chemical properties of the
greenhouse soil, while adding the soil microbiota occurring
at the field subsites. Pots were then planted with seeds of
either the forb D. candida or the grass P. smithii. Green-
house conditions were maintained at 23.9 °C (day, 16 h)
and 20 °C (night, 8 h), with light supplemented with mer-
cury vapour lamps (165 lmolm
2
s
1
). To assess changes
in relative growth rate as a function of PSF, half of the
plants were harvested at 55 d after planting and the
remaining plants were harvested at 80 d after planting.
The shoot to root ratio, S:R, was calculated as above,
and relative growth rate (RGR) was calculated as:
RGR ¼lnðB2ÞlnðB1Þ
T2T1
ð1Þ
where B
1
was the total biomass (both roots and shoots) at
the initial harvest, B
2
was the total biomass at the final
harvest, T
1
is the number of days until the initial harvest,
and T
2
is the number of days until the final harvest. Total
biomass and S:R were log transformed prior to analysis to
meet the assumptions of normality. Differences in S:R,
RGR and total biomass across soil treatments and species
with subsites nested within water bodies were evaluated
Table 1. Soil chemistry at sites with and without Tamarix.
Soil parameter Tamarix status Water body
Fort peck reservoir Bighornriver Yellowstone river
2009 soil samples
EC (dSm
1
)Absent 0.48 (0.21) 0.85 (0.42) 1.16 (0.72)
Present 0.97(0.78) 1.01 (0.30) 1.50 (1.22)
pH (standard units) Absent 7.84 (0.31) 7.54 (0.18) 7.78 (0.19)
Present 7.77 (0.32) 7.41 (0.21) 7.64 (0.16)
Ca
2+
(mmol l
1
)Absent 1.96 (1.37) 4.62 (0.68) 7.92 (5.88)
Present 5.32(6.60) 5.58 (1.62) 9.44 (8.97)
K
+
(mmol l
1
)Absent 0.56 (0.36) 2.34 (0.80) 0.45 (0.36)
Present 1.02(0.75) 2.56 (0.69) 0.56 (0.34)
Na
+
(mmol l
1
) Absent 1.17 (1.25) 0.29 (0.15) 1.43 (1.04)
Present 0.95 (0.75) 0.76 (0.46) 2.63 (3.47)
Mg
2+
(mmol l
1
)Absent 1.05 (0.61) 1.63 (0.30) 2.92 (2.27)
Present 2.89(3.43) 2.20 (0.67) 3.30 (3.04)
SAR Absent 1.29 (1.79) 0.17 (0.08) 0.63 (0.23)
Present 0.52 (0.38) 0.39 (0.25) 0.89 (0.92)
2010 soil samples
NO
3(mgkg
1
) Absent 3.50 (1.80) 11.17 (1.89) 2.83 (1.53)
Present 18.67 (27.15) 24.67 (2.52) 14.83 (15.46)
K
+
(mgkg
1
) Absent 229.67 (111.38) 281.33 (100.07) 150.33 (51.03)
Present 224.67 (57 .47) 530.00 (83.86) 220.67 (85.34)
P(mgkg
1
) Absent 4.00 (2.00) 6.67 (0.58) 6.67 (3.06)
Present 6.33 (4.93) 27.00 (4.00) 9.00 (2.65)
OM (%) Absent 1.20 (0.36) 2.50 (0.79) 0.77 (0.21)
Present 1.37 (0.21) 3.37 (1.66) 1.07 (0.64)
Means are presented with SD in parentheses. Bold indicates significant differences (P=0.05) of soil property between Tamarix-occupied and unoccupied
subsites.Data adapted from: Lehnhoff et al. (2012).
Applied Vegetation Science
Doi: 10.1111/avsc.12011©2012 International Association for Vegetation Science 441
E.A. Lehnhoff & F.D. Menalled Tamarix-affected soil
with generalized linear models (GLMs) because of missing
values (i.e. plants that did not grow). Model simplification
was implemented by removing non-significant terms and
conducting ANOVA between the models with a Chi-
squared distribution. If the P-value was >0.05, the test indi-
cated that removing the terms did not decrease the model’s
explanatory power, and the simpler model was retained.
Results
Plant growth in field-collected soil
From the germination test, the percentage (±SD) of seeds
that germinated for each species at 11 d was: L. cinereus,
41.7 ±17.6; A. hymenoides,0.0±0.0; E. lanceolatus,
30.0 ±13.2; P. smithii,28.3±7.6; D. candida,45.0±0.0;
A. cicer,3.3±2.9; T. pratense,58.3±7.6; and R. columnif-
era,95.0±0.0. The fact that A. hymenoides did not germi-
nate in the germination test (which was conducted on
germination paper rather than soil) indicated that the
seeds were either dormant or had very low viability.
A. hymenoides also exhibited very poor emergence across
all subsites in subsequent experiments; therefore, data for
this species were not included in further analysis. A. cicer
also had low germination at 11 d, but more seeds germi-
nated over time (data not shown); therefore it was
included in further analysis.
For the number of days to 50% emergence of seedlings
in the conetainers, the three-way interaction of species,
water body and subsite (i.e. the presence or absence of
Tamarix) was significant (F
18,462
=2.43, P=0.001). To
investigate these complex interactions, the species at indi-
vidual water bodies were examined separately. Only
L. cinereus (F
1,22
=337.5, P=0.011) and D. candida
(F
1,22
=580.2, P=0.004) at the Bighorn River, and
D. candida (F
1,22
=408.4, P=0.023) at Fort Peck Reservoir
exhibited differences in emergence time between subsites,
emerging a mean of 7.5, 9.8 and 8.3 d earlier, respectively,
in soil from subsites with Tamarix present.
PlantsgrowninsoilfromsiteswhereTamarix was pres-
ent generally had a higher S:R and produced more total
biomassthanthosegrowninsoilfromsubsiteswhere
Tamarix was absent (Table 2), although there were interac-
tions between species and water bodies (Table 3). Again, to
explore these interactions, species at individual water
bodies were examined separately. At the Bighorn River,
D. candida (F
1,20
=7.76, P=0.011) and T. pratense
(F
1,22
=12.75, P=0.002) allocated more resources to
above-ground biomass in soil collected from subsites with
Tamarix present, while the allocation pattern was the oppo-
site for A. cicer (F
1,21
=5.12, P=0.034) at the Yellowstone
River. At Fort Peck Reservoir, all species except T. pratense
(F
1,21
=0.17, P=0.684) allocated more resources to
above-ground biomass in soils from subsites with Tamarix.
At the water body level, with all species included, more
biomass was produced in soil from subsites where Tamarix
was present than where it was absent for all three water
bodies, but the difference was higher at FortPeck Reservoir
(+147%) than at the Bighorn (+68%) or Yellowstone
Rivers (+50%) (Table 2). At Fort Peck Reservoir, S:R was
also higher at the subsites with Tamarix than at those with-
out it.
For individual species, more total biomass was produced
in soils from subsites with Tamarix at the Bighorn River for
L. cinereus (F
1,22
=18.10, P<0.001), E. lanceolatus (F
1,22
=
7.84, P=0.010), P. smithii (F
1,22
=13.05, P=0.002) and
A. cicer (F
1,19
=17.40, P=0.001) than from susbsites
where Tamarix was absent. At the Yellowstone River,
A. cicer (F
1,21
=15.49, P=0.001), R. columnifera (F
1,21
=
8.80, P=0.007), L. cinereus (F
1,22
=18.26, P<0.001) and
E. lanceolatus (F
1,22
=18.86, P<0.001) also produced
more biomass in soils from subsites with Tamarix.This
same pattern was true at Fort Peck Reservoir for T. pratense,
R. columnifera,L. cinereus,E. lanceolatus and P. smithii.For
soils only from subsites where Tamarix was present, species
biomass production differed by water body (F
2,228
=47.03,
P<0.001), with biomass generally being the highest at
the Bighorn River, although there were interactions
between site and species (F
12,228
=4.52, P<0.001; Fig. 2).
Within water bodies at Tamarix-present subsites, there
were few differences in species performance, although
T. pratense produced more biomass than the other species
at the Bighorn (F
6,75
=7.21, P<0.001) and Yellowstone
River (F
6,77
=11.86, P<0.001) sites, and D. candida pro-
duced less biomass than E. lanceolatus or P. smithii
(F
6,76
=3.00, P=0.011) at Fort Peck Reservoir (Fig. 3).
Plant-soil feedback study
The ANOVA comparisons between RGR models indicated
that removing interaction terms between subsite (Tamarix
presence or absence) nested within water body and steam
pasteurization and species did not reduce the explanatory
power of the model (P=0.603, df
full model
=145, df
reduced
model
=160), and thus the simpler model was retained.
Removing steam pasteurization from the simplified model
reduced its explanatory power (P=0.040, df
model with
steam
=160, df
model without steam
=162), indicating that
steam pasteurization of soil inoculum affected RGR. Pas-
teurization of the soil increased RGR (P=0.020) from
0.064 to 0.084 gg
1
d
1
, suggesting that soil microbes
negatively affected plant RGR. However, the lack of signifi-
cance in interaction terms indicated that the presence or
absence of Tamarix at subsites did not affect the growth of
either species when soil from the subsites was used as inoc-
ulum for greenhouse soils. ANOVA results for S:R models
showed that neither steam pasteurization (P=0.140,
Applied Vegetation Science
442 Doi:10.1111/avsc.12011 ©2012 International Association for Vegetation Science
Tamarix-affected soil E.A. Lehnhoff & F.D. Menalled
df
model with steam
=160, df
model without steam
=162), nor the
interaction terms (P=0.626, df
full model
=145, df
reduced
model
=160) were significant. Results were similar for total
biomass, with non-significant steam pasteurization
(P=0.893, df
model with steam
=160, df
model without
steam
=162) and the interaction (P=0.502, df
full
model
=145, df
reduced model
=160) terms. These results indi-
cate that soils from subsites with and without the presence
of Tamarix did not have different effects on plant S:R or
total biomass growth.
Discussion
Plant growth results from the conetainer experiment are
not consistent with the hypothesis that plants grown in
Tamarix-affected soil would grow more poorly than in
non-affected soil; rather, they indicate either that Tamarix
conditioned the soil in a manner to make it more suitable
for growth of other plants, or that Tamarix had colonized
and occupied more fertile soils. The former possibility is
supported by numerous studies showing that NIS have leg-
acy effects on soils (Ehrenfeld 2010). For example, N-fixing
NIS can alter ecosystem nutrient dynamics by directly
increasing N levels (Vitousek & Walker 1989), litter
decomposition rates can be increased (Liao et al. 2008)
providing more nutrients, and soil chemistry can be chan-
ged through altered pH or redistribution of salts from the
lower soil profile (Vivrette & Muller 1977; Conser & Con-
nor 2009). Soil samples collected from the subsites
(Table 1) showed that concentrations of NO
3,K
+
and P at
the Bighorn River, and Ca
2+
,K
+
and Mg
2+
at Fort Peck Res-
ervoir, were over twice as high at subsites with Tamarix
than without it. Yellowstone River soils showed a similar
pattern of nutrient concentrations between Tamarix pres-
ent and absent sites, although the results were not statisti-
cally significant. Higher nutrient concentrations under
Tamarix stands have been observed by others (Bagstad
et al. 2006; Xu et al. 2006; Yin et al. 2010) and may
explain differences in plant growth in our conetainer
study. Native plant species germination and growth in
Tamarix-affected soils may be negatively affected by elevated
soil salinity [measured as electrical conductivity (EC)]. For
example, plant communities separated along salinity
gradients (EC up to 15 and 12.8 dSm
1
) in north-central
Table 2. Mean ratio of shoots to roots (S:R) and total biomass produced by plants grown in soils from subsites with or without Tamarix present.
Species S:R Total Biomass (g)
Tamarix absent Tamarix present Tamarix absent Tamarix present
Species differences
Dalea candida 1.27 (0.36) 1.56 (0.48) 0.10 (0.05) 0.12 (0.06)
Leymus cinereus 0.92 (0.36) 0.99 (0.33) 0.07 (0.05) 0.19 (0.13)
Astragalus cicer 1.41 (0.58) 1.41 (0.52) 0.07 (0.05) 0.14 (0.09)
Trifolium pratense 1.25 (0.49) 1.42 (0.40) 0.27 (0.23) 0.42 (0.40)
Ratibida columnifera 0.66 (0.22) 0.64 (0.25) 0.12 (0.09) 0.17 (0.11)
Elymus lanceolatus 0.91 (0.34) 1.10 (0.47) 0.12 (0.08) 0.21 (0.10)
Pascopyrum smithii 0.90 (0.32) 1.00 (0.33) 0.10 (0.06) 0.21 (0.12)
Combined vegetation 1.04 (0.46) 1.16 (0.50) 0.12 (0.12) 0.21 (0.20)
Water body differences
Bighorn river 1.16 (0.43) 1.35 (0.65) 0.19 (0.16) 0.32 (0.27)
Yellowstone river 1.09 (0.50) 1.00 (0.35) 0.10 (0.10) 0.15 (0.13)
Fort peck reservoir 0.87 (0.39) 1.13 (0.40) 0.07 (0.03) 0.17 (0.13)
SD in parentheses; bold indicates significant biomass differences (P0.05) between Tamarix-occupied and -unoccupied subsites.
Table 3. ANOVA table for shoot to root ratio (S:R) and total biomass of
restoration plant species (Dalea candida, Leymus cinereus, Astragalus
cicer, Trifolium pratense, Ratibida columnifera, Elymus lanceolatus and
Pascopyrum smithii) grown in soil collected from subsites with or without
Tamarix at three water bodies (Bighorn River, Fort Peck Reservoir and
Yellowstone River).
Source of variation df Sum
squares
Mean
square
FP
S:R
Species 6 33.775 5.629 61.021 <0.001
Water body 2 4.136 2.068 22.420 <0.001
Water body 9soil 3 3.303 1.101 11.935 <0.001
Species 9sater body 12 1.987 0.166 1.795 0.047
Species 9water body
9soil
18 2.261 0.126 1.361 0.146
Residuals 450 41.512 0.092
Total biomass
Species 6 49.664 8.227 21.490 <0.001
Water body 2 68.689 34.345 89.168 <0.001
Water body 9Soil 3 51.636 17.212 44.687 <0.001
Species 9water body 12 49.132 4.094 10.630 <0.001
Species 9water b ody 18 14.565 0.809 2.101 0.005
Residuals 459 173.325 0.385
Applied Vegetation Science
Doi: 10.1111/avsc.12011©2012 International Association for Vegetation Science 443
E.A. Lehnhoff & F.D. Menalled Tamarix-affected soil
Utah and on the Colorado River, respectively (Carman &
Brotherson 1982; Busch & Smith 1995). However, EC lev-
els at the water bodies we studied were all less than 2.0
dSm
1
, which is considered below the level at which
plants experience detrimental effects (Taylor & McDaniel
1998). In accordance, seedling emergence in our study
Fig. 2. Comparison of individual plant species biomass growing insoils collected from sites where Tamarix was present at three water bodies in Montana.
Shaded barsare the mean and error barsrepresent 1 SE. Within species, siteslabelled with different letters have significantly different (P<0.05) means.
Fig. 3. Comparison of plant species biomass within individual water bodies. Plants grown in conetainers in soil collected from subsites where Tamarix was
present at three water bodies in Montana. Shaded bars are the mean and error bars represent 1 SE. Within sites, species labelled with different letters have
significantly different (P<0.05) means.
Applied Vegetation Science
444 Doi:10.1111/avsc.12011 ©2012 International Association for Vegetation Science
Tamarix-affected soil E.A. Lehnhoff & F.D. Menalled
was not negatively impacted when growing in Tamarix-
invaded soils.
Results from the PSF study indicated that negative feed-
backs from Tamarix on the growth of native species did not
exist. PSF may be observed deeper in the soil profile where
soils are more affected by Tamarix roots than by plant litter.
However, PSF in the 015 cm depth that we investigated
are more directly related to restoration, as this is the depth
important for initial establishment and subsequent rooting
of the forb and grass restoration species. The results gener-
ally indicated that any biological differences in the soil
between sites where Tamarix was present and absent were
not enough to stimulate differences in plant growth. A bio-
logical factor that could potentially alter plant growth is
the presence or absence of AMF in the soil. Pringle et al.
(2009) showed that plantmycorrhizal symbioses are
affected by NIS, with implications for the plant-soil ecosys-
tems including alteration of the microbial community
(Mummey & Rillig 2006), influences on nutrient availabil-
ity, and disruption of symbiosis (Stinson et al. 2006).
Beauchamp et al. (2005) showed that Tamarix is non-my-
cotrophic, suggesting that sites occupied by Tamarix would
have lower levels of AMF propagules. Lehnhoff et al.
(2012) found that soils with Tamarix present had reduced
numbers of mycorrhizal propagules in the soil, although
this did not appear to cause differences in plant growth in
this study.
Our study addressed Tamarix impacts on soil chemistry
and microbially mediated PSF, and results indicated that
all species studied (with the exception of A. hymenoides,
which had poor germination) would be suitable for resto-
ration planting in Tamarix-impacted soil. However, water
body differences in soil and hydrology may also be impor-
tant in species success. Nutrient concentrations were gen-
erally higher at the Bighorn River than at the other water
bodies (Lehnhoff et al. 2012), and these differences could
be driving the biomass differences observed. Site hydrol-
ogy, and specifically water availability during the growing
season, was not addressed in this study, but should be con-
sidered when choosing restoration plant species. Bay &
Sher (2008) showed that proximity to perennial water was
an important factor in establishment of native species in
Tamarix site restoration. In sites located at a considerable
distance from perennial water, such as at the Fort Peck
Reservoir when the reservoir is in a period of drawdown,
drought-tolerant species would be required for restoration.
Flood-tolerant species would be necessary for the free-
flowing systems such as the Yellowstone River. Along the
dam-controlled Bighorn River, where flooding does not
occur annually, a mix of drought- and flood-tolerant spe-
cies may be more appropriate. In future research, the spe-
cies used in this study should be tested at field sites to
evaluate their performance.
Finally, it should be recognized that Tamarix is a rela-
tively new invader in the northern and northwestern
USA, being present in Montana for ca. 50 yr, and for only
1537 yr at the study sites, so the time Tamarix has had to
affect soils is limited. Soil impacts, including both chemical
and PSF, could be important factors in the future. None-
theless, with the annual flooding of the Yellowstone River,
the periodic flooding of the Bighorn River and the periodic
rise and fall of the water level in Fort Peck Reservoir,
Tamarix effects on the soil are expected to be minimal as
compared with rivers in the southwestern USA where
stream flow patterns have been greatly altered and flows
reduced (Stromberg et al. 2007). Overall results indicate
that Tamarix-affected soil does not inhibit the growth of
other plant species at sites in its northern range where it is
a relatively new invader.
Acknowledgements
This work was funded by a grant from the Montana Nox-
ious Weed Trust Fund (MDA-061 G). We thank Cathy
Zabinski for review of an earlier version of this manuscript,
and Dan Campbell and Kimberley Taylor for assistance
with soil collection and greenhouse work. Patricia Gilbert
(US Army Corps of Engineers, Fort Peck Reservoir) pro-
vided invaluable assistance in accessing sites and collecting
soil samples at Fort Peck Reservoir. We also thank two
anonymous reviewers and the associate editor for com-
ments that improved this manuscript.
References
Bagstad, K.J., Lite, S.J. & Stromberg, J.C. 2006. Vegetation, soils,
and hydrogeomorphology of riparian patch types of a dry-
land river. Western North American Naturalist 66: 2344.
Bay, R.F. & Sher, A.A. 2008. Success of active revegetation after
Tamarix removal in riparian ecosytems of the southwestern
United States: a quantitiative assessment of past restoration
projects. Restoration Ecology 16: 113128.
Beauchamp, V.B., Stromberg, J.C. & Stutz, J.C. 2005. Interac-
tions between Tamarix ramosissima (saltcedar), Populus
fremontii (cottonwood), and mycorrhizal fungi: effects on
seedling growth and plant species coexistence. Plant and Soil
275: 221231.
Bever, J.D., Westover, K.M. & Antonovics, J. 1997. Incorporat-
ing the soil community into plant population dynamics:
the utility of the feedback approach. Journal of Ecology 85:
561573.
Busch, D.E. & Smith, S.D. 1995. Mechanisms associated with
decline of woody species in riparian ecosystems of the south-
western US. Ecological Monographs 65: 347–370.
Callaway, R.M., Thelen, G.C., Rodriguez, A. & Holben, W.E.
2004. Soil biota and exotic plant invasion. Nature 427: 731
733.
Applied Vegetation Science
Doi: 10.1111/avsc.12011©2012 International Association for Vegetation Science 445
E.A. Lehnhoff & F.D. Menalled Tamarix-affected soil
Carman, J.G. & Brotherson, J.D. 1982. Comparisons of sites
infested and not infested with saltcedar (Tamarix pentandra)
and Russian olive (Elaeagnus angustifloia). Weed Science 30:
360–364.
Clark, B.R., Hartley, S.E., Suding, K.N. & de Mazancourt, C.
2005. The effect of recycling on plant competitive hierar-
chies. American Naturalist 165: 609622.
Conser, C. & Connor, E.F. 2009. Assessing the residual effects of
Carpobrotus edulis invasion, implications for restoration. Bio-
logical Invasions 11: 349358.
Davis, M.A., Grime, P.J. & Thompson, K. 2000. Fluctuating
resources in plant communities: a general theory of invasi-
bility. Journal of Ecology 88: 528534.
Ehrenfeld, J.G. 2010. Ecosystem consequences of biological
invasion s. Annual Review of Ecology, Evolution, and Systematics
41: 5980.
Ehrenfeld, J.G., Kourtev, P. & Huang, W.Z. 2001. Changes in
soil functions following invasions of exotic understory
plants in deciduous forests. Ecological Applications 11: 1287
1300.
French, K., Mason, T.J. & Sullivan, N. 2011. Recruitment limita-
tion of native species in invaded coastal dune communities.
Plant Ecology 212: 601609.
Friedman, J.M., Auble, G.T., Shafroth, P.B., Scott, M.L., Merigli-
ano, M.F., Preehling, M.D. & Griffin, E.K. 2005. Dominance
of non-native riparian trees in western USA. Biological Inva-
sions 7: 747751.
Gaskin, J.F. & Schaal, B.A. 2002. Hybrid Tamarix widespread in
US invasion and undetected in native Asian range. Proceedings
of the National Academy of Sciences of the United States of America
99: 1125611259.
Hamel, C., Vujanovic, V., Jeannotte, R., Nakano-Hylander, A. &
St-Arnaud, M. 2005. Negative feedback on a perennial crop:
fusarium crown and root rot of asparagus is related to
changes in soil microbial community structure. Plant and Soil
268: 7587.
Harms, R.S. & Hiebert, R.D. 2006. Vegetation response following
invasive tamarisk (Tamarix spp.) removal and implications
for riparian restoration. Restoration Ecology 14: 461472.
Harrison, K.A. & Bardgett, R.D. 2010. Influence of plant species
and soil conditions on plant-soil feedback in mixed grassland
communities. Journal of Ecology 98: 384395.
Hobbie, S.E. 1992. Effects of plant species on nutrient cycling.
Trends In Ecology & Evolution 7: 336339.
Knops, J.M.H., Bradley, K.L. & Wedin, D.A. 2002. Mechanisms
of plant species impacts on ecosystem nitrogen cycling.
Ecology Letters 5: 454466.
Kulmatiski, A. & Beard, K.H. 2011. Long-term plant growth
legacies overwhelm short-term plant growth effects on soil
microbial community structure. Soil Biology & Biochemistry
43: 823830.
Ladenburger, C.G., Hild, A.L., Kazmer, D.J. & Munn, L.C. 2006.
Soil salinity patterns in Tamarix invasions in the Bighorn
Basin, Wyoming, USA. Journal of Arid Environments 65: 111
128.
Lehnhoff, E., Menalled, F.D. & Rew, L.J. 2011. Tamarisk (Tamarix
spp.) establishment in its most Northern Range. Invasive Plant
Science and Management 4: 5865.
Lehnhoff, E.A., Rew, L.J., Zabinski, C.A. & Menalled, F.D. 2012.
Reduced impacts or a longer lag phase? Tamarix in the north-
western USA. Wetlands 32: 497508.
Liao,C., Peng, R., Luo,Y., Zhou, X., Wu,X., Fang, C., Chen,J. & Li,
B. 2008. Altered ecosystem carbon and nitrogen cycles by
plantinvasion: a meta-analysis.New Phytologist177: 706714.
McDaniel, K.C. & Taylor, J.P. 2003. Saltcedar recovery after her-
bicide-burn and mechanical clearing practices. Journal
of Range Management 56: 439445.
Meinhardt, K.A. & Gehring, C.A. 2012. Disrupting mycorrhizal
mutualisms: a potential mechanism by which exotic tama-
risk outcompetes native cottonwoods. Ecological Applications
22: 532549.
Menalled, F.D., Buhler, D.D. & Liebman, M. 2005. Composted
swine manure effects on germination and early growth of
crop and weed species under greenhouse conditions. Weed
Technology 19: 784789.
Moseman, S.M., Zhang, R., Qian, P.Y. & Levin, L.A. 2009. Diver-
sity and functional responses of nitrogen-fixing microbes to
three wetland invasions. Biological Invasions 11: 225239.
Mummey, D.L. & Rillig, M.C. 2006. The invasive plant species
Centaurea maculosa alters arbuscular mycorrhizal fungal com-
munities in the field. Plant and Soil 288: 8190.
Pearce, C.M. & Smith, D.G. 2007. Invasive saltcedar (Tamarix):
its spread from the American southwest to the northern
great plains. Physical Geography 28: 507530.
Pringle, A., Bever, J.D., Gardes, M., Parrent, J.L., Rillig, M.C. &
Klironomos, J.N. 2009. Mycorrhizal Symbioses and plant invasions.
Annual Review of Ecology, Evolution and Systematics 40: 699715.
Seastedt, T.R., Hobbs, R.J. & Suding, K.N. 2008. Management of
novel ecosystems: are novel approaches required? Frontiers
in Ecology and the Environment 6: 547553.
Shafroth, P.B. & Briggs, M.K. 2008. Restoration ecology and
invasive riparian plants: an introduction to the special sec-
tion on Tamarix spp. in western North America. Restoration
Ecology 16: 9496.
Shafroth, P.B., Friedman, J.M. & Ischinger, L.S. 1995. Effects of
salinity on establishment of Populus fremontii (cottonwood)
and Tamarix ramosissima (saltcedar) in southwestern United
States. Great Basin Naturalist 55: 5865.
Stinson, K.A., Campbell, S.A., Powell, J.R., Wolfe, B.E., Call-
away, R.M., Thelen, G.C., Hallett, S.G., Prati, D. & Klirono-
mos, J.N. 2006. Invasive plant suppresses the growth of
native tree seedlings by disrupting belowground mutualisms.
PLoS Biology 4: 727731.
Stromberg, J.C., Lite, S.J., Marler, R., Paradzick, C., Shafroth,
P.B., Shorrock, D., White, J.M. & White, M.S. 2007. Altered
stream-flow regimes and invasive plant species: the Tamarix
case. Global Ecology and Biogeography 16: 381393.
Taylor, J.P. & McDaniel, K.C. 1998. Restoration of saltcedar
(Tamarix sp.)-infested floodplains on the Bosque del Apache
National Wildlife Refuge. Weed Technology 12: 345352.
Applied Vegetation Science
446 Doi:10.1111/avsc.12011 ©2012 International Association for Vegetation Science
Tamarix-affected soil E.A. Lehnhoff & F.D. Menalled
USDA. 2010. Re-vegetation practices and species for Russian olive and
saltcedar management. United States Department of Agriculture,
Natural Resources Conservation Service, Invasive Species Technical
Note no. MT-30 (attachment A),Bozeman,MT,USA.
Vinton, M.A. & Burke, I.C. 1995. Interactions between individ-
ual plant species and soil nutrient status in shortgrass steppe.
Ecology 76: 11161133.
Vitousek, P.M. & Walker, L.R. 1989. Biological invasion by
Myrica faya in Hawaii plant demography, nitrogen fixation,
ecosystem effects. Ecological Monographs 59: 247265.
Vivrette, N.J. & Muller, C.H. 1977. Mechanism of invasion and
dominance of coastal grasslands by Mesembryanthemum
crystallinum.Ecological Monographs 47: 301318.
Weidenhamer, J.D. & Callaway, R.M. 2010. Direct and indirect
effects of invasive plants on soil chemistry and ecosystem
function. Journal of Chemical Ecology 36: 5969.
Wolfe, B.E., Husband, B.C. & Klironomos, J.N. 2005. Effects of a
belowground mutualism on an aboveground mutualism.
Ecology Letters 8: 218223.
Xu, W., Luo, G. & Chen, X. 2006. Soil properties under shrubs in
arid area of oasisdesert transition belt. Ying yong Sheng tai
Xue bao 17: 583586.
Yin, C.H., Feng, G., Zhang, F.S., Tian, C.Y. & Tang, C.X. 2010.
Enrichment of soil fertility and salinity by tamarisk in saline
soils on the northern edge of the Taklamakan Desert. Agricul-
tural Water Management 97: 19781986.
Applied Vegetation Science
Doi: 10.1111/avsc.12011©2012 International Association for Vegetation Science 447
E.A. Lehnhoff & F.D. Menalled Tamarix-affected soil
... Tamarix leaves have previously been noted to have more biologically available N compounds than co-occurring Salix, Populus, and Fraxinus (Going and Dudley 2008;Kennedy and Hobbie 2004). These results are in line with those of Lehnhoff and Menalled (2013), who showed that eight plant species commonly used in riparian restorations in Montana grew better on Tamarix-affected soil than on non-Tamarix soil. We also cannot rule out that increased water retention due to litter might have affected plant performance. ...
... For example, although in this experiment Aristida and Sporobolus were poor competitors compared with the exotic species, they both performed adequately and began forming flowers (R.A. Sherry, personal observation). Also, Lehnhoff and Menalled (2013) grew eight noninvasive species commonly used in Montana riparian restorations in Tamarix and non-Tamarix soils and found no negative effects of the Tamarix soils on the natives, despite increased salinity and fertility of the Tamarix soil. In some places, often immediately adjacent to water, willows and other native perennials are poised to take advantage of Tamarix dieback. ...
Article
Full-text available
Introductions of biocontrol beetles (tamarisk beetles) are causing dieback of exotic tamarisk in riparian zones across the western United States, yet factors that determine plant communities that follow tamarisk dieback are poorly understood. Tamarisk-dominated soils are generally higher in nutrients, organic matter, and salts than nearby soils, and these soil attributes might influence the trajectory of community change. To assess physical and chemical drivers of plant colonization after beetle-induced tamarisk dieback, we conducted separate germination and growth experiments using soil and litter collected beneath defoliated tamarisk trees. Focal species were two common native (red threeawn, sand dropseed) and two common invasive exotic plants (Russian knapweed, downy brome), planted alone and in combination. Nutrient, salinity, wood chip, and litter manipulations examined how tamarisk litter affects the growth of other species in a context of riparian zone management. Tamarisk litter, tamarisk litter leachate, and fertilization with inorganic nutrients increased growth in all species, but the effect was larger on the exotic plants. Salinity of 4 dS m 1 benefitted Russian knapweed, which also showed the largest positive responses to added nutrients. Litter and wood chips generally delayed and decreased germination; however, a thinner layer of wood chips increased growth slightly. Time to germination was lengthened by most treatments for natives, was not affected in exotic Russian knapweed, and was sometimes decreased in downy brome. Because natives showed only small positive responses to litter and fertilization and large negative responses to competition, Russian knapweed and downy brome are likely to perform better than these two native species following tamarisk dieback.
... Smooth brome readily establishes in sites with high resource availabilities and its cover and density increases with increasing disturbance magnitude Kenkel 2008, 2010). Alternatively, western wheatgrass grows more slowly than smooth brome grass (Launchbaugh 1964) yet can emerge earlier in soils disturbed by invasive riparian species such as salt cedar compared to noninvaded soils (Lehnhoff and Menalled 2013). The maternal environment can strongly affect the competitiveness of offspring of native and invasive perennial grasses (Chun and others 2007;Dyer and others 2010). ...
Article
Native perennial grasses in riparian areas are threatened by the invasion of weedy competitors, such as smooth brome grass (Bromus inermis Leyss. [Poaceae]). Smooth brome actively invades disturbed riparian zones through high seed production and fast seedling establishment. Restoring native perennial grasses to these regions is challenging because continual disturbances along riparian corridors open niches for further plant invasions. Identifying growth traits of common native perennial grasses, such as western wheatgrass (Pascopyrum smithii (Rydb.) Á. Löve [Poaceae]), as well as for smooth brome, should provide managers with necessary information about smooth brome invasion and native plant restoration. In this study, we sought to identify how the germination timing and rates of smooth brome and western wheatgrass were affected by seed site location and light availability. Regardless of site or light dynamics, we found that smooth brome grass germinated 2 d earlier and at a 5× higher rate than western wheatgrass. Sites that were frequently disturbed by human activity also had higher seedling germination rates than sites with little disturbance for both smooth brome and western wheatgrass. In addition, when light was restricted for the first few growing days, seedling germination of both species was higher than in diurnal light environments. These findings suggest the use of 2 management tools for enhancing western wheatgrass populations in smooth brome–invaded riparian zones: high seeding rates to improve western wheatgrass competitiveness and deep planting methods such as drill-seeding to improve germination rates and percentage.
Article
O estudo das particularidades de espécies vegetais em substrato degradado de aterro sanitário pode subsidiar o desenvolvimento de métodos que avaliem o sucesso da remediação ambiental da cobertura final, sobretudo em se tratando de avaliação do estabelecimento de plantas via sistema radicular. O trabalho avaliou o comportamento de Cajanus cajan (C.cajan), Crotalaria brefiflora (C. brefiflora), Crotalaria juncea (C. juncea) e Crotalaria ochroleuca (C. ochroleuca) em substrato degradado de aterro sanitário. O experimento foi desenvolvido no Aterro Sanitário de Belo Horizonte, MG, Brasil. Após 30 dias avaliou-se a densidade de plantas e após 12 meses o porte individual da parte área (comprimento e fitomassa seca/úmida) e da raiz (fitomassa seca/úmida e densidade radicular). A umidade e a densidade do solo foram monitoradas mensalmente. C. cajan, C. ochroleuca e C. juncea obtiveram uniformidade e crescimento inicial em 60 dias.C. juncea, C. ochroleuca e C. brefiflora atingiram todos os estágios fenológicos até o quinto mês. C. ochroleuca e C. juncea obtiveram maior comprimento de parte aérea (193.17cm e 177.56cm). A maior fitomassa seca acumulada foi obtida em C. cajan, 14.35g.ind-1 na parte aérea e 38.18g.dm-3 na raiz. C. cajan e C. brefiflora obtiveram os maiores valores de densidade de plantas (81 e 61 ind.m-2). Em caráter adaptativo, o melhor desempenho foi obtido com a espécie C. brefiflora que acumulou 30.70g.dm-3 de fitomassa na raiz, com densidade do solo 1.2g.cm-3, e C. juncea, que obteve a segunda maior produção de fitomassa na parte aérea, 64 g.ind-1 em somente 11% de umidade do solo.
Article
Invasive Australian acacias can alter soil chemistry and microbial communities in areas they invade. After clearing invasive acacias, these changes can persist, and previously invaded areas can become dominated by nitrophilic weedy species. Restoration of viable native plant communities in cleared sites often fails due to a lack of native species re-establishment. Therefore, to improve restoration outcomes, it is important to understand the effects of soil chemical and biotic legacies, and of nitrophilic weedy species, on native species re-establishment. To investigate the effect of soil chemical legacies, we germinated and grew Protea repens seedlings (a native proteoid shrub) as an indicator species in soil taken from areas cleared of Acacia saligna in lowland fynbos, as well as from non-invaded areas under controlled conditions. To investigate the effect of soil biotic legacies, we sterilized half the soil from each cleared or non-invaded area. We grew Ehrharta calycina (a native nitrophilic weedy grass species) in half of each treatment and measured the effect of treatments on P. repens germination and growth. Germination percentage, root and shoot dry mass of P. repens did not significantly differ between altered and native soil chemistry. The germination percentage of P. repens was significantly greater (93%) in the presence of soil microbial communities than in their absence. The presence of E. calycina significantly increased (29%) the root-to-shoot ratio of P. repens than their absence. Since the legacy of altered soil chemistry did not have a direct negative effect on P. repens germination and growth; we conclude that restoration efforts do not always have to manage altered soil chemistry after clearing invasive A. saligna.
Article
Full-text available
The exotic shrub Tamarix ramosissima (saltcedar) has replaced the native Populus fremontii (cottonwood) along many streams in southwestern United States. We used a controlled outdoor experiment to examine the influence of river salinity on germination and first-year survival of P. fremontii var. wislizenii (Rio Grande cottonwood) and T. ramosissima on freshly deposited alluvial bars. We grew both species from seed in planters of sand subjected to a declining water table and solutions containing 0, 1, 3, and 5 times the concentrations of major ions in the Rio Grande at San Marcial, NM (1.2, 10.0, 25.7, and 37.4 meq 1⁻¹; 0.11, 0.97, 2.37, and 3.45 dS m⁻¹). Germination of P. fremontii declined by 35% with increasing salinity (P = .008). Germination of T. ramosissima was not affected. There were no significant effects of salinity on mortality or above-and belowground growth of either species. In laboratory tests the same salinities had no effect on P. fremontii germination. P. fremontii germination is more sensitive to salinity outdoors than in covered petri dishes, probably because water scarcity resulting from evaporation intensifies the low soil water potentials associated with high salinity. River salinity appears to play only a minor role in determining relative numbers of P. fremontii and T. ramosissima seedlings on freshly deposited sandbars. However, over many years salt becomes concentrated on floodplains as a result of evaporation and salt extrusion from saltcedar leaves. T. ramosissima is known to be more tolerant of the resulting extreme salinities than P. fremontii. Therefore, increases in river salinities could indirectly contribute to decline of P. fremontii forests by exacerbating salt accumulation on floodplains.
Article
Full-text available
Saltcedar (Tamarix pentandra Pall.) and Russian olive (Elaeagnus angustifolia L.) invade moist pastures and rangeland and cause serious forage-production and soilwater losses. Our objective was to develop criteria for classifying sites relative to the likelihood of infestation by saltcedar and Russian olive, based on comparisons of soil and vegetation characteristics of infested and adjacent uninfested sites. Discriminant analyses indicated that Russian olive occurs on soils with low to medium concentrations of soluble salts (100-3500 ppm), whereas saltcedar occurs on soils with much higher soluble salt concentrations (700-15 000 ppm). Characteristics of the herbaceous vegetation on sites infested with saltcedar or Russian olive differed distinctly from each other and from adjacent, uninfested sites. Frequency of occurrence of certain herbaceous understory species provided the most accurate basis for discrimination of infested and uninfested areas. Discriminant analysis may be of value in the development of infestation-proneness indices.
Article
Full-text available
Tamarix spp. (tamarisk) have caused ecological impacts in the southwestern United States; however, such impacts have not been extensively studied in the Northwest where tamarisk is a relatively new invader. Here we present the results of soil, arbuscular mycorrhizal fungi, and vegetation studies from tamarisk-occupied and unoccupied areas on the dammed Bighorn River, Fort Peck Reservoir, and the free flowing Yellowstone River, in Montana. Soil sample results indicated that at Fort Peck Reservoir soil salinity was twice as high at occupied sites compared to unoccupied ones, and at the Bighorn River occupied sites nitrate, phosphorus and potassium were 2.2, 4, and 1.9 times higher, respectively, than at unoccupied sites. No soil differences were observed on the Yellowstone River. Mycorrhizal infectivity potential was high in both occupied and unoccupied soils, with a slight reduction (from 73% to 65% colonization) in tamarisk occupied soils. These impacts were statistically but not ecologically significant and did not extend to other metrics of impact such as richness, Simpson’s diversity or composition of plant communities. Our results indicate that either tamarisk has minimal impacts in the northwest, or it is still in a lag phase.
Article
Saltcedar ( Tamarix pentandra Pall.) and Russian olive ( Elaeagnus angustifolia L.) invade moist pastures and rangeland and cause serious forage-production and soil-water losses. Our objective was to develop criteria for classifying sites relative to the likelihood of infestation by saltcedar and Russian olive, based on comparisons of soil and vegetation characteristics of infested and adjacent uninfested sites. Discriminant analyses indicated that Russian olive occurs on soils with low to medium concentrations of soluble salts (100–3500 ppm), whereas saltcedar occurs on soils with much higher soluble salt concentrations (700–15000 ppm). Characteristics of the herbaceous vegetation on sites infested with saltcedar or Russian olive differed distinctly from each other and from adjacent, uninfested sites. Frequency of occurrence of certain herbaceous understory species provided the most accurate basis for discrimination of infested and uninfested areas. Discriminant analysis may be of value in the development of infestation-proneness indices.
Article
Vegetation development bordering the Middle Rio Grande, as with most major southwestern U.S. tributaries, has historically undergone rapid and dynamic change. The introduction of saltcedar (or Tamarisk, genus Tamarix) and other exotic species into this environment within the 20th century has contributed to this process. These plants are now an integral component of the riparian vegetation mix. Manpower, logistics, and financial resources constrain the degree to which a desired riparian habitat can be restored from saltcedar thickets on the Bosque del Apache National Wildlife Refuge near Socorro, NM. Saltcedar clearing is accomplished using a combination of herbicide, burning, and mechanical control techniques costing from $750 to $1,300/ha. Soil salinity and depth to water are the principal physical features limiting revegetation efforts. Cottonwood and black willow plantings and natural regeneration after timed irrigations have produced diverse habitats that support a wide array of faunal species in areas previously occupied by homogeneous saltcedar.
Article
The invasion of Mesembryanthemum crystallinum into coastal grassland was observed at Surf on the central coast of California, USA. The species became established in areas unoccupied or sparsely occupied by other plants, and then grew into surrounding areas. The following season, few grassland seedlings established beneath dried Mesembryanthemum as compared to adjacent grassland. The reduction in numbers of grassland seedlings did not appear to be the result of limiting levels of moisture, light or macronutrients which were found in lower levels in the grassland, nor to be due to grazing by small mammals, since the pattern persisted in the absence of grazing. The differential pattern of seedling establishment was correlated with high levels of salt found in the soil beneath dried Mesembryanthemum. Mesembryanthemum crystallinum is an annual plant which accumulates salt throughout its lifespan. After the plant dies, this salt is released with leaching by fog and rain. The salt produces a detrimental osmotic environment preventing growth of nontolerant species. The salt does not appear to have a direct toxic influence on grassland species. Osmotic interference resulting from accumulation and release of salt appears to be the means by which Mesembryanthemum crystallinum dominates areas previously occupied by grassland.
Article
Mechanical clearing and herbicide-burn treatments were compared to evaluate salteedar (Tamarix chinensis Lour.) control and recovery along the Rio Grande on the Bosque del Apache National Wildlife Refuge, Socorro, N.M. The herbicide-burn treatment included an aerial application of imazapyr (+/-)-2-[4,5dihydro-4-methyl-4-(1-methylethyl)-5-oxo-1H-imidazol-2-yl]-3- pyridinecarboxylic acid] + glyphosate [N-(phosphonomethyl)glycine] (0.6 + 0.6 kg ai hat rate) followed 3 years later by a prescription broadcast fire that eliminated > 99% of the standing dead stems. Six years after initial herbicide application, saitcedar mortality was 93%. Mechanical saltcedar clearing entailed removing aerial (trunks and stems) growth by blading, stacking and burning debris, followed by removal of underground plant portions (root crowns) by plowing, raking, and burning stacked material. Saitcedar mortality 3 years after mechanical clearing averaged 70%, which was deemed unsatisfactory. Thus, root plowing, raking, and pile burning was repeated. Three years later, after the second mechanical clearing, saitcedar mortality was 97%. Costs for the herbicide-burn treatment averaged $283 ha(-1), whereas mechanical control costs were $884 ha(-1) for the first surface and root clearing and an additional $585 ha(-1) for the second root clearing. Riparian managers should consider environmental conditions and restoration strategies prior to selecting a saltcedar control approach. Although control costs were significantly lower for the herbicide-burn treatment compared to mechanical clearing in this study, the choice of methods should always consider alternative control strategies for saitcedar. Frequently, combinations of methods result in more efficient, cost-effective results.
Article
Aim To test the hypothesis that anthropogenic alteration of stream-flow regimes is a key driver of compositional shifts from native to introduced riparian plant species. Location The arid south-western United States; 24 river reaches in the Gila and Lower Colorado drainage basins of Arizona. Methods We compared the abundance of three dominant woody riparian taxa (native Populus fremontii and Salix gooddingii, and introduced Tamarix) between river reaches that varied in stream-flow permanence (perennial vs. intermittent), presence or absence of an upstream flow-regulating dam, and presence or absence of municipal effluent as a stream water source. Results Populus and Salix were the dominant pioneer trees along the reaches with perennial flow and a natural flood regime. In contrast, Tamarix had high abundance (patch area and basal area) along reaches with intermittent stream flows (caused by natural and cultural factors), as well as those with dam-regulated flows. Main conclusions Stream-flow regimes are strong determinants of riparian vegetation structure, and hydrological alterations can drive dominance shifts to introduced species that have an adaptive suite of traits. Deep alluvial groundwater on intermittent rivers favours the deep-rooted, stress-adapted Tamarix over the shallower-rooted and more competitive Populus and Salix. On flow-regulated rivers, shifts in flood timing favour the reproductively opportunistic Tamarix over Populus and Salix, both of which have narrow germination windows. The prevailing hydrological conditions thus favour a new dominant pioneer species in the riparian corridors of the American Southwest. These results reaffirm the importance of reinstating stream-flow regimes (inclusive of groundwater flows) for re-establishing the native pioneer trees as the dominant forest type.