ChapterPDF Available

Plant mineral nutrition

Authors:

Abstract and Figures

Plant life in the kwongan has evolved on some of the world’s most nutrient-impoverished sandy soils. The availability of phosphorus (P) is particularly low on these sandy soils, but soil nitrogen (N), potassium (K) and micronutrients are also notoriously scarce (McArthur, 1991). The extreme infertility of most kwongan soils is primarily due to the low nutrient content of the parent material that gave rise to the sand and to their old age and strong degree of weathering. Over time, weathering leads to the loss of key rock-derived nutrients (e.g., P) in the absence of major soil-rejuvenating processes (e.g., glaciations, volcanic eruptions) (Walker & Syers, 1976; Laliberté et al., 2012). On the other hand, nitrogen, a nutrient derived from the atmosphere, is continuously lost from the system, predominantly as a result of fire, when most N is volatilised (Orians & Milewski, 2007). Nitrogen fixation is therefore crucially important to compensate for these losses (see also Chapter 1). As a result, widespread agricultural development of these infertile sandy soils in the 1950s commenced only after the critical need for application of P, sulfur (S), K and micronutrients, particularly copper (Cu) and zinc (Zn), for pasture establishment, had been established (Yeates, 1993).
Content may be subject to copyright.
101
CHAPTER 4: PLANT MINERAL NUTRITION
Hans Lambers, Michael W. Shane, Etienne Laliberté,
Nigel Swarts, François Teste, Graham Zemunik
INTRODUCTION
PLANT LIFE IN THE KWONGAN has evolved on some of the worlds most nutrient-impoverished sandy
soils. e availability of phosphorus (P) is particularly low on these sandy soils, but soil nitrogen (N), potassium
(K) and micronutrients are also notoriously scarce (McArthur, 1991). e extreme infertility of most kwongan
soils is primarily due to the low nutrient content of the parent material that gave rise to the sand and to their old
age and strong degree of weathering (see chapter 1A for more information). Over time, weathering leads to the
loss of key rock-derived nutrients (e.g., P) in the absence of major soil-rejuvenating processes (e.g., glaciations,
volcanic eruptions) (Walker & Syers, 1976; Laliber et al., 2012). On the other hand, N, a nutrient derived
from the atmosphere, is continuously lost from the system, predominantly as a result of re, when most N
is volatilised (Orians & Milewski, 2007). Nitrogen xation is therefore crucially important to compensate
for these losses (see also chapter 1). As a result, widespread agricultural development of these infertile sandy
soils in the 1950s commenced only after the critical need for application of P, sulfur (S), K and micronutrients,
particularly copper (Cu) and zinc (Zn), for pasture establishment, had been established (Yeates, 1993).
Given that extreme soil infertility imposes a severe constraint to plant growth, one might expect the
kwongan ora to show low diversity, composed of only a restricted number of plant species that evolved
the necessary adaption(s) to successfully grow on these infertile soils. Yet the exact opposite is actually
found, and a key feature of kwongan is its exceptionally high degree of oristic and functional diversity
(Lambers et al., 2010). Interestingly, the greatest biodiversity on the sandplains is found on the most
severely P-impoverished soils (Fig. 1).
In this chapter, we present the main nutrient-acquisition strategies displayed by kwongan species and
discuss their functioning. First, we focus on non-mycorrhizal species with specialised root adaptations
to acquire P, as they are especially abundant on the sandplains (Lambers et al., 2010). Many of these
specialised root adaptations would also enhance the acquisition of micronutrients, as discussed below.
Second, we present the dierent types of mycorrhizal strategies that are found in kwongan, focusing on
possible relationships between soil fertility and mycorrhizal type. ird, we present several symbiotic
systems that contribute to N
2
xation, including the nodules of the legume-rhizobium symbiosis (Lange,
1961), the rhizothamnia of sheoaks and Frankia, an actinomycete (Becking, 1970), and the coralloid
roots of cycads and cyanobacteria (Halliday & Pate, 1976). Finally, the specialised nutrient-acquisition
strategies of the many carnivorous and parasitic species in the kwongan are discussed.
KWONGAN PLANT LIFE
1 0 2
Figure 1. Plant diversity and soil phosphorus
(P) status in south-western Australia’s
global biodiversity hotspot. Note the relative
abundance of non-mycorrhizal species on
soils with the lowest P content / total P
concentration. Plant diversity and soil data
are from a comprehensive floristic survey
of over 1000 quadrats in the wheatbelt of
Western Australia by Gibson
et al.
(2004).
Modified after Lambers
et al.
(2010).
ADAPTATIONS TO COPE WITH A LOW P AVAILABILITY
Non-mycorrhizal strategies
On the most severely P-impoverished soils in south-western Australia, plant diversity is greatest (Lambers
et al., 2010). at is where non-mycorrhizal species rule. e number of non-mycorrhizal species
is greatest when soil P level is lowest (Fig. 1), and this is where the relative cover of non-mycorrhizal
Proteaceae species is highest (Fig. 2). Conversely, on less P-impoverished soils, mycorrhizal species
dominate (Lambers et al., 2006; 2010). is is in contrast with much younger landscapes, where non-
mycorrhizal species belonging to certain families, e.g., Amaranthaceae, Brassicaceae, Caryophyllaceae,
Chenopodiaceae, Polygonaceae and Urticaceae (Tester et al., 1987; Wang & Qiu, 2006) tend to occupy
disturbed and relatively nutrient-rich sites (Allen & Allen, 1980; Francis & Read, 1994). ese families
that are typically associated with nutrient-rich soils are poorly represented or totally absent on the south-
western Australian sandplains (Lambers & Teste, 2013).
A plant family that is richly represented on severely P-impoverished soils in the kwongan and similar
landscapes in South Africa are the Proteaceae (Cowling & Lamont, 1998; Pate & Bell, 1999). Most of
the species in this family produce proteoid roots (Fig. 3) (Purnell, 1960) and almost all of them are non-
mycorrhizal (Shane & Lambers, 2005a); an exception is the mycorrhizal species Hakea verrucosa, which
grows on the ultramac rocks of Bandalup Hill which are rich in nickel (Ni) (Boulet & Lambers, 2005).
e carboxylate-releasing strategy, which is typical for species with proteoid roots, would mobilise Ni on
ultramac soils (Robinson et al., 1996), and thus cause Ni toxicity (Lambers et al., 2008a). Proteoid roots
are dense clusters of rootlets of limited growth; the rootlets develop numerous root hairs (Lamont, 2003;
Shane & Lambers, 2005a). Since proteoid roots are not restricted to Proteaceae, the term ‘cluster roots’ is
commonly used as an alternative (Shane & Lambers, 2005a).
CHAPTER 4: PLANT MINERAL NUTRITION
103
Figure 2. The canopy cover of Proteaceae and
non-Proteaceae species as dependent on soil
phosphorus (P) concentration along the Jurien
Bay chronosequence. Proteaceae are generally
only found on P-impoverished soils, where
their cover is about half of all other species
combined. Note that the values do not add up
to 100%, because of the presence of bare soil.
On the sandplains, cluster roots are predominantly produced near the surface, just under the litter layer or
an ash bed after a re; however, when organic matter is present at greater depth, cluster roots can also be
found lower in the soil prole (Lamont, 1973). is suggests that cluster roots are produced in response
to slightly elevated levels of P, but they are suppressed at higher P supply (Lamont et al., 1984; Shane
& Lambers, 2005a). Cluster roots release vast amounts of carboxylates in an ‘exudative burst’ (Watt &
Evans, 1999; Shane et al., 2004a); the carboxylates mobilise P that is sorbed onto soil particles; the P
then becomes available for uptake by plant roots (Fig. 4) (Lambers et al., 2006). erefore, cluster roots
eectively ‘mine’ P that is unavailable for plants without this strategy. Outside the Proteaceae, cluster roots
in kwongan species can be expected in some Casuarinaceae (Reddell et al., 1997), Fabaceae (Lamont,
1972; Brundrett & Abbott, 1991) and Restionaceae species (Lambers et al., 2006), but it should be noted
that in Casuarinaceae and Restionaceae so far only deal with species outside Western Australia. Both the
Restionaceae and Anarthriaceae also produce capillaroid roots (Lamont, 1982), but whether capillaroid
roots are associated with ecient P acquisition remains to be explored.
Dauciform (i.e. carrot-shaped) roots occur in some tribes of the Cyperaceae (Davies et al., 1973;
Lamont, 1974; Shane et al., 2006b), another common predominantly non-mycorrhizal family on nutrient-
impoverished soils in the kwongan. Dauciform roots are morphologically very dierent from cluster roots
(Fig. 5), but functionally quite similar; they also release vast amounts of carboxylates in an exudative
burst, and are suppressed at high P supply (Shane et al., 2006a).
Signicant amounts of carboxylates can also be exuded by kwongan species in the absence of
specialised structures like cluster roots and dauciform roots, e.g., in a range of Kennedia species (Fabaceae)
from south-western Australia (Ryan et al., 2012). ese species dier vastly in the amount of carboxylates
they release. However, they have in common that their carboxylate exudation is reduced in the presence of
arbuscular mycorrhizal fungi, which colonise the roots of some Kennedia species. Other Kennedia species
are non-mycorrhizal (Ryan et al., 2012). is suggests a trade-o in plant carbon allocation to either
mycorrhizal fungi or carboxylate release for P acquisition.
KWONGAN PLANT LIFE
1 0 4
Figure 3. Variation in cluster-root morphology of different woody species of Proteaceae. (a) ‘Simple’ proteoid roots, typical of,
e
.
g
.,
Hakea
species, develop a
bottlebrush-like morphology. The main root is perennial and proteoid rootlet initiation (far left) to senescence (far right) occurs over approximately 21 days in
Hakea
prostrata
grown in hydroponics at extremely low [Pi] 1 µM (photos: Michael W. Shane). (b) and (c) Relatively large volumes of soil become tightly bound to maturing
proteoid roots of field-grown
Hakea
ceratophylla
(photos: Michael W. Shane). (d) ‘Compound’ proteoid roots typical of
Banksia
species develop a Christmas-tree-like
morphology. Hydroponically-grown
Banksia attenuata
showing proteoid rootlet initiation (far left) to maturity (far right) over 15 days (photos: Michael W. Shane). (e) A
thick proteoid root-mat typically develops just beneath leaf litter of field-grown
Banksia
attenuata
. (f) Rootlets of
Banksia attenuata
growing in an ash bed after a fire
(photos: Marion L. Cambridge). Scale bars in mm (a) 13; (b) 30; (c) 7; (d) 22; (e) 34; (f) 15.
a
b
e
d
c
f
CHAPTER 4: PLANT MINERAL NUTRITION
105
Figure 4. Effects of carboxylates (and other
exudates) on inorganic (P
i
) and organic P (P
o
)
mobilisation in soil. Carboxylates (organic
anions) are released via an anion channel.
The exact way in which phosphatases are
released is not known. Carboxylates mobilise
both inorganic and organic P
1
and P
o
, which
both sorb onto soil particles. The carboxylates
effectively take the place of P, thus pushing it
in solution. Phosphatases hydrolyse organic P
compounds, once these have been mobilised
by carboxylates. Carboxylates will also chelate
some of the cations that bind P, especially iron
(Fe), and other micronutrients. Chelated Fe
moves to the root surface, where it is reduced
followed by uptake by the roots, via a Fe
2+
transporter. This transporter is not specific,
but also transports other micronutrients,
such as manganese (Mn), copper (Cu) and
zinc (Zn), which have been mobilised by
carboxylates in soil. The carboxylates allow P
to be ‘mined’, as opposed to the ‘scavenging’
strategy of mycorrhizas. For further
explanation, see text; modified after Lambers
et al.
(2008a).
e leaves of Proteaceae in general (Ja, 1979; Rabier et al., 2008; Fernando et al., 2009), including
species in the kwongan (Shane & Lambers, 2005b), produce carboxylate-releasing cluster roots, and
contain relatively high levels of manganese (Mn) in their leaves. is has also been found for Fabaceae
species with cluster roots, e.g., Lupinus albus (Gardner et al., 1982) and Aspalathus linearis (Morton,
1983). is is explained by the ability of cluster roots to mobilise Mn (Gardner et al., 1981; Grierson
& Attiwill, 1989; Dinkelaker et al., 1995). Leaf Mn levels might therefore provide an indication of the
extent to which roots rely on carboxylate release to acquire P (Shane & Lambers, 2005b). Following this
approach, we recently found that non-mycorrhizal species along a 2-million-year chronosequence near
Jurien Bay (chapter 1; Laliber et al., 2012) have higher leaf Mn concentrations than their neighbours
without any of the carboxylate-releasing strategies discussed above, regardless of soil age. High Mn levels
were found in leaves that produce sand-binding roots, similar to cluster-rooted Proteaceae and dauciform-
rooted Cyperaceae (Hayes et al., 2013). It is therefore likely that the signicance of sand-binding roots in
Haemodoraceae, Restionaceae and Anarthriaceae (Fig. 5) (Shane et al., 2011; Smith et al., 2011) is also,
at least partly, that of mobilisation of P due to the release carboxylates or compounds with a similar eect.
is warrants further investigation.
KWONGAN PLANT LIFE
1 0 6
Figure 5. Variation in root morphology of different grass-like species. (a and b) Cyperaceae. (a) Field-grown
Schoenus unispiculatus
showing a shoot with attached
perennial roots and ephemeral ‘dauciform’ (= carrot-shaped) roots (arrow). Dauciform roots tightly bind numerous sand grains (inset). (b) Root axes of hydroponically-
grown plants showing numerous, relatively large, dauciform roots (
Schoenus unispiculatus
, left axis) compared with smaller dauciform roots developed in series
(three arrow heads) (
Carex fascicularis
, right axis). (c – f) Anarthriaceae (c) Field-grown
Lyginia
barbata
showing relatively large sand-covered perennial main
root axis (sand-binding root) with small ephemeral branch roots attached (capillaroid roots, arrows). Individual sand-binding roots were collected after one year’s
growth initially directed (arrow) into 500-mm PVC tubes (20 mm diameter) containing native soil. Tubes were open at the top and sealed at the bottom (inset). (d)
Remarkable development of numerous, long roots hairs on a main root, and of fine capillaroid roots of a
L
.
barbata
plant grown in hydroponics at extremely low [P]
1 µM. (e) Field-grown perennial sand-binding roots of
L
.
barbata
and (f) a thick sheath of sand grains tightly bound at the root surface (photos: Michael W. Shane).
Scale bars in mm (a) 7 (inset) 3; (b) 3; (c) 12 (inset) 52; (d) 28; (e) 21; (f) 1.5.
a
e
c
b
f
d
CHAPTER 4: PLANT MINERAL NUTRITION
107
e high abundance of Proteaceae in severely P-impoverished landscapes is not exclusively accounted for
by their very ecient P-acquisition strategy. In addition, at least three other traits determine their high P
eciency and their success on P-impoverished soils. First, their leaf P concentrations are extremely low
(Pate & Dell, 1984; Denton et al., 2006), and their rate of photosynthesis per unit leaf P is among the
highest ever recorded (Wright et al., 2004; Denton et al., 2007) which is partly accounted for by extensive
replacement of phospholipids by lipids that do not contain P (Lambers et al., 2012). Second, their long-
lived leaves are very ecient and procient at remobilising P during leaf senescence (Denton et al., 2007,
Hayes, 2014). ird, their seeds, unlike their vegetative tissues, contain very high concentrations of P
(Denton et al., 2007; Groom & Lamont, 2010). ese nutrient reserves, rather than the reserves of carbon,
determine the early growth of Hakea seedlings (Lamont & Groom, 2002). In Banksia hookeriana, half of
all the P in aboveground plant parts may be in their seeds, which comprise only half a per cent of all the
aboveground biomass (Witkowski & Lamont, 1996). Very similar results have been found for Banksia
species in South Australia (Groves et al., 1986). Seed set in Banksia species in kwongan tends to be very
low, especially in species that resprout after re; commonly only a few per cent of all the owers produce
seeds (Fuss & Sedgley, 1991; Lamont & Wiens, 2003). Since seed set can be increased by addition of
nutrients (Stock et al., 1989), low seed set appears to be a mechanism allowing seedlings to grow without
an external P source for a prolonged period (Hocking, 1982; Milberg & Lamont, 1997). Some of the
P-eciency traits that are common in Proteaceae may well occur in other kwongan species, but these have
received far less attention.
While Proteaceae and several other species that are endemic to south-western Australia are very good
at acquiring P and using it eciently, many are extremely sensitive to P-toxicity symptoms (Shane et al.,
2004b; Standish et al., 2007; Hawkins et al., 2008; Lambers et al., 2013). Even a slight increase of the low
P concentration that is common in their natural environment is enough to severely disturb their growth
and may cause death. While it is relatively common in P-impoverished landscapes, P sensitivity is by
no means universal among species from P-impoverished habitats, and even some Proteaceae species are
insensitive to elevated P supply, e.g., Grevillea crithmifolia (Shane & Lambers, 2006). e physiological
mechanism accounting for P toxicity in higher plants is their very low capacity to down-regulate their P
uptake system (Shane et al., 2004c; Shane & Lambers, 2006; de Campos et al., 2013a). eir low capacity
to down-regulate P uptake is associated with a high capacity to remobilise P from senescing leaves, and
vice versa (de Campos et al., 2013a), but whether this association is based on a mechanistic link involving
the control of P transporters is not known.
Mycorrhizal symbioses
Mycorrhizal associations can enhance P acquisition from low-P soils by their ‘scavenging’ strategy
(Lambers et al., 2008b; Smith & Read, 2008). e mycorrhizal hyphae extend the surface available for
P uptake beyond that of root hairs, exploring sites that the root hairs cannot reach. All mycorrhizal
symbioses are capable of this, including the most widespread and ancient arbuscular mycorrhizal
symbiosis. e other main mycorrhizal symbioses include ectomycorrhizas, ericoid mycorrhizas, and
orchid mycorrhizas.
Unlike ectomycorrhizas, arbuscular mycorrhizal fungi give rise to intracellular structures, that are
invariably outside the root cells’ plasma membrane (Fig. 6; Cairney, 2000). Arbuscular mycorrhizas are
KWONGAN PLANT LIFE
1 0 8
widespread on the south-western Australian sandplains, but not quite as common as in the rest of the
world (Fig. 7; Brundrett, 2009). is may reect the fact that most south-western Australian soils are very
old and thus depleted in P. At very low soil P concentrations, arbuscular mycorrhizas are not as eective in
enhancing plant P uptake (Partt, 1979), because the strategy relies mostly on ‘scavenging’ easily-available
P, and not on enhancing P availability through chemical alteration of the mycorhizosphere (Lambers et
al., 2006; 2008b). Following uptake, P is transported via hyphae towards the roots; inside the root cortex,
P is released from the hyphae in intracellular structures called arbuscules, and subsequently taken up by
root cells (Parniske, 2008). In exchange, root cells provide photosynthetically xed carbon to the fungal
partner (Smith & Read, 2008; Kiers et al., 2011).
Figure 6. Mycorrhizal structures. (ac)
Arbuscular mycorrhizal structures, including
a mycorrhizal spore (b) and an arbuscule
in a root of
Spyridium globulosum
(c). (d)
Hartig net of an ectomycorrhizal fungus
in a root of
Spyridium globulosum
. (e–f)
Ericoid mycorrhizal structures; (e)
Astroloma
xerophyllum
hair root stained with lactophenol
cotton blue. (g) Orchidaceous mycorrhizal
fungi of the genus
Pterostylis
; seedling
stem of
Pterostylis sanguinea
with hyphal
outgrowths on water agar. Photos: a–b,
François P. Teste; c–d, Graham Zemunik; e–h,
Kingsley W. Dixon.
a
e
c
g
b
f
d
CHAPTER 4: PLANT MINERAL NUTRITION
109
Figure 7. The relative importance of
specialised nutritional strategies for flowering
plants in Western Australia (
i.e.
not just
the south-west, but the south-west would
represent ¾ of the total flora) in comparison
with the whole world. Data are the ratio of
actual over expected numbers of species.
Categories of plants with bars extending to
the right of the vertical broken line are more
diverse in Western Australia than they are
on a global scale. Modified after Brundrett
(2009). EM: ectomycorrhizal species; NM all
non-mycorrhizal species; Monocot RC SB: all
species with dauciform roots (Cyperaceae),
capillaroid roots (Restionaceae and
Anarthriaceae) and sand-binding roots; AM,
arbuscular mycorrhizal.
Ectomycorrhizal symbioses are far more common in the ancient landscapes of Western Australia than
they are elsewhere (Fig. 7; Brundrett, 2009). ere is evidence that ectomycorrhizas are more eective
than arbuscular mycorrhizas at enhancing plant P uptake when soil P concentrations are lower (e.g.,
older, more strongly-weathered soils; Lambers et al., 2006; 2008b). Ectomycorrhizas function not only
as ‘scavengers’ of P, like arbuscular mycorrhizas, but they may also release carboxylates and enzymes that
give them access to organic forms of both P and N (Landeweert et al., 2001; Van Hees et al., 2006; Van
Schöll et al., 2008). e fungal hyphae usually penetrate the roots intercellularly to form the Hartig net,
where exchange of nutrients acquired by the fungal hyphae and carbon provided by the plant take place
(Smith & Read, 2008) (Fig. 6). Ectomycorrhizas are common in species belonging to Casuarinaceae,
Fabaceae, Myrtaceae and Rhamnaceae, but some of these can also form arbuscular mycorrhizal symbioses
(Brundrett, 2009) or produce cluster roots, e.g., Viminaria juncea (Lamont, 1972; de Campos et al., 2013b).
Ericoid mycorrhizas are typical for species belonging to Ericaceae (Fig. 6) (Brundrett, 2009). Similar
to arbuscular mycorrhizas, a large fraction of the fungal tissues is within the root cortical cells (Smith
& Read, 2008). Ericoid mycorrhizal roots, like ectomycorrhizas, also have access to additional chemical
pools of P. ey may release phosphatases, which enhance the availability of organic P, and exude
carboxylates, which increase the availability of sparingly soluble P (Landeweert et al., 2001; Van Leerdam
et al., 2001; Van Hees et al., 2006). As a result, it has been suggested that ericoid mycorrhizas should be
particularly eective at enhancing plant P uptake on very old, P-impoverished soils, where organic P can
represent an important fraction of the total soil P pool and where P can be strongly sorbed to soil minerals
(Lambers et al., 2008b).
About 50 plant species from the genus ysanotus form mycorrhizas that have a unique morphology.
e ‘ysanotus mycorrhizas’ are characterised by hyphae that penetrate between epidermal cells and
ramify between cortex and epidermis (McGee, 1988). Supercially, it appears like an altered morphology,
or one that is neither arbuscular mycorrhizal nor ectomycorrhizal. Interestingly, some of the fungi
responsible for ysanotus mycorrhizas do form arbuscular mycorrhizas and ectomycorrhizas on other
host plants, and some growth promotion has been observed (McGee, 1988). Studies on the Jurien Bay
chronosequence (chapter 1A), have provided evidence of additional altered mycorrhizal morphologies on
other plant genera(G. Zemunik, unpubl.). We suspect there is likely even greater diversity of nutrient-
acquisition strategies in kwongan than previously thought.
KWONGAN PLANT LIFE
11 0
Finally, orchid mycorrhizas are conned to the family Orchidaceae (Fig. 6) (Brundrett, 2009).
Like in arbuscular mycorrhizas, a large fraction of the fungal tissues is within the root cortical
cells, as fungal coils, rather than arbuscules (Smith & Read, 2008). As soon as the orchid seeds have
germinated, the seedlings, which have very few reserves, depend on nutrients contained in the organic
matter soil layer which are supplied via the mycorrhizal fungus. Most terrestrial orchid mycorrhizal
fungi are basidiomycetes and belong to the form-genus Rhizoctonia, a diverse group of asexual fungi also
comprising plant pathogens and saprophytes. Rhizoctonia species may form mycorrhizal associations
with both orchids and conifers; however, the very few conifers in kwongan have not yet been studied
in this context. Orchids with a limited capacity to photosynthesise (mycoheterotrophs), such as orchids
from south-western Australia, are generally considered parasitic on the fungus, where the association
between host and fungus does not appear to be mutually benecial (Leake, 2004). Even orchids that
have the ability to photosynthesise may form an ectomycorrhizal association with forest trees, and their
stable nitrogen- and carbon-isotope signatures indicate a dependency on ectomycorrhizas (Bidartondo et
al., 2004). However, more recent research is challenging this notion, with some photosynthetically active
orchids demonstrating net exchange of carbon between plant and fungus (Cameron et al., 2008). In the
underground orchid, Rhizanthella gardneri, which remains non-photosynthetic during its entire life cycle,
the fungus continues to play this role (Batty et al., 2004).
Orchids appear to be substantially under-represented in the Southwest Australian Floristic Region
(SWAFR), relative to plants with other nutrient-acquisition strategies (Fig. 7). is may seem surprising,
given that the region is well known for its large orchid diversity (approx. 400 species) (Brown et al.,
2008). However, the Orchidaceae contain over 25,000 species worldwide, with the vast majority of
orchids being either tree-dwelling or rock-dwelling in equatorial rainforests. e lack of epiphytes (both
orchids and non-orchids) on the south-western Australian sandplains might reect the absence of suitable
climatic conditions for this life form. e hot, dry summers and cool conditions in winter and spring
are much more suited to terrestrial species, which live underground as tubers in summer, re-sprouting
in autumn and winter, when conditions are cooler and soils moist. Epiphytes may also require some of
the traits that succulents have; this is another life style that is poorly represented in kwongan (chapter
5). Yet, it is unlikely that only these selection pressures or the competition for mycorrhizal partners to
acquire nutrients have contributed to the vast diversity in the Orchidaceae. Rather, the specialisation of
pollination systems through modication of oral signals (chapter 7B), particularly in the tropics where
competition for pollinators is ercest, have been integral to orchid diversication (Johnson & Steiner,
2000). In summary, the under-representation of orchids in kwongan (Fig. 7) is most likely due to the
scarcity of epiphytic species in kwongan, rather than specic aspects of their nutrient-acquisition strategy.
SYMBIOTIC NITROGEN FIXATION
In any terrestrial ecosystem, N is continuously lost through leaching and denitrication. In kwongan,
losses as a result of combustion of nitrogenous compounds during res are also a major component of
the N cycle at the ecosystem level. ese losses have to be compensated for the cycle to continue. While
lightning does generate nitrous oxides, this component is small (Hill et al., 1984). Global estimates of N
CHAPTER 4: PLANT MINERAL NUTRITION
111
xed by lightning are less than 10 Tg per year (Tg = 1012 g or 106 metric tons) (Galloway et al., 1995).
Biological N
2
xation accounts for most of the input into ecosystems, estimated at 90130 Tg per year on
the continents (Galloway et al., 1995; Vitousek et al., 1997).
Two techniques are commonly used to assess biological N
2
xation. e rst is qualitative; it is based
on the principle that the enzyme that reduces N
2
to NH
3
also reduces acetylene (C
2
H
2
) to ethylene (C
2
H
4
),
which is readily measured by gas chromatography (Dilworth, 1966). is technique demonstrates whether
certain structures are capable of xing N
2
or not, but it cannot reliably be used to assess how much they
are xing, because nodule activity rapidly declines in the presence of acetylene (Minchin et al., 1983). e
second is more quantitative, provided basic assumptions are met; it uses the stable isotope
15
N, which can
be measured by mass spectrometry (Minchin et al., 1983; Hansen et al., 1987). Since the
15
N abundance
of N
2
in air and of soil N tend to dier, the natural
15
N abundance is frequently used to estimate the
contribution of biological N
2
xation to the total amount of N acquired by a plant (Shearer & Kohl, 1986).
A non-xing reference plant is required to determine the
15
N abundance in the absence of N
2
xation, and
therein lays a problem in that the basic assumption is that both species access the same source of soil N.
If the mycorrhizal partners of the two plants dier, i.e. one being arbuscular mycorrhizal and the other
ectomycorrhizal, then the natural
15
N abundance technique may give erroneous results (Högberg, 1990).
If the plants that are compared access their N from dierent depth and the
15
N abundance diers with
depth, then the
15
N abundance technique cannot be used either (Robinson, 2001). Despite these problems the
natural
15
N abundance technique is a valuable tool in ecological research (Handley & Scrimgeour, 1997).
Legumes in association with rhizobia play a prominent role in N
2
xation on the south-western
Australian sandplains (Fig. 8), especially during winter after a re (Hingston et al., 1982; Hansen & Pate,
1987). Water stress is the main reason for a decrease in N
2
xation in summer (Hansen & Pate, 1987).
A decline in symbiotic N
2
xation with increase in time after a re is at least partly accounted for by a
decrease in P availability (Hingston et al., 1982; Hansen et al., 1991). Acacia pulchella, a species that is
very common following a re, sharply declines to 30% of its original high density immediately after a re
by year four, and to less than 8% 13 years after a re (Monk et al., 1981). Progressive death of the plants
in the populations returns 1.9 kg N per ha per yr; the remainder is provided by litter (1 kg N per ha per
yr) and shed seed (1 kg N per ha per yr). While this re-following legume acquires most N from soil
immediately after a re, about 70% is derived from symbiotic N
2
xation by year four after a re (Monk et
al., 1981). Legumes tend to have a high demand for P (Hartwig, 1998), which is at least partly accounted
for by rapid rates of turnover of oxygen-damaged nitrogenase, the enzyme responsible for converting N
2
into NH
3
. Nitrogenase is very sensitive to oxygen, and its repair by turnover requires ribosomal RNA,
which represents a major fraction of P in nodules. Raven (2012) estimated that a maximum of 12% of
non-storage P could occur in RNA associated with replacement of damaged nitrogenase and/or oxygen-
damage-avoidance mechanism in N
2
-xing organisms. While the dogma is that legumes have a high
demand for P, this may be biased by most studies focusing on crop species (Sprent, 1999). Moreover,
many legumes on south-western Australian sandplains also have adaptations to acquire P from low-P
soils, including cluster roots in Viminaria juncea (Lamont, 1972) and Daviesia species (Brundrett &
Abbott, 1991), release of phosphatases in Lotus australis and Kennedia prorepens (Denton et al., 2006),
and exudation of carboxylates in the absence of specialised structures, e.g., in several Kennedia species
(Pang et al., 2010; Ryan et al., 2012). Most legumes in the kwongan also establish mycorrhizal symbioses
KWONGAN PLANT LIFE
11 2
(Brundrett & Abbott, 1991).
Figure 8. Structures involved in symbiotic N
2
fixation in species of (a) sheoak, (b) legume, and (c – d) cycad. (a) Rhizothamnia (arrow heads) composed of
dichotomously branched nodules on a perennial root of
Allocasuarina humilis
collected in Lesueur National Park. Nodules contain Actinobacteria (
Frankia
). (b)
Senesced remains of short-lived cluster roots (white arrows), and living nodules (orange arrows) on a perennial root of
Viminaria juncea
(broom bush). Reddish
coloration in a cross-section of a nodule (inset) shows location of symbionts. (c) Intact apogeotropic coralloid roots, freshly unearthed and attached to (d)
Macrozamia
fraseri
where coralloid roots encircle the base of the plant and are exposed underneath a thin layer of soil or litter. (e) Blue–green coloration in a transverse section
through a coralloid root shows location of symbionts inside the root (arrow heads) (photos: Michael W. Shane). Scale bars in mm (a) 14; (b) 18; (inset) 1.5; (c) 5; (d)
24; (e) 12.
a
c
b
ed
CHAPTER 4: PLANT MINERAL NUTRITION
113
In addition to numerous legumes, sheoaks (Casuarina and Allocasuarina) also form a N
2
-xing symbiosis,
with Frankia (Actinomycota) as the microsymbiont (Fig. 8) (Bond, 1957; Becking, 1970). e symbiotic
structures in actinorhizal plants are commonly referred to as rhizothamnia (Mowry, 1933) or nodules
(Torrey & Racette, 1989). Most studies on symbiotic N
2
xation in Casuarinaceae species focused on
those that are not endemic in Western Australia, and these are the studies summarised here. ere is
distinct host specicity among Casuarina and Allocasuarina species, with some Frankia strains capable
of forming nodules on many species and others on very few (Reddell & Bowen, 1985; Torrey & Racette,
1989). As with the growth of legumes, growth of Casuarina equisetifolia, when dependent on symbiotically
xed N
2
, is more sensitive to low levels of P than is growth of seedlings supplied with nitrate; at higher
levels of P, the growth-response curves are similar for both N-fertilised and inoculated plants (Sanginga
et al., 1989). Under glasshouse conditions, nodulated plants of C. cunninghamiana and C. equisetifolia grow
vigorously in nutrient solution free of an N source other than N
2
(Bond, 1957). Under eld conditions, C.
equisetifolia can obtain up to 67% of its N from symbiotic N
2
xation (Parrotta et al., 1994). Since studies
on Casuarinaceae species in the kwongan are rare, we can only extrapolate from studies on species that
occur elsewhere, leading to the impression that sheoaks could contribute signicantly to biological N
2
xation on the south-western Australian sandplains.
Coralloid roots are structures arising from the roots of cycads (Grobbelaar et al., 1971; Lindblad et al.,
1985) (Fig. 8). ese structures can host cyanobacterial symbionts of the genus Nostoc and Calothrix, with
little evidence for host specialisation (Gehringer et al., 2010). Precoralloid structures form in the absence
of a microsymbiont (Wittmann et al., 1965; Nathanielsz & Sta, 1975; Ahern & Sta, 1994). Like other
cyanobacteria, those in coralloid roots produce neurotoxins, and this accounts for the severe toxicity of
cycad tissues (Lindblad et al., 1990; Charlton et al., 1992). e microsymbonts in coralloid roots may also
produce other toxins, including nodularin (Gehringer et al., 2012). e toxicity of Macrozamia riedlei to
cattle was well-known to early settlers, and known as ‘rickets’ or ‘wobbles’ (Gardiner & Bennetts, 1956;
Gabbedy et al., 1975). Coralloid roots x N
2
at physiologically signicant rates, mainly during the wet
season in winter, capable of doubling plant N content every 811 years (Halliday & Pate, 1976). Grove
et al. (1980) found that the ratio of weight of coralloid roots to weight of boles of M. riedlei plants was
greatest on a recently burnt site. Concentrations of N in coralloid roots were signicantly higher in plants
growing on a recently burnt site. is suggests that rates of N
2
xation are driven by the demand of rapidly
regrowing leaves after a re.
By examining trade-os inherent in plant carbon, N and P capture, Houlton et al. (2008) suggest
a clear advantage to symbiotic N
2
-xing species in some P-limited systems in that these species invest
N into P-acquisition, i.e. phosphatases, which provide them access to organic P. is can be a major
component of total soil P in old soils (Turner et al., 2013). Indeed, in their global comparison, Houlton
et al. (2008) found that soil phosphatase activities under plants known to be capable of symbiotic N
2
xation are three times higher than those in soil sampled beneath non-xing species. is claim could
be broadened in that some N
2
-xing species, including kwongan species, also invest N in Rubisco to
assimilate carbon, of which a large fraction is released as carboxylates from cluster roots, e.g., in Viminaria
juncea (Lamont, 1972) and Casuarina species (Reddell et al., 1997), or from non-cluster roots, e.g., in
Cullen, Glycine, Lotus and Kennedia species (Denton et al., 2006; Pang et al., 2010; Ryan et al., 2012).
Symbiotic N
2
xation may be an important trait to enhance release of phosphatases and carboxylates in
P-limited ecosystems (Houlton et al., 2008), but this remains to be evaluated in the kwongan.
In addition to N
2
xation occurring in clearly dened symbiotic structures involving higher plants
(Fig. 8) and in soil crusts (Pate, 1998), there is increasing evidence that this process also occurs in the
absence of such structures by endophytic bacteria. is can be a signicant source of N for sugarcane in
Brazil (Boddey et al., 2003), and likely also plays a role in trees (Bal et al., 2012) and other plants growing
on N-poor substrates (Reinhold-Hurek & Hurek, 2011). We envisage that non-symbiotic N
2
xation may
also play a role in kwongan, especially on very young dunes where the soil N levels are very low, without
a prominent presence of the symbiotic systems, discussed above, contributing to the system when soil N
accumulates (Laliber et al., 2012; Hayes et al., 2013).
ACQUISITION OF MICRONUTRIENTS
Soils on south-western Australia’s sandplains are notoriously low in micronutrients, especially Cu and Zn
(Yeates, 1993). With the exception of Mn and Ni, micronutrient levels in Banksia leaves are less than what
is generally considered ‘sucient’ for crop growth (Denton et al., 2007). High levels of Ni are remarkable,
since this is a micronutrient for legumes and a few other species that metabolise urea (Broadley et al.,
2012), and we are unaware of any function of Ni in Proteaceae. Leaf Mn concentrations are most likely
high because the activity of cluster roots that mobilises P through release of carboxylates and protons
will also enhance the mobility of Mn (Godo & Reisenauer, 1980; Jauregui & Reisenauer, 1982). Perhaps
the same mechanism accounts for high Ni concentrations (Lee et al., 1978). Release of citrate can also
mobilise Zn, but its eect depends on soil type (Duner et al., 2012) (Fig. 4). Exudates released from
tomato and spinach roots mobilise both Zn and Cu (Degryse et al., 2008). It is therefore highly likely that
the exudates released by cluster roots of Proteaceae grown under low-P conditions will not only mobilise
P, but also micronutrients. Since P is the major limiting macronutrient in kwongan soil (Laliber et al.,
2012), it is unlikely that specic adaptations evolved in Proteaceae to acquire micronutrients.
Despite a likely capacity to mobilise Zn in the rhizosphere, Zn concentrations in leaves of a range
of Banksia and other species (Denton et al., 2007; Hayes et al., 2014) are lower than what is considered
sucient. e same is true for co-occurring Calothamnus quadridus, Allocasuarina humilis, Banksia sessilis
and Xanthorrhoea preissii, but not for Eucalyptus todtiana (Myrtaceae) and Jacksonia furcella (Fabaceae)
(Pate & Dell, 1984). We have to bear in mind that low nutrient concentrations partly result from a
dilution eect’ by large amounts of sclerenchymatic tissue in kwongan plants. ere is no information
in the literature to indicate how these plants can function at such remarkably low Zn levels. Zinc plays a
role in a very wide range of enzymes in all organisms; in humans, 10% of all proteins require Zn (Broadley
et al., 2012). Zinc can play both a catalytic and a structural role in proteins. Zinc deciency reduces the
activity of a Zn-dependent superoxide dismutase, which play a role in scavenging toxic reactive oxygen
species (Cakmak, 2000).
Since species without specialised carboxylate-releasing roots coexist with those that produce cluster
roots or dauciform roots in kwongan, can the latter enhance nutrient acquisition by plants that lack
specialised roots? Muler et al. (2014), indeed, obtained evidence for a positive eect of Banksia attenuata
(Proteaceae) on the growth of co-occurring Scholtzia involucrata (Myrtaceae). is positive eect of
KWONGAN PLANT LIFE
11 4
Banksia attenuata cannot be explained by mobilisation of P, since addition of P did not enhance the
growth of Scholzia involucrata. e Mn uptake and leaf Mn concentration, and possibly those of other
micronutrients, are enhanced in S. involucrata when grown together with the cluster-rooted B. attenuata.
It is likely that facilitation is the result of the nutrient-mobilising strategy of B. attenuata, based on release
of carboxylates, as discussed above. It has yet to be explored how common facilitation by carboxylate-
releasing species on severely P-impoverished soils is.
CARNIVOROUS PLANTS
Carnivorous and protocarnivorous plants are relatively common on the south-western Australian
sandplains (Fig. 7). In Western Australia there are 5.5 times more carnivorous species than expected
based on the total number of species (Brundrett, 2009). Carnivorous plants tend to be non-mycorrhizal
and use their carnivorous habit as an alternative nutrient-acquisition strategy (Brundrett, 2009). In
kwongan, carnivorous species belong to the genera Byblis, Cephalotus, Drosera and Utricularia (Table 1).
e only known protocarnivorous genus is Stylidium (Darnowski et al., 2006); protocarnivorous species
lack the easily recognisable insect-trapping structures of true carnivorous species, but they do obtain
nutrients from captured prey.
Table 1. Genera of the different groups of parasitic species of the south-western Australian sandplains and the families they belong to. For references, see text.
GENUS FAMILY COMMON NAME PARASITIC HABIT
Cassytha Lauraceae Dodder laurel Holoparasitic stem parasite
Cuscuta Convolvulaceae Dodder Holoparasitic stem parasite
Pilostyles Apodanthaceae Holoparasitic stem parasite
Orobanche Orobanchaceae Broomrape Holoparasitic root parasite
Amyema Loranthaceae Mistletoe Hemiparasitic stem parasite
Euphrasia Orobanchaceae Eye–bright Hemiparasitic root parasite
Exocarpus Santalaceae Ballart Hemiparasitic root parasite
Leptomeria Santalaceae Currant bush Hemiparasitic root parasite
Nuytsia Loranthaceae Christmas tree Hemiparasitic root parasite
Olax Olacaceae Hemiparasitic root parasite
Santalum Santalaceae Quandong, sandalwood Hemiparasitic root parasite
CHAPTER 4: PLANT MINERAL NUTRITION
115
Figure 9. Carnivorous plants in the kwongan, and their flypaper, pitcher or suction traps. (a)
Drosera erythrorhiza
; (b)
Drosera pallida
; (c)
Drosera heterophylla
; (d)
Cephalotus follicularis
(Albany pitcher plant); (e)
Byblis gigantea
; (f):
Utricularia
menziesii
and (g) its belowground suction traps (utricles); photos: b, c, Graham
Zemunik; a and d–g, Hans Lambers.
a b c
d
f
e
g
KWONGAN PLANT LIFE
11 6
Byblis and Drosera species produce ypaper traps, i.e. leaves with glandular emergences (tentacles) that
secrete glistening, adhesive glue drops for attracting and capturing prey (Fig. 9). Watson et al. (1982)
concluded that captured arthropods could provide all the N and P that was required for growth of Drosera
erythrorhiza, but only a negligible proportion of K. Feeding
15
N-labelled Drosophila ies to leaf rosettes
of the tuberous Drosera erythrorhiza leads to 76% transfer of the
15
N in the prey to the plants. Most of
the acquired N is transferred to the tubers, and then to the new rosette during the next season (Dixon
et al., 1980). Drosera closterostigma and D. glanduIigera show an increase in growth, N and P content, and
reproductive performance from articial feeding of arthropods, but no apparent benets from minerals
alone or additive eects of minerals above that due to prey (Karlsson & Pate, 1992). Using natural
15
N
abundance of a range of Drosera species in their natural habitats, neighbouring non-carnivorous plants
and arthropods near or on each Drosera species, Schulze et al. (1991) estimated that for self-supporting
erect and climbing species 50% of their N was derived from their prey. Lower values were found for
rosette species.
Cephalotus follicularis, the Albany pitcher plant, is the only species in the family of Cephalotaceae and
not related to other pitcher plants (Chase et al., 2009). Using a similar
15
N natural abundance method as
discussed above for Drosera, Schulze et al. (1997) showed a dependence on soil N until four pitchers had
opened. Beyond that stage, plant size increased with the number of catching pitchers, but the fraction of
soil N remained high. Large pitcher plants derive a quarter of their N from insects.
Utricularia is a genus of carnivorous plants from wet habitats, producing suction traps under water or
in very wet soil (Chase et al., 2009). e bladders have a lower hydrostatic pressure than the surrounding
water, so when the trap door opens, upon touching a sensitive hair near the entrance, the prey is sucked
in (Sydenham & Findlay 1975). Using labelled N and P, it has been shown that the traps export 30% of
the N from captured prey to growing leaves within two days (Friday & Quarmby, 1994). e plants also
heavily depend on prey as a source of P (Adamec, 2013).
Stylidium is considered a protocarnivorous genus, because some species in this genus trap small insects
such as gnats and midges, using mucilage-secreting glandular hairs on their inorescences and stems, very
similar to the ypaper traps referred to above (Darnowski et al., 2006; Chase et al., 2009). Many species of
the genus form ectomycorrhiza-like structures (Warcup, 1980) as well as forming arbuscular mycorrhizas
(Brundrett & Abbott, 1991).
Given that P, rather than N, is the major limiting nutrient in the kwongan, it would be interesting
to know what fraction of their total P is derived from the prey they caught and digested. To assess this,
we need to know the N:P ratio of plant and insect material. For tubers of Drosera erythrorhyza, the N:P
is 3.810 (Pate & Dixon, 1978). For the insects that were captured, this ratio is about 8 (Pate & Dixon,
1978). If we use these gures to get a rough idea about the fraction of plant P that is derived from prey,
then we would conclude that it is perhaps similar to the fraction of N derived from prey. However, it is
also likely that the N locked up in the preys chitin is not available for the plant, so perhaps the fraction is
even greater.
CHAPTER 4: PLANT MINERAL NUTRITION
117
PARASITIC PLANTS
Although there are many non-mycorrhizal parasitic plants on the south-western Australian sandplains,
as a fraction of the total ora there are actually only about half as many species when compared with
global gures (Fig. 7). is is in stark contrast with other non-mycorrhizal plants with cluster roots, of
which there are 8.5 times more than expected based on global data (Brundrett, 2009). As discussed above,
the non-mycorrhizal species in kwongan tend to be the ones on the most P-impoverished soils, and this
also pertains to parasites (Fig. 10). Although parasitism is generally considered an alternative nutrient-
acquisition strategy, the under-representation of parasites in terms of number of species in the kwongan
ora suggests that it is not a very successful one, possibly because of its inherent risks, when compared
with other strategies. After discussing the mechanisms that parasites use to acquire nutrients from their
host, we will explore what those risks might be.
Figure 10. Parasitic plant species richness
as dependent on total soil phosphorus (P)
concentration. Plant diversity and soil data
are from a comprehensive floristic survey
of over 1000 quadrats in the wheatbelt of
Western Australia by Gibson
et al.
(2004).
Parasitic plants in the kwongan belong to holoparasitic stem parasites (Cassytha, Cuscuta, Pilostyles),
holoparasitic root parasites (Orobanche, athough it is not quite clear if it is native or alien), hemiparasitic stem
parasites (Amyema) or hemiparasitic root parasites (Euphrasia, Exocarpus, Leptomeria, Nuytsia, Olax, Santalum)
(http://orabase.dpaw.wa.gov.au/) (Table 1; Fig. 11). Pilostyles species in kwongan are unusual in that they live
as endophytes, embedded in the stems of Daviesia, Gastrolobium or Jacksonia species, except for their owers
and fruits (Dell et al., 1982; iele et al., 2008) (Fig. 12). All other parasitic plants on the sandplains form
a haustorium, which connects to the phloem (in holoparasites) or the xylem (in hemiparasites) (Fig. 13)
(Lambers et al., 2008a). ere are no studies on the functioning of kwongan holoparasites, which are not as
common as hemiparasites on the sandplains, so what follows is based on information on other holoparasitic
species. Holoparasites do not contain chlorophyll and thus depend on their host for a supply of both carbon and
nutrients. In their haustoria, they increase leakage of sugars and nutrients from the phloem, and rapidly take up
what is released. In contrast, hemiparasites do contain chlorophyll and connect via their haustoria to the xylem.
KWONGAN PLANT LIFE
11 8
Figure 11. Hemiparasitic and holoparasitic species in the kwongan. (a) The root hemiparasite
Nuytsia floribunda
(the Western Australian Christmas tree) in Lesueur
National Park (photo: Marion L. Cambridge). (b)
Nuytsia floribunda
, covered by the stem holoparasite
Cassytha
(dodder laurel) (photo: Graham Zemunik). (c)
Cassytha
attached to a twig of
Melaleuca
(photo: H. Lambers). (d) The stem hemiparasite
Amyema miquelii
(mistletoe) on
Corymbia calophylla
(marri) (photo: H. Lambers). (e)
The root hemiparasite
Santalum acuminatum
(quandong). (f) Two individuals of the hyperparasitic stem hemiparasite
Amyema miraculosa
on the root hemiparasite
Santalum acuminatum
. (g) Young individual of
Amyema miraculosa
on
Santalum acuminatum
(photos: Hans Lambers); (h) flowering
Amyema miraculosa
(photo: G.
Zemunik).
ey require a more negative water potential (greater suction tension) in their xylem than that of their host
in order to provide a gradient for water to move in the xylem towards the parasite (chapter 5). e parasite
thus imports the nutrients dissolved in the xylem sap. is may include some carbon, in the form of amino
acids and organic acids, but no sugars (Tennakoon et al., 1997b). Unlike holoparasites, hemiparasites can
photosynthesise, though rates may be low (Lambers et al., 2008).
b dc
e f hg
a
CHAPTER 4: PLANT MINERAL NUTRITION
119
Figure 12. Flowers (A) and fruits (B) of
Pilostyles hamiltonii
growing on stems of
Daviesia angulata in situ
(Thiele
et al.
, 2008)
(photos: Kevin R. Thiele).
Haustoria allow holoparasites to tap into their hosts phloem (Fig. 13). e haustoria of most root
hemiparasites look like suction cups (Fig. 13), but those of Nuytsia oribunda (Western Australian
Christmas tree) have a remarkably dierent appearance (Herbert, 1919; Fineran, 1985). A
sclerenchymatous ‘horn’ or ‘prong’ formed within the haustorium acts as a sickle-like cutting device,
which transversely severs the host root and then becomes lodged in haustorial collar tissue directly
opposite to that where it originated (Calladine & Pate, 2000).
Both stem and root hemiparasites that are common on the south-western Australian sandplains
typically have a more negative water potential (greater suction tension) than their hosts: Amyema
tzgeraldii (Davidson & Pate, 1992), A. linophyllum (Davidson et al., 1989), A. miquelii (Whittington &
Sinclair, 1988; Miller et al., 2003), Olax phyllanthi (Pate et al., 1990), Nuytsia oribunda (Calladine &
Pate, 2000) and Santalum acuminatum (Loveys et al., 2001). e greater suction tension is generally due to
high transpiration rates of the hemiparasite, a signicant resistance to the water ow in the haustoria, or a
combination of both (Hellmuth, 1971; Whittington & Sinclair, 1988; Davidson & Pate, 1992; Cernusak
et al., 2004). Rapid transpiration rates without similarly rapid rates of photosynthesis accounts for the low
water-use eciency of hemiparasites (Davidson et al., 1989; Davidson & Pate, 1992).
e host range of stem hemiparasites (mistletoes) tends to be narrow (Davidson et al., 1989; Davidson
& Pate, 1992), but that of root hemiparasites is generally very wide (Pate et al., 1990; Tennakoon et al.,
1997a; Calladine et al., 2000). Santalum acuminatum (quandong) has a wide host range, but, based on
the similarity in natural
15
N abundance, it appears to predominantly parasitise N
2
-xing hosts (legumes
and Allocasuarina) (Tennakoon et al., 1997a). Haustoria are not simply the organ that allows access to
the host, but may also allow selectivity of what is derived from the host and even convert some of the
compounds that arrive in the xylem (Lamont & Southall, 1982; Pate, 2001).
e strategy deployed by the many parasitic species in the kwongan appears to be a highly specialised
way to obtain nutrients; yet parasites are under-represented in the kwongan ora, whereas other nutrient-
acquisition strategies are abundant (Fig. 7). Why is that so? We surmise that there are major costs
associated with the parasitic lifestyle. For hemiparasites to maintain a high suction tension requires either
a b
KWONGAN PLANT LIFE
1 2 0
rapid transpiration rates or a low haustorial or stem conductance as well as accumulation of large amounts
of osmotic solutes, e.g., mannitol in Santalum acuminatum (quandongs) (Loveys et al., 2001). ere are also
risks. Stomatal control to allow transpiration to continue at much more negative water potentials than
occur in their hosts may cause dehydration, xylem cavitation and possibly death when the hosts can no
longer provide sucient water. For mistletoes, there is the additional risk due to re, which may require
re-establishment following re-introduction of seed from populations that were not aected by re. To
some extent, this may also be true for dodder laurels (Cassytha), but these can also germinate in soil, and
do not require a host for germination.
Figure 13. Root haustoria of hemiparasitic species. (a) Collar-like morphology of haustorium (white) of the hemiparasitic tree,
Nuytsia floribunda
(Western Australian
Christmas tree), encircling a host root. (bd) Root haustoria of hemiparasitic
Santalum acuminatum
(quandong, Santalaceae). (b) Cap-like morphology of haustorium
(orange) latched onto host root of a
Melaleuca
sp. A single fine root (arrow) connects the haustorium to the parasitic plant (arrow). (c and d) Hand-cut sections of
fresh tissue stained with phloroglucinol/HCl. Xylem vessel elements stained red. (c) Longitudinal section of haustorium in (b) with attached host-root (hr) in cross-
section. A pair of overarching bundles, each composed of numerous strands of xylem (red), originate close to the point of fine root attachment at the dorsal surface of
the haustorium (arrow in (b)). Bundles fan out over outermost layers of the host root (hr) xylem vessels. (d) Higher magnification of (c) show xylem strands made up
of short vessel elements joined in series up to the interface where xylem elements of the haustorium join with the host root xylem vessels (photos: Michael W. Shane).
Scale bars in mm (a) 4; (b) 2; (c) 1; (d) 0.3.
b
d
a
c
CHAPTER 4: PLANT MINERAL NUTRITION
121
CONCLUDING REMARKS
e ora on the sandplains exhibits an astounding number of specialised nutrient-acquisition and
nutrient-use strategies on some of the worlds most nutrient-impoverished soils. While mycorrhizal
associations are common, non-mycorrhizal species are prominently present on the most infertile sites,
because of their strategy to ‘mine’ P and micronutrients, rather than ‘scavenge’ for it, as mycorrhizas do.
Fascinating discoveries have been made since the original book covered plant life on the sandplain (Pate
& Beard, 1984). Much is still to be learned, especially if we aim to apply some of the knowledge gained
for agriculture in a world with a growing population and a diminishing availability of rock phosphate
(Cordell et al., 2009; Lambers et al., 2011).
REFERENCES
Adamec L. 2013. Foliar mineral nutrient uptake in carnivorous
plants: what do we know and what should we know? Frontiers in
Plant Science 4, article 10.
Ahern CP, Sta IA. 1994. Symbiosis in cycads: e origin
and development of coralloid roots in Macrozamia communis
(Cycadaceae). American Journal of Botany 81: 1559–1570.
Allen EB, Allen MF. 1980. Natural re-establishment of vesic-
ular-arbuscular mycorrhizae following stripmine reclamation in
Wyoming. Journal of Applied Ecology 17: 139147.
Bal A, Anand R, Berge O, Chanway CP. 2012. Isolation and
identication of diazotrophic bacteria from internal tissues
of Pinus contorta and uja plicata. Canadian Journal of Forest
Research 42: 807–813.
Batty AL, Dixon KW, Brundrett MC, Sivasithamparam K.
2004. Orchid conservation and mycorrhizal associations. In:
Sivasithamparam K, Dixon KW, Barrett RL eds. Microorganisms
in plant conservation and biodiversity. Dordrecht: Kluwer
Academic Publishers, 195–226.
Becking J. 1970. Plant-endophyte symbiosis in non-leguminous
plants. Plant and Soil 32: 611654.
Bidartondo MI, Burghardt B, Gebauer G, Bruns TD, Read
DJ. 2004. Changing partners in the dark: isotopic and molecu-
lar evidence of ectomycorrhizal liaisons between forest orchids
and trees. Proceedings of the Royal Society B: Biological Sciences 271:
1799–1806.
Boddey R, Urquiaga S, Alves BR, Reis V. 2003. Endophytic
nitrogen xation in sugarcane: present knowledge and future
applications. Plant and Soil 252: 139–149.
Bond G. 1957. e development and signicance of the root nod-
ules of Casaurina. Annals of Botany 21: 373–380.
Boulet F, Lambers H. 2005. Characterisation of arbuscu-
lar mycorrhizal fungi colonisation in cluster roots of shape
Hakea verrucosa F. Muell (Proteaceae), and its eect on growth
and nutrient acquisition in ultramac soil. Plant and Soil 269:
357–367.
Broadley M, Brown P, Cakmak I, Rengel Z, Zhao F. 2012.
Function of nutrients: micronutrients. In: Marschner P ed.
Marschner’s Mineral Nutrition of Higher Plants, ird Edition.
London: Academic Press, 191–248.
Brown AP, Dundas P, Dixon KW, Hopper SD. 2008. Orchids of
Western Australia. Perth: University of Western Australia Press
Brundrett MC. 2009. Mycorrhizal associations and other means
of nutrition of vascular plants: understanding the global diversity
of host plants by resolving conicting information and develop-
ing reliable means of diagnosis. Plant and Soil 320: 3777.
Brundrett MC, Abbott LK. 1991. Roots of jarrah forest plants.
I. Mycorrhizal associations of shrubs and herbaceous plants.
Australian Journal of Botany 39: 445457.
Cairney JWG. 2000. Evolution of mycorrhiza systems.
Naturwissenschaften 87: 467–475.
Cakmak I. 2000. Possible roles of zinc in protecting plant cells
from damage by reactive oxygen species. New Phytologist 146:
185–205.
Calladine A, Pate JS. 2000. Haustorial structure and func-
tioning of the root hemiparastic tree Nuytsia oribunda (Labill.)
R.Br. and water relationships with its hosts. Annals of Botany 85:
723–731.
Calladine A, Pate JS, Dixon KW. 2000. Haustorial develop-
ment and growth benet to seedlings of the root hemiparasitic
tree Nuytsia oribunda (Labill.) R.Br. in association with various
hosts. Annals of Botany 85: 733–740.
Cameron DD, Johnson I, Read DJ, Leake JR. 2008. Giving and
receiving: measuring the carbon cost of mycorrhizas in the green
orchid, Goodyera repens. New Phytologist 180: 176–184.
Cernusak L, Pate J, Farquhar G. 2004. Oxygen and carbon iso-
tope composition of parasitic plants and their hosts in southwest-
ern Australia. Oecologia 139: 199–213.
Charlton TS, Marini AM, Markey SP, Norstog K, Duncan
MW. 1992. Quantication of the neurotoxin 2-amino-3-(me-
thylamino)-propanoic acid (BMAA) in cycadales. Phytochemistry
31: 3429–3432.
Chase MW, Christenhusz MJM, Sanders D, Fay MF. 2009.
Murderous plants: Victorian Gothic, Darwin and modern
insights into vegetable carnivory. Botanical Journal of the Linnean
Society 161: 329–356.
KWONGAN PLANT LIFE
1 2 2
Cordell D, Drangert J-O, White S. 2009. e story of phos-
phorus: global food security and food for thought. Global
Environmental Change 19: 292–305.
Cowling RM, Lamont BB. 1998. On the nature of Gondwanan
species ocks: Diversity of Proteaceae in Mediterranean
south-western Australia and South Africa. Australian Journal of
Botany 46: 335–355.
Darnowski DW, Carroll DM, Płachno B, Kabano E,
Cinnamon E. 2006. Evidence of protocarnivory in triggerplants
(Stylidium spp.; Stylidiaceae). Plant Biology 8: 805812.
Davidson NJ, Pate JS. 1992. Water relations of the mistle-
toe Amyema tzgeraldii and its host Acacia acuminata. Journal of
Experimental Botany 43: 1549–1555.
Davidson NJ, True KC, Pate JS. 1989. Water relations of the
parasite: host relationship between the mistletoe Amyema lino-
phyllum (Fenzl) Tieghem and Casuarina obesa Miq. Oecologia 80:
321–330.
Davies J, Briarty LG, Rieley JO. 1973. Observations on the
swollen lateral roots of the Cyperaceae. New Phytologist 72:
167–174.
de Campos MCR, Pearse SJ, Oliveira RS, Lambers H. 2013a.
Down-regulation of net phosphorus uptake capacity is inversely
related to leaf phosphorus resorption in four species from a
phosphorus-impoverished environment. Annals of Botany 111:
445454.
de Campos MCR, Pearse SJ, Oliveira RS, Lambers H. 2013b.
Viminaria juncea does not vary its shoot phosphorus concentra-
tion and only marginally decreases its mycorrhizal colonization
and cluster-root dry weight under a wide range of phosphorus
supplies. Annals of Botany 111: 801–809.
Degryse F, Verma VK, Smolders E. 2008. Mobilization of Cu
and Zn by root exudates of dicotyledonous plants in resin-bu-
ered solutions and in soil. Plant and Soil 306: 6984.
Dell B, Kuo J, Burbidge A. 1982. Anatomy of Pilostyles hamiltonii
C. A. Gardner (Raesiaceae) in stems of Daviesia. Australian
Journal of Botany 30: 1–9.
Denton MD, Sasse C, Tibbett M, Ryan MH. 2006. Root
distributions of Australian herbaceous perennial legumes in
response to phosphorus placement. Functional Plant Biology 33:
1091–1102.
Denton MD, Veneklaas EJ, Freimoser FM, Lambers H. 2007.
Banksia species (Proteaceae) from severely phosphorus-impover-
ished soils exhibit extreme eciency in the use and re-mobiliza-
tion of phosphorus. Plant, Cell and Environment 30: 15571565.
Dilworth MJ. 1966. Acetylene reduction by nitrogen-x-
ing preparations from Clostridium pasteurianum. Biochimica et
Biophysica Acta 127: 285–294.
Dinkelaker B, Hengeler C, Marschner H. 1995. Distribution
and function of proteoid roots and other root clusters. Botanica
Acta 108: 193–200.
Dixon K, Pate J, Bailey W. 1980. Nitrogen nutrition of the tuber-
ous sundew Drosera erythrorhiza Lindl. with special reference
to catch of arthropod fauna by its glandular leaves. Australian
Journal of Botany 28: 283297.
Duner A, Hoand E, Temminghof EJM. 2012.
Bioavailability of zinc and phosphorus in calcareous soils as
aected by citrate exudation. Plant and Soil 361: 165–175.
Fernando DR, Baker AJM, Woodrow IE. 2009. Physiological
responses in Macadamia integrifolia on exposure to manganese
treatment. Australian Journal of Botany 57: 406413.
Fineran BA. 1985. Graniferous tracheary elements in haustoria
of root parasitic angiosperms. e Botanical Review 51: 389441.
Francis R, Read DJ. 1994. e contributions of mycorrhizal
fungi to the determination of plant community structure. Plant
and Soil 159: 11–25.
Friday L, Quarmby C. 1994. Uptake and translocation of
prey-derived 15
N
and 32
P
in Utricularia vulgaris L. New
Phytologist 126: 273–281.
Fuss AM, Sedgley M. 1991. Pollen tube growth and seed set of
Banksia coccinea R.Br. (Proteaceae). Annals of Botany 68: 377–384.
Gabbedy BJ, Meyer EP, Dickson J. 1975. Zamia palm
(Macrozamia reidlei) poisoning of sheep. Australian Veterinary
Journal 51: 303–305.
Galloway JN, Schlesinger WH, Levy H, II, Michaels A,
Schnoor JL. 1995. Nitrogen xation: anthropogenic enhance-
ment-environmental response. Global Biogeochemical Cycles 9:
235–252.
Gardiner CA, Bennetts HW. 1956. e Toxic Plants of Western
Australia. Perth: W.A. Newspapers Ltd.
Gardner WK, Parbery DG, Barber DA. 1981. Proteoid root
morphology and function in Lupinus albus. Plant and Soil 60:
143–147.
Gardner WK, Parbery DG, Barber DA. 1982. e acquisition of
phosphorus by Lupinus albus L. II. e eect of varying phospho-
rus supply and soil type on some characteristics of the soil/root
interface. Plant and Soil 68: 3341.
Gehringer MM, Adler L, Roberts AA, Mott MC, Mihali
TK, Mills TJT, Fieker C, Neilan BA. 2012. Nodularin, a cya-
nobacterial toxin, is synthesized in planta by symbiotic Nostoc sp.
e ISME Journal.
Gehringer MM, Pengelly JJL, Cuddy WS, Fieker C, Forster
PI, Neilan BA. 2010. Host selection of symbiotic Cyanobacteria
in 31 species of the Australian cycad genus: Macrozamia
(Zamiaceae). Molecular Plant-Microbe Interactions 23: 811–822.
Gibson N, Keighery GJ, Lyons MN, Webb A. 2004. Terrestrial
ora and vegetation of the Western Australian wheatbelt. Records
of the Western Australian Museum 67: 139–189.
Godo GH, Reisenauer HM. 1980. Plant eects on soil man-
ganese availability. Soil Science Society of America Journal 44:
993–995.
Grierson PF, Attiwill PM. 1989. Chemical characteristics of the
proteoid root mat of Banksia integrifolia L. Australian Journal of
Botany 37: 137–143.
Grobbelaar N, Strauss J, Groenewald E. 1971. Non-leguminous
seed plants in Southern Africa which x nitrogen symbiotically.
Plant and Soil 35: 325–334.
Groom PK, Lamont BB. 2010. Phosphorus accumulation in
Proteaceae seeds: a synthesis. Plant and Soil 334: 61–72.
Grove TS, O’connell AM, Malajczuk N. 1980. Eects of re
on the growth, nutrient content and rate of nitrogen xation of
the cycad Macrozamia riedlei. Australian Journal of Botany 28:
271–281.
Groves RH, Hocking PJ, McMahon A. 1986. Distribution of
biomass, nitrogen, phosphorus and other nutrients in Banksia
marginata and B. ornata shoots of dierent ages after re.
Australian Journal of Botany 34: 709–725.
CHAPTER 4: PLANT MINERAL NUTRITION
123
Halliday J, Pate J. 1976. Symbiotic nitrogen xation by coral-
loid roots of the cycad Macrozamia riedlei: physiological charac-
teristics and ecological signicance. Functional Plant Biology 3:
349–358.
Handley LL, Scrimgeour CM. 1997. Terrestrial plant ecology
and
15
N natural abundance: e present limits to interpretation
for uncultivated systems with original data from a Scottish old
eld. Advances in Ecological Research 27: 133–212.
Hansen A, Pate JS, Hansen AP. 1991. Growth and reproductive
performance of a seeder and a resprouter species of Bossiaea as a
function of plant age after re. Annals of Botany 67: 497–509.
Hansen AP, Pate JS. 1987. Comparative growth and symbiotic
performance of seedlings of Acacia spp. in dened pot culture or
as natural understorey components of a eucalypt forest ecosystem
in S.W. Australia. Journal of Experimental Botany 38: 13–25.
Hansen AP, Pate JS, Atkins CA. 1987. Relationships between
acetylene reduction activity, hydrogen evolution and nitrogen
xation in nodules of Acacia spp.: experimental background to
assaying xation by acetylene reduction under eld conditions.
Journal of Experimental Botany 38: 1–12.
Hartwig UA. 1998. e regulation of symbiotic N
2
xation: a
conceptual model of N feedback from the ecosystem to the gene
expression level. Perspectives in Plant Ecology, Evolution and
Systematics 1: 92–120.
Hawkins H-J, Hettasch H, Mesjasz-Przybylowicz J,
Przybylowicz W, Cramer MD. 2008. Phosphorus toxicity in
the Proteaceae: a problem in post-agricultural lands. Scientia
Horticulturae 117: 357–365.
Hayes P, Turner BL, Lambers H, Laliberté E. 2013. Foliar
nutrient concentrations and resorption eciency in plants of
contrasting nutrient-acquisition strategies along a 2-million-year
dune chronosequence. Journal of Ecology 102: 396-410.
Hellmuth EO. 1971. Eco-physiological studies on plants in arid
and semi-arid regions in Western Australia: IV. Comparison of
the eld physiology of the host, Acacia gasbyi and its hemipara-
site, Amyema nestor under optimal and stress conditions. Journal of
Ecology 59: 351–363.
Herbert DA. 1919. Nuytsia oribunda (e Christmas Tree) –
Its structure and parasitism. Journal and Proceedings of e Royal
Society of Western Australia 5: 72–88.
Hill RD, Rinker RG, Coucouvinos A. 1984. Nitrous oxide
production by lightning. Journal of Geophysical Research 89:
1411–1421.
Hingston FJ, Malajczuk N, Grove TS. 1982. Acetylene reduc-
tion (N
2
-xation) by jarrah forest legumes following re and
phosphate application. Journal of Applied Ecology 19: 631645.
Hocking P. 1982.e nutrition of fruits of two Proteaceous
shrubs, Grevillea wilsonii and Hakea undulata, from south-west-
ern Australia. Australian Journal of Botany 30: 219–230.
Högberg P. 1990.
15
N natural abundance as a possible marker of
the ectomycorrhizal habit of trees in mixed African woodlands.
New Phytologist 115: 483–486.
Houlton BZ, Wang Y-P, Vitousek PM, Field CB. 2008. A uni-
fying framework for dinitrogen xation in the terrestrial bio-
sphere. Nature 454: 327–330.
Jaré T. 1979. Accumulation du manganèse par les Protéacées
de Nouvelle Calédonie. Comptes Rendus de lAcadémie des Sciences
(Paris), D 289: 425–428.
Jauregui MA, Reisenauer HM. 1982. Dissolution of oxides of
manganese and iron by root exudate components. Soil Science
Society of America Journal 46: 314317.
Johnson SD, Steiner KE. 2000. Generalization versus special-
ization in plant pollination systems. Trends in Ecology & Evolution
15: 140143.
Karlsson PS, Pate JS. 1992. Contrasting eects of supplemen-
tary feeding of insects or mineral nutrients on the growth and
nitrogen and phosphorous economy of pygmy species of Drosera.
Oecologia 92: 813.
Kiers ET, Duhamel M, Beesetty Y, Mensah JA, Franken O,
Verbruggen E, Fellbaum CR, Kowalchuk GA, Hart MM,
Bago A, Palmer TM, West SA, Vandenkoornhuyse P, Jansa J,
Bücking H. 2011. Reciprocal rewards stabilize cooperation in
the mycorrhizal symbiosis. Science 333: 880–882.
Laliberté E, Turner BL, Costes T, Pearse SJ, Wyrwoll K-H,
Zemunik G, Lambers H. 2012. Experimental assessment of
nutrient limitation along a 2-million year dune chronosequence
in the south-western Australia biodiversity hotspot. Journal of
Ecology 100: 631–642.
Lambers H, Ahmedi I, Berkowitz O, Dunne C, Finnegan PM,
Hardy GESJ, Jost R, Laliberté E, Pearse SJ, Teste FP. 2013.
Phosphorus nutrition of P-sensitive Australian native plants:
threats to plant communities in a global biodiversity hotspot.
Conservation Physiology 1: 10.1093/conphys/cot1010.
Lambers H, Brundrett MC, Raven JA, Hopper SD. 2010. Plant
mineral nutrition in ancient landscapes: high plant species diver-
sity on infertile soils is linked to functional diversity for nutri-
tional strategies. Plant and Soil 334: 11–31. 36: 1911–2070
Lambers H, Cawthray GR, Giavalisco P, Kuo J, Laliberté E,
Pearse SJ, Scheible W-R, Stitt M, Teste F, Turner BL. 2012.
Proteaceae from severely phosphorus-impoverished soils exten-
sively replace phospholipids with galactolipids and sulfolipids
during leaf development to achieve a high photosynthetic phos-
phorus-use eciency. New Phytologist 196: 1098–1108.
Lambers H, Chapin FS, Pons TL. 2008a. Plant Physiological
Ecology, second edition. New York: Springer.
Lambers H, Finnegan PM, Laliberté E, Pearse SJ, Ryan MH,
Shane MW, Veneklaas EJ. 2011. Phosphorus nutrition of
Proteaceae in severely phosphorus-impoverished soils: are there
lessons to be learned for future crops? Plant Physiology 156:
1058–1066.
Lambers H, Raven JA, Shaver GR, Smith SE. 2008b. Plant
nutrient-acquisition strategies change with soil age. Trends in
Ecology and Evolution 23: 95–103.
Lambers H, Shane MW, Cramer MD, Pearse SJ, Veneklaas EJ.
2006. Root structure and functioning for ecient acquisition of
phosphorus: matching morphological and physiological traits.
Annals of Botany 98: 693–713.
Lambers H, Teste FP. 2013. Interactions between arbuscular
mycorrhizal and non-mycorrhizal plants: do non-mycorrhizal
species at both extremes of nutrient-availability play the same
game? Plant Cell and Environment.
Lamont B. 1973. Factors aecting the distribution of proteoid
roots within the root systems of two Hakea species. Australian
Journal of Botany 21: 165187.
Lamont B. 1974. e biology of dauciform roots in the sedge
Cyathochaete avenacea. New Phytologist 73: 985–996.
Lamont B. 1982. Mechanisms for enhancing nutrient uptake in
plants, with particular reference to Mediterranean South Africa
and Western Australia. e Botanical Review 48: 597–689.
KWONGAN PLANT LIFE
1 2 4
Lamont B, Wiens D. 2003. Are seed set and speciation rates
always low among species that resprout after re, and why?
Evolutionary Ecology 17: 277–292.
Lamont BB. 1972.Proteoid’ roots in the legume Viminaria
juncea. Search 3: 90–91.
Lamont BB. 2003. Structure, ecology and physiology of root
clusters – a review. Plant and Soil 248: 1–19.
Lamont BB, Brown G, Mitchell DT. 1984. Structure, envi-
ronmental eects on their formation, and function of proteoid
roots in Leucadendron laureolum (Proteaceae). New Phytologist 97:
381–390.
Lamont BB, Groom PK. 2002. Green cotyledons of two Hakea
species control seedling mass and morphology by supplying min-
eral nutrients rather than organic compounds. New Phytologist
153: 101–110.
Lamont BB, Southall KJ. 1982. Distribution of mineral nutri-
ents between the mistletoe, Amyema preissii, and its host, Acacia
acuminata. Annals of Botany 49: 721–725.
Landeweert R, Hoand E, Finlay RD, Kuyper TW, van
Breemen N. 2001. Linking plants to rocks: ectomycorrhizal
fungi mobilize nutrients from minerals. Trends in Ecology &
Evolution 16: 248–254.
Lange RT. 1961. Nodule bacteria associated with the indige-
nous Leguminosae of south-western Australia. Journal of General
Microbiology 26: 351–359.
Leake JR. 2004. Myco-heterotroph/epiparasitic plant interac-
tions with ectomycorrhizal and arbuscular mycorrhizal fungi.
Current Opinion in Plant Biology 7: 422428.
Lee J, Reeves RD, Brooks RR, Jaré T. 1978. e relation
between nickel and citric acid in some nickel-accumulating
plants. Phytochemistry 17: 1033–1035.
Lindblad P, Bergman B, Hofsten AV, Hallbom L, Nylund JE.
1985. e Cyanobacterium-Zamia symbiosis: An ultrastructural
study. New Phytologist 101: 707–716.
Lindblad P, Tadera K, Yagi F. 1990. Occurrence of azoxyglyco-
sides in Macrozamia riedlei and their eects on the free-living iso-
lated Nostoc PCC 73102. Environmental and Experimental Botany
30: 429434.
Loveys BR, Loveys BR, Tyerman SD. 2001. Water relations
and gas exchange of the root hemiparasite Santalum acuminatum
(quandong). Australian Journal of Botany 49: 479486.
McArthur WM. 1991. Reference soils of southwestern Australia.
South Perth: Department of Agriculture Western Australia.
McGee PA. 1988. Growth response to and morphology of
mycorrhizas of ysanotus (Anthericaceae Monocotyledonae).
New Phytologist 109: 459463.
Milberg P, Lamont BB. 1997. Seed/cotyledon size and nutrient
content play a major role in early performance of species on nutri-
ent-poor soils. New Phytologist 137: 665672.
Miller AC, Watling JR, Overton IC, Sinclair R. 2003. Does
water status of Eucalyptus largiorens (Myrtaceae) aect infec-
tion by the mistletoe Amyema miquelii (Loranthaceae)? Functional
Plant Biology 30: 1239–1247.
Minchin FR, Witty JF, Sheehy JE, Müller M. 1983. A major
error in the acetylene reduction assay: decreases in nodular nitro-
genase activity under assay conditions. Journal of Experimental
Botany 34: 641649.
Monk D, Pate JS, Loneragan WA. 1981. Biology of Acacia pul-
chella R.br. with special reference to symbiotic nitrogen xation.
Australian Journal of Botany 29: 579592.
Morton JF. 1983. Rooibos tea, Aspalathus linearis, a caeineless,
low-tannin beverage. Economic Botany 37: 164–173.
Mowry H. 1933. Symbiotic nitrogen xation in the genus
Casuarina. Soil Science 36: 409–426.
Muler AL, Oliveira RS, Lambers H, Veneklaas EJ. 2013. Does
cluster-root activity of Banksia attenuata (Proteaceae) benet
phosphorus or micronutrient uptake and growth of neighbouring
shrubs? Oecologia 174: 23–31.
Nathanielsz CP, Sta IA. 1975. A mode of entry of blue-green
algae into the apogeotropic roots of Macrozamia communis.
American Journal of Botany 62: 232–235.
Orians GH, Milewski AV. 2007. Ecology of Australia: e
eects of nutrient-poor soils and intense res. Biological Reviews
82: 393423.
Pang J, Ryan MH, Tibbett M, Cawthray GR, Siddique KHM,
Bolland MDA, Denton MD, Lambers H. 2010. Variation
in morphological and physiological parameters in herbaceous
perennial legumes in response to phosphorus supply. Plant and
Soil 331: 241–255.
Partt RL. 1979. e availability of P from phosphate-goethite
bridging complexes. Desorption and uptake by ryegrass. Plant
and Soil 53: 5565.
Parniske M. 2008. Arbuscular mycorrhiza: the mother of plant
root endosymbioses. Nature Reviews Microbiology 6: 763–775.
Parrotta JA, Baker DD, Fried M. 1994. Application of
15
N-enrichment methodologies to estimate nitrogen xation
in Casuarinaequisetifolia. Canadian Journal of Forest Research 24:
201–207.
Pate JS, Dixon KW. 1978. Mineral nutrition of Drosera eryth-
rorhiza Lindl. with special reference to Its tuberous habit.
Australian Journal of Botany 26: 455464.
Pate JS. 2001. Haustoria in action: case studies of nitrogen acqui-
sition by woody xylem-tapping hemiparasites from their hosts.
Protoplasma 215: 204217.
Pate JS, Beard JS, eds. 1984. Kwongan. Plant Life of the Sandplain.
Nedlands: University of Western Australia Press.
Pate JS, Bell TL. 1999. Application of the ecosystem mimic con-
cept to the species-rich Banksia woodlands of Western Australia.
Agroforestry Systems 45: 303341.
Pate JS, Davidson NJ, Kuo J, Milburn JA. 1990. Water rela-
tions of the root hemiparasite Olax phyllanthi (Labill) R.Br.
(Olacaceae) and its multiple hosts. Oecologia 84: 186–193.
Pate JS, Dell B. 1984. Economy of mineral nutrients in sand-
plain species. In: Pate JS, Beard JS eds. Kwongan. Plant Life of the
Sandplain Nedlands: University of Western Australia Press.
Pate JS, Unkovich, M. J., Erskine, P. D., Stewart, GR. 1998.
Australian mulga ecosystems 13
C
and
15
N natural abundances of
biota components and their ecophysiological signicance. Plant,
Cell and Environment 21: 1231–1242.
Purnell H. 1960. Studies of the family Proteaceae. I. Anatomy
and morphology of the roots of some Victorian species. Australian
Journal of Botany 8: 38–50.
Rabier J, Laont-Schwob I, Notonier R, Fogliani B, Bouraïma-
Madjèbi S. 2008. Anatomical element localization by EDXS
in Grevillea exul var. exul under nickel stress. Environmental
Pollution 156: 1156–1163.
CHAPTER 4: PLANT MINERAL NUTRITION
125
Raven JA. 2012. Protein turnover and plant RNA and phospho-
rus requirements in relation to nitrogen xation. Plant Science
188–189: 25–35.
Reddell P, Bowen GD. 1985. Frankia source aects growth,
nodulation and nitrogen xation in Casuarina species. New
Phytologist 100: 115–122.
Reddell P, Yun Y, Shipton WA. 1997. Cluster roots and mycor-
rhizae in Casuarina cunninghamiana: their occurrence and forma-
tion in relation to phosphorus supply. Australian Journal of Botany
45: 41–51.
Reinhold-Hurek B, Hurek T. 2011. Living inside plants: bacte-
rial endophytes. Current Opinion in Plant Biology 14: 435443.
Robinson BH, Brooks RR, Kirkman JH, Gregg PE, Gremigni
P. 1996. Plant-available elements in soils and their inuence
on the vegetation over ultramac (“serpentine”) rocks in New
Zealand. Journal of the Royal Society of New Zealand 26: 457–468.
Robinson D. 2001.
15
N as an integrator of the nitrogen cycle.
Trends in Ecology & Evolution 16: 153–162.
Ryan MH, Tibbett M, Edmonds-Tibbett T, Suriyagoda LDB,
Lambers H, Cawthray GR, Pang J. 2012. Carbon trading for
phosphorus gain: the balance between rhizosphere carboxyl-
ates and mycorrhizal symbiosis in plant phosphorus acquisition.
Plant, Cell and Environment 35: 2061–2220.
Sanginga N, Danso S, Bowen G. 1989. Nodulation and growth
response of Allocasuarina and Casuarina species to phosphorus
fertilization. Plant and Soil 118: 125–132.
Schulze ED, Gebauer G, Schulze W, Pate JS. 1991.e utiliza-
tion of nitrogen from insect capture by dierent growth forms of
Drosera from Southwest Australia. Oecologia 87: 240–246.
Schulze W, Schulze ED, Pate JS, Gillison AN. 1997. e nitro-
gen supply from soils and insects during growth of the pitcher
plants Nepenthes mirabilis, Cephalotus follicularis and Darlingtonia
californica. Oecologia 112: 464471.
Shane MW, Cawthray GR, Cramer MD, Kuo J, Lambers H.
2006a. Specialized ‘dauciform’ roots of Cyperaceae are struc-
turally distinct, but functionally analogous with ‘cluster’ roots.
Plant, Cell and Environment 29: 1989–1999.
Shane MW, Cramer MD, Funayama-Noguchi S, Cawthray
GR, Millar AH, Day DA, Lambers H. 2004a. Developmental
physiology of cluster-root carboxylate synthesis and exudation in
harsh hakea. Expression of phosphoenolpyruvate carboxylase and
the alternative oxidase. Plant Physiology 135: 549–560.
Shane MW, Dixon KW, Lambers H. 2006b.e occurrence of
dauciform roots amongst Western Australian reeds, rushes and
sedges, and the impact of phosphorus supply on dauciform-root
development in Schoenus unispiculatus (Cyperaceae). New
Phytologist 165: 887–898.
Shane MW, Lambers H. 2005a. Cluster roots: a curiosity in con-
text. Plant and Soil 274: 101–125.
Shane MW, Lambers H. 2005b. Manganese accumulation in
leaves of Hakea prostrata (Proteaceae) and the signicance of clus-
ter roots for micronutrient uptake as dependent on phosphorus
supply. Physiologia Plantarum 124: 441–450.
Shane MW, Lambers H. 2006. Systemic suppression of clus-
ter-root formation and net P-uptake rates in Grevillea crithmifolia
at elevated P supply: a proteacean with resistance for develop-
ing symptoms of ‘P toxicity. Journal of Experimental Botany 57:
413–423.
Shane MW, McCully ME, Canny MJ, Pate JS, Lambers H.
2011. Development and persistence of sandsheaths of Lyginia
barbata (Restionaceae): relation to root structural development
and longevity. Annals of Botany 108: 13071322.
Shane MW, McCully ME, Lambers H. 2004b. Tissue and
cellular phosphorus storage during development of phosphorus
toxicity in Hakea prostrata (Proteaceae). Journal of Experimental
Botany 55: 1033–1044.
Shane MW, Szota C, Lambers H. 2004c. A root trait account-
ing for the extreme phosphorus sensitivity of Hakea prostrata
(Proteaceae). Plant, Cell and Environment 27: 991–1004.
Shearer G, Kohl D. 1986. N
2
-xation in eld settings:
Estimations based on natural
15
N abundance. Functional Plant
Biology 13: 699–756.
Smith RJ, Hopper SD, Shane MW. 2011. Sand-binding roots in
Haemodoraceae: global survey and morphology in a phylogenetic
context. Plant and Soil 348: 453470.
Smith SE, Read DJ. 2008. Mycorrhizal Symbiosis. London:
Academic Press and Elsevier.
Sprent JI. 1999. Nitrogen xation and growth of non-crop
legume species in diverse environments. Perspectives in Plant
Ecology, Evolution and Systematics 2: 149–162.
Standish RJ, Stokes BA, Tibbett M, Hobbs RJ. 2007. Seedling
response to phosphate addition and inoculation with arbuscu-
lar mycorrhizas and the implications for old-eld restoration in
Western Australia. Environmental and Experimental Botany 61:
5865.
Stock WD, Pate JS, Kuo J, Hansen AP. 1989. Resource control
of seed set in Banksia laricina C. Gardner (Proteaceae). Functional
Ecology 3: 453460.
Sydenham PH, Findlay GP. 1975. Transport of solutes and water
by resetting bladders of Utricularia. Functional Plant Biology 2:
335–351.
Tennakoon KU, Pate JS, Arthur D. 1997a. Ecophysiological
aspects of the woody root hemiparasite Santalum acuminatum (R.
Br.) A. DC and its common hosts in South Western Australia.
Annals of Botany 80: 245–256.
Tennakoon KU, Pate JS, Stewart GR. 1997b. Haustorium-
related uptake and metabolism of host xylem solutes by the
root hemiparasitic shrub Santalum acuminatum (R. Br.) A. DC.
(Santalaceae). Annals of Botany 80: 257–264.
Tester M, Smith SE, Smith FA. 1987. e phenomenon of “non-
mycorrhizal” plants. Canadian Journal of Botany 65: 419431.
iele KR, Wylie SJ, Maccarone L, Hollick P, McComb JA.
2008. Pilostyles coccoidea (Apodanthaceae), a new species from
Western Australia described from morphological and molecular
evidence. Nuytsia 18: 273–284.
Torrey JG, Racette S. 1989. Specicity among the Casuarinaceae
in root nodulation by Frankia. Plant and Soil 118: 157–164.
Turner BL, Lambers H, Condron LM, Cramer MD, Leake JR,
Richardson AE, Smith SE. 2013. Soil microbial biomass and
the fate of phosphorus during long-term ecosystem development.
Plant and Soil 367: 225–234.
Van Hees PAW, Rosling A, Essen S, Godbold DL, Jones DL,
Finlay RD. 2006. Oxalate and ferricrocin exudation by the
extramatrical mycelium of an ectomycorrhizal fungus in symbio-
sis with Pinus sylvestris. New Phytologist 169: 367–378.
Van Leerdam D, Williams P, Cairney J. 2001. Phosphate-
solubilising abilities of ericoid mycorrhizal endophytes of Woollsia
pungens (Epacridaceae). Australian Journal of Botany 49: 75–80.
KWONGAN PLANT LIFE
1 2 6
Van Schöll L, Kuyper TW, Smits MM, Landeweert R, Hoand
E, Van Breemen N. 2008. Rock-eating mycorrhizas: eir role
in plant nutrition and biogeochemical cycles. Plant and Soil 303:
3547.
Vitousek PM, Aber JD, Howarth RW, Likens GE, Matson PA,
Schindler DW, Schlesinger WH, Tilman DG. 1997. Human
alteration of the global nitrogen cycle: sources and consequences.
Ecological Applications 7: 737–750.
Walker TW, Syers JK. 1976. e fate of phosphorus during
pedogenesis. Geoderma 15: 19.
Wang B, Qiu Y-L. 2006. Phylogenetic distribution and evolu-
tion of mycorrhizas in land plants. Mycorrhiza 16: 299–363.
Warcup JH. 1980. Ectomycorrhizal associations of Australian
indigenous plants. New Phytologist 85: 531–535.
Watson AP, Matthiessen JN, Springett BP. 1982. Arthropod
associates and macronutrient status of the red-ink sundew
(Drosera erythrorhiza Lindl.). Australian Journal of Ecology 7:
13–22.
Watt M, Evans JR. 1999. Proteoid roots. Physiology and devel-
opment. Plant Physiology 121: 317–323.
Whittington J, Sinclair R. 1988. Water relations of the mis-
tletoe, Amyema miquelii, and its host Eucalyptus fasciculosa.
Australian Journal of Botany 36: 239–255.
Witkowski ETF, Lamont BB. 1996. Disproportionate allocation
of mineral nutrients and carbon between vegetative and repro-
ductive structures in Banksia hookeriana. Oecologia 105: 3842.
Wittmann W, Bergersen F, Kennedy G. 1965.e coralloid
roots of Macrozamia communis L. Johnson. Australian Journal of
Biological Sciences 18: 1129–1134.
Wright IJ, Reich PB, Westoby M, Ackerly DD, Baruch Z,
Bongers F, Cavender–Bares J, Chapin T, Cornelissen JHC,
Diemer M, Flexas J, Garnier E, Groom PK, Gulias J, Hikosaka
K, Lamont BB, Lee T, Lee W, Lusk C, Midgley JJ, Navas M-L,
Niinemets Ü, Oleksyn J, Osada N, Poorter H, Poot P, Prior L,
Pyankov VI, Roumet C, omas SC, Tjoelker MG, Veneklaas
EJ, Villar R. 2004. e worldwide leaf economics spectrum.
Nature 428: 821–827.
Yeates JS. 1993. Soils and fertilizer use in southwestern
Australia. Fertilizer Research 36: 123–125.
CHAPTER 4: PLANT MINERAL NUTRITION
127

Supplementary resource (1)

... Yet, the exact opposite is actually found, and a key feature of sandplains are their exceptionally high degree of floristic and functional diversity (Lambers et al. 2010). Interestingly, the greatest plant diversity on the sandplains is found on the most severely P-impoverished soils Lambers et al. 2014; Zemu-Drosera sewelliae growing in heavy laterite. Photo taken near Muchea, WA. ...
... Based on the total number of species in the Southwest Australian Biodiversity Hotspot, one can calculate how many carnivorous species one might expect, based on global averages. However, there are in excess of four times more carnivorous species than expected (Lambers et al. 2014). Carnivorous species have diversified tremendously during the course of evolution in south-western Australia, and the Pygmy Drosera species are an excellent example of this diversification in action. ...
Article
Full-text available
There are currently 52 species of Pygmy Drosera known to science (Lowrie 2014), many of which have only been described in the past few decades. Most of these are endemic to south-west Western Australia with one species, Drosera pygmaea, also found in South Australia, Victoria, Tasmania, New South Wales, Queensland, and New Zealand.
... Prevalence of low soil pH (and thus low P availability: Huston, 1979;Lambers et al., 2014;Zemunik et al., 2016) towards the tropics has been proposed as the cause of negative soil pHplant richness relationships at low latitudes (Partel, 2002). The rationale behind this argument is that because in the tropics most habitats are low in pH, there has been more speciation of tropical and subtropical plants at acidic settings driving this relationship, which differs from that at high latitudes. ...
... We acknowledge that there are more P-acquisition strategies in P-impoverished landscapes than discussed in this review, especially carnivory, which is very common in P-impoverished habitats, coprophagy and parasitism, which is relatively uncommon in these habitats if they are seasonally dry. Since these strategies have been discussed in detail before (Cramer et al. 2014;Lambers et al. 2014;Oliveira et al. 2016), we do not deal with them in this review. Table 1 summarises the P-acquisition strategies we discuss in this review. ...
Article
Full-text available
Background Unveiling the diversity of plant strategies to acquire and use phosphorus (P) is crucial to understand factors promoting their coexistence in hyperdiverse P-impoverished communities within fire-prone landscapes such as in cerrado (South America), fynbos (South Africa) and kwongan (Australia). Scope We explore the diversity of P-acquisition strategies, highlighting one that has received little attention: acquisition of P following fires that temporarily enrich soil with P. This strategy is expressed by fire ephemerals as well as fast-resprouting perennial shrubs. A plant’s leaf manganese concentration ([Mn]) provides significant clues on P-acquisition strategies. High leaf [Mn] indicates carboxylate-releasing P-acquisition strategies, but other exudates may play the same role as carboxylates in P acquisition. Intermediate leaf [Mn] suggests facilitation of P acquisition by P-mobilising neighbours, through release of carboxylates or functionally similar compounds. Very low leaf [Mn] indicates that carboxylates play no immediate role in P acquisition. Release of phosphatases also represents a P-mining strategy, mobilising organic P. Some species may express multiple strategies, depending on time since germination or since fire, or on position in the landscape. In severely P-impoverished landscapes, photosynthetic P-use efficiency converges among species. Efficient species exhibit rapid rates of photosynthesis at low leaf P concentrations. A high P-remobilisation efficiency from senescing organs is another way to use P efficiently, as is extended longevity of plant organs. Conclusions Many P-acquisition strategies coexist in P-impoverished landscapes, but P-use strategies tend to converge. Common strategies of which we know little are those expressed by ephemeral or perennial species that are the first to respond after a fire. We surmise that carboxylate-releasing P-mobilising strategies are far more widespread than envisaged so far, and likely expressed by species that accumulate metals, exemplified by Mn, metalloids, such as selenium, fluorine, in the form of fluoroacetate, or silicon. Some carboxylate-releasing strategies are likely important to consider when restoring sites in biodiverse regions as well as in cropping systems on P-impoverished or strongly P-sorbing soils, because some species may only be able to establish themselves next to neighbours that mobilise P.
... In light of the lower provided P and naturally P-impoverished soils that the current study species are found in, the higher percentage P uptake may reflect the differing abilities of ORMs to forage, capture and utilize the small amounts of P that are available to them (Nurfadilah et al., 2013;Mujica et al., 2016), with tight coupling of P remobilization from senescing parts of these orchids enabling high levels of P conservation in planta (Pate and Dixon, 1982). Phosphorus limitation increases plant diversity and has fuelled the proliferation of a variety of P-acquisition strategies in the SWAFR, although there has been a resulting decline in the abundance of mycorrhizal plants (Brundrett, 2009;Lambers et al., 2014;Zemunik et al., 2015Zemunik et al., , 2016. The terrestrial orchids of the SWAFR have retained their mycorrhizal lifestyle, becoming one of the few groups of plants that are obligately dependent on their fungal partners (Smith and Read, 2008), allowing niche differentiation (Nurfadilah et al., 2013). ...
Article
Full-text available
Background and aims: Many terrestrial orchids have an obligate dependence on their mycorrhizal associations for nutrient acquisition, particularly during germination and early seedling growth. Though important in plant growth and development, phosphorus (P) nutrition studies in mixotrophic orchids have been limited to only a few orchid species and their fungal symbionts. For the first time, we demonstrate the role of a range of fungi in the acquisition and transport of inorganic P to four, phylogenetically distinct, green-leaved terrestrial orchid species (Diuris magnifica, Disa bracteata, Pterostylis sanguinea and Microtis media subsp. media) which naturally grow in P-impoverished soils. Methods: Mycorrhizal P uptake and transfer to orchids was determined and visualised using agar microcosms with a diffusion barrier between P source ( 33P orthophosphate) and orchid seedlings, allowing extramatrical hyphae to reach the source. Key results: Extramatrical hyphae of the studied orchid species were effective in capturing and transporting inorganic P into the plant. Following seven days of exposure, between 0.5% (D. bracteata) and 47% (D. magnifica) of the P supplied was transported to the plants (at rates between 0.001 and 0.097fmol h -1). This experimental approach was capable of distinguishing species based on their P-foraging efficiency, and highlighted the role that fungi play in P nutrition during early seedling development. Conclusions: Our study shows that orchids occurring naturally on P-impoverished soils can obtain significant amounts of inorganic P from their mycorrhizal partners, and significantly more uptake of P supplied than previously shown in other green-leaved orchids. These results provide support for differences in mycorrhiza-mediated P acquisition between orchid species and fungal symbionts in green-leaved orchids at the seedling stage. The plant-fungus combinations of this study also provide evidence for plant -mediated niche differentiation occurring, with ecological implications in P-limited systems.
... OCBILs contain a high proportion of relictual plants, and are more likely to support small populations of long-lived plant taxa, exhibiting maintenance of genetic diversity (Bezemer et al., 2016), and specialist traits for sitespecific nutrient procurement in phosphorus-limited landscapes, e.g. carnivory, cluster and sand-binding roots (Lambers et al., 2011(Lambers et al., , 2012(Lambers et al., , 2014. In addition, traits such as long juvenile periods, small and/or disjunct populations, soil-type specificity and limited dispersal capability suggest vulnerability of OCBIL flora to frequent fire (Holmes et al., 2000;Barrett et al., 2009;Enright et al., 2014), and other natural and human disturbance (Arianoutsou et al., 2013;Leão et al., 2014). ...
Article
Occurring across all southern hemisphere continents except Antarctica, old, climatically buffered, infertile landscapes (OCBILs) are centres of biological richness, often in biodiversity hotspots. Among a matrix of young, often disturbed, fertile landscapes (YODFELs), OCBILs are centres of endemism and diversity in the exceptionally rich flora of the south-west Australian global biodiversity hotspot, home to Noongar peoples for ≥ 48 000 years. We analysed contemporary traditional Noongar knowledge of adjacent OCBILs (e.g. granite outcrops) and YODFELs (e.g. creekline fringes) both at a single site and in two larger areas to test whether patterns of disturbance dictated by Noongar custom align with OCBIL theory. We found that Noongar traditional knowledge reflects a regime of concentrated YODFEL rather than OCBIL disturbance—a pattern which aligns with maximal biodiversity preservation. SIMPER testing found traditional Noongar OCBIL and YODFEL activities are 64–75% dissimilar, whereas Pearson’s chi-square tests revealed camping, burning, travelling through country and hunting as primarily YODFEL rather than OCBIL activities. We found that Noongar activities usually avoid OCBIL disturbance. This combined with high floristic diversity following enduring First Peoples’ presence, suggests that traditional Noongar knowledge is valuable and necessary for south-west Australian biodiversity conservation. Similar cultural investigations in other OCBIL-dominated global biodiversity hotspots may prove profitable.
Article
Full-text available
Phosphorus (P), a critical macronutrient for plant growth and reproduction, is primarily acquired and translocated in the form of inorganic phosphate (Pi) by roots. Pi deficiency is widespread in many natural ecosystems, including forest plantations, due to its slow movement and easy fixation in soils. Plants have evolved complex and delicate regulation mechanisms on molecular and physiological levels to cope with Pi deficiency. Over the past two decades, extensive research has been performed to decipher the underlying molecular mechanisms that regulate the Pi starvation responses (PSR) in plants. This review highlights the prospects of Pi uptake, transport, and signaling in woody plants based on the backbone of model and crop plants. In addition, this review also highlights the interactions between phosphorus and other mineral nutrients such as Nitrogen (N) and Iron (Fe). Finally, this review discusses the challenges and potential future directions of Pi research in woody plants, including characterizing the woody-specific regulatory mechanisms of Pi signaling and evaluating the regulatory roles of Pi on woody-specific traits such as wood formation and ultimately generating high Phosphorus Use Efficiency (PUE) woody plants.
Article
Full-text available
The expansion of residential areas has increased the demand for exotic and increasingly compact landscape plants. In this context, this work aimed to evaluate the effects of paclobutrazol (PBZ) and the timing of ethylene application on growth reduction and flowering anticipation in ornamental pineapples grown in pots. The randomized block design was used with factorial arrangement (2 x 5) and 4 replications and 4 plants per plot. The primary treatments are the presence and absence of PBZ. The secondary treatments were five times of floral induction with ethylene: 90; 120; 150; 180; 210 days after transplanting (DAT) the seedlings into the pots. It was evaluated the variables: height ratio between the heights of the pot and the plant; rosette diameter; leaves length ‘D’ and flowering index. At 255 DAT, although the plants did not respond to floral induction with ethylene, only those treated with PBZ were more compact and had different characteristics such as smaller size and a ratio below 1/3 of the height of the vase in relation to the plant that to favor its commercialization, in pot.
Article
Full-text available
Diet is the main intake source of selenium (Se) in the body. Southern Jiangxi is the largest navel orange-producing area in China, and 25.98% of its arable land is Se-rich. However, studies on the Se-rich characteristics and Se dietary evaluation of navel orange fruits in the natural environment of southern Jiangxi have not been reported. This study was large-scale and in situ samplings (n = 492) of navel oranges in southern Jiangxi with the goal of investigating the coupling relationships among Se, nutritional elements, and quality indicators in fruits and systematically evaluating Se dietary nutrition to the body. The results indicated that the average content of total Se in the flesh was 4.92 μg⋅kg–1, and the percentage of Se-rich navel oranges (total Se ≥ 10 μg⋅kg–1 in the flesh) was 7.93%, of which 66.74% of the total Se was distributed in the pericarp and 33.26% in the flesh. The average content of total Se in the flesh of Yudu County was the highest at 5.71 μg⋅kg–1. There was a significant negative correlation (p < 0.05) between Se, Cu, and Zn in the Se-rich flesh. According to the Se content in the flesh, the Se dietary nutrition evaluation was carried out, and it was found that the Se-enriched navel orange provided a stronger Se nutritional potential for the human body. These findings will help to identify Se enrichment in navel orange fruit in China’s largest navel orange-producing area and guide the selection of Se-rich soils for navel orange production in the future.
Article
Carnivorous plants have access to several potential sources of nitrogen, including root uptake, predation, litterfall, atmospheric deposition and defecation by mutualistic animals. Our aim was to assess the relative importance of different N sources so as to better understand the ecology of these physiologically diverse plants that include many genera and species inhabiting terrestrial and aquatic environments worldwide. Plant physiology and habitat were the major determinants of the relative importance of N source. Our secondary aim was to examine protocarnivorous plants that do not fit the exact definition for carnivory. Several protocarnivorous plants were classified as carnivorous based on specialised trapping mechanisms, isotopic data and mixing models. Several carnivorous plants can transfer their functions of prey capture and digestion to mutualistic animal partners, which is termed ecological outsourcing. Outsourcing arthropod prey capture and digestion to mutualistic bats is a beneficial strategy for the carnivorous plant Nepenthes hemsleyana.
Article
Full-text available
Nutrient‐poor ecosystems globally exhibit high plant diversity. One mechanism enabling species coexistence in such ecosystems is facilitation among plants with contrasting nutrient‐acquisition strategies. The ecophysiological processes underlying these interactions remain poorly understood. We hypothesised that root positioning plays a role between sympatric species in nutrient‐poor vegetation. We investigated how growth traits of the focal mycorrhizal non‐cluster rooted Hibbertia racemosa change when grown in proximity of non‐mycorrhizal Banksia attenuata, which produces cluster roots that increase nutrient availability, compared with growth with conspecifics. Focal plants were placed in the centre of rhizoboxes, and we assessed biomass allocation, root system architecture, specific root length and leaf nutrient concentration. When grown with B. attenuata, focal plants decreased root investment, increased root growth towards B. attenuata, and positioned their roots near B. attenuata cluster roots. Specific root length was greater, and the degree of localised root investment correlated positively with B. attenuata cluster‐root biomass. Total nutrient contents in the focal individuals were greater when grown with B. attenuata. Focal plants directed their root growth towards the putatively facilitating neighbour's cluster roots, modifying root traits and investment. Preferential root positioning and root morphological traits play important roles in positive plant‐plant interactions. This article is protected by copyright. All rights reserved.
Article
The entry of blue-green algae into an aerial root of Macrozamia communis via a break in the dermal layers and leading through a continuous cortical channel into the algal zone is reported. Penetration of the algae appears to be mainly in the form of filaments. Intercellular migration of the algae is found to occur after dissolution of host cell middle lamellae, but host cortical cells are also destroyed near some of the algae. Intracellular migration is also suggested to be a likely pathway for the algae as they proceed to the algal zone.
Article
Macrozamia communis exhibits root dimorphism, possessing both normal and coralloid roots. The latter are pneumatophores in which an algal zone may be present or absent. In the coralloid roots the root cap tissue was interpreted as forming a secondary cortex which persisted throughout the life of the root. Underlying the secondary cortex was a transformed epidermis, the cells of which, following the establishment of algae within the intercellular spaces, elongated radially to form an algal zone easily visible to the eye. Nostoc-like blue-green algae were isolated from this zone. In the absence of algal infection the transformed epidermis remained inconspicuous. The root cap anatomy of the normal roots was similar to that described for other gymnosperms. The tips of the alga-free coralloid roots were characteristically located near the soil surface. The alga-containing coralloid roots were less strictly negatively geotropic and were found at depths up to 30 cm below the soil surface.
Article
Studies of the reproductive biology of the Proteaceae have often assumed that pollinator activity and possibly resources limit reproductive output of an individual. To date no studies have demonstrated that members of this family could produce more progeny if provided with more resources while many studies have shown that seed production is often limited by pollinator availability and seed predator activity. In this study the reproductive capacity (ovules/plant) of a rare West Australian Banksia, Banksia laricina C. Gardner, was determined as was the percentage of this capacity which was fulfilled. Only 2.6% of the reproductive capacity was realized, which is a low value typical of most hermaphroditic Proteaceae. The stages at which seed production was reduced were also investigated and this species appears to regulate maternal investment at all reproductive stages. It was also found that removal of 50% of the pollinated inflorescences of an individual did not affect final seed set. This suggested that some endogenous factor, probably resources, limited final seed output. Manipulation of resources available to the plant by nutrient addition was undertaken and increased seed production was evident upon the addition of either nitrogen, phosphorus or a balanced nutrient solution. Thus, it appears that low seed set in B. laricina is the result of selection for obligate outcrossing and high seed mineral composition so as to maintain genetic diversity of a restricted number of high quality propagules suitable for establishment in the low nutrient soils of their natural habitat.
Article
The water relations and leaf gas-exchange characteristics of the root hemiparasite quandong (Santalum acuminatum (R.Br.) A.DC) and its neighbouring plants were examined at three field sites in central Southern Australia. This paper examines the role of water potential and osmotic gradients in facilitating the movement of water from host plants to quandong. Quandong exhibited a significantly more negative water potential than the neighbouring plant species at both field sites during summer and winter. A significant proportion of the osmotic potential was accounted for by mannitol, Na + , K + and Cl - . A water potential difference of 1.7 MPa was maintained between quandong and its putative host over a measurement period of 24 h. Xylem sap and leaves of quandong contained considerable concentration (0.1–0.4 mol (kg tissue water) –1 ) of mannitol. Stomatal conductance and assimilation of quandong were lower than those of the neighbouring plants at both Middleback and Aldinga during both summer and winter measurements. Measurements of transpiration for quandong differed between the two sites. The lower transpirational water loss resulted in quandong at Middleback having an instantaneous water-use efficiency higher (0.13–2.2 µmol (CO 2 ) mmol –1 (H 2 O)) than the neighbouring plants. Daily sap flow and calculated hydraulic conductivity were not significantly different between quandong and putative host plant.
Article
The ability of 13 mycorrhizal endophytes from Woollsia pungens (Cav.) F.Muell. (Epacridaceae) and three Hymenoscyphus isolates to solubilise hydroxyapatite and fluorapatite was assessed in axenic culture under a range of conditions. All Hymenoscyphus isolates and most of the endophytes from W. pungens solubilised hydroxyapatite in unbuffered media containing NH 4+ as the sole nitrogen source, but were unable to solubilise this substrate in buffered media or when NO 3–was the sole nitrogen source. None of the isolates solubilised fluorapatite under any of the conditions tested. The internal phosphorus status of the mycelia had no influence on the ability of the isolates to solubilise either substrate. The data suggest that the phosphate-solubilising activities of mycorrhizal endophytes of Epacridaceae in soil may be confined to the less chemically resistant phosphorus sources.
Article
This paper reviews a growing body of literature on the use of variations in the natural abundance of 15N to estimate the fractional contribution of N2-fixation to N2-fixing systems. This method is based on the small difference in 15NN abundance which frequently occurs between N derived from N2-fixation and N derived from other sources. The requirement of the method is that this difference be significant. Whether this requirement is met is site specific and must be empirically established at each site of interest. Advantages and disadvantages of this method are compared with those of more conventional methods. Sources of error, including heterogeneity of 15NN abundance of non-atmospheric N sources are considered. Tests of the method, under both greenhouse and field conditions, are described. Estimates based on this method compare favourably with other methods for field evaluation of N2-fixation, provided that the site and the sampling strategy are appropriate for application of the method. Applications of the method in several ecosystems are described.
Article
'Coralloid' roots containing blue-green algae occur commonly on the upper root stocks of M. riedlei in natural habitat in Western Australia. Each coralloid mass persists for several seasons; replacement sets form at irregular intervals, especially after fire. 15N2 and acetylene reduction assays demonstrate that coralloid roots fix nitrogen at physiologically significant rates. C2H2 reduction rates by coralloid roots are higher in winter than in summer. Performance is positively correlated with rainfall; soil temperature appears to be of lesser importance. Diurnal fluctuations in nitrogenase activity occur. Calibration using 15N2 gives a molar ratio of C2H2 reduced : N2 fixed of 5.8 : 1. The seasonal average of C2H2 reduction of 14.8 nmol per g fresh wt coralloid root per min is then equivalent to 37.6 g N per kg fresh wt per year, a fixation rate potentially capable of doubling coralloid root nitrogen once in every 8 weeks, and whole plant nitrogen every 8-11 years. Returns of fixed nitrogen in two natural populations of Macrozamia are estimated by compounding measurements of biomass of host and symbiotic organs with the seasonal average for coralloid fixation rate. The values obtained (18.8 and 18.6 kg N ha-1 year-1) indicate that Macrozamia contributes significantly to the nitrogen economy of its ecosystem.
Article
The bladder of the insectivorous aquatic plant Utricularia, after stimulation and consequent sudden increase in volume, slowly resets, transporting solutes, mainly Na+, K+ and C1-, and water, from the lumen to the outside solution. The resetting process, which involves the movement of both water and chloride against the direction of passive driving forces, requires energy. As the resetting rate is diminished by the application of respiratory inhibitors and by lowered temperature, but is not affected by darkness, it is clear that respiration and not photosynthesis provides the energy for transport. Experiments with bladders resetting under paraffin oil show that the transported luminal fluid emerges from the mouth region of the bladder. The transport of solution from lumen to outside is essentially a one-way process. Changes of osmotic potential in the luminal solution affect transport rate, but similar changes in the outside solution do not. Bladders which are fully set appear to be almost completely impermeable to water. The transport process can be accounted for as an active transport of chloride ions, with sodium as the predominant accompanying ion, from the lumen to some intermediate region in the cellular bladder wall. Between this region and the lumen an osmotic gradient could be set up that would cause a passive flow of water from the lumen. An increase in hydrostatic pressure in the space could then expel solution, possibly through a subcuticular space to the outside.