Sie sind auf Seite 1von 16

Quaternary Science Reviews 112 (2015) 17e32

Contents lists available at ScienceDirect

Quaternary Science Reviews


journal homepage: www.elsevier.com/locate/quascirev

Invited review

The role of palaeoecological records in assessing ecosystem services


Elizabeth S. Jeffers a, *, Sandra Nogue
 a, b, Katherine J. Willis a, b, c
a
Long Term Ecology Laboratory, Biodiversity Institute, Oxford Martin School, Department of Zoology, University of Oxford, Oxford OX1 3PS, UK
b
Department of Biology, University of Bergen, All
egaten 41, N-5007 Bergen, Norway
c
Royal Botanic Gardens, Kew, Richmon, TW9 3AE, UK

a r t i c l e i n f o a b s t r a c t

Article history: Biological conservation and environmental management are increasingly focussing on the preservation
Received 10 July 2013 and restoration of ecosystem services (i.e. the benefits that humans receive from the natural functioning
Received in revised form of healthy ecosystems). Over the past decade there has been a rapid increase in the number of palae-
12 December 2014
oecological studies that have contributed to conservation of biodiversity and management of ecosystem
Accepted 19 December 2014
Available online
processes; however, there are relatively few instances in which attempts have been made to estimate the
continuity of ecosystem goods and services over time. How resistant is an ecosystem service to envi-
ronmental perturbations? And, if damaged, how long it does it take an ecosystem service to recover?
Keywords:
Ecosystem services
Both questions are highly relevant to conservation and management of landscapes that are important for
Palaeoecology ecosystem service provision and require an in-depth understanding of the way ecosystems function in
Long-term ecological records space and time. An understanding of time is particularly relevant for those ecosystem services e be they
Final ecosystem services supporting, provisioning, regulating or cultural services that involve processes that vary over a decadal
Ecosystem goods (or longer) timeframe. Most trees, for example, have generation times >50 years. Understanding the
Proxies response of forested ecosystems to environmental perturbations and therefore the continuity of the
Dendroecology ecosystem services they provide for human well-being e be it for example, carbon draw-down (regu-
lating service) or timber (provisioning service) e requires datasets that reflect the typical replacement
rates in these systems and the lifecycle of processes that alter their trajectories of change. Therefore, data
are required that span decadal to millennial time-scales. Very rarely, however, is this information
available from neo-ecological datasets and in many ecosystem service assessments, this lack of a tem-
poral record is acknowledged as a significant information gap.
This review aims to address this knowledge gap by examining the type and nature of palaeoecological
datasets that might be critical to assessing the persistence of ecosystem services across a variety of time
scales. Specifically we examine the types of palaeoecological records that can inform on the dynamics of
ecosystem processes and services over time e and their response to complex environmental changes.
We focus on three key areas: a) exploring the suitability of palaeoecological records for examining
variability in space and time of ecosystem processes; b) using palaeoecological data to determine the
resilience and persistence of ecosystem services and goods over time in response to drivers of change;
and c) how best to translate raw palaeoecological data into the relevant currencies required for
ecosystem service assessments.
© 2014 Elsevier Ltd. All rights reserved.

1. Introduction services are derived from ecosystem processes i.e. the physical,
chemical and biological interactions between organisms and their
There is growing appreciation for the many goods and services environment (Fig. 1a). The Millennium Ecosystem Assessment (MA)
provided to people by well-functioning ecosystems, including food, evaluated the current state of ecosystem service provision world-
fuel, climate regulation and spiritual values. These goods and wide and found that the majority of ecosystems are becoming
increasingly degraded, which threatens the long-term supply of
ecosystem service delivery (MA, 2005). Preserving (Chan et al.,
2006) and restoring (Palmer and Filoso, 2009) ecosystems for the
* Corresponding author. Tel.: þ44 1865 281 851. services they provide to people requires informed land and natural
E-mail address: elizabeth.jeffers@zoo.ox.ac.uk (E.S. Jeffers).

http://dx.doi.org/10.1016/j.quascirev.2014.12.018
0277-3791/© 2014 Elsevier Ltd. All rights reserved.
18 E.S. Jeffers et al. / Quaternary Science Reviews 112 (2015) 17e32

Fig. 1. Schematic of the ecosystem services framework used by the UK National Ecosystem Assessment to demonstrate the links between ecosystem processes, final ecosystem
services and goods (a). Copyright 2011 UK National Ecosystem Assessment. Environmental change (e.g. climate warming) will alter the ability of ecosystems to provide goods and
services. Predicting the impacts of environmental change on human well-being requires knowledge of how ecosystem processes respond to direct drivers of change and how these
changes will cascade through each step in the production of ecosystem services and goods (b).

resource decision making, particularly in landscapes beyond pro- state exists), information on the alternative stable states of that
tected areas, where human activities are reducing biodiversity and ecosystem, and an understanding of what happens when the sys-
impeding natural ecosystem processes at increasing rates tem is perturbed. Mapping dynamical ecosystem response to
(Balmford and Bond, 2005). change therefore demands ecological records that span intervals in
Managing ecosystems for the continued supply of goods and time where such responses can be observed. It has been acknowl-
services they provide for human well-being depends upon the edged a number of times that palaeoecological records can provide
availability of information about their variation across space (de some of these data (e.g. Dawson et al., 2011; Dearing et al., 2012)
Groot et al., 2010) and over time as systems respond to on-going and some excellent case-studies have demonstrated the utility of
environmental change and short-term environmental perturba- palaeoecological records in this respect (e.g. Dearing et al., 2012;
tions (Carpenter et al., 2009). Ecosystem management efforts are Colombaroli and Tinner, 2013; Gosling and Williams, 2013;
often being aimed at a moving target (Dawson et al., 2011) and McLauchlan et al., 2013a). But in order for palaeoecological re-
therefore require an understanding of how ecosystem components cords to become more widely used in the determination of
and processes respond to direct and indirect drivers of change (Diaz ecosystem service provision over time, there are fundamental
et al., 2007; Tylianakis et al., 2008) and of the cascading effects questions that the palaeo-ecological community at large need to
these changes have on the supply of ecosystem services and goods ask. In some ways, these are similar to those asked when consid-
to humans (Fig. 1b). ering the use of palaeoecological records in biodiversity conserva-
Since the publication of the MA, the field of ecosystem service tion and management (e.g. Willis and Birks, 2006; Froyd and Willis,
research has made significant progress in developing methodolog- 2008) and include i) what length of temporal record is needed? ii)
ical approaches for determining the variation over space in what proxies should be used to reconstruct ecosystem processes?
ecosystem service provision and how this might change in the future iii) what datasets should be utilised to reconstruct ecosystem ser-
under particular management scenarios (e.g. UK National Ecosystem vice provision? iv) what is the relevant spatial scale at which to
Assessment, UKNEA). In addition, attempts have been made to
provide spatial displays of ecosystem service provision (Eigenbrod
et al., 2010; Lavorel et al., 2011). Output from these exercises vary
in complexity from simple land cover maps (e.g. Naidoo et al., 2008)
to complex models incorporating interactions between ecosystem
components and processes (e.g. Goldstein et al., 2012). The aim of
these approaches is to identify hotspots of ecosystem service pro-
vision that should be protected from development (Fig. 2). However,
these approaches represent ecosystem service provision at a ‘fixed’,
static point in time. What is still lacking, and represents a significant
knowledge gap is the continuity of these services over time (Dawson
et al., 2011; Mace et al., 2012) particularly in response to drivers of
change (e.g. species introductions, climate change, land-use change
Nelson et al., 2005). Fig. 2. Spatially-explicit biodiversity and ecosystem data are synthesized in order to
To understand ecosystem service provision over time and re- generate maps of ecosystem service provision. These maps identify hotspots of
ecosystem service provision that should be prioritized for conservation and ecosystem
sponses of different services to environmental perturbations re-
management action, modified from de Groot et al. (2010). Copyright 2010 Ecological
quires knowledge of the baseline context of an ecosystem (if such a Complexity.
E.S. Jeffers et al. / Quaternary Science Reviews 112 (2015) 17e32 19

reconstruct the ecosystem changes? v) what are the most useful environment which support the provision of ecosystem services),
ways to analyse and display the data to make them accessible to including soil formation, nutrient cycling, primary production,
(and useful for) environmental managers and policy makers alike? pollination and biomass production; ii) Final Ecosystem Service (i.e.
In this review, we highlight the different types of proxy records the outputs from ecosystems that directly deliver gains and losses
that appear to be most appropriate, and the spatial and temporal in human welfare), which includes crops and clean water supply;
scales over which these records provide information relevant for and iii) Goods (i.e. the outcomes of ecosystem services that are
ecosystem management (Table 1). We then move on to assess the valued by humans) such as timber, storm protection, soil erosion.
utility of palaeoecological records in determining the persistence of The research reviewed here demonstrates that recent analytical
ecosystem services in response to external drivers of change. and methodological advances, in combination with the growing
Finally, we consider how this information should be analysed and number of palaeoecological databases (which provide ecological
displayed in order to be of greatest use in natural resource planning data at high spatial and temporal resolution), can yield the infor-
and how palaeoecological reconstructions are translated into the mation required to sustainably manage landscapes for ecosystem
currencies required for ecosystem service assessments. service provision given expected future environmental changes.
The ecosystem services framework that we have adopted in this
review is that used by the UK National Ecosystem Services 2. Ecosystem processes
Assessment (UKNEA, 2011 and described in Mace et al., 2012)
which is itself a modification of the MA framework (MA, 2005). The Ecosystem processes underpin the provision of final ecosystem
NEA framework classifies ecosystems (Fig. 1a) into i) Ecosystem services and an understanding of how they vary in time and space
Processes (i.e. the interactions between biota and their is critical for managing current and future ecosystem services.

Table 1
Examples of dynamical indicators of ecosystem processes, final ecosystem services and goods provided by palaeoecological proxies.

Proxy Dynamical indicator(s) Citation

Nutrient cycling d15N of foliage, wood and lake sediments Tree- and ecosystem-scale variation in Elliot and Brush, 2006;
N availability, at annual to millennial McLauchlan et al., 2007, 2010,
scale resolution (respectively). 2013b
Soil formation and Elemental (e.g. ICP-MS) and magnetic susceptibility Rate (accumulation year1) of soil (a) Willis et al., 1997
stabilization analysis of lake sediments evolution (a) and stabilization within a (b) Hu et al., 2001
catchment (b)
Biomass (a) Pollen accumulation rates in lake sediments Rate of change in biomass per year for (a) Magri, 1994; Sepp€
a et al.,
(b) Landscape Reconstruction Algorithm applied to each taxa (a), the proportion of taxa on 2009
proportional pollen data the landscape (b) and proportion of taxa (b) Sugita, 2007a, b
(c) Bayesian inference model of proportional pol- within each 12 km2 cell (c), at annual, (c) Paciorek and McLachlan,
len data and modern plant occurrence data centennial, and millennial resolution 2009
from stand to catchment scale
Crops (a) Species distribution model constrained by pol- Probability of occurrence (%) of plant (a) Alba-Sanchez et al., 2010;
len assemblages and palaeo-climate data crop taxa (a) and rate of change in grain Macias-Fauria and Willis,
(b) Size measurements (length, breadth and thick- weight (mg year1) (b), at centennial 2013
ness) of fossilized crop grains and millennial resolution from local to (b) Ferrio et al., 2004, 2005
continental scales
Water supply (a) Assemblages of invertebrate (Coleoptera, Tri- Surface flow rates in rivers and streams (a) Howard et al., 2009, 2010
choptera and Chironomidae) species and their (m s1) (a), million acre feet (MAF) yr1 (b) Gangopadhyay et al., 2009
known association with modern river flow (b) where MAF ¼ 1.233  109 m3, and (c) Bennion et al., 2004
regimes total phosphorus concentrations (mg
(b) Non-parametric analysis (K nearest neighbour) l1) in lake water at annual to
of tree ring chronologies and historic river or centennial resolution (c)
stream flow records
(c) Diatom-inferred total phosphorus in lakes
Climate regulation Ecosystem process model (e.g. Land Processes and Terrestrial carbon storage (Pg C) within Prentice et al., 2011
eXchanges) of terrestrial carbon storage biomes for target time periods
constrained by pollen assemblage data
Timber (a) Average cumulative radii of each tree derived Volume (m3 yr1) of standing and dead (a) Metsaranta and Lieffers,
from wood ring analysis of living and sub-fossil trees (a) and minimum cutting cycle (b), 2009
trees at the stand scale with annual (b) Rozendaal and Zuidema,
(b) Tree age and lifetime growth trajectory inferred resolution; the % reduction in average 2011
from dendrochronological analysis of incre- maximum and periodic radial growth
ment bores of living trees per pest outbreak, average interval
(c) Severity and timing of pest outbreaks for timber between outbreaks (years) and
stands from tree ring width chronologies duration of reduced growth (years); (c) Swetnam and Lynch, 1993
(d) Fire return interval and intensity from fire scar and the occurrence of individual fires
records in tree rings and charcoal accumulation (year), time in between occurrences
rates in lake sediments (years) and spatial extent of the
disturbance (ha) (d) Higuera et al., 2011
Soil erosion protection (a) Radiometric (210Pb and 137Cs) dating and Sediment loss (t ha1 yr1) through (a) Dearing et al., 1987, 2012
palaeo-magnetic analysis (from magnetic time (a) and sediment discharge rates (b) Coulthard et al., 2002
susceptibility) of lake sediments (m3 yr1) into the lake from each 2
(b) Cellular automata model of coupled hydro- e50 m2 cell within the catchment (b)
geomorphological processes constrained by
lake sediment influx rates and grain size
Coastal protection Amount of time following a disturbance before Recovery rate (years) of mangroves Gonz
alez et al., 2010
mangrove communities (inferred from fossil pollen) following disturbances at the local to
return to a previous state of cover (in % of total regional scale
pollen sum)
20 E.S. Jeffers et al. / Quaternary Science Reviews 112 (2015) 17e32

Within the ecosystem service framework the processes themselves et al., 2007) at each of these relevant scales. Sedimentary d15N
are not directly valued in monetary terms; instead their contribu- values in particular reflect the full array of processes acting on N
tion is assessed in terms of their impact on the quality and quantity molecules as they cycle through the terrestrial environment,
of the final services they support. Key questions in relation to these especially soil denitrification (Houlton and Bai, 2009), and enter the
ecosystem processes is how do they vary over time, what is the lake basin through surface runoff (McLauchlan et al., 2013b). In
most appropriate timescale to consider for obtaining data and how some cases, anthropogenically-derived N molecules (e.g. from
do they respond to environmental perturbations? There are at least agricultural and human waste), which tend to have largely enriched
three ecological processes where palaeoecological records have an (i.e. higher) d15N values relative to natural sources of N, can leave a
important contribution to make in terms of understanding these unique ‘signature’ in stream runoff (Harrington et al., 1998; Burns
questions namely nutrient cycling, soil formation and biomass and Kendall, 2002) which is ultimately recorded in sediments.
change. Each will be considered in turn. Thus, sedimentary d15N records are able to reliably track the
transfer of N from landscapes to aquatic ecosystems (i.e. leaching)
2.1. Nutrient cycling and the atmosphere (i.e. gaseous losses through denitrification).
When N is lost from the terrestrial landscape through leaching
A key research challenge is to determine how changes in (i.e. nitrates), it enters aquatic ecosystems, typically through small
nutrient cycling in space and time influence ecosystem services. streams. It is widely known that surplus N can result in eutrophi-
Nutrients are essential for the growth of all living organisms and cation in aquatic systems causing hypoxia and anoxia, leading to
the growth rates of plants are typically limited by the availability of the creation of dead zones within coastal ecosystems (Diaz and
either nitrogen (N) or phosphorus (P). As such, the availability and Rosenberg, 2008). However, what is unclear from short-term
cycling of nutrients underpins many other ecosystem services such (<10 years) records, is how long this impact persists within the
as food, timber and fuel provision (Lavelle et al., 2005). For much of ecosystem. In a study by Elliott and Brush (2006), N isotope analysis
human history, people have attempted to overcome nutrient limi- of sedimented organic matter retrieved from a wetland adjacent to
tation of beneficial plant species, particularly crops, by applying the Chesapeake Bay (Maryland, USA), was used to trace the impact
manure and bones to soils. In the early 1900's the invention (and of increasing N inputs from the surrounding agricultural and sub-
mass industrialisation) of the HabereBosch process enabled the urban landscape on the freshwater system over the last 350 years.
large-scale production of synthetic fertilizers. This revolution in They found progressive enrichment of sedimentary d15N over the
agriculture allowed humans to overcome persistent nutrient limi- last few centuries was associated with forest clearing for agricul-
tation of food crops; however, since nutrients are highly mobile in ture and increasing nutrient inputs from human waste, particularly
the soil profile, much of the N and P added to soils through syn- in the last 50 years (Elliott and Brush, 2006). This historical
thetic fertilizers are lost to surface water (i.e. leaching) and the perspective on N cycling was used to provide a sound evidence-
atmosphere (i.e. gaseous losses). This occurs when the amount base to argue for the restoration of wetlands due to their ability
applied exceeds the amount immediately required by the target to reduce N pollution in coastal ecosystems by increasing rates of
plants (or micro-organisms in the soil). Greater amounts of readily denitrification (i.e. the return of reactive N compounds to inert N2
available N in the environment as a result of excess fertilizer use are gas in the atmosphere) on the landscape (Brush, 2009), thereby
now a major source of pollution in some systems; this can lead to providing a key ecosystem service to people in terms of pollution
changes in ecosystem structure and functioning, and reduce the control.
provision of multiple ecosystem services (Compton et al., 2011; Ammonia (NH3) is lost to the atmosphere from agricultural
Sutton, 2011). Where does excess N applied to agricultural land- landscapes through the process of volatilization, particularly when
scapes go? How do increased supplies of reactive N (i.e. the forms of N supply exceeds plant and soil microbe demand. The redeposition
N that can be directly utilized by plants) affect terrestrial and of this reactive form of N on landscapes can have cascading effects
aquatic ecosystems? And how long do the effects of N pollution on the structure and function of terrestrial and aquatic ecosystems
persist in the system? (Sutton, 2011). The ability of plants and soil organisms to retain
Answering these questions requires information on the climatic, surplus N is controlled by their requirements (or demands) for N at
biotic and geologic controls on the long-term fluxes of N through the time of deposition. Under enriched atmospheric CO2 condi-
ecosystems (McLauchlan et al., 2013a). The response of ecosystems tions, plant demand for N is increased in order to support higher
to additional supplies of N is determined to a large extent by the rates of photosynthesis. The progressive N limitation (PNL) hy-
balance between the supply of reactive N and the amount pothesis in particular posits that as atmospheric CO2 concentra-
demanded by plants and soil organisms. While nutrients can limit tions increase, plant productivity and the ability of plants to
productivity, plants are also known to exert high levels of control assimilate CO2 could become limited by available N (Luo et al.,
over the availability and cycling of a number of nutrients (Hobbie, 2004; Reich et al., 2006). Thus, the deposition of reactive N could
1992; Chapman et al., 2006). Soil microbes can also lock away in fact be supporting higher rates of primary productivity and
(i.e. immobilize) N, reducing the amount returned to soils for up- increased carbon storage in terrestrial ecosystems, which in turn
take by plants (Chapman et al., 2006). Furthermore, these re- reduces the amount of N lost from terrestrial ecosystems. For
lationships vary through time in response to changes in example, an estimated 11.2 Pg of carbon stored in terrestrial eco-
temperature, precipitation, N deposition and CO2 concentrations on systems since 1860 can be attributed to 1.3 Pg of N created by
timescales ranging from a few years to decades (Wardle et al., anthropogenic activities since that time (Zaehle, 2013). Deter-
2004). Thus, understanding variability in nutrient cycling in mining whether terrestrial ecosystems provide a net terrestrial sink
response to climate change requires indices of nutrient availability or source for the additional N created by humans, and the time-
that integrate processes acting on the micro- (i.e. seconds to days) scale over which this occurs, is a key research priority for palaeo-
to landscape (i.e. years to centuries) and regional (i.e. ~1 million and neo-ecosystem ecologists (Lewis et al., 2014).
years) scales (Walker and Wardle, 2014). To address this question, McLauchlan et al. (2007) conducted an
Stable isotope analysis of N (d15N) in organic matter provides a analysis of tree wood d15N sampled from 857 tree ring segments
reliable way to track the availability of N across space (e.g. from (22 trees), which provided a ~170 year record of N cycling in the
foliage, Craine et al., 2009) and time (e.g. from plant macrofossils, northern hardwood forests surrounding Mirror Lake (northeastern
Wolfe et al., 2013; from wood and lake sediments, McLauchlan USA). Results from this study intriguingly indicated that trees (and
E.S. Jeffers et al. / Quaternary Science Reviews 112 (2015) 17e32 21

lake sediments) showed declining d15N values from ~1920 to the to provide a long-term sink for increasing supplies of
present and thus reduced N losses from the terrestrial environ- anthropogenically-derived N, although the persistence of this
ment, supporting the N limitation hypothesis described above. A process is largely determined by local or regional factors. It also
study of herbarium specimens of grass species collected across makes clear the importance of considering the links between the
eastern Kansas between 1876 and 2008 also showed progressive carbon and N cycles when attempting to predict future levels of
declines in foliar d15N from 1940 and declining concentrations of N plant productivity and the provision of key services that are derived
in foliage from 1926 (Fig. 3), despite relatively high levels of N from this (e.g. crops, quality of herbage for grazing and timber
deposition in the region in recent decades (McLauchlan et al., 2010). production and climate regulation).
In contrast, Hietz et al. (2011) presented two series of wood d15N
studies from Barro Colorado Island, Panama and the Huai Kha 2.2. Soil formation
Khaeng Reserve in Thailand which both demonstrate rising d15N
from 1968 to 1920 respectively. Thus, the trajectory of change in N Soils regulate a number of ecosystem processes (e.g. nutrient
availability to terrestrial plants as indicated by wood d15N appears cycling) and support the provision of final services such as crops
to vary by biome. and clean water supply (Dominati et al., 2010); as such, it has been
In order to understand the nature of the relationship between suggested that they can be an essential determinant of the eco-
the dynamics of N cycling and climatic change (including altered nomic status of countries (Daily, 1997). In fact, the United Nations
atmospheric concentrations), a much longer record is needed, General Assembly declared in 2013 the creation of World Soil Day
ideally pre-dating large scale human alteration of the N cycle and in recognition of the important contribution they make to human
spanning contrasting biomes and ecosystems. A global synthesis of wellbeing. Despite the clear importance of soils, current under-
86 lake sediment d15N records spanning the last 15,000 years was standing of their natural capital is incomplete (Dominati et al.,
therefore undertaken by McLauchlan et al. (2013b). Interestingly 2010) mostly due to the static nature of existing global soil maps
this study indicated a more complex picture than had been previ- (Grunwald et al., 2011). Soil formation is a function of parent ma-
ously assumed. In the early postglacial period when atmospheric terial, climate, biota, topography and time; with recognition that
CO2 and temperatures both increased, N availability declined soils can take 1000s of years to develop (Jenny, 1941). Existing
globally and this phenomenon continued for nearly 8000 years global soil maps provide only a snapshot of current soil type and
while terrestrial environments accumulated carbon, which sup- lack information on the processes regulating soil formation; yet this
ports the progressive N limitation hypothesis. However, over the information is critical for assessing both the impacts of climate and
past 500 years, no globally consistent trends are apparent in d15N also land-use change on soil losses, the amount of time it takes to
suggesting that local conditions have had an overriding influence replenish soil stocks and the impact of changing vegetation type in
on the ability of terrestrial ecosystems to absorb and retain N. These determining the type of soil that develops. Efforts are currently
findings together demonstrate the ability of terrestrial ecosystems being made to develop ecosystem process models (e.g. the STEP-
AWBH space-time model for digital soil mapping) that can pro-
duce global soil maps which account for these dynamic processes
(Grunwald et al., 2011); this is an area where palaeoecological re-
cords have much to offer.
Palaeoecological proxies can both inform and validate soil for-
mation models by demonstrating variation in the rate of soil evo-
lution (i.e. pedogenic change over time, sensu Huggett, 1998) in
response to intrinsic (e.g. vegetation cover) and extrinsic (e.g.
climate) changes. Pennington (1971, 1986) presented the first at-
tempts at reconstructing soil processes from elemental analysis of
postglacial lake sediments in the English Lake District. More
recently, Willis et al. (1997) used inductively coupled plasma
emission atomic spectrometry of bulk sediments to reconstruct the
timing of a transition from an acidic podzol to an alkaline brown
earth soil in northeastern Hungary. Their interpretation of the
elemental concentrations as indicators of shifting soil types on the
catchment was validated by a comparable analysis of extant leaf
litter and soils from the basin. The results demonstrated that the
development of a brown earth soil from podzol occurred at least
600 years after a shift from a coniferous to an oak-dominated forest
(Willis et al., 1997). This study provided the first clear and direct
evidence of plant modulation of soil chemistry occurring over
millennia from lake sediment records.
Geochemical data can also be used to infer changes in soil
erosion and influx of mineral matter from the catchment into the
Fig. 3. Scatter plot of stable isotopic values of nitrogen (d15N, ‰ relative to air), a proxy lake as well as soil acidification associated with vegetation and
for ecosystem N availability, and total N (mg g1) in leaf tissue from 545 herbarium
climatic changes (Hu et al., 2001). For example, in a study of Ho-
specimens collected in Kansas (USA) between 1876 and 2008. Decreasing values of
d15N after 1940 show that N availability in plants declined concurrently with rising locene ecosystem changes in Southwestern Alaska, Hu et al. (2001)
fossil fuel burning and thus concentrations of atmospheric carbon dioxide. Piecewise used allogenic concentrations of elements (measured with induc-
linear regression identified inflection points in the trajectories of foliar d15N at 1940 (a) tively coupled argon plasma/atomic emission spectroscopy) in
and foliar N at 1926 (b). It is likely that these observations reflect progressive re- sediments to demonstrate progressive soil stabilization from
ductions in N availability to plants despite anthropogenic N deposition in recent de-
cades. Long-term declines in N availability could result in subsequent declines in the
10,000 yrs BP associated with widespread colonization of the her-
ecosystem services and goods that depend upon regular nutrient availability. Repro- baceous tundra by Betula following postglacial warming. The
duced with permission from McLauchlan et al. (2010). Copyright 2010 New Phytologist. integration of such temporal records across a region can therefore
22 E.S. Jeffers et al. / Quaternary Science Reviews 112 (2015) 17e32

provide a wealth of information on rates and directions of change in vegetation abundance and LOVE (LOcal Vegetation Estimates)
soils and their variation over space, particularly in light of antici- (Sugita, 2007a, b). These models translate raw pollen counts into
pated climate changes. vegetation abundance by taxa at the regional and local scales,
respectively (Sugita, 2007a, b). Testing of these models has
2.3. Biomass demonstrated their vast potential for land cover reconstruction.
Overballe-Petersen et al. (2013), for example, tested the ability of
Plant biomass, a product of primary productivity, is treated as an the LRA models to reconstruct vegetation cover surrounding a
ecosystem process (following Mace et al., 2012) because it is Danish forest hollow by comparing the model output with
fundamentally linked to the provision of many key ecosystem observed vegetation data from historical records (vegetation maps
services and goods (e.g. climate regulation, soil erosion protection, and inventory data from management plans) at selected points in
clean water provision and timber and crop yields). Tracking the time over the last 150 years. The LRA output corresponded very
response rates of plant biomass to climatic perturbations is well with inventory data up to 200 m from the edge of the forest
essential for predicting how the provision of these services may hollow and provided a more accurate representation of vegetation
vary in the future; doing so requires information on the current cover than pollen percentage data (Overballe-Petersen et al., 2013).
amount of biomass (i.e. the ‘standing crop’), its spatial distribution Bayesian hierarchical models of forest composition have also
and how this varies over annual to millennial time scales. Below we been built to reconstruct historic variations in vegetation abun-
present three areas where palaeoecological data can provide dance from fossil pollen percentage data (Paciorek and McLachlan,
important information on temporal and spatial changes in plant 2009). This approach estimates the relationship between modern
biomass: vegetation cover and the representation of these taxa within fossil
pollen assemblages in nearby ponds. Paciorek and McLachlan
2.3.1. Determining changes in amount of plant biomass over time (2009) demonstrated the utility of this approach to generate pre-
When determining changes in the amount of plant biomass over dicted abundances of forest taxa across central New England at
time, fossil pollen accumulation rates (PAR) in sedimentary basins different points in time. The authors applied this modelling
can provide a useful proxy (Magri, 1994). For example, PAR has been approach to pollen data collected from 23 ponds where the sur-
utilised to demonstrate changes in overall plant biomass on gla- rounding vegetation was known for both the present and colonial
cialeinterglacial timescales (Magri, 1994), Holocene changes in era (1635e1800 AD). Cross-validation analysis of their results
biomass for individual plant species (Giesecke and Fontana, 2008; showed that the model provided good a prediction of actual forest
Seppa € et al., 2009), and Holocene biomass changes associated composition at regional (i.e. >50 km) scales (Paciorek and
with changing community and population dynamics over time McLachlan, 2009).
(Sepp€ €ppa and colleagues (2009)
a et al., 2009). In the latter study, Se
demonstrated the potential of using PAR to determine the changes 2.3.3. Mapping biomass response to climate change over space and
that would occur in standing biomass with a climate-driven change time
in community composition. To do this they first compared varia- Currently, species niche or bioclimatic envelope modelling is the
tions in current biomass (tons ha1) of Pinus, Picea and Betula most commonly used approach to predict the expected future
populations at varying distances (0e4500 m, 0e1000 m, 0e500 m, distribution and abundance of plant species in response to climate
and 0e100 m) from the shores of study lakes in Finland with PAR change (Araújo and Peterson, 2012). This approach correlates the
values measured from surface samples from those lakes and from present day distribution of target species with a number of climatic
others where these taxa were not present. Results indicated a close parameters associated with their distribution to create a climate
correspondence between PAR and local biomass at regional scales envelope model. The climate envelope model is then run again
(i.e. up to 4500 m from the lake shore); therefore lake sites with using estimated future climatic conditions to predict the potential
high biomass of Pinus, Picea and Betula in the surrounding catch- distributional changes due to climate change. However, it is widely
ment had correspondingly high influx of pollen from these tree taxa acknowledged that many of these climate envelope models are
and conversely those sites with low or no biomass, had similarly highly sensitive to the algorithms used (Hijmans and Graham,
low PAR. This relationship was then used to examine the variability 2006; Araújo and Peterson, 2012) and a full knowledge of the cli-
of biomass of different populations (in particular Picea abies) during matic sensitivity of the plant taxa involved is required in order to
the Holocene in response to climate change (Sepp€ a et al., 2009). make realistic predictions about future range shifts (Clark et al.,
Results indicated that maximum Picea PAR values observed in 2011). In particular, the distribution of many European plant taxa
Finland during the mid-Holocene are indicative of Picea biomass of is strongly influenced by historical factors including past human
about 50e60 tonnes per hectare, which suggests that modern Picea and climate activities (Willis and Birks, 2006; Froyd and Willis,
biomass in the region is roughly 35e40% of its pre-anthropogenic 2008; Willis et al., 2010). Furthermore, the species occurrence re-
extent. Thus, using fossil pollen data to assess biomass responses cords that are often used to build climate niche models are static
to climate change provides information on a greater extent of and may not represent the true range of climatic conditions that
environmental conditions than is possible with modern observa- species can tolerate. Quantifying the speed at which species will
tions alone. respond to climate change, and projecting the potential future
ranges of key plant species requires occurrence records that
2.3.2. Spatial distribution of plant biomass correspond with the generation time of the major organisms under
Fossil pollen data e in their raw form e have often been criti- investigation. Without this, climate envelope models are unable to
cised for their inability to represent changes in land cover (Gaillard capture the full range of variability of the species under
et al., 2008). In the past decade, a number of approaches have been investigation.
developed to translate pollen percentage and accumulation rate Fossil pollen data therefore have an important role in testing
data into estimates of relative abundance of individual plant taxa in these models through the process of hindcasting. This is where the
the surrounding catchment and thus provide changes in the spatial species envelope model is run back in time (using palaeolimatic
distribution of plant biomass over time. Examples of two Landscape data) and the predicted distribution of the species is examined
Reconstruction Algorithms (LRA) that have been developed for this against the actual distribution as evidenced in the fossil pollen
purpose include REVEALS which provides a regional estimate of records. A good example of this was provided by Alba-Sanchez et al.
E.S. Jeffers et al. / Quaternary Science Reviews 112 (2015) 17e32 23

(2010) in their use of the species distribution model MaxEnt warming period, Little Ice Age and present). While the crop species
(maximum entropy machine-learning algorithm) to estimate past were the most difficult to model e due to human management
distribution changes in Iberian Abies (fir) under changing climate effects on plant occurrence e the modelled distributions yielded
scenarios including the Last Glacial Maximum (21 kya) and the highly accurate predictions (AUC > 0.90) when compared to the
mid-Holocene (6 kya) across the Iberian Peninsula. From this study fossil output. However, the authors also demonstrated that a key
they concluded that the integration of modern occurrence records factor affecting the predictability of crop taxa distribution was the
of Abies species, current ecological-niche characteristics associated cultural aspects driving cultivation patterns, such as changes in
with these occurrences, and palaeoecological evidence of past V. vinifera planting under Muslim versus Christian rule in southern
occurrence provided a more robust bioclimatic niche model, Europe around the 10th century (CE) (Macias-Fauria and Willis,
particularly for identifying the locations of climatic refugia. 2013). Thus, historical e as well as the environmental e contexts
can decouple species responses to climatic variables (Jackson et al.,
3. Final ecosystem services 2009) and future advances in species distribution modelling of long
term ecological data should be focused on the development of new
The outputs of ecosystems from which people derive direct approaches that can incorporate the impacts of discrete historical
benefits in terms of welfare gains and/or losses, such as crops, a events within the prediction of species ranges.
clean water supply and an amiable climate, are classified as While species distribution modelling can provide information
ecosystem services (UKNEA, 2011). However, the knowledge on the areas that are likely to provide the optimal climatic condi-
required to understand and manage these ecosystem services tions for crop growth, novel methods in fossil crop grain research
ranges from local scale considerations (e.g. farm fields for crops) to can provide additional information about changes in the yields of
whole river basin catchments (e.g. clean water provision), to global species within that range. In cereal crops, grain weight is an
scale climatic processes. Each of these processes range over a important indicator of grain yield, since larger grain weights
similar variety of temporal scales (i.e. seasonal to millennial). It is directly correspond with higher yields in the end-products of
extremely difficult to predict the complex ways in which current wheat, such as flour (Henry and Kettlewell, 1996). Cereal grains are
and expected future environmental changes will shift the spatial often found in archaeological sites; however the process of pres-
and temporal mosaic of ecosystem system service provision ervation (i.e. carbonisation) acts to reduce grain mass and dimen-
(Mooney et al., 2009) and it is a research priority to find datasets sion (Ferrio et al., 2004). In order to compare the weight of ancient
that can aid the development of a more detailed understanding of grains to those grown in modern conditions, it is necessary to
environmental change impacts on ecosystem service provision and correct for this effect. Ferrio and colleagues (2004) used morpho-
how this will affect human well-being into the future (Pereira et al., logical analysis of grains obtained from archaeological deposits
2010). Examples of three key provisioning services where palae- across Catalonia, Spain, dated to 3900e2200 cal. yrs BP to develop a
oecological records provide important temporal data to enhance regression model that provides an estimate of the original grain
this understanding are crops, water supply and quality, and climate weight (mg) at the time of burial given data on the length (mm),
regulation. breadth (mm) and thickness (mm) of experimentally carbonized
modern grains of wheat and barley (Ferrio et al., 2004). By
3.1. Crops comparing the reconstructed grain weights with independent
sources of environmental data (i.e. temperature or moisture), the
The provision of food is one of the most direct ways in which authors were able to tease out the effects of increased water
ecosystems provide services to people and a major global challenge availability due to beneficial climate change versus improvements
is to determine current and future impacts of environmental in grain weight from human management activities such selective
change on food availability e particularly crops (Godfray et al., breeding for heavier grains by early farmers.
2010). In order to meet expected food demand in 2050, scientists Another approach for measuring the direct impacts of climate
need to be able to predict the areas for optimal crop growth given change, in particular water availability, on grain size is to use stable
likely climate scenarios (Foresight, 2011). This requires dynamic isotope analysis of carbon (d13C) from individual grains. In a study
data on crop yields and their relationship to changes in climate as by Araus et al. (1997) on fossil cereal grains from Catalonia and
well as nutrient and moisture availability. Currently, the temporal Andalusia (SE Spain) dated between 7000 and 2200 cal. yrs BP, for
extent of global data on crop production and climate change used in example, the authors demonstrated that carbon isotope discrimi-
ecosystem assessments is limited to ~30 years (Lobell et al., 2011). A nation (D13C) in individual grains (i.e. the difference between the
number of recent palaeoecological and archaeological studies have d13C values of air and that of the grains) corresponds well with
started to demonstrate the utility of fossil pollen and stable isotope water availability to ancient crops (Araus et al., 1997). Thus carbon
analysis in extending this record much further back in time. isotope discrimination can be used as a proxy reflecting both
Similar to modelling biomass distributions through time, the changing climatic conditions and irrigation practices (Ferrio et al.,
typical method for determining future distribution of crops in 2005). Such insights can provide a clearer understanding of early
response to climate change, is through the use of species envelope agricultural practices, and their long-term impacts on biodiversity
models. However, as mentioned above, these are strongly influ- and ecosystem processes (Whitehouse and Kirleis, 2014) which
enced by the algorithms used to create the models and therefore have relevance for assessing and potentially improving modern
independent validation under varying conditions is essential. Fossil pastoral crop management practices particularly in the face of cli-
records have great potential to do this through the comparison of matic changes.
species distribution model outputs (run backwards in time) with Stable isotopic values of N within fossilized grains can also
actual distribution as displayed in fossil datasets (Fig. 4). For provide a reliable record of variations in soil fertility at the time of
example, Macias-Fauria and Willis (2013) aggregated data from plant growth, a factor that can have a major impact on crop yield.
fossil pollen assemblages, landscape features (i.e. relief heteroge- For example, in a study of grain weight from charred grain remains
neity) and global palaeo-climate model output in order to generate of wheat and barley in Granada, SE Spain spanning 1500 years
the probability of occurrence of seven economically valuable spe- following the establishment of agriculture in the region around
cies, including the crops Olea europea and Vitis vinifera, across 6000 cal. yrs BP, Aguilera et al. (2008) found that grain weight
Europe for three climatically distinct time periods (the medieval declined dramatically while D13C values (indicating water
24 E.S. Jeffers et al. / Quaternary Science Reviews 112 (2015) 17e32

Fig. 4. Species distribution maps of important crop species Olea europea and Vitis vinifera describe the past ranges of these taxa during historic periods of climatic warming (i.e.
Medieval Warm Period) and cooling (Little Ice Age) episodes relative to their modern distributions (i.e. 20th Century Warming). Fossil pollen data for these species were used to
constrain climate- and landscape feature-based models of these crop species' range in each period of time. Copyright 2012 Global Ecology and Biogeography.

availability) remained stable. They found that the observed decline chronologies to extend the temporal extent of past flow regimes
in grain weight was linked with declining soil fertility (as indicated (Jain et al., 2002). A good example of this approach can be found in a
by decreasing values of grain d15N), not changing climate. Thus, a project that determined long-term variability in the flow of the
progressive loss of soil fertility resulted in a halving of cereal grain Upper Colorado River (Gangopadhyay et al., 2009). In this study, the
weights over 1000 years despite the presence of favourable climatic authors compared the instrumental record of past river flows
conditions for growth (Aguilera et al., 2008). By reconstructing the available from the Lees Ferry Gauge data (spanning the interval
changes through time in the N and water status of crops along with 1906e2005) with tree-ring widths spanning the same interval of
measures of yield (via grain weight), the authors were therefore time from a number of living trees within the catchment. A clear
able to identify the long-term effects of early unsustainable farming relationship between actual-flow data and tree-ring width was
practices on the provision of important food crops (Aguilera et al., demonstrated (Fig. 5), which enabled a transfer function to be
2008). Such data have the potential to provide an empirical basis developed that translated tree-ring chronologies into estimated
for future scenario models of food supply changes in response to rates of river flow (million acre-feet, MAF, yr1). This transfer
climate change and changing land management techniques function was then applied to tree-ring records dating back to 1482
(Pereira et al., 2010). to determine the long-term variability in the river flow. This is in
itself an interesting study on baseline conditions; however, the
3.2. Water supply and quality authors also used the data to build a river flow regime model
capable of predicting the probability of a regime shift between
Access to clean, reliable freshwater supplies is a fundamental persistently wet or dry states in the river given the number of years
requirement for human life, yet only 25% of global total river runoff since the previous shift (Gangopadhyay and McCabe, 2010). This
and groundwater recharge is available for human use and 40% of provides an exceptionally useful tool for managers aiming to allo-
this amount is already being removed solely for this purpose cate water to the many users across the basin.
(Heathwaite, 2010). Demand for water will increase with popula- Palaeoecological data from sediment buried in rivers and lakes
tion growth, yet climate warming and pollution are reducing can also provide important baseline information on variation in
existing fresh water supplies (Vorosmarty et al., 2000; Heathwaite, water supply and quality. This is particularly important evidence for
2010). Water resource planners need to know how water supply government agencies tasked with the implementation of clean
varies under different environmental conditions relative to current water legislation. The European Water Framework Directive (WFD),
and projected demands. However, historical records of river flow for example requires that biological, hydromorphological and
(in selected regions where such records were regularly collected) chemical indicators of water quality should be based on the degree
are typically limited to the last 30e50 years (IPCC, 2001). The to which present-day conditions deviate from those expected in the
temporal and spatial extent of the instrumental record is therefore absence of human influence. The Lotic invertebrate Index for Flow
too short to capture natural cycles of drought and recovery over Evaluation (LIFE) and the Palaeo-LIFE indices, for example, have
millennia within the hydrological system, particularly in light of the been developed which use extant and fossil macroinvertebrate
anticipated effects on water availability by climate change species assemblages (from the Coleoptera, Trichoptera and Chiro-
(Kundzewicz et al., 2008). nomidae orders) to characterise past river flow regimes (Extence
There are a number of palaeoecological proxies that are partic- et al., 1999; Greenwood et al., 2006; Monk et al., 2006; Howard
ularly useful for indicating vulnerability in water supply under et al., 2009, 2010).
potential climate change scenarios including using tree-ring
E.S. Jeffers et al. / Quaternary Science Reviews 112 (2015) 17e32 25

increases in TP and one loch had a decline in TP since pre-Industrial


times (Bennion et al., 2004). Thus the individual nature and history
of land-management practices around each site has strongly
influenced the changes in water quality in response to nutrient
loading. It was concluded that restoration to pre-Industrial condi-
tions, as required by the WFD, must proceed based on loch-specific
conditions, not a simplistic relationship between TP loading and
water quality.

3.3. Climate regulation

The ability of woody plants to take up carbon dioxide (CO2) from


the atmosphere and store it within woody biomass provides a
benefit to people in the form of climate regulation (Sedjo and
Sohngen, 2012). Payment for ecosystem service (PES) schemes are
being developed (e.g. Reducing Emissions from Deforestation and
forest Degradation, REDD) that aim to assign a monetary value to
forests that are left intact. The value of forests is determined by
their ability to remove and store anthropogenic emissions of CO2
Fig. 5. Scatter plot of reconstructed and measured flow rates (million acre feet, MAF,
(Kindermann et al., 2008). The success of PES schemes for carbon
year1) of the Colorado River at the Lees Ferry, Arizona (USA) gauge for the years
1906e2005. Annual stream flow was reconstructed from a network of tree ring storage, therefore, depends upon obtaining accurate assessments of
chronologies across the Upper Colorado River Basin. The least squares fit line shows the current and future ability of forest plots across the globe to store
that a very strong relationship exists between the reconstructed and observed values carbon (Asner et al., 2010). Since the carbon storage ability of
of stream flow, which allows researchers to infer palaeo-stream flows from dendro-
woody plants is affected by dynamic environmental factors such as
chronologies that pre-date instrumental records. Reproduced from Gangopadhyay
et al. (2009). Copyright 2009 Water Resources Research. climate, nutrients, and fire, assigning a carbon storage value to a
forest is particularly challenging and presents an important
research challenge (Bonan, 2008).
In a study by Monk et al. (2006), for example, ten years Recent work, for example, has indicated that although between
(1990e2000) of stream flow measurements taken from 83 rivers in 2000 and 2007 the global forest carbon sink removed 2.5 billion
England and Wales were used to establish classes of river flow tonnes of carbon per year from the atmosphere (Pan et al., 2011),
velocities (e.g. <20 cm s1, 20e100 cm s1 and >100 cm s1). The the amount drawn down by different forest types varies greatly.
macroinvertebrate species associated with these different veloc- The most draw-down was contributed by tropical forests (1.3
ities were measured and then a stepwise, multiple linear regression billion tonnes per year), followed by temperate forests (0.8 billion
model of the data was used to translate the presence of macro- tonnes per year) and boreal forests (0.5 billion tonnes per year).
invertebrate species within each flow regime into an index of flow Probably most surprising was the finding that within the tropical
velocity. The resulting index provides a reliable approach for zone secondary forests (i.e. those that are recovering from past
reconstructing previously unknown river flow velocities from deforestation and logging) provide the largest sink, ~1.7 billion
known invertebrate assemblages. The method was also used by tonnes of carbon per year due to their relatively rapid early forest
Howard et al. (2009, 2010) to reconstruct palaeo-flow regimes growth (Pan et al., 2011). An understanding of how these different
(PalaeoLIFE index) from sub-fossil assemblages of Coleoptera, Tri- forest types respond to climatic perturbations and how quickly
choptera and Chironomidae species buried in exposed fluvial se- they recover from previous disturbance are therefore important
quences of the River Trent in England during the middle to late factors that need to be taken into consideration when modelling
Holocene. In this study Howard et al., found a shift from moderate potential draw-down.
to slow flow velocities that was coincident with increased erosion Palaoecological records have much to offer in this respect e not
from Neolithic clearing and animal grazing (Howard et al., 2010). only providing evidence of the type of forest which is likely to re-
In addition, the use of palaeoecological data to reconstruct water turn following disturbance (be it human activities or climatic per-
quality baselines for standing water-bodies is illustrated in a study turbations) but also the amount of time required for a particular
by Bennion et al. (2004) that used fossil diatom assemblages to forest type to recover to previous levels of biomass (and thus car-
assess eutrophication impacts and baseline reference conditions for bon storage capacity). In a study by Cole et al. (2014) for example,
Scottish freshwater lochs. In this, they collected sedimentary cores which examined rates of recovery of tropical forest following pre-
from 26 Scottish freshwater basins (lochs) spanning the time in- vious disturbance events using fossil pollen sequences from the
terval c 1850 (i.e. pre-Industrial) to the present and reconstructed four major areas of tropical forest globally (South America, Central
the changes that occurred in the fossil diatom assemblages. By America, Africa and Southeast Asia), results indicated huge varia-
comparing the base and top samples for each core they were able to tion in recovery rates. In some tropical regions, recovery of the
determine the amount of turnover and floristic change in the lochs, forests occurred within 30 years while in other regions it took up to
a factor which would have been strongly influenced by the nutrient 500 years. There were also significant differences in rates of re-
status of the lake. They found that periods of time with high covery between regions. Forests in Central America recovered the
turnover rates were indicative of high nutrient enrichment and quickest whereas South American forests recovered the slowest.
thus low water quality. Given the tight coupling between diatom The type of disturbance also affected recovery rates. Recovery was
compositional changes and levels of phosphorus concentrations in the quickest following natural disturbances (e.g. hurricanes, fires)
the lake, the authors used a transfer function to translate diatom compared to human disturbance (Cole et al., 2014). This global
species assemblages into concentrations of total phosphorus (TP) in analysis provides important baseline information for modelling
the water (mg l1). This latter measure is particularly relevant for variation in carbon budgets in response to perturbations.
management plans because the results indicated that although 19 Palaeoecological data has also been used to constrain global
lochs have had increases in TP, 6 lochs had no change or negligible models of past amounts (Pg C) of carbon storage in the terrestrial
26 E.S. Jeffers et al. / Quaternary Science Reviews 112 (2015) 17e32

biosphere (Peng et al., 1998). For example, Prentice and colleagues


used a dynamic global vegetation model called Land Processes and
eXchanges (LPX) to estimate global terrestrial carbon uptake at the
Late Glacial Maximum (LGM) relative to the pre-Industrial Holo-
cene period (Prentice et al., 2011). Pollen-based reconstructions of
biome distributions derived from BIOME 6000 (Prentice et al.,
2000) at each time period showed overall agreement with the
LPX simulated biome distributions, thus validating the ability of the
model to reconstruct biome shifts. Then the validated LPX model
was used to show that carbon storage in vegetation and soils at the
LGM was 620 Pg C less than in the pre-industrial Holocene period
and this was largely due to reduced carbon density in forest biomes
at that time. The model-based difference in terrestrial carbon
storage between the LGM and the pre-industrial Holocene was
found to fall within the range of values (i.e. 300e700 Pg C) inferred
from stable isotope analysis of carbon d13C in marine sediments
(Prentice and Harrison, 2009).
Forest fires are also known to affect the ability of terrestrial
ecosystems to store carbon in forest soils (Nave et al., 2011), yet
existing PES schemes such as the Reducing Emissions from Defor-
estation and Degradation (REDDþ) strategy have not yet incorpo-
rated the effect of forest fires on carbon storage (Barlow et al.,
2012). Given the evidence for recent increases in wildfires
(Westerling et al., 2006), it is essential to understand how long-
term changes in fire regime may affect the ability of forests to
provide this climate regulation service. Estimates of carbon release
from terrestrial ecosystems due to fire (derived from charcoal
accumulation rates, see more about this in Section 4.1) show that
biomass burning in the Holocene has had significant impacts on
global carbon dynamics (Carcaillet et al., 2002). This is an important
area of future research and resources such as the Global Paleofire
Database (Power et al., 2010) have the potential to provide an
invaluable resource for an integrated analysis of the complex re-
lationships between fire regime and carbon storage capacity of Fig. 6. Logs of Diptocarpaceae tree species harvested from natural forests in Sabah,
Malaysia (a). Tree ring chronologies can be reconstructed from cores taken from living
terrestrial ecosystems. or fallen trees (b), which provides high temporal resolution data on wood growth rates
(volume m3 year1) and how this varies with respect to intrinsic and extrinsic drivers
4. Goods of change. Images reproduced, with permission, from (a) Jake Snaddon and (b) Sandra
Nogue .

Ecosystem ‘goods’ are the outcomes that final ecosystem ser-


vices produce and the values that these generate for people. The
value of ecosystem goods can be based on a monetary measure (e.g. account the variability in wood production associated with these
the market cost of timber) or a measurable improvement in health drivers. Foresters have traditionally relied upon repeated mea-
(e.g. reduction in mortality rate) or personal security (e.g. reduced surements of tree diameter at the breast-height (DBH) of the
loss of property). The production of ecosystem goods typically in- researcher as a proxy for increases in timber volume over time
volves some form of human investment (Mace et al., 2012), such as because this information can be retrieved quickly from a number of
a production process that transforms trees into timber for building trees in a research sample plot. However, repeated DBH measure-
purposes. Three examples of ecosystem ‘goods’ where the long- ments over time taken from experimental forest plots have known
term perspective provided by palaeoecological records can limitations namely: 1) observations are limited to sampling periods
greatly enhance knowledge of its value and variation over time are and miss the variability in annual and sub-annual growth rates that
timber, soil erosion protection and coastal erosion protection. occur between these (Metsaranta and Lieffers, 2009) and 2) the
temporal extent of sampling efforts tend to be short relative to the
4.1. Timber lifespan of the study trees (Rozendaal and Zuidema, 2011).
Wood ring growth analysis of increment bores taken from living
Human demand for timber resources is increasing and much of trees (Fig. 6b) provide precise, annualized data on variability in
this demand is being met from managed plantations (i.e. 35% of wood growth which is not captured in DBH measurements.
harvested roundwood), although some premium woods continue Metsaranta and Lieffers (2009) for example, used dendrochrono-
to be harvested from natural, old-growth forests (Fig. 6a) (Sampson logical analyses of radial growth in trees to assess differences in
et al., 2005). In order to set sustainable harvest rates for these high- estimates of stand volume growth, mortality and stem wood
value forests, managers typically develop timber harvest plans that biomass between DBH measurements taken at 10-year intervals
require information on current stocks (i.e. standing volume) and and wood cores and disks taken from living and dead trees
predictions of future timber availability given the expected growth (respectively) from an unmanaged jack pine (Pinus banksiana)
rate of existing trees and new recruitment. Since individual tree stand in the Canadian boreal forest. They found that DBH mea-
growth is affected by intrinsic factors such as age and competition surements taken on 10-year intervals miss significant annual
with neighbours, as well as extrinsic factors including climate and variation in volume growth, mortality and snag-fall, and thus do
disturbances from fire and pests, predictions need to take into not provide an adequate measure of the stochastic processes acting
E.S. Jeffers et al. / Quaternary Science Reviews 112 (2015) 17e32 27

on wood growth in the boreal forest which are recorded in annual Where insect remains are poorly preserved, such as in alkaline
growth rings. lakes, fossil pollen assemblages and geochemical analysis of bulk
Dendroecological studies can also be used to predict the amount material in lake sediment sequences can be used to indicate the
of time required for high-value forest trees to reach the minimum timing of past outbreaks of insect defoliators (Morris et al., 2010,
diameter for cutting (i.e. to set the minimum cutting cycle); this is 2013). Morris and Brunelle (2012), for example, showed that ra-
essential baseline information for forest managers aiming to opti- tios of host to non-host taxa pollen accumulation rates can provide
mise their harvests. In a study by Therrell et al. (2007), for example, a conservative estimate of past, large magnitude outbreaks of
the researchers obtained tree ring measurements and tree age spruce bark beetles (Dendroctonus rufipennis Kirby). Furthermore, a
determinations from increment bores to show that Pterocarpus key strength of lake sediment-based reconstructions of past pest
angolensis requires 100 years of growth before it will reach the outbreaks is the ability to reconstruct the cascading effects of in-
minimum cutting diameter in southern Africa (Therrell et al., 2007). festations on ecosystem processes such as soil erosion (and the soil
What has also been demonstrated is that these species-specific erosion protection service provided by this process) and N leaching
measurements vary strongly with respect to site conditions (and the service of nutrient pollution control) when combined with
(Rozendaal and Zuidema, 2011). additional biogeochemical proxies such as sedimentary geochem-
While climate change may affect the ability of valuable timber istry and stable isotope analysis (see Sections 2.1 and 2.2). These
species to persist in their current locations over the long-term factors in turn determine forest response to disturbance (Morris
(Macias-Fauria and Willis, 2013), more immediate threats to et al., 2013; McLauchlan et al., 2014); therefore unpicking the
standing stocks of timber may be provided by disturbances such as feedback mechanisms between them is an important area of cur-
pests (e.g. bark beetle, Raffa et al., 2008), pathogens (e.g. ash dieback, rent research (see the Novus Research Coordination Network for
Kowalski, 2006) and increased wildfires (e.g. in boreal forest, Macias more information: https://novusrcn.wordpress.com/tag/nitrogen/).
Fauria and Johnson, 2008) as well as the interactive effects of these Wildfires, like pest outbreaks, reduce tree growth, destroy the
disturbances on nutrient supplies to forests (McLauchlan et al., commercial value of forest stands and can limit the future yields of
2014). Pest outbreaks in particular can greatly reduce timber timber species (MacLean, 1990). What is also known however, is
yields; therefore, forest managers need to know the extent to which that many forests have a regular cycle of burning related to both
timber volume is reduced by outbreaks of varying intensity and how abiotic drivers (e.g. seasonality and lightening strikes), and biotic
this varies under different environmental conditions. factors (e.g. the amount of standing dry mass e and thus flamma-
There are now a number of on-line tools available that can bility of the system). Understanding this burning regime is critical
predict the impact of an infestation on timber yield. The Spruce to the management of forests e especially if burning is used as a
Budworm Decision Support System (SBW DSS), for example, pro- tool to create forest gaps and encourage more plant diversity
vides forest managers with an estimated benefit (in terms of (Turner et al., 1994; Colombaroli and Tinner, 2013). Dendroeco-
marginal timber supply, m3/ha) of protecting stands against spruce logical and micro-fossil data recovered from lake sediments can
budworm (Choristoneura occidentalis) defoliation (MacLean et al., therefore provide important temporal information on the timing
2001). In this model, the predicted future benefit of management and intensity of past fire events. Higuera and colleagues, for
interventions is determined by historical defoliation impacts. example, have shown that information from fire scars on extant
However, these estimates are derived from empirical records of trees can be integrated with charcoal accumulation rates (CHAR) in
previous defoliation impacts spanning a period of only five years. forest hollows (Higuera et al., 2005) and lake sediments (Higuera
Incorporation of longer-term records would greatly improve the et al., 2011) in order to establish precise fire return intervals
predictive ability of such tools. Dendroecological records in (years), estimates of area (ha) burned per fire (a proxy for fire in-
particular can provide information on past outbreak impacts on tensity) and how this varies with respect to stand age and pre-
tree growth (Swetnam and Lynch, 1993) and the effect of infesta- vailing climate over centennial to millennial timescales. A
tion on the radial growth of trees (Swetnam et al., 1985). Swetnam validation study conducted on a 300-yr tree-ring based fire history
and Lynch (1993) for example, used this approach to show that past spanning a 128,840 ha area within central Yellowstone National
western spruce budworm outbreaks in the southern Rocky Park showed that peaks in CHAR represented known fire occur-
Mountains led to a 50% decrease in the average maximum radial rences within 1.2e3.0 km of their study lakes while total CHAR
growth of host trees during each outbreak that occurred between values provided a reliable prediction of the area burned per fire
1700 and 1983 (with an average duration of 12.9 years). Further within 6e51 km of the lakes. This is a particularly good example
work showed that outbreaks in this region were associated with where multiple lines of palaeoecological evidence were combined
wet periods, not drought years as was previously believed with statistical analysis in order to translate fossil data into the
(Swetnam and Betancourt, 1998). values or currencies required by ecosystem managers. In fact, fossil
In another study, Simard et al. (2002, 2006) used macrofossil records of past fire regime dynamics (i.e. pollen and charcoal) were
remains of insect pests such as spruce budworm faeces and Lepi- the only palaeoecological indicators included in the Millennium
doptera head capsules buried in thick humus deposits and mires in Ecosystem Assessment (MA, 2005).
Quebec, Canada to extend the chronology of past pest outbreaks
beyond the lifespan of living trees. They demonstrated that the 4.2. Soil erosion protection
abundance of fossil insect remains corresponded in time with
changes in local plant species composition (inferred from fossil Soil erosion is a natural process that can be accelerated by
pollen assemblages); however, intense budworm activity as seen in climate change and human activities; this can lead to significant
the 20th century was found to be a rare occurrence during the economic costs to people due to the loss of productive land for
Holocene and these large outbreak events tended to coincide with agriculture (de Groot et al., 2002) and impacts on water quality and
periods of low fire recurrence and warm climatic conditions supply (see Section 3.2). Ecosystem components (i.e. vegetation
(Simard et al., 2006). The mechanisms underlying the interaction cover) and processes (i.e. water and sediment retention by plants)
between fire recurrence and insect outbreaks in forests however act to reduce the flow of sediments off of the landscape and into
remains unclear, and is a priority area for research into the factors water bodies, where additional losses can be incurred in terms of
determining forest resistance and resilience to disturbance (Flower water quality changes. Afforestation policies are widely promoted
et al., 2014). in order to reduce erosion; however this management approach is
28 E.S. Jeffers et al. / Quaternary Science Reviews 112 (2015) 17e32

not appropriate for all ecosystems. For example, expensive affor- from a Caribbean mangrove to assess the time required for
estation projects have been undertaken in arid and semi-arid areas mangrove species to recover (i.e. to be present in the fossil pollen
in China (i.e. 100 billion USD since 2000), particularly in natural record) following disturbances from sea level change, hurricanes,
grass and shrublands, in an effort to reduce erosion and stop the climate change and human disturbances. They showed that re-
effects of desertification. However, these projects have instead had covery of mangrove species (e.g. Rhizophora, Laguncularia)
undesirable feedback effects on natural vegetation cover, soil following a major hurricane took up to 100 years while recovery
quality and water availability (Farley et al., 2005; Cao et al., 2011), from clearance due to changes in land use (from coconut planta-
such that in many places the water table has dropped 30e50%, tions to commerce) took less than 10 years (Gonza lez et al., 2010).
resulting in increased rates of erosion and desertification due to the The management benefit of this information is clear: the data show
poor selection of tree species for afforestration projects. that the costs of mangrove clearing (i.e. loss of fishing income and
What has been learnt from these studies is that there is a critical land stabilization services) will persist for at least a decade,
knowledge gap on how changes in climate and vegetation cover therefore the short-term benefits of clearance must exceed the net
relate to changes in the ability of the landscape to retain sediment. present value of the expected costs in terms of lost ecosystem
Filling this gap requires information on the flows of soil from land services over this period.
into sedimentary basins over decadal-to-millennial timescales. A
palaeoecological technique that is particularly relevant in this 5. Future directions of palaeo-ecosystem service research
respect is the measurement of magnetic susceptibility of sedi-
mentary sequences where the ability of a landscape to regulate Human impacts and natural environmental change are altering
sediment flows through time can be reconstructed from frequency the ability of ecosystems to provide essential goods and services
dependent magnetic susceptibility (MS) values of lake sediments (MA, 2005). The ability of land use planning and ecosystem man-
(Dearing et al., 1987, 2012); high values of MS equate to high rates of agement tools to deliver sustainable management solutions in the
sediment loss (measured as tons ha1 year1) within the catchment face of on-going environmental change is dependent upon the in-
and thus low levels of erosion regulation. A good example of using clusion of information on the long-term variation in ecosystem
this approach can be found in Dearing et al. (2012) where they service provision; this has yet to be achieved. This review has
mapped the provision of this service from 1800 to 2006 to show illustrated that (1) palaeoecological data are capable of being used
that agricultural intensity in the Lower Yangtze basin was coinci- to reconstruct the baseline dynamics of a large number of essential
dent with a progressive loss of erosion protection and other ecosystem processes, final services and goods, and (2) these data
services. are available at the high spatial and temporal resolution that is
Variation in soil erosion protection can be visualized over space required for managing biodiversity and ecosystem services at the
as well as time with the use of a cellular automaton model. landscape and regional scales. However, three major developments
Coulthard et al. (2002) describe a hydro-geomorphological model are now required in palaeoecological research in order to underpin
called CAESAR that can simulate changes in the hydrological and the utility of long-term data for understanding and mapping vari-
sediment regime of a catchment given input data on detrital sedi- ation in ecosystem service provision over time and space (Fig. 7) as
ment influx and sedimentation rates obtained from lake sediments. follows:
Model outputs provide spatially explicit information on the de- 1) Firstly, it needs to be recognised that novel analytical tech-
livery of sediments into a lake and how these vary with respect to niques in palaeoecology and their validation in modern and
past land use and climate changes. The CAESAR model is able to experimental studies have greatly advanced our ability to recon-
simulate sediment discharge rates (m3 year1) over short and long struct changes in ecosystem processes and function over time. A
time scales (i.e. 10e100,000 years) at a fine temporal (i.e. hourly to few good examples of where this has recently been accomplished is
annually) and spatial (i.e. 2e50 m2 cells) resolution. It can also with sedimentary d15N as a proxy for N availability (McLauchlan
separate the effect of multiple drivers such as rainfall and land use et al., 2007), dung fungal spore accumulation rates (Gill et al.,
change on the movement of sediment through the catchment 2013) and beetle assemblages (Smith et al., 2014) in sediments as
(Welsh et al., 2009). This model can be used to anticipate potential proxies for large herbivore biomass, and PAR as a proxy for plant
future changes in sediment transport given different management biomass dynamics (Sepp€ a et al., 2009). These developments are
strategies and potential future climate change scenarios, which will rapidly increasing the amount of information about ecosystem
be invaluable for informing restoration activities before costly processes and services that can be derived from palaeoecological
mistakes are made. studies. Furthermore, by demonstrating precisely how modern
measurements equate to their fossil proxies, these developments
4.3. Coastal protection are enabling researchers to cross boundaries between neo- and
palaeo-ecological research, which is essential in order for palae-
Coastal ecosystems produce disproportionately more ecosystem oecological data to be included in ecosystem assessments. As new
services than most other systems yet these valuable resources are palaeoecological proxies are developed, it is essential that they are
experiencing the most rapid environmental changes (Agardy et al., validated against modern ecological measures used in ecosystem
2005). Intact mangrove habitats in particular are very important for assessments.
marine productivity, local shrimp and fish production, coastal land 2) Secondly, recent developments in modelling techniques have
stabilization and storm protection; however they are also cut down also advanced our ability to translate raw palaeoecological data into
to provide wood to meet local fuel and construction demands units that are comparable with neo-ecological observations and/or
(Walters et al., 2008). There is thus an inherent trade-off between compatible with ecosystem service mapping tools (Table 1). For
the short-term benefits of mangrove clearing (i.e. fuel and con- example, statistical techniques allow us to translate raw pollen data
struction material provision) and the long-term benefits of into changes in vegetation density over time and space (Sugita,
mangrove protection in terms of increased shrimp and fish harvest 2007a, b; Paciorek and McLachlan, 2009). Another advancement
income and increased protection to coastal properties (McNally is in the use of palaeoecological data to build and validate
et al., 2011). However, to fully understand this trade-off, it is ecosystem process models, which simulate complex ecosystem
essential to determine the amount of time required for mangroves dynamics at a variety of spatial scales and allow us to determine
to regenerate. Gonza lez et al. (2010) used fossil pollen data taken how changes in one or more variables cascades through the system
E.S. Jeffers et al. / Quaternary Science Reviews 112 (2015) 17e32 29

Fig. 7. Conceptual diagram showing how palaeoecological data taken from individual study sites can be integrated across space and time to generate maps that highlight areas that
have sustained a relatively high level of ecosystem service provision over time.

and leads to changes in plant community dynamics and ecosystem availability, etc.) at specific time slices (e.g. every 50 or 100 years)
functions (Anderson et al., 2006). Models such as these are used in from the data stored in these databases (Fig. 7). Here we have
ecosystem assessments (e.g. the MA and UKNEA) to generate demonstrated the availability of palaeoecological data to meet this
possible future trajectories of ecosystem change given alternative need; what remains to be achieved is the translation of these data
management and environmental change scenarios (MA, 2005). into map-able information layers. Achieving this requires the
When this is done within a Bayesian Belief Network model (BBN) development of a common framework for the interpretation of
(Chen and Pollino, 2012), the predictions can be generated along palaeoecological proxy data, their chronologies and spatial
with an estimate of probability (UKNEA, 2011). These BBNs are also boundaries.
able to utilize a mix of qualitative and quantitative data from This review has illustrated a small sub-section of the cutting-
multiple sources as inputs for ecosystem model development edge research that is already being done in the developing field
(McCann et al., 2006). Thus, palaeo- and neo-ecological data (either of palaeo-ecosystem service research; there is a huge amount of
as continuous data or discrete events) are able to be integrated relevant data e but the task is now to illustrate its usefulness to
within a holistic model of current and future ecosystem dynamics. non-palaeoecologists. This can be achieved by utilising current (and
In order for long-term data to be utilised within the tools and future) site-scale records to generate spatially and temporally
modelling approaches used by landscape and ecosystem managers, explicit tools for managing biodiversity and ecosystem services
the palaeoecological community must take the lead in demon- across landscapes so that they might be sustained into the future
strating the benefits (in terms of improved ecosystem service despite impending global environmental change.
outcomes) that can be derived from including these data within
existing ecosystem modelling frameworks. Achieving this requires Acknowledgements
increased training in the use of such modelling methods by
palaeoecologists and wider inclusion of such modelling approaches ESJ was funded by a James Martin Research Fellowship from the
within traditional palaeoecological studies. Oxford Martin School. SN acknowledges financial support from a
3) Lastly, sustainable land use planning requires the integration Postdoctoral fellowship from the Spanish Ministry of Education
of ecological information over large spatial scales (de Groot et al., (EX2009-0669) and the Norwegian Academy of Science and Letters
2010). The synthesis of palaeoecological records from palae- through the VISTA programme. Thanks to Jake Snaddon for giving
oecological databases allows for the reconstruction of ecosystem advice on figure design and providing the picture of harvested trees
dynamics over space as well as time. The recent proliferation of in Sabah.
palaeoecological databases (NEOTOMA, Global Pollen Database,
Global Charcoal Database, NCDC etc.) has facilitated the use of past
reconstructions to address pressing questions about the impacts of References
regional, continental and global environmental change on
Agardy, T., Alder, J., Dayton, P., Curran, S., Kitchingman, A., Wilson, M., Catenazzi, A.,
ecosystem service provision (McMahon et al., 2011; Brewer et al., Restrepo, J., Birkeland, C., Blaber, S., Saifullah, S., Branch, G., Boersma, D.,
2012). However, the usefulness of these datasets for managing Nixon, S., Dugan, P., Davidson, N., Vo € ro
€ smarty, C., 2005. Coastal systems. In:
ecosystems could be constrained by their relative inaccessibility to Hassan, R., Scholes, R., Ash, N. (Eds.), Ecosystems and Human Well-being:
Current State and Trends. Island Press, Washington, DC, pp. 513e549.
non-specialists. What is required is an ability to produce maps of Aguilera, M., Araus, J.L., Voltas, J., Rodriguez-Ariza, M.O., Molina, F., Rovira, N.,
ecosystem characteristics (e.g. forest composition, soil type, N Buxo, R., Ferrio, J.P., 2008. Stable carbon and nitrogen isotopes and quality traits
30 E.S. Jeffers et al. / Quaternary Science Reviews 112 (2015) 17e32

of fossil cereal grains provide clues on sustainability at the beginnings of de Groot, R.S., Wilson, M.A., Boumans, R.M.J., 2002. A typology for the classification,
Mediterranean agriculture. Rapid Commun. Mass Spectrom. 22, 1653e1663. description and valuation of ecosystem functions, goods and services. Ecol.
Alba-Sanchez, F., Lopez-Saez, J.A., Benito-de Pando, B., Linares, J.C., Nieto-Lugilde, D., Econ. 41, 393e408.
Lopez-Merino, L., 2010. Past and present potential distribution of the Iberian Dearing, J.A., Håkansson, H., Liedberg-Jo € nsson, B., Persson, A., Skansjo €, S.,
Abies species: a phytogeographic approach using fossil pollen data and species Widholm, D., El-Daoushy, F., 1987. Lake sediments used to quantify the
distribution models. Divers. Distrib. 16, 214e228. erosional response to land use change in southern Sweden. Oikos 50, 60e78.
Anderson, N.J., Bugmann, H., Dearing, J.A., Gaillard, M.J., 2006. Linking palae- Dearing, J.A., Yang, X., Dong, X., Zhang, E., Chen, X., Langdon, P.G., Zhang, K.,
oenvironmental data and models to understand the past and to predict the Zhang, W., Dawson, T.P., 2012. Extending the timescale and range of ecosystem
future. Trends Ecol. Evol. 21, 696e704. services through paleoenvironmental analyses, exemplified in the lower
Araújo, M.B., Peterson, A.T., 2012. Uses and misuses of bioclimatic envelope Yangtze basin. Proc. Natl. Acad. Sci. 109, E1111eE1120.
modeling. Ecology 93, 1527e1539. Diaz, R.J., Rosenberg, R., 2008. Spreading dead zones and consequences for marine
Araus, J.L., Febrero, A., Buxo, R., Camalich, M.D., Martin, D., Molina, F., ecosystems. Science 321, 926e929.
RodriguezAriza, M.O., Romagosa, I., 1997. Changes in carbon isotope discrimi- Diaz, S., Lavorel, S., de Bello, F., Quetier, F., Grigulis, K., Robson, M., 2007. Incorpo-
nation in grain cereals from different regions of the western Mediterranean rating plant functional diversity effects in ecosystem service assessments. Proc.
Basin during the past seven millennia. Palaeoenvironmental evidence of a Natl. Acad. Sci. U. S. A. 104, 20684e20689.
differential change in aridity during the late Holocene. Glob. Change Biol. 3, Dominati, E., Patterson, M., Mackay, A., 2010. A framework for classifying and
107e118. quantifying the natural capital and ecosystem services of soils. Ecol. Econ. 69,
Asner, G.P., Powell, G.V.N., Mascaro, J., Knapp, D.E., Clark, J.K., Jacobson, J., Kennedy- 1858e1868.
Bowdoin, T., Balaji, A., Paez-Acosta, G., Victoria, E., Secada, L., Valqui, M., Eigenbrod, F., Armsworth, P.R., Anderson, B.J., Heinemeyer, A., Gillings, S., Roy, D.B.,
Hughes, R.F., 2010. High-resolution forest carbon stocks and emissions in the Thomas, C.D., Gaston, K.J., 2010. The impact of proxy-based methods on map-
Amazon. Proc. Natl. Acad. Sci. 107, 16738e16742. ping the distribution of ecosystem services. J. Appl. Ecol. 47, 377e385.
Balmford, A., Bond, W., 2005. Trends in the state of nature and their implications for Elliott, E.M., Brush, G.S., 2006. Sedimented organic nitrogen isotopes in freshwater
human well-being. Ecol. Lett. 8, 1218e1234. wetlands record long-term changes in watershed nitrogen source and land use.
Barlow, J., Parry, L., Gardner, T.A., Ferreira, J., Aragao, L.E.O.C., Carmenta, R., Environ. Sci. Technol. 40, 2910e2916.
Berenguer, E., Vieira, I.C.G., Souza, C., Cochrane, M.A., 2012. The critical impor- Extence, C.A., Balbi, D.M., Chadd, R.P., 1999. River flow indexing using British
tance of considering fire in REDDþ programs. Biol. Conserv. 154, 1e8. benthic macroinvertebrates: a framework for setting hydroecological objec-
Bennion, H., Fluin, J., Simpson, G.L., 2004. Assessing eutrophication and reference tives. Regul. Rivers Res. Manag. 15, 545e574.
conditions for Scottish freshwater lochs using subfossil diatoms. J. Appl. Ecol. 41, Farley, K.A., Jobbagy, E.G., Jackson, R.B., 2005. Effects of afforestation on water yield:
124e138. a global synthesis with implications for policy. Glob. Change Biol. 11,
Bonan, G.B., 2008. Forests and climate change: forcings, feedbacks, and the climate 1565e1576.
benefits of forests. Science 320, 1444e1449. Ferrio, J.P., Alonso, N., Voltas, J., Araus, J.L., 2004. Estimating grain weight in
Brewer, S., Jackson, S.T., Williams, J.W., 2012. Paleoecoinformatics: applying geo- archaeological cereal crops: a quantitative approach for comparison with cur-
historical data to ecological questions. Trends Ecol. Evol. 27, 104e112. rent conditions. J. Archaeol. Sci. 31, 1635e1642.
Brush, G.S., 2009. Historical land use, nitrogen, and coastal eutrophication: a Ferrio, J.P., Araus, J.L., Buxo, R., Voltas, J., Bort, J., 2005. Water management practices
paleoecological perspective. Estuaries Coasts 32, 18e28. and climate in ancient agriculture: inferences from the stable isotope compo-
Burns, D.A., Kendall, C., 2002. Analysis of d15N and d18O to differentiate NO3 sition of archaeobotanical remains. Veg. Hist. Archaeobot. 14, 510e517.
sources in runoff at two watersheds in the Catskill Mountains of New York. Flower, A.G., Gavin, D., Heyerdahl, E.K., Parsons, R.A., Cohn, G.M., 2014. Western
Water Resour. Res. 38, 9-1e9-11. spruce budworm outbreaks did not increase fire risk over the last three cen-
Cao, S.X., Chen, L., Shankman, D., Wang, C.M., Wang, X.B., Zhang, H., 2011. Excessive turies: a dendrochronological analysis of inter-disturbance synergism. Plos One
reliance on afforestation in China's arid and semi-arid regions: lessons in 9, e114282.
ecological restoration. Earth-Sci. Rev. 104, 240e245. Foresight, 2011. The Future of Food and Farming: Challenges and Choices for Global
Carcaillet, C., Almquist, H., Asnong, H., Bradshaw, R.H.W., Carrion, J.S., Gaillard, M.J., Sustainability. Final Project Report. The Government Office for Science, London.
Gajewski, K., Haas, J.N., Haberle, S.G., Hadorn, P., Muller, S.D., Richard, P.J.H., Froyd, C.A., Willis, K.J., 2008. Emerging issues in biodiversity & conservation
Richoz, I., Rosch, M., Goni, M.F.S., von Stedingk, H., Stevenson, A.C., Talon, B., management: the need for a palaeoecological perspective. Quat. Sci. Rev. 27,
Tardy, C., Tinner, W., Tryterud, E., Wick, L., Willis, K.J., 2002. Holocene biomass 1723e1732.
burning and global dynamics of the carbon cycle. Chemosphere 49, 845e863. Gaillard, M.-J., Sugita, S., Bunting, M.J., Middleton, R., Brostro €m, A., Caseldine, C.,
Carpenter, S.R., Mooney, H.A., Agard, J., Capistrano, D., DeFries, R.S., Diaz, S., Dietz, T., Giesecke, T., Hellman, S.V., Hicks, S., Hjelle, K., Langdon, C., Nielsen, A.-B.,
Duraiappah, A.K., Oteng-Yeboah, A., Pereira, H.M., Perrings, C., Reid, W.V., Poska, A., von Stedingk, H., Veski, S., 2008. The use of modelling and
Sarukhan, J., Scholes, R.J., Whyte, A., 2009. Science for managing ecosystem simulation approach in reconstructing past landscapes from fossil pollen
services: beyond the millennium ecosystem assessment. Proc. Natl. Acad. Sci. U. data: a review and results from the POLLANDCAL network. Veg. Hist.
S. A. 106, 1305e1312. Archaeobot. 17, 419e443.
Chan, K.M.A., Shaw, M.R., Cameron, D.R., Underwood, E.C., Daily, G.C., 2006. Con- Gangopadhyay, S., McCabe, G.J., 2010. Predicting regime shifts in flow of the Colo-
servation planning for ecosystem services. PLoS. Biol. 4, 2138e2152. rado River. Geophys. Res. Lett. 37, L20706.
Chapman, S.K., Langley, J.A., Hart, S.C., Koch, G.W., 2006. Plants actively control Gangopadhyay, S., Harding, B.L., Rajagopalan, B., Lukas, J.J., Fulp, T.J., 2009.
nitrogen cycling: uncorking the microbial bottleneck. New Phytol. 169, 27e34. A nonparametric approach for paleohydrologic reconstruction of annual
Chen, S.H., Pollino, C.A., 2012. Good practice in Bayesian network modelling. En- streamflow ensembles. Water Resour. Res. 45, W06417.
viron. Model. Softw. 37, 134e145. Giesecke, T., Fontana, S.L., 2008. Revisiting pollen accumulation rates from Swedish
Clark, J.S., Bell, D.M., Hersh, M.H., Nichols, L., 2011. Climate change vulnerability of lake sediments. Holocene 18, 293e305.
forest biodiversity: climate and competition tracking of demographic rates. Gill, J.L., McLauchlan, K.K., Skibbe, A.M., Goring, S., Zirbel, C.R., Williams, J.W., 2013.
Glob. Change Biol. 17, 1834e1849. Linking abundances of the dung fungus Sporormiella to the density of bison:
Cole, L.E.S., Bhagwat, S.A., Willis, K.J., 2014. Recovery and resilience of tropical for- implications for assessing grazing by megaherbivores in palaeorecords. J. Ecol.
ests after disturbance. Nat. Commun. 5. 101, 1125e1136.
Colombaroli, D., Tinner, W., 2013. Determining the long-term changes in biodiver- Godfray, H.C.J., Beddington, J.R., Crute, I.R., Haddad, L., Lawrence, D., Muir, J.F.,
sity and provisioning services along a transect from Central Europe to the Pretty, J., Robinson, S., Thomas, S.M., Toulmin, C., 2010. Food security: the
Mediterranean. Holocene 23, 1625e1634. challenge of feeding 9 billion people. Science 327, 812e818.
Compton, J.E., Harrison, J.A., Dennis, R.L., Greaver, T.L., Hill, B.H., Jordan, S.J., Goldstein, J.H., Caldarone, G., Duarte, T.K., Ennaanay, D., Hannahs, N., Mendoza, G.,
Walker, H., Campbell, H.V., 2011. Ecosystem services altered by human changes Polasky, S., Wolny, S., Daily, G.C., 2012. Integrating ecosystem-service tradeoffs
in the nitrogen cycle: a new perspective for US decision making. Ecol. Lett. 14, into land-use decisions. Proc. Natl. Acad. Sci. 109, 7565e7570.
804e815. Gonz alez, C., Urrego, L.E., Martinez, J.I., Polania, J., Yokoyama, Y., 2010. Mangrove
Coulthard, T.J., Macklin, M.G., Kirkby, M.J., 2002. A cellular model of Holocene up- dynamics in the southwestern Caribbean since the ‘Little Ice Age’: A history of
land river basin and alluvial fan evolution. Earth Surf. Process. Landf. 27, human and natural disturbances. Holocene 20, 849e861.
269e288. Gosling, W.D., Williams, J.J., 2013. Ecosystem service provision sets the pace for pre-
Craine, J.M., Elmore, A.J., Aidar, M.P.M., Bustamante, M., Dawson, T.E., Hobbie, E.A., Hispanic societal development in the central Andes. Holocene 23, 1619e1624.
Kahmen, A., Mack, M.C., McLauchlan, K.K., Michelsen, A., Nardoto, G.B., Greenwood, M.T., Wood, P.J., Monk, W.A., 2006. The use of fossil caddisfly assem-
Pardo, L.H., Penuelas, J., Reich, P.B., Schuur, E.A.G., Stock, W.D., Templer, P.H., blages in the reconstruction of flow environments from floodplain paleo-
Virginia, R.A., Welker, J.M., Wright, I.J., 2009. Global patterns of foliar nitrogen channels of the River Trent, England. J. Paleolimnol. 35, 747e761.
isotopes and their relationships with climate, mycorrhizal fungi, foliar nutrient Grunwald, S., Thompson, J.A., Boettinger, J.L., 2011. Digital soil mapping and
concentrations, and nitrogen availability. New Phytol. 183, 980e992. modeling at continental scales: finding solutions for global issues. Soil Sci. Soc.
Daily, G.C., 1997. Nature's Services: Societal Dependence on Natural Ecosystems. Am. J. 75, 1201e1213.
Island Press, Washington, DC. Harrington, R.R., Kennedy, B.P., Chamberlain, C.P., Blum, J.D., Folt, C.L., 1998. 15N
Dawson, T.P., Jackson, S.T., House, J.I., Prentice, I.C., Mace, G.M., 2011. Beyond pre- enrichment in agricultural catchments: field patterns and applications to
dictions: biodiversity conservation in a changing climate. Science 332, 53e58. tracking Atlantic salmon (Salmo salar). Chem. Geol. 147, 281e294.
de Groot, R.S., Alkemade, R., Braat, L., Hein, L., Willemen, L., 2010. Challenges in Heathwaite, A.L., 2010. Multiple stressors on water availability at global to catch-
integrating the concept of ecosystem services and values in landscape planning, ment scales: understanding human impact on nutrient cycles to protect water
management and decision making. Ecol. Complex 7, 260e272. quality and water availability in the long term. Freshw. Biol. 55, 241e257.
E.S. Jeffers et al. / Quaternary Science Reviews 112 (2015) 17e32 31

Henry, R.J., Kettlewell, P.S., 1996. Cereal Grain Quality. Chapman & Hall, London. McLauchlan, K.K., Craine, J.M., Oswald, W.W., Leavitt, P.R., Likens, G.E., 2007.
Hietz, P., Turner, B.L., Wanek, W., Richter, A., Nock, C.A., Wright, S.J., 2011. Long-term Changes in nitrogen cycling during the past century in a northern hardwood
change in the nitrogen cycle of tropical forests. Science 334, 664e666. forest. Proc. Natl. Acad. Sci. U. S. A. 104, 7466e7470.
Higuera, P.E., Sprugel, D.G., Brubaker, L.B., 2005. Reconstructing fire regimes with McLauchlan, K.K., Ferguson, C.J., Wilson, I.E., Ocheltree, T.W., Craine, J.M., 2010.
charcoal from small-hollow sediments: a calibration with tree-ring records of Thirteen decades of foliar isotopes indicate declining nitrogen availability in
fire. Holocene 15, 238e251. central North American grasslands. New Phytol. 187, 1135e1145.
Higuera, P.E., Whitlock, C., Gage, J.A., 2011. Linking tree-ring and sediment-charcoal McLauchlan, K.K., Williams, J.J., Engstrom, D.R., 2013a. Nutrient cycling in the
records to reconstruct fire occurrence and area burned in subalpine forests of palaeorecord: fluxes from terrestrial to aquatic ecosystems. Holocene 23,
Yellowstone National Park, USA. Holocene 21, 327e341. 1635e1643.
Hijmans, R.J., Graham, C.H., 2006. The ability of climate envelope models to predict McLauchlan, K.K., Williams, J.J., Craine, J.M., Jeffers, E.S., 2013b. Changes in global
the effect of climate change on species distributions. Glob. Change Biol. 12, nitrogen cycling during the Holocene epoch. Nature 495, 352e355.
2272e2281. McLauchlan, K.K., Higuera, P.E., Gavin, D.G., Perakis, S.S., Mack, M.C., Alexander, H.,
Hobbie, S.E., 1992. Effects of plant species on nutrient cycling. Trends Ecol. Evol. 7, Battles, J., Biondi, F., Buma, B., Colombaroli, D., Enders, S.K., Engstrom, D.R.,
336e339. Hu, F.S., Marlon, J.R., Marshall, J., McGlone, M., Morris, J.L., Nave, L.E., Shuman, B.,
Houlton, B.Z., Bai, E., 2009. Imprint of denitrifying bacteria on the global terrestrial Smithwick, E.A.H., Urrego, D.H., Wardle, D.A., Williams, C.J., Williams, J.J., 2014.
biosphere. Proc. Natl. Acad. Sci. 106, 21713e21716. Reconstructing disturbances and their biogeochemical consequences over
Howard, L., Wood, P., Greenwood, M., Rendell, H., 2009. Reconstructing riverine multiple timescales. Bioscience 64, 105e116.
paleo-flow regimes using subfossil insects (Coleoptera and Trichoptera): the McMahon, S.M., Harrison, S.P., Armbruster, W.S., Bartlein, P.J., Beale, C.M.,
application of the LIFE methodology to paleochannel sediments. J. Paleolimnol. Edwards, M.E., Kattge, J., Midgley, G., Morin, X., Prentice, I.C., 2011. Improving
42, 453e466. assessment and modelling of climate change impacts on global terrestrial
Howard, L.C., Wood, P.J., Greenwood, M.T., Rendell, H.M., Brooks, S.J., biodiversity. Trends Ecol. Evol. 26, 249e259.
Armitage, P.D., Extence, C.A., 2010. Sub-fossil Chironomidae as indicators of McNally, C.G., Uchida, E., Gold, A.J., 2011. The effect of a protected area on the
palaeoflow regimes: integration into the PalaeoLIFE flow index. J. Quat. Sci. tradeoffs between short-run and long-run benefits from mangrove ecosystems.
25, 1270e1283. Proc. Natl. Acad. Sci. U. S. A. 108, 13945e13950.
Hu, F.S., Finney, B.P., Brubaker, L.B., 2001. Effects of holocene Alnus expansion on Metsaranta, J.M., Lieffers, V.J., 2009. Using dendrochronology to obtain annual data
aquatic productivity, nitrogen cycling, and soil development in southwestern for modelling stand development: a supplement to permanent sample plots.
Alaska. Ecosystems 4, 358e368. Forestry 82, 163e173.
Huggett, R.J., 1998. Soil chronosequences, soil development, and soil evolution: a Millennium Ecosystem Assessment, 2005. Ecosystems and Human Well-being:
critical review. CATENA 32, 155e172. Current State and Trends. Island Press, Washington, DC, 917 pp.
IPCC (Intergovernmental Panel on Climate Change), 2001. Climate Change 2001: Monk, W.A., Wood, P.J., Hannah, D.M., Wilson, D.A., Extence, C.A., Chadd, R.P., 2006.
Impacts, Adaptation, and Vulnerability, Technical Summary. IPCC Secretariat, Flow variability and macroinvertebrate community response within riverine
World Meteorological Organization, Geneva, 56 pages. systems. River Res. Appl. 22, 595e615.
Jackson, S.T., Betancourt, J.L., Booth, R.K., Gray, S.T., 2009. Ecology and the ratchet of Mooney, H., Larigauderie, A., Cesario, M., Elmquist, T., Hoegh-Guldberg, O.,
events: climate variability, niche dimensions, and species distributions. Proc. Lavorel, S., Mace, G.M., Palmer, M., Scholes, R., Yahara, T., 2009. Biodiversity,
Natl. Acad. Sci. U. S. A. 106, 19685e19692. climate change, and ecosystem services. Curr. Opin. Environ. Sustain. 1,
Jain, S., Woodhouse, C.A., Hoerling, M.P., 2002. Multidecadal streamflow regimes in 46e54.
the interior western United States: implications for the vulnerability of water Morris, J.L., Brunelle, A.R., 2012. Pollen accumulation in lake sediments during
resources. Geophys. Res. Lett. 29, 2036. historic spruce beetle disturbances in subalpine forests of southern Utah, USA.
Jenny, H., 1941. Factors of Soil Formation: a System of Quantitative Pedology. Holocene 22, 961e974.
McGraw-Hill, New York; London. Morris, J.L., Brunelle, A.R., Munson, A.S., 2010. Pollen evidence of historical
Kindermann, G., Obersteiner, M., Sohngen, B., Sathaye, J., Andrasko, K., forest disturbance on the Wasatch Plateau, Utah. West. North Am. Nat. 70,
Rametsteiner, E., Schlamadinger, B., Wunder, S., Beach, R., 2008. Global cost 175e188.
estimates of reducing carbon emissions through avoided deforestation. Proc. Morris, J.L., le Roux, P.C., Macharia, A.N., Brunelle, A., Hebertson, E.G., Lundeen, Z.J.,
Natl. Acad. Sci. U. S. A. 105, 10302e10307. 2013. Organic, elemental, and geochemical contributions to lake sediment de-
Kowalski, T., 2006. Chalara fraxinea sp. associated with dieback of ash (Fraxinus posits during severe spruce beetle (Dendroctonus rufipennis) disturbances. For.
excelsior) in Poland. For. Pathol. 36, 264e270. Ecol. Manag. 289, 78e89.
Kundzewicz, Z.W., Mata, L.J., Arnell, N.W., Do € ll, P., Jimenez, B., Miller, K., Oki, T., Naidoo, R., Balmford, A., Costanza, R., Fisher, B., Green, R.E., Lehner, B., Malcolm, T.R.,
ŞEn, Z., Shiklomanov, I., 2008. The implications of projected climate change for Ricketts, T.H., 2008. Global mapping of ecosystem services and conservation
freshwater resources and their management. Hydrol. Sci. J. 53, 3e10. priorities. Proc. Natl. Acad. Sci. 105, 9495e9500.
Lavelle, P., Dugdale, R., Scholes, R., Berhe, A.A., Carpenter, E., Codispoti, L., Izac, A.-M., Nave, L.E., Vance, E.D., Swanston, C.W., Curtis, P.S., 2011. Fire effects on temperate
guer, P., Ward, B., 2005. Nutrient cycling.
Lemoalle, J., Luizao, F., Scholes, M.C., Tre forest soil C and N storage. Ecol. Appl. 21, 1189e1201.
In: Hassan, R., Scholes, R., Ash, N. (Eds.), Ecosystems and Human Well-being: Nelson, G.C., Bennett, E., Berhe, A.A., Cassman, K.G., DeFries, R., Dietz, T., Dobson, A.,
Current State and Trends. Island Press, Washington, pp. 331e353. Dobermann, A., Janetos, A., Levy, M., Marco, D., Nakicenovic, N., O'Neill, B.,
Lavorel, S., Grigulis, K., Lamarque, P., Colace, M.P., Garden, D., Girel, J., Pellet, G., Norgaard, R., Petschel-Held, G., Ojima, D., Pingali, P.L., Watson, R., Zurek, M.B.,
Douzet, R., 2011. Using plant functional traits to understand the landscape 2005. Drivers of change in ecosystem condition and services. In: Carpenter, S.R.,
distribution of multiple ecosystem services. J. Ecol. 99, 135e147. Pingali, P.L., Bennett, E., Zurek, M.B. (Eds.), Ecosystems and Human Well-being:
Lewis, D.B., Castellano, M.J., Kaye, J.P., 2014. Forest succession, soil carbon accu- Scenarios. Island Press, Washington, pp. 173e222.
mulation, and rapid nitrogen storage in poorly remineralized soil organic Overballe-Petersen, M.V., Nielsen, A.B., Bradshaw, R.H.W., 2013. Quantitative
matter. Ecology 95, 2687e2693. vegetation reconstruction from pollen analysis and historical inventory data
Lobell, D.B., Schlenker, W., Costa-Roberts, J., 2011. Climate trends and global crop around a Danish small forest hollow. J. Veg. Sci. 24, 755e771.
production since 1980. Science 333, 616e620. Paciorek, C.J., McLachlan, J.S., 2009. Mapping ancient forests: Bayesian inference for
Luo, Y., Su, B.O., Currie, W.S., Dukes, J.S., Finzi, A., Hartwig, U., Hungate, B., Mc spatio-temporal trends in forest composition using the fossil pollen proxy re-
Murtrie, R.E., Oren, R.A.M., Parton, W.J., Pataki, D.E., Shaw, M.R., Zak, D.R., cord. J. Am. Stat. Assoc. 104, 608e622.
Field, C.B., 2004. Progressive nitrogen limitation of ecosystem responses to Palmer, M.A., Filoso, S., 2009. Restoration of ecosystem services for environmental
rising atmospheric carbon dioxide. Bioscience 54, 731e739. markets. Science 325, 575e576.
Mace, G.M., Norris, K., Fitter, A.H., 2012. Biodiversity and ecosystem services: a Pan, Y., Birdsey, R.A., Fang, J., Houghton, R., Kauppi, P.E., Kurz, W.A., Phillips, O.L.,
multilayered relationship. Trends Ecol. Evol. 27, 19e26. Shvidenko, A., Lewis, S.L., Canadell, J.G., Ciais, P., Jackson, R.B., Pacala, S.W.,
Macias-Fauria, M., Willis, K.J., 2013. Landscape planning for the future: using fossil McGuire, A.D., Piao, S., Rautiainen, A., Sitch, S., Hayes, D., 2011. A Large and
records to independently validate bioclimatic envelope models for economi- Persistent Carbon Sink in the World's Forests. Science 333, 988e993.
cally valuable tree species in Europe. Glob. Ecol. Biogeogr. 22, 318e333. Peng, C.H., Guiot, J., van Campos, E., 1998. Estimating changes in terrestrial vege-
Macias Fauria, M., Johnson, E.A., 2008. Climate and wildfires in the North American tation and carbon storage: using palaeoecological data and models. Quat. Sci.
boreal forest. Philos. Trans. R. Soc. B Biol. Sci. 363, 2315e2327. Rev. 17, 719e735.
MacLean, D.A., 1990. Impact of forest pests and fire on stand growth and timber Pennington, W., Tutin, M.T.G., Lishman, J.P., 1971. Iodine in lake sediments in
yield: implications for forest management planning. Can. J. For. Res. 20, northern England and Scotland. Biol. Rev. 46, 279e313.
391e404. Pennington, W., 1986. Lags in adjustment of vegetation to climate caused by the
MacLean, D.A., Erdle, T.A., MacKinnon, W.E., Porter, K.B., Beaton, K.P., Cormier, G., pace of soil development - evidence from Britain. Vegetatio 67, 105e118.
Morehouse, S., Budd, M., 2001. The Spruce Budworm Decision Support System: Pereira, H.M., Leadley, P.W., Proenca, V., Alkemade, R., Scharlemann, J.P.W., Fer-
forest protection planning to sustain long-term wood supply. Can. J. For. Res. 31, nandez-Manjarres, J.F., Araujo, M.B., Balvanera, P., Biggs, R., Cheung, W.W.L.,
1742e1757. Chini, L., Cooper, H.D., Gilman, E.L., Guenette, S., Hurtt, G.C., Huntington, H.P.,
Magri, D., 1994. Late-quaternary changes of plant biomass as recorded by pollen- Mace, G.M., Oberdorff, T., Revenga, C., Rodrigues, P., Scholes, R.J., Sumaila, U.R.,
stratigraphical data: a discussion of the problem at Valle di Castiglione, Italy. Walpole, M., 2010. Scenarios for global biodiversity in the 21st Century. Science
Rev. Palaeobot. Palynol. 81, 313e325. 330, 1496e1501.
McCann, R.K., Marcot, B.G., Ellis, R., 2006. Bayesian belief networks: applications in Power, M.J., Marlon, J.R., Bartlein, P.J., Harrison, S.P., 2010. Fire history and the Global
ecology and natural resource management. Can. J. For. Res. Rev. Can. Rech. For. Charcoal Database: a new tool for hypothesis testing and data exploration.
36, 3053e3062. Palaeogeogr. Palaeoclimatol. Palaeoecol. 291, 52e59.
32 E.S. Jeffers et al. / Quaternary Science Reviews 112 (2015) 17e32

Prentice, I.C., Harrison, S.P., 2009. Ecosystem effects of CO2 concentration: evidence Swetnam, T.W., Lynch, A.M., 1993. Multicentury, regional-scale patterns of western
from past climates. Clim. Past 5, 297e307. spruce budworm outbreaks. Ecol. Monogr. 63, 399e424.
Prentice, I.C., Harrison, S.P., Bartlein, P.J., 2011. Global vegetation and terrestrial Therrell, M.D., Stahle, D.W., Mukelabai, M.M., Shugart, H.H., 2007. Age, and radial
carbon cycle changes after the last ice age. New Phytol. 189, 988e998. growth dynamics of Pterocarpus angolensis in southern Africa. For. Ecol. Manag.
Prentice, I.C., Jolly, D., participants, B., 2000. Mid-Holocene and glacial-maximum 244, 24e31.
vegetation geography of the northern continents and Africa. J. Biogeogr. 27, Turner, M.G., Hargrove, W.W., Gardner, R.H., Roe, W.H., 1994. Effects of fire on
507e519. landscape heterogeneity in Yellowstone National Park, Wyoming. J. Veg. Sci. 5,
Raffa, K.F., Aukema, B.H., Bentz, B.J., Carroll, A.L., Hicke, J.A., Turner, M.G., 731e742.
Romme, W.H., 2008. Cross-scale drivers of natural disturbances prone to Tylianakis, J.M., Didham, R.K., Bascompte, J., Wardle, D.A., 2008. Global change and
anthropogenic amplification: the dynamics of bark beetle eruptions. Bioscience species interactions in terrestrial ecosystems. Ecol. Lett. 11, 1351e1363.
58, 501e517. UK National Ecosystem Assessment, World Conservation Monitoring Centre, 2011.
Reich, P.B., Hobbie, S.E., Lee, T., Ellsworth, D.S., West, J.B., Tilman, D., Knops, J.M.H., UK National Ecosystem Assessment: Technical Report. United Nations Envi-
Naeem, S., Trost, J., 2006. Nitrogen limitation constrains sustainability of ronment Programme World Conservation Monitoring Centre, Cambridge.
ecosystem response to CO2. Nature 440, 922e925. Vorosmarty, C.J., Green, P., Salisbury, J., Lammers, R.B., 2000. Global water resources:
Rozendaal, D.M.A., Zuidema, P.A., 2011. Dendroecology in the tropics: a review. vulnerability from climate change and population growth. Science 289,
Trees-Struct. Funct. 25, 3e16. 284e288.
Sampson, R.N., Bystriakova, N., Brown, S., Gonzalez, P., Irland, L.C., Kauppi, P., Walker, L.R., Wardle, D.A., 2014. Plant succession as an integrator of contrasting
Sedjo, R., Thompson, I.D., 2005. Timber, fuel and fiber. In: Hassan, R., Scholes, R., ecological time scales. Trends Ecol. Evol. 29, 504e510.
Ash, N. (Eds.), Ecosystems and Human Well-being: Current State and Trends. Walters, B.B., Ronnback, P., Kovacs, J.M., Crona, B., Hussain, S.A., Badola, R., Primavera, J.H.,
Island Press, Washington, D.C, pp. 243e269. Barbier, E., Dahdouh-Guebas, F., 2008. Ethnobiology, socio-economics and man-
Sedjo, R., Sohngen, B., 2012. Carbon sequestration in forests and soils. Annu. Rev. agement of mangrove forests: a review. Aquat. Bot. 89, 220e236.
Resour. Econ. 4, 127e144. Wardle, D.A., Bardgett, R.D., Klironomos, J.N., Setala, H., van der Putten, W.H.,
Seppa€, H., Alenius, T., Muukkonen, P., Giesecke, T., Miller, P.A., Ojala, A.E.K., 2009. Wall, D.H., 2004. Ecological linkages between aboveground and belowground
Calibrated pollen accumulation rates as a basis for quantitative tree biomass biota. Science 304, 1629e1633.
reconstructions. Holocene 19, 209e220. Welsh, K.E., Dearing, J.A., Chiverrell, R.C., Coulthard, T.J., 2009. Testing a cellular
Simard, I., Morin, H., Lavoie, C., 2006. A millennial-scale reconstruction of spruce modelling approach to simulating late-Holocene sediment and water transfer
budworm abundance in Saguenay, Quebec, Canada. Holocene 16, 31e37. from catchment to lake in the French Alps since 1826. Holocene 19, 785e798.
Simard, I., Morin, H., Potelle, B., 2002. A new paleoecological approach to recon- Westerling, A.L., Hidalgo, H.G., Cayan, D.R., Swetnam, T.W., 2006. Warming and
struct long-term history of spruce budworm outbreaks. Can. J. For. Res. Rev. Can. earlier spring increase western U.S. forest wildfire activity. Science 313,
Rech. For. 32, 428e438. 940e943.
Smith, D., Nayyar, K., Schreve, D., Thomas, R., Whitehouse, N., 2014. Can dung Whitehouse, N.J., Kirleis, W., 2014. The world reshaped: practices and impacts of
beetles from the palaeoecological and archaeological record indicate herd early agrarian societies. J. Archaeol. Sci. 51, 1e11.
concentration and the identity of herbivores? Quat. Int. 341, 119e130. Willis, K.J., Birks, H.J.B., 2006. What is natural? The need for a long-term perspective
Sugita, S., 2007a. Theory of quantitative reconstruction of vegetation I: pollen from in biodiversity conservation. Science 314, 1261e1265.
large sites REVEALS regional vegetation composition. Holocene 17, 229e241. Willis, K.J., Bailey, R.M., Bhagwat, S.A., Birks, H.J.B., 2010. Biodiversity baselines,
Sugita, S., 2007b. Theory of quantitative reconstruction of vegetation II: all you need thresholds and resilience: testing predictions and assumptions using palae-
is LOVE. Holocene 17, 243e257. oecological data. Trends Ecol. Evol. 25, 583e591.
Sutton, M.A., 2011. The European Nitrogen Assessment: Sources, Effects and Policy Willis, K.J., Braun, M., Sumegi, P., Toth, A., 1997. Does soil change cause vegetation
Perspectives. Cambridge Unversity Press, Cambridge. change or vice versa? A temporal perspective from Hungary. Ecology 78, 740e750.
Swetnam, T.W., Thompson, M.A., Sutherland, E.K., 1985. Spruce Budworms Hand- Wolfe, A.P., Hobbs, W.O., Birks, H.H., Briner, J.P., Holmgren, S.U., Ingo 
 lfsson, O.,
book: Using Dendrochronology to Measure Radial Growth of Defoliated Trees. Kaushal, S.S., Miller, G.H., Pagani, M., Saros, J.E., Vinebrooke, R.D., 2013. Strati-
US Department of Agriculture. Agricultural Handbook No. 639, 39 pages. graphic expressions of the HoloceneeAnthropocene transition revealed in
Swetnam, T.W., Betancourt, J.L., 1998. Mesoscale disturbance and ecological sediments from remote lakes. Earth-Sci. Rev. 116, 17e34.
response to decadal climatic variability in the American Southwest. J. Clim. 11, Zaehle, S., 2013. Terrestrial nitrogenecarbon cycle interactions at the global scale.
3128e3147. Philos. Trans. R. Soc. B Biol. Sci. 368, 1621.

Das könnte Ihnen auch gefallen