ArticlePDF Available

Data-driven regionalization of forested and non-forested ecosystems in coastal British Columbia with LiDAR and RapidEye imagery

Authors:
  • Hakai Institute and Simon Fraser University

Abstract and Figures

Traditionally, forest inventory and ecosystem mapping at local to regional scales rely on manual interpretation of aerial photographs, based on standardized, expert-driven classification schemes. These current approaches provide the information needed for forest ecosystem management but constrain the thematic and spatial resolution of mapping and are infrequently repeated. The goal of this research was to demonstrate the utility of an unsupervised, quantitative technique based on Light Detection And Ranging (LiDAR) data and multi-spectral satellite imagery for mapping local-scale ecosystems over a heterogeneous landscape of forested and non-forested ecosystems. We derived a range of metrics characterizing local terrain and vegetation from LiDAR and RapidEye imagery for Calvert and Hecate Islands, British Columbia. These metrics were used in a cluster analysis to classify and quantitatively characterize ecological units across the island. A total of 18 clusters were derived. The clusters were attributed with quantitative summary statistics from the remotely sensed data inputs and contextualized through comparison to ecological units delineated in a traditional expert-driven mapping method using aerial photographs. The 18 clusters describe ecosystems ranging from open shrublands to dense, productive forest and include a riparian zone and many wetter and wetland ecosystems. The clusters provide detailed, spatially-explicit information for characterizing the landscape as a mosaic of units defined by topography and vegetation structure. This study demonstrates that using various types of remotely sensed data in a quantitative classification can provide scientists and managers with multivariate information unique from that which results from traditional, expert-based ecosystem mapping methods.
Content may be subject to copyright.
1
Data-driven regionalization of forested and non-forested ecosystems in coastal British
Columbia with LiDAR and RapidEye imagery
Shanley D. Thompson a,1,*, Trisalyn A. Nelson a,2, Ian Giesbrecht b,3, Gordon Frazer b,4, Sari C.
Saundersc,5
a Spatial Pattern Analysis and Research Lab, Department of Geography, University of
Victoria, PO BOX 3060, Victoria, British Columbia, Canada, V8W 3R4
b Hakai Institute, Box 309, Heriot Bay, British Columbia, Canada, V0P 1H0
c BC Ministry of Forests, Lands and Natural Resources, 2100 Labieux Rd., Nanaimo
British Columbia, Canada, V9T 6E9
Email: sdthomps@uvic.ca (* Corresponding author)
Pre-print of published version.
Reference: Thompson, S.D., Nelson, T.A., Giesbrecht, I., Frazer, G., & Saunders, S.C.
(2016). Data-driven regionalization of forested and non-forested ecosystems in coastal
British Columbia with LiDAR and RapidEye imagery. Applied Geography 69: 35-50.
DOI: http://dx.doi.org/10.1016/j.apgeog.2016.02.002
Journal link:
http://www.sciencedirect.com/science/article/pii/S0143622816300108
Disclaimer:
The pdf document is a copy of the final version of this manuscript that was subsequently
accepted by the journal for publication. The paper has been through peer review, but it has
not been subject to any additional copy-editing or journal-specific formatting. As such, it will
look different from the final version of record, which may be accessed by following the DOI
or journal link above.
2
Data-driven regionalization of forested and non-forested ecosystems in coastal British
Columbia with LiDAR and RapidEye imagery
Abstract
Traditionally, forest inventory and ecosystem mapping at local to regional scales rely on manual
interpretation of aerial photographs, based on standardized, expert-driven classification schemes.
These current approaches provide the information needed for forest ecosystem management but
constrain the thematic and spatial resolution of mapping and are infrequently repeated. The goal
of this research was to demonstrate the utility of an unsupervised, quantitative technique based
on Light Detection And Ranging (LiDAR) data and multi-spectral satellite imagery for mapping
local-scale ecosystems over a heterogeneous landscape of forested and non-forested ecosystems.
We derived a range of metrics characterizing local terrain and vegetation from LiDAR and
RapidEye imagery for Calvert and Hecate Islands, British Columbia. These metrics were used in
a cluster analysis to classify and quantitatively characterize ecological units across the island. A
total of 18 clusters were derived. The clusters were attributed with quantitative summary
statistics from the remotely sensed data inputs and contextualized through comparison to
ecological units delineated in a traditional expert-driven mapping method using aerial
photographs. The 18 clusters describe ecosystems ranging from open shrublands to dense,
productive forest and include a riparian zone and many wetter and wetland ecosystems. The
clusters provide detailed, spatially-explicit information for characterizing the landscape as a
mosaic of units defined by topography and vegetation structure. This study demonstrates that
using various types of remotely sensed data in a quantitative classification can provide scientists
and managers with multivariate information unique from that which results from traditional,
expert-based ecosystem mapping methods.
Highlights
Statistical clustering of remotely sensed data produced a map of 18 ecosystems
Spatially-explicit topographic and vegetation structural data quantify each cluster
The approach complements traditional methods based on air photo interpretation
Keywords
ecosystem, ecoregion, classification, cluster analysis, ecosystem structure, terrain
1 Introduction
An ecosystem, as defined by the Convention on Biological Diversity (CBD) and the
Millennium Ecosystem Assessment (MA), is a dynamic complex of biotic components, and the
interaction between these components and their physical environment. Ecosystems can be
conceptualized at a variety of spatial scales, in a hierarchical manner (Bailey, 1987; Franklin,
2013). As reviewed by Whittaker (1967) and Kent et al. (1997), boundaries between ecosystems
are, in reality, most often gradual and fuzzy; attributes such as species composition overlap along
environmental gradients. The science of delineating regions (e.g., ecological or biogeographical)
in geographical space is referred to as regionalization (Loveland & Merchant, 2004; Olstad,
3
2012) and is important for the understanding and management of the natural world (Mackey,
Berry, & Brown, 2007; McMahon, Wiken, & Gauthier, 2004). Envisioned use of regions often
determines the appropriate scale for mapping and thus which and how many regions can or will
be delineated for a given project. National- or regional-scale conservation planning, resource
management, and ecosystem services assessments benefit from maps that delineate spatial units
on the basis of regional climate, large landforms, land cover class, and/or patterns of primary
productivity (e.g., Handcock and Csillag 2002; Leathwick et al. 2003; Sayre et al. 2009).
Differences in vegetation composition and structure, as influenced by micro-scale soil moisture
and nutrient conditions, define ecological communities at plot to landscape scales, supporting
more specific local and regional science and management activities (e.g., Banner et al. 1996).
Many jurisdictions rely on the manual interpretation of aerial photographs and field
observations to delineate ecological regions and forest attributes at local to national scales.
Interpreters use a set of methodological standards (e.g., Resources Inventory Committee 1998;
Canadian Forest Service 2001; Resource Information Management Branch 2005) that provide
the information desired for forest or ecosystem management, but inherently limit the spatial and
thematic resolutions of the resulting map. For instance, forest stand or ecosystem polygons are
delineated and attributed with information regarding composition and structure that is considered
representative of the entire polygon (Wulder et al., 2006). Alternative methods for mapping that
incorporate other types of remotely sensed imagery in a quantitative classification may be used
to delineate local regions with measurable attributes at improved spatial precisions. Further,
relative to the manual delineation of ecological regions, a quantitative approach may be
automated or semi-automated, offering increased consistency, repeatability, and cost-efficiency
for monitoring over time and across large areas (MacMillan, Moon, & Coupé, 2007; Morgan,
Gergel, & Coops, 2010).
Remote sensing can capture various structural, compositional, and functional aspects of
ecosystems. Multispectral imagery at high to moderate spatial resolutions is frequently used to
distinguish among vegetation life form and age class (e.g., Wulder et al. 2004; Johansen et al.
2007; Valeria et al. 2014). Multispectral data are also used generate indices such as the
Normalized Difference Vegetation Index (NDVI) (Rouse, Haas, Schell, & Deering, 1974), which
correlates well with patterns of net primary productivity (Goward, Tucker, & Dye, 1985) and has
been useful for wetland delineation (e.g., Dechka et al. 2002; Barron et al. 2014; White et al.
2015). Hyperspectral imagery is well suited to species composition mapping (e.g., Dalponte et al.
2013, Feret & Asner 2013). Light Detection And Ranging (LiDAR) technology captures highly
accurate, direct measurements of three-dimensional vegetation structure (Lim, Treitz, Wulder,
St-Onge, & Flood, 2003; van Leeuwen & Nieuwenhuis, 2010). In particular, LiDAR provides
excellent measurements of forest structural properties such as tree height and canopy closure
(Coops et al., 2007; Holmgren, 2004; Lefsky et al., 1999). The high spatial resolution of LiDAR-
derived elevation data may (Hogg & Holland, 2008; Maxa & Bolstad, 2009) or may not (Knight
et al. 2013) lead to improved wetland delineation compared to use of aerial photography,
especially in areas of low topographic variation.
As no single sensor can capture all of the structural, functional and compositional
characteristics of terrestrial ecosystems at the spatial resolution and extent typically desired by
environmental managers, researchers often combine multiple types of remotely sensed data
together to generate improved and more complex maps (e.g., Ke et al. 2010, Wulder et al. 2007,
4
Jones et al. 2010). Combining various types of remotely sensed data in a quantitative (statistical)
regionalization, researchers have effectively captured environmental domains over large regions
(e.g., Handcock and Csillag 2002; Coops et al. 2009; Fitterer et al. 2012; Powers et al. 2012).
Drawing on the high spatial resolution and information content of remotely sensed data,
regionalizations are also possible that highlight ecosystem patterns, gradients, and ecotones at
local scales (Hargrove & Hoffman, 2004; Kupfer, Gao, & Guo, 2012; Long, Nelson, & Wulder,
2010; Olstad, 2012). These smaller regions have higher internal homogeneity, a desirable
property for local-scale management (Bryan, 2006) and for serving as strata for field-based
sampling. As computing power increases and is more able to accommodate very large datasets,
quantitative methods that combine numerous remotely sourced datasets at high to moderate
spatial resolutions, including LiDAR, can produce information-rich (i.e., multivariate)
categorical maps for environmental research and management. investigated., Researchers
continue to evaluate new techniques and data for regionalization in ecological contexts (e.g., Xu
et al. 2013, Niesterwiscz et al. 2013), however, research incorporating vegetation structural data
from LiDAR with other remotely sensed data for local-scale regionalization has been limited.
The goal of this research was to demonstrate the utility of LiDAR and high spatial
resolution multispectral imagery for mapping and characterizing the variety of local-scale
ecosystems over a complex landscape of forested and non-forested ecosystems on the outer coast
of British Columbia. Given that different ecosystem maps of the same area are possible
depending on data and methodology, the objective was to use an unsupervised, quantitative
regionalization method to determine what types of ecosystem units would be captured when
using high resolution remotely sensed measures representing vegetation structure and
topography. Following attribution of the resulting regions with the remotely sensed data, we
contextualize our findings by comparing the results of the unsupervised classification to those of
an expert-driven classification. We conclude with a discussion of implications for ecosystem
management and recommendations for local-scale ecosystem mapping that will be useful in a
variety of remote regions characterized by heterogeneous landscapes where field data collection
is logistically challenging.
2 Methods
2.1 Study Area
Calvert and Hecate Islands (total 37,433 ha) are remote islands on the central coast of
British Columbia, Canada (Figure 1). The islands are dominated by fairly subdued topography
from low to moderate elevations in a biogeoclimatic unit classified as the Coastal Western
Hemlock zone, Very Wet, Hypermaritime subzone, Central variant (CWHvh2). Elevations in the
study area reach ~1000 m; these areas classified as the Mountain Hemlock zone, Wet
Hypermaritime Subzone (MHwh1). Gridded climate data for the region (ClimateBC v5 - Wang
et al. 2012) indicate that average (19902012) annual temperature is 7.6C°C for the CWHvh2
and 4.8°C for the MHwh. Average annual precipitation in these biogeoclimatic units is 3512.2
mm and 5140.7 mm, respectively. High precipitation, abundant fog, and low evapotranspiration
result in an abundance of wet soils, wetland ecosystems, and relatively unproductive forests
(Banner, LePage, Moran, & de Groot, 2005).
5
Figure 1 Study area in coastal British Columbia, Canada
6
2.2 Remotely Sensed Data
The quantitative regionalization used in this study relied on a combination of LiDAR data
and multispectral satellite imagery with differing native spatial resolutions. All data were
aggregated to a common spatial resolution of 20 m, a scale widely used in LiDAR-based forest
inventories, soil mapping, and terrain analysis (Brosofske et al., 2014; Gillin et al., 2015).
Airborne LiDAR data were acquired across Calvert and Hecate Islands in August 2012. Mounted
on a fixed-wing aircraft flying at 1150 m above ground level with a maximum scan angle of
±26°, the discrete-return (4 returns/pulse), small-footprint (0.3 mrad) V-Gen LiDAR system
acquired data with an average point density of 2.32 pt/m2 and a standard deviation of 1.07 pt/m2.
Relative and absolute vertical accuracies at one standard deviation were estimated by the data
provider to be ±15 cm and ± 30 cm, respectively. A small number of data voids in the LiDAR
coverage were present and were excluded from the analysis.
Several data processing steps were required to extract candidate topographic indices and
vegetation metrics from the LiDAR dataset. First, LiDAR returns were separated into ground and
non-ground feature classes using TerraScan’s (Terrasolid Ltd.) automated ground-filtering
routine. Manual refinement of the point classification was undertaken in areas of steep, complex
topography to correct any obvious classification errors created by the automated ground filter
(Merrick Advanced Remote SensingMARSsoftware suite). Second, we used the BLAST
extension pack (blast2dem) of LAStools (rapidlasso GmbH) to construct a gridded one-meter,
‘bare-earth’ Digital Terrain Model (DTM) from a Triangulated Irregular Network (TIN) of
ground-classified points. The gridded one-meter DTM was subsequently resampled to a spatial
resolution of 20 meters using the cell-area weighted mean elevation prior to computing several
topographic indices (System for Automated Geoscientific AnalysesSAGA). Third, we
converted (i.e., height normalized) each elevation in the LiDAR point cloud to an estimated
height in meters above the ground surface using the difference in elevation between a point and
its corresponding elevation on the TIN of ground-classified points (LAStools). We extracted the
following terrain indices from the 20 m DTM: slope (%), the Topographic Radiation ASPect
(TRASP) (Roberts & Cooper, 1989), the Topographic Position Index (TPI) normalized to the
local standard deviation in elevation as per De Reu et al. (2013), and the Topographic Wetness
Index (TWI) calculated using the D-infinity algorithm (Tarboton, 1997) (Table 1). Elevation,
slope, TRASP, TPI, and TWI are considered useful inputs in automated ecosystem classification
and predictive vegetation mapping because these variables directly and indirectly influence plant
growth and community composition via effects on temperature, precipitation, soil moisture, soil
nutrients, and wind exposure (Franklin, 1995). These types of topographic derivatives are
commonly used in predictive ecosystem mapping (e.g., MacMillan et al. 2007; Chastain and
Struckhoff 2008; Dobrowski et al. 2008; Fraser et al. 2012), and standard ecosystem
classification in British Columbia at the local scale relies heavily on topographic concepts
(Banner et al., 1996).
We used the height normalized LiDAR point cloud to compute several area-based
(gridded) canopy height and density metrics on the same 20 m grid shared by the DTM and
derivative topographic indices. Area-based height metrics included estimates of the mean,
maximum, standard deviation, coefficient of variation (CV), and percentiles of canopy height
computed using all laser points with heights greater than or equal to 2 m within a single grid cell
7
(Magnussen and Boudewyn, 1998; Næsset, 2002). A measure of canopy density or gap fraction
was estimated for each grid cell as the ratio of the number laser points found below 2 m to the
total number of laser points located within the cell (Hopkinson and Chasmer, 2009). All grid
cells devoid of laser returns greater than 2 m in height were considered to be non-forested in this
study.
Multispectral reflectance data (5 m spatial resolution) were acquired from the RapidEye
satellite sensor (BlackBridge) in 2011. Using PCI Geomatica, the RapidEye imagery was first
converted to Top of Atmosphere units to adjust for sun angle and earth-sun distance and further
corrected for atmospheric noise using the Dark Object Subtraction method (Chavez, 1996). We
calculated the NDVI as an overall proxy for vegetation greenness or productivity. The NDVI
was resampled to 20 m spatial resolution to match the LiDAR data. Finally, as our focus was on
terrestrial systems, we used a provincial water body database
1
, as well as the NDVI layer (values
< 0.2) to remove water and un-vegetated pixels.
1
The Freshwater Atlas (FWA): http://geobc.gov.bc.ca/base-mapping/atlas/fwa/index.html
8
Table 1. Description of remotely sensed variables used in analyses
Variable
Derivation
Elevation
Elevation above sea level in metres, derived
from DEM.
Gap fraction
The ratio of total number of laser points < 2 m
in height to the total number of laser points
within a grid cell.
Height
coefficient of
variation
2nd central moment about the mean
Height
maximum
LiDAR Canopy Height Model = value of
maximum laser point return per grid cell.
Height mean
Arithmetic mean height (m) of laser points ≥
the minimum height threshold (2 m in this
study).
Normalized
Difference
Vegetation
Index (NDVI)
NDVI = (NIR reflectance red reflectance) /
(NIR reflectance +red reflectance) (Rouse et
al. 1974).
Slope
Slope in percent rise, derived from the DEM
Topographic
Position Index
(TPI)
TPI = zo - zavg where zo = elevation of focus
pixel and zavg = average elevation neighbouring
pixels (Weiss 2001). Calculated here for a 100
m neighbourhood and standardized to z-scores
as per De Reu et al. (2013).
Topographic
Radiation
ASPect
(TRASP)
TRASP = (1 - cos((3.1416/180)(aspect-30)))/2
where aspect is in degrees (Roberts and
Cooper 1989).
Topographic
Wetness Index
(TWI)
TWI = ln(specific catchment area / tan(slope)).
Calculated using the D-infinity algorithm
(Tarboton 1997).
9
2.3 Existing Ecosystem Data
The study area was recently mapped using British Columbia’s provincial standard of
Terrestrial Ecosystem Mapping (TEM) (Green, 2014; Resources Inventory Committee, 1998).
TEM uses a well-defined hierarchical classification system referred to as Biogeoclimatic
Ecosystem Classification (BEC) (Banner et al., 1996; Pojar, Klinka, & Meidinger, 1987). At the
regional level, vegetation, soils, and topography are used to infer climatic zones. Within these
regional climate zones, subzones are defined by the climax plant association typical of a zonal
site (zonal sites are those on which vegetation is primarily influenced by climate, not edaphic
features). Variants may be used to distinguish finer-scaled variation in climate within the
Subzone. Within each subzone or variant, a set of ecosystems is defined, which are referred to as
site series. The site series (each given a unique two letter and two- or three-digit code) are
classified with reference to established rules that associate each with a particular combination of
terrain, soil and vegetation characteristics.
TEM relies on analysts to interpret aerial photographs, topographic maps, and other
available geospatial information to identify site series. Interpreted TEM polygons may contain
up to three site series each, where individual patches of these site series are smaller than the
minimum mapping unit (typically 0.52.0 ha). In these complex polygons, the relative proportion
of each site series is noted, although the exact location is not. The TEM for the study area
identified 38 site series, including six non-vegetated site series, seven non-forested site series,
and 25 forested site series (Green, 2014). Ground sampling is used to calibrate and assess the
TEM. Field surveys were conducted in 2013 and 2014 by provincial ecologists and Hakai
Institute and affiliated researchers, primarily in the northern part of Calvert Island and the
southern part of Hecate Island following provincial BEC standards (BC Ministry of Forests and
Range and BC Ministry of Environment, 2010). Field plots identified 10 ecosystem types,
including several sub-types not explicitly captured in the TEM (e.g., shallow minerotrophic
blanket bogs, and deep ombrotrophic blanket bogs).
2.4 Unsupervised regionalization
Multivariate clustering was used to determine the types of local-scale ecosystems that are
distinguishable on Calvert and Hecate Islands with key metrics derived from LiDAR and
RapidEye data. Clustering is an unsupervised, multivariate technique, which can be used to
group observations or sample units that are similar with respect to the variables used to define
them. These techniques generally seek to minimize within-group variability and maximize
between-group variability.
A Spearman’s rank correlation analysis was conducted to assess collinearity among all
the aforementioned terrain (slope, TRASP, TWI, and TPI) and vegetation variables (NDVI,
mean height, maximum height, and gap fraction). Correlations were fairly low (< 0.5) between
the terrain variables and the vegetation variables, as well as among the terrain variables (Table
2). Correlations were moderate to high (> 0.7) among the vegetation variables. From these four
vegetation variables, we chose to retain one of the LiDAR vegetation variables (mean height)
and the multispectral vegetation variable (NDVI) as inputs for the clustering.
With each of the chosen variables having been clipped and resampled to a consistent
spatial extent and spatial resolution (20 m), and being projected to the same datum (NAD 83), a
“raster stack” was built (using the “raster” package in R) to combine the variables into a single
data frame for clustering. Clustering was conducted separately for the forested and non-forested
10
regions of the island, since the average height and coefficient of variation of height from LiDAR
were only available for the forested region (defined here as vegetation > 2 m in height). All data
were standardized to z-scores prior to clustering in order to ensure approximately equal weight
for all variables. A clustering methodology called TwoStep, employing a probability-based
distance measure, was implemented in SPSS (v22). The TwoStep approach was chosen because
it is well suited to very large datasets. It begins with an initial partition of the data, followed by a
hierarchical clustering of these partitions. The number of resulting clusters may be explicitly
specified by the analyst. To select an appropriate number, we first considered several statistical
criteria: the Bayesian Information Criterion (BIC), the CalinskiHarabasz (CH) criterion
(Calinski and Harabasz 1974) and the Average Silhouette Width (ASW) (Table 4.3). Whereas
the BIC is calculated internally during the TwoStep clustering procedure, we calculated the CH
and ASW statistics based on k-means partitioning using the pamk() function from the fpc library
in R (v3.1.2). In choosing the number of clusters, we also took into consideration the number of
site series identified in the existing ecosystem map and in field surveys conducted in 2013 and
2014.
Once the desired number of clusters was determined, the clusters were mapped in
geographic space and a majority filter of 4 x 4 pixels was applied to remove speckle in the final
product. Otherwise, no minimum mapping unit was imposed. Clusters were attributed with
summary statistics for each data layer input into the cluster analysis (e.g., NDVI, elevation, and
slope). Additionally, the LiDAR-derived variables of gap fraction and maximum height
(excluded from the cluster analysis because of high collinearity with each other as well as with
NDVI and average mean height) were also used to describe the clusters.
11
Table 2. Spearman’s rank correlation coefficients
Elevation
(m)
Gap
fraction
Height
coefficient of
variation
Height -
maximum
(m)
Height
mean
(m)
NDVI
Slope
(%)
TPI
TRASP
TWI
Elevation (m)
1.00
Gap fraction
-0.02
1.00
Height
coefficient of
variation
-0.06
-0.32
1.00
Height -
maximum (m)
-0.03
-0.90
0.38
1.00
Height mean
(m)
-0.06
-0.92
0.41
0.93
1.00
NDVI
0.09
-0.76
0.30
0.73
0.75
1.00
Slope (%)
0.48
-0.44
0.17
0.40
0.39
0.41
1.00
TPI
0.11
0.07
-0.18
-0.12
-0.15
-0.09
0.03
1.00
TRASP
0.04
-0.09
0.06
0.10
0.12
0.11
0.08
0.03
1.00
TWI
-0.14
0.20
-0.02
-0.15
-0.13
-0.15
-0.46
-0.59
-0.02
1.00
NDVI = Normalized Difference Vegetation Index (NDVI), TPI = Topographic Position Index, TRASP = Topographic Radiation
ASPect, TWI = Topographic Wetness Index
12
2.5 Map Comparisons
We used several approaches to contextualize our unsupervised ecosystem map relative to
the expert-driven TEM. First, each cluster was attributed with field survey information by
overlaying the field survey points on the clusters in ArcGIS, extracting the cluster value at each
point, and summarizing by cluster type. Second, TEM information was attributed to each cluster
as follows: the majority (modal) cluster value occurring within each TEM polygon was extracted
and joined to the TEM database. Next, the total area of each TEM site series within each cluster
type was derived using information on the proportional area of each site series within each TEM
polygon. We used pie charts to summarize the detailed composition of our clusters in terms of
TEM units and the composition of TEM units in terms of our clusters.
Third, we calculated the diversity of different TEM site series comprising each cluster, as
well as the inverse (diversity of clusters within each TEM site series) to determine
correspondence between the two mapping methodologies. More diverse clusters or TEM site
series represent poor agreement between the two maps. We assessed this relationship both
graphically (via pie-charts) and numerically, via Simpson’s Diversity Index, which is commonly
used in ecological studies to indicate species diversity in a way that captures both richness (a
count of species) and evenness (relative abundance of each species). The Simpson’s Diversity
Index is calculated using the formula
D = 1 [sum(n/N)2]
where n is the number of individuals of a particular species, and N is the total number of
all individuals for a given location. In our case, rather than species diversity, we calculated the
index to measure the diversity of site series in each cluster and again to measure the diversity of
clusters within each site series. The values ranged from 0 to 1, with higher values indicating a
poor association between the two mapping methodologies. We note, however, that the index of
diversity used to compare the maps at this level does not take into account ecological similarity
among site series and that a cluster containing many similar site series may not be as severely
mismatched to the TEM as a cluster containing the same number of ecologically dissimilar site
series. To avoid bias due to the variation in the size and number of each TEM polygon and each
TEM site series, TEM site series that are rare on the landscape were excluded (those below the
5th percentile in number of polygons as well as area). Individual TEM polygons below the 5th
percentile in area, regardless of site series, were also removed.
Finally, to allow direct comparison of the two maps with a single classification scheme,
we assigned each TEM site series and each cluster to one of six broad ecosystem classes. We
used broad ecosystem classes suggested by the local TEM authors as well as the regional
handbook for ecosystem classification (Green & Klinka, 1994). Clusters were assigned to broad
classes based on interpretation of the calculated terrain metrics, particularly topographic wetness,
slope and topographic position. The generalized classes maintain distinctions between forested
and non-forested ecosystems and emphasize differences along a moisture gradient. To compare
the unsupervised regionalization with the expert-driven TEM, we calculated the spatial extent of
each of these generalized ecosystem types for both systems.
13
3 Results
3.1 Unsupervised regionalization
We identified 12 forested clusters (Clusters 112) and six non-forested clusters (Clusters
1318) (Figure 2 and 3), drawing on several lines of information to judge an appropriate and
useful number of clusters in each group. The statistically optimal number of clusters varied from
two to three for non-forested regions and two to 25 for forested regions (Table 3), while the TEM
and field plot data suggested upwards of 32 vegetated site series may be present on the island.
Upon visual inspection of the resulting map, 18 clusters were chosen to provide the desired
spatially-detailed characterization of the landscape without sub-dividing beyond the limits of the
data or our ability to interpret, describe, andpotentiallysample the landscape mosaic in the
field.
Table 3. The statistical optimum number of clusters according to three different criteria
Criterion
Method Details
Optimum cluster
number
(Forested model)
Optimum cluster
number
(Non-Forested
model)
Average
Silhouette
Width (ASW)a
ASW = [ ∑Si/n]
The similarity (e.g., distance) of each data point i
to other points in the same cluster a, relative to
distance to points in other clusters, b averaged for
all points. In an optimal solution, values are close
to 1.0, meaning clusters are very homogenous and
well separated.
25
3
Calinski
Harabasz
(CH)a
BSS(K-1)/ WSS(n-K), where BSS is between
group sum of squares, WSS is within group sum
of squares, K is number of clusters and n is
number of data points. Optimal number of
clusters maximizes this index.
2
2
Bayesian
Information
Criterion (BIC)b
BIC identifies the model (cluster solution) that
would most likely produce the observed data,
penalizing for complexity (number of clusters *
number of observations). Optimal number of
clusters maximizes this index (i.e., BIC will be
higher for smaller numbers of clusters).
2
2
a ASW and CH were calculated in R based on k-means clustering.
b BIC was calculated in SPSS based on Two-Step clustering.
14
Figure 2. Eighteen clusters representing a range of forested and non-forested terrestrial
ecosystems on Calvert and Hecate Islands. Clusters 1 through 12 are forested and Clusters
13 through 18 are non-forested. White areas within the land mass are data voids.
15
Figure 3. Zoomed in view of 18 clusters on Calvert and Hecate Islands. Clusters 1 through
12 are forested and Clusters 13 through 18 are non-forested.
16
Cluster 1 described a forest of intermediate wetness (mean TWI = 6.07) and moderate
productivity (mean NDVI = 0.75, mean height = 7.7 m), relative to the study area (Table 4, Fig.
4 and 5). It occurred in the middle of the range of elevations found in the study area (average of
292 m) on steep slopes (average of 44%), and cool aspects (mean TRASP = 0.16). Cluster 2 was
more productive, and had relatively tall trees (mean NDVI = 0.78, mean maximum height =
20.3, mean height = 13 m). It was found at low elevations (average of 86 m). Clusters 3 and 4
were the most abundant classes across the study area (3903 ha and 3696 ha, or 11.2% and 10.6%
of the study area, respectively). They were characterized as forests of intermediate to high
wetness (mean TWI = 6.52 and 6.93, respectively) with moderate productivity (mean NDVI =
0.74). Both occurred at low elevations (average < 82 m). Cluster 3 occurred on warm aspects
(mean TRASP = 0.83), while Cluster 4 occurred on cool aspects (mean TRASP = 0.16).
Cluster 5 occurred primarily on ridges (mean TPI = 1.52) at low elevations (average 119
m) and was predicted to be the driest cluster (mean TWI = 4.4). Productivity was moderate
(mean NDVI = 0.75) and tree heights averaged 9 m. Clusters 6 and 8 had a similar topographic
wetness to Clusters 3 and 4, but were considerably less productive (mean NDVI values were
0.64 and 0.65 for Clusters 6 and 8, respectively). Both had open canopies (gap fractions of 0.69)
and short trees (average mean and max heights < 6 m). Cluster 6 represented cool aspects, while
Cluster 8 was associated with warm aspects. Cluster 7 was found at low elevations (mean =
85m), at fairly level sites (mean slope = 8%), and at lower slope positions (mean TPI = -0.88). It
was predicted to have very high wetness (mean TWI = 12.89) with moderate productivity (mean
NDVI = 0.74). Cluster 9 was predicted to be relatively dry, similar to Cluster 5 (mean TWI =
4.59), with moderate productivity (mean NDVI = 0.71) and short vegetation (mean height 5.73
m). Cluster 9 was found at the highest slope positions among our clusters (mean TPI = 1.55), at
high elevations (506 m on average) on very steep slopes (51% on average).
Cluster 10 was found in moderately productive forest (mean NDVI = 0.73) with average
tree heights of 6.5 m. It occurred at middle elevations (average 311 m), on moderate slopes
(average 33%) of distinctly warm aspect (mean TRASP = 0.86), and was predicted to be found
on sites with moderate wetness (mean TWI = 5.96). Clusters 11 and 12 were the tallest, densest,
and most productive forests on the island: maximum tree heights averaged > 21 m, mean gap
fractions were ≤ 0.16, and NDVI averaged 0.79 and 0.82, respectively. Cluster 11 occurred at
low elevations (average of 120 m) and moderate slopes (average 33%), while Cluster 12 was
found at higher elevations (average of 461 m) on very steep slopes (average 65%).
The remaining six clusters were characterized as non-forested (mean and maximum
vegetation height less than 2 m). All non-forested clusters had low mean NDVI values (≤ 0.6)
and very high mean gap fractions (≥ 0.93), relative to the forested clusters. Despite this low
NDVI and high gap fraction, NDVI boxplots suggest that some non-forested pixels reach forest-
like levels of productivity (Fig. 5). Cluster 13 was relatively scarce (849 ha of the study area) and
was found across a wide range of elevations (mean 361 m) and at relatively steep slopes (38%).
It was predicted to be relatively dry, similar to Clusters 5 and 9 (mean TWI = 4.86). Cluster 14
was also relatively scarce (696 ha in total) and generally restricted to low elevations (79 m
average). It had the lowest productivity and lowest canopy cover of any cluster (average NDVI
of 0.42 and average gap fraction of 0.98). Cluster 15 was a wetter cluster (mean TWI = 7.03) that
occurred at high elevations (average 413 m) across a wide range of aspects. Cluster 16 was
predicted to have very high wetness (mean TWI = 11.95), and it was uncommon across the study
17
site (720 ha). Generally it occurred at low elevations (99 m average) and on level sites (mean
slope 4%). Cluster 17 and 18 were predicted to have moderately high wetness (mean TWI ranged
from 6.82 to 7.20) and were found on very gentle slopes (8% average). Cluster 17 occurred on
cool aspects (mean TRASP = 0.16), and was extensive across the island (covering 2449 ha),
while Cluster 18 was less abundant (1674 ha) and occurred on warm aspects (mean TRASP =
0.80).
18
Figure 4. Distribution of clusters across each LiDAR-derived terrain index. Clusters 1
through 12 are forested, and Clusters 13 through 18 are non-forested.
19
Figure 5. Distribution of clusters across LiDAR and RapidEye vegetation data. Clusters 1
through 12 are forested, and Clusters 13 through 18 are non-forested.
20
Table 4. Mean values of remotely sensed inputs and descriptors of 18 clusters. Clusters 1 through 12 are forested and Clusters
13 through 18 are non-forested
Cluster
Extent
(ha)
Elevation
(m)
Gap
fraction
Height
coefficient of
variation
Height
maximum (m)
Height
mean (m)
NDVI
Slope (%)
TPI
TRASP
TWI
1
1,936
292
0.39
0.46
12.08
7.72
0.75
44
-0.41
0.16
6.07
2
1,160
86
0.24
0.52
20.30
13.00
0.78
43
-1.65
0.47
6.29
3
3,903
81
0.42
0.46
12.80
8.48
0.74
18
-0.20
0.83
6.52
4
3,696
73
0.41
0.44
12.65
8.41
0.74
15
-0.38
0.16
6.93
5
2,900
119
0.32
0.42
13.74
9.04
0.75
33
1.52
0.39
4.39
6
2,908
113
0.69
0.37
5.99
4.73
0.64
12
0.08
0.18
6.67
7
1,473
85
0.40
0.45
14.29
9.77
0.74
8
-0.88
0.49
12.89
8
1,921
100
0.69
0.38
5.88
4.76
0.65
11
0.20
0.80
6.32
9
1,020
506
0.48
0.38
7.97
5.73
0.71
51
1.55
0.41
4.59
10
2,039
311
0.50
0.45
9.67
6.51
0.73
34
0.07
0.86
5.96
11
2,206
120
0.16
0.36
23.38
16.61
0.79
33
-0.10
0.69
6.13
12
1,326
463
0.12
0.33
21.11
15.25
0.82
65
-0.13
0.63
5.54
13
849
361
0.93
n/a
1.89
n/a
0.59
38
1.10
0.38
4.86
14
696
79
0.98
n/a
0.83
n/a
0.42
13
0.45
0.32
5.73
15
1,902
413
0.95
n/a
1.60
n/a
0.57
18
-0.11
0.52
7.03
16
720
99
0.94
n/a
1.57
n/a
0.57
4
-0.44
0.45
11.95
17
2,449
105
0.93
n/a
1.93
n/a
0.60
8
-0.02
0.16
7.20
18
1,674
76
0.93
n/a
1.91
n/a
0.60
8
0.03
0.80
6.82
NDVI = Normalized Difference Vegetation Index (NDVI), TPI = Topographic Position Index, TRASP = Topographic Radiation
ASPect, TWI = Topographic Wetness Index
21
3.2 Map Comparisons
The clusters contained information that was unique from that contained in the TEM. Most
clusters in our established set of 18 were comprised of a variety of TEM site series (Table 5,
Supplementary Figure A.1) and each TEM site series contained multiple clusters (Supplementary
Figure A.2). Measured in terms of relative area, Cluster 14 was the most homogenous, with two
site series (bedrock and blanket bogs) comprising 84% its total area, with few site series
comprising the remaining area. When measured in terms of the Simpson’s diversity index, which
was based on counts (i.e., number of TEM site series within each cluster, and number of cluster
types within each TEM site series), Cluster 16 was the least diverse (most homogenous), with a
Diversity Index value of 0.53 (Figure 6). Less extensive classes (defined by either method) were
often less diverse, presumably in part because the chance of including additional classes is
smaller when area is very limited. Cluster 9 was the most diverse, followed by Cluster 7, with
values of 0.89 and 0.87, respectively (Figure 6).
Areas mapped as rivers (RI) in the TEM were entirely captured by Cluster 7 (Fig. A.2,
Fig. 7), which we have identified as a riparian ecosystem; Cluster 7 thus incorporated the actual
river as well as the riparian vegetation. Aside from rivers, the most homogenous TEM site series
were high elevation ecosystem types occurring in the Mountain Hemlock (MHwh1)
biogeoclimatic subzone (e.g., MHwh1/01(MB), MHwh1/06(MD), MHwh1/02(MM), MHwh1
03(MR), MHwh1/00(MS), and MHwh1/09(YC). Across their full spatial extent, these site series
consistently overlapped the same one to three cluster types (particularly Cluster 9). When
measured in terms of counts, Simpson’s diversity indices for these montane classes were also
low, ranging from 0 to 0.44 (Fig. 6). Conversely, the lower elevation forested site series such as
CWHvh2/00(TS), CWHvh2/12(LS), and CWHvh2/02(LR), overlapped with many different
cluster types, likely due in part to their larger extents. Despite the heterogeneous composition of
the 18 clusters in terms of TEM classes and vice versa, a generalized (six-class) comparison
showed greater similarity (Table 6, Fig. 7). Specifically, the islands were dominated by wet and
wetland forests with low productivity (~35% of the total area). Drier forests, and especially drier
non-forested ecosystems, comprised considerably less area (~15% of the total area combined).
Both mapping approaches indicated that both fresh to very moist forests and shrub/herb wetland
ecosystems occupy ~22 % to 27% of the islands.
Table 5. Comparison of 18 clusters to existing Terrestrial Ecosystem Mapping data
Cluster
Dominant TEM unita,b (% of cluster)
1
Zonal forest (36)
2
Zonal forest (49)
3
Bog forest (42)
4
Bog forest (41)
5
Cedar - salal forest (38)
6
Bog woodland (42)
7
Zonal forest (28)
8
Bog woodland (52)
22
9
Cedar - salal (31)
10
Bog forest (38)
11
Zonal forest (47)
12
Zonal forest (35)
13
Bedrock (52)
14
Bedrock (42), Blanket bog (42)
15
Blanket bog (65)
16
Fen (28)
17
Blanket bog (65)
18
Blanket bog (56)
a Zonal forest = CWHvh2/01(HS); Blanket bog = CWHvh2/00(TS); Bog forest = CWHvh2/11(YG); Cedar - salal
forest = CWHvh2/03(RS); Bog woodland = CWHvh2/12(LS), Fen = CWHvh2/00(FS), Bedrock = RO.
b See pie charts in Fig A.1 for further details.
Figure 6. Diversity of TEM site series (with respect to cluster composition) and diversity of
clusters (with respect to TEM site series). Lower values of the Simpson’s Diversity Index
represent greater similarity between TEM and cluster types
23
Table 6. Extent of generalized ecosystem classes on Calvert and Hecate Islands, British
Columbia, using an unsupervised classification of remotely sensed data (cluster analysis)
and an expert-driven classification (Terrestrial Ecosystem Mapping)
General
ecosystem class
Corresponding
clusters
Corresponding BEC units present
in study area TEMa,
Percentage of
study area
(clusters)
Percentage of
study area
(TEM)
Shrub/herb
upland
13
MHwh1/00(MS)
CWHvh2/00(SA)
1.9
0.4
Shrub/herb
wetland
14, 15, 16, 17, 18
CWHvh2/00(TS)
CWHvh2/00(BG)
CWHvh2/00(FS)
MHwh1/00(TS)
CWvh2/00(WI)
27.2
21.5
Wetter and
wetland forests
3, 4, 6, 7, 8
CWHvh2/00(RH)
CWHvh2/11(YG)
CWHvh2/12(LS)
CWHvh2/13(RC)
MHwh1/06(MD)
MHwh1/08(YS)
MHwh1/09(YC)
34.7
34.4
Drier forests
5, 9
CWHvh2/00(RM)
CWHvh2/02(LR)
CWHvh2/03(RS)
CWHvh2/14(SS)
CWHvh2/15(SK)
CWHvh2/16(SR)
MHwh1/00(YR)
MHwh1/02(MM)
12.3
14.9
Fresh to very
moist forests
1, 2, 10, 11, 12
CWHvh2/00(RD)
CWHvh2/01(HS)
CWHvh2/04(HM)
CWHvh2/06(SF)
CWHvh2/07(SD)
CWHvh2/17(SW)
CWHvh2/18(SE)
MHwh1/01(MB)
MHwh1/03(MR)
MHwh1/05(YT)
23.9
22.4
Non-vegetated
Excluded from
analysis to extent
possible
Bedrock (RO), rivers (RI), ponds
(PD), lakes (LA), shallow open
water (OW), exposed soil (ES)
n/a
6.2
a MHwh1 = Mountain Hemlock zone (Wet Hypermaritime Subzone); CWHvh2 = Coastal Western Hemlock zone
(Very Wet Hypermaritime Subzone, Central variant). The two-digit numbers (and associated two-letter acronyms)
following the forward slash refer to Site Series; those with numbers 00 are generally non-forested or not formalized
in the current version of the provincial classification for the area (Banner et al. 1993; Green and Klinka 1994;
MacKenzie and Moran 2004).
24
Figure 7. Generalized ecosystem classes as depicted by the expert-based classification
(Terrestrial Ecosystem Map) in the left panel, and the unsupervised regionalization on the
right.
25
4 Discussion
Our analysis generated a total of 18 clusters (ecological regions) including open, shrub-herb
dominated ecosystems (e.g., Clusters 13 to 18), a riparian zone (Cluster 7), wetter and wetland
ecosystems (e.g., Clusters 3, 4, 6 to 8, 14 through 18), and less wet, more productive forests (e.g.,
Clusters 2, 11, 12). We assessed what unique ecological information is contained in these
clusters derived from a variety of high spatial resolution remotely sensed data and what
information complements an expert-driven classification. For a generalized (six-class)
classification, both the quantitative classification and the expert-driven classification tell a
similar story. Namely, there is a high abundance of wetter and wetland ecosystems on the island,
and productive forests are limited in extent.
At a more detailed level, the 18 clusters complement or augment the expert-driven map.
First, whereas the TEM for our study area does not explicitly differentiate based on aspect,
several of the 18 clusters were differentiated primarily by aspect. For instance, Clusters 3 and 4
were found to have similar values of elevations, height, NDVI, slope, and TWI, but are different
in terms of TRASP (Table 4, Figure 4 and 5). Likewise, Cluster 6 and 8 had similar values of
elevation, slope, gap fractions, similar height, and NDVI profiles, and were predicted to be
similar in topographic wetness. However, Cluster 6 was associated with cooler aspects while
Cluster 8 was associated with warmer aspects. Another example is Clusters 17 and 18, which
were similar with respect to the majority of remotely sensed variables but were quite different
with respect to TRASP. We included TRASP as an input variable because it is known to
influence ecological patterns and processes. In particular, in temperate regions, localized aspect-
dependent differences in sun exposure, temperature, and moisture can lead to differences in plant
growth, plant composition and structure, growth and response to disturbance (Åström, Dynesius,
Hylander, & Nilsson, 2007; Diggins & Catterlin, 2014; Fekedulegn, Hicks, & Colbert, 2003;
Holland & Steyn, 1975), as well as to soil microbial communities (Carletti et al., 2008).
Another unique and promising aspect of the quantitative regionalization demonstrated in
this research relates to the delineation of wetland (defined here to include wet forest) ecosystems.
Wetlands provide a range of important ecological processes and ecosystem services, including
critical habitat, water quality and quantity regulation, and nutrient cycling and climate regulation,
which may vary according to wetland extent, type, and location within a watershed (Brinson,
1993; Emili, 2003; Fennessy, 2014; Zedler & Kercher, 2005). Consequently, a mapping
methodology that readily identifies wetland distributions at high spatial resolutions is desirable.
Yet these ecosystems are often poorly mapped remotely because of their small size relative to the
spatial resolution of commonly used satellite imagery such as Landsat (30 m) (Congalton, Birch,
Jones, & Schriever, 2002; Ozesmi & Bauer, 2002) or relative to a standard minimum mapping
unit, such as the 0.5 to 2 ha unit typical of TEM. Further, classification of wetlands with
multispectral imagery is driven by vegetation or land cover spectral response (Dechka et al.,
2002). Additional features used to classify wetlands to a particular class or site series in the field
such as acidity/alkalinity and magnitude of lateral and vertical water flow (MacKenzie & Moran,
2004) are not apparent in multispectral imagery, particularly under heavy vegetation cover
(Dechka et al., 2002; Rosenqvist, Finlayson, Lowry, & Taylor, 2007). Terrain indices such as the
Topographic Wetness Index may be useful for wetland delineation by estimating surface water
flow and accumulation but cannot capture all the hydrological properties typically used to
classify wetland type in the field. In this study, we predicted the location of wetlands using
terrain and vegetation structural indices generated from the LiDAR data, combined with multi-
26
spectral imagery from RapidEye. Our study demonstrates that in addition to multispectral
imagery and terrain indices provided by a 20 m LiDAR-derived DEM, 3D vegetation structural
information provided by the LiDAR data may also be of great value. For instance, in our study
the wetland vegetation was shown to be quite distinct from less wet vegetation with regards to
productivity and structure (e.g., canopy height and openness). The clusters we assigned to the
broad wetland classes (Clusters 3, 4, 6, 7, 8, and 15 through 18) overlapped considerably with
various types of BEC-TEM (expert-delineated) wetlands or water bodies (Figure 7, and A.1).
Thus, although a precise ecological interpretation of these clusters is challenging in the absence
on-the-ground verification, we can be reasonably confident that they do represent wetlands (e.g.,
Figures 2, 3, 7 and A.2). Future research could examine the potential of using even higher spatial
resolution (e.g., < 5m) LiDAR DEMs, the effect of neighbourhood sizes for metrics such as TPI,
and alternative terrain metrics for generating variables that are a stronger proxy for hydrological
processes, and thus for enabling more detailed wetland mapping and classification.
A final key and uniquely informative aspect of the quantitative regionalization approach
is the ability to quantify and map spatial variation in forest height and productivity at 20 m
spatial resolution. Forest inventories generally do measure stand height, basal area, canopy
cover, and so on, however these attributes are aggregated over larger polygons. A system in
British Columbia known as Site Index by BEC Site Series (SIBEC) estimates heights at age 50
for particular tree species within each site series. The SIBEC system adds to the utility of TEM
by allowing a comparison of productivity across sites, yet when mapped, is constrained by the
same spatial ambiguity as is TEM, and is based on a model rather than direct measurement. Our
study has shown that tree heights for the study area are quite low (most clusters have a mean
height <10 m). Low tree heights were also measured and predicted for the CWHvh2 in a nearby
SIBEC study within the CWHvh2, where mean tree heights in old-growth stands ranged from 5.2
m for Western redcedar (Thuja plicata) in the site series 04, to 2.0 m for Western hemlock
(Tsuga heterophylla) in site series 12 (Banner et al., 2005). Our research also indicated that
canopies are generally very open (forested clusters have, on average, a canopy closure of <60%),
and are variable in terms of vertical heterogeneity. Combined with NDVI, a measure of
vegetation greenness, our clusters contained rich and detailed structural and productivity
information.
Understanding the unique information contained within an expert-driven compared to a
quantitative ecological classification using newer geographic technologies has been the subject
of several contemporary studies (e.g., Thomas et al. 2002). There are several reasons for
differences between an expert-driven and a quantitative remotely sensed classification, including
the fact that expert systems may be better at separating ecologically meaningful information
from within large amounts of (non-ecologically meaningful) variation (Schmidtlein, Tichý,
Feilhauer, & Faude, 2010). Experts manually interpreting aerial photos for ecosystem mapping
inherently incorporate multiple characteristics such as tone or colour, shape, size, pattern,
texture, shadow, and landscape context (Morgan et al., 2010). Various types of information can
be automatically extracted from remotely sensed imagery to approximate but not replicate some
of these same visual cues for ecosystem classification, including spectral reflectance, image
texture, shape information derived from object-based classifiers, as well as landscape pattern
indices. With high spatial resolution LiDAR data, an additional dimension of data is available for
interpretation that is not present in standard multispectral satellite imagery. The three-
dimensional forest structure provided by LiDAR and the high spatial resolution terrain models
27
providing a range of ecological mapping applications (Lefsky, Cohen, Parker, & David, 2002;
Vierling, Vierling, Gould, Martinuzzi, & Clawges, 2008). Through this study we aim to
demonstrate the utility of a quantitative, unsupervised remote-sensing based regionalization, and
to assess the unique and complementary information content of mapping, relative to
conventional expert-driven ecosystem mapping. Future work should focus on the use of high-
spatial resolution remotely sensed data (especially LiDAR) in a supervised classification
approach (e.g., Predictive Ecosystem Mapping, (MacMillan et al. 2007)) to assess the ability to
model and map ecosystem distributions, including BEC site series. Such an approach should also
focus on incorporating data and metrics chosen to capture compositional attributes of
ecosystems. In particular, it may be useful to incorporate hyperspectral remotely sensed data,
which is well suited to mapping vegetation composition (e.g., Jones et al. 2010).
5 Conclusion
This study has demonstrated that LiDAR and high spatial resolution multi-spectral
imagery can be combined using a quantitative regionalization to map a variety of ecosystem
types across a heterogeneous landscape. Given the quantitative nature of our approach, it is
transparent and easily repeatable, and likely to succeed in other regions where topography is
subdued and forest structure complex. The ecosystem classes highlight vegetation structure and
productivity, refined by topography, and may have a variety of applications beyondyet
complementary tothose provided by expert-based ecosystem mapping from aerial
photography. For instance, we have identified clusters that are attributed with quantitative
vegetation height information (mean height and variation in mean height) and are spatially
explicit. Height information can be used for carbon accounting, because carbon storage in
vegetation can be estimated through allometric equations that use variables such as tree height to
estimate biomass. As well, estimates regarding relative primary productivity are readily available
in our dataset because of the NDVI incorporated into the clustering. Productivity is an essential
ecosystem function that supports a range of ecosystem services; while approximations of site
productivity (or potential) within traditional mapped polygons can be made based on species
composition, our methodology may facilitate monitoring because of the increased spatial
precision with which it is estimated. As LiDAR and high-spatial resolution multispectral imagery
become increasingly available in the future, they will be important complementary datasets in
local-scale ecosystem mapping.
Acknowledgments
We are grateful to the Tula Foundation for supporting this research. Additional funding was
provided by the Natural Sciences and Engineering Research Council of Canada (PGS-D and
Discovery grants). The LiDAR data were acquired by Terra Remote Sensing Inc. for the Hakai
Institute. Thank you to Keith Holmes and Luba Reshitnyk of the Hakai Institute, members of the
Coastal Ecosystem Dune Dynamics (CEDD) lab at the University of Victoria, and Richard
Fournier (Université de Sherbrooke) for additional data processing, preparation, and
management. Will MacKenzie (Provincial BEC Ecologist) summarized the ClimateBC data for
the BEC subzones. Finally, thank you to Ken Lertzman (Simon Fraser University), Trevor Lantz
(University of Victoria), Nicholas Coops (University of British Columbia), and Mike Wulder
(Natural Resources Canada) for valuable comments and discussions.
28
References
Åström, M., Dynesius, M., Hylander, K., & Nilsson, C. (2007). Slope aspect modifies community
responses to clear-cutting in boreal forests. Ecology, 88(3), 749758. http://doi.org/10.1890/06-0613
Bailey, R. G. (1987). Suggested hierarchy of criteria for multi-scale ecosystem mapping. Landscape and
Urban Planning, 14, 313319.
Banner, A., LePage, P., Moran, J., & de Groot, A. (2005). The HyP3 Project: Pattern, process and
productivity in hypermaritime Forests of coastal British Columbia.
Banner, A., Meidinger, D. V., Lea, E. C., Maxwell, R. E., & Sacken, B. C. (1996). Ecosystem mapping
methods for British Columbia. Environmental Monitoring and Assessment, 39(1-3), 97117.
http://doi.org/10.1007/BF00396139
Banner, A., W. MacKenzie, S. Haeussler, S. Thomson, J. Pojar and R. Trowbridge. 1993. A field guide to
site identification and interpretation for the Prince Rupert Forest Region. Land Manage. Handb. 26. BC
Min. For., Victoria, BC.
Barron, O. V., Emelyanova, I., Van Niel, T. G., Pollock, D., & Hodgson, G. (2014). Mapping
groundwater-dependent ecosystems using remote sensing measures of vegetation and moisture dynamics.
Hydrological Processes, 28(2), 372385. http://doi.org/10.1002/hyp.9609
BC Ministry of Forests and Range and BC Ministry of Environment. (2010). Field Manual for Describing
Terrestiral Ecosystems. 2nd edition. Land Management Handbook 25. Victoria, BC.
Brinson, M. (1993). Changes in the functioning of wetlands along environmental gradients. Wetlands,
13(2), 6574. http://doi.org/10.1007/BF03160866
Brosofske, K. D., Froese, R. E., Falkowski, M. J., & Banskota, A., (2014). A review of methods for
mapping and prediction of inventory attributes for operational forest management. Forest Science, 60(4),
733-756. http://dx.doi.org/10.5849/forsci.12-134
Bryan, B. A. (2006). Synergistic techniques for better understanding and classifying the environmental
structure of landscapes. Environmental Management, 37(1), 126140. http://doi.org/10.1007/s00267-004-
0058-1
Calinski, T., & Harabasz, J. (1974). A dendrite method for cluster analysis. Communications in Statistics,
3(1), 127.
Canadian Forest Service. (2001). Canada’s National Forest Inventory - National standards for photo
plots: Compilation procedures (Vol. Version 1.). Victoria, BC: Natural Resources Canada.
Carletti, P., Vendramin, E., Pizzeghello, D., Concheri, G., Zanella, A., Nardi, S., & Squartini, A. (2008).
Soil humic compounds and microbial communities in six spruce forests as function of parent material,
slope aspect and stand age. Plant and Soil, 315(1-2), 4765. http://doi.org/10.1007/s11104-008-9732-z
Chastain, R. J., & Struckhoff, M. (2008). Mapping vegetation communities using statistical data fusion in
the Ozark National Scenic Riverways, Missouri, USA. Photogrammetric Engineering and Remote
Sensing, 74(2), 247264. http://doi.org/http://dx.doi.org/10.14358/PERS.74.2.247
Chavez, P. (1996). Image-based atmospheric corrections-revisited and improved. Photogrammetric
Engineering and Remote Sensing, 62(9), 10251036. http://doi.org/10.1016/S0168-1699(02)00108-4
29
Congalton, R., Birch, K., Jones, R., & Schriever, J. (2002). Evaluating remotely sensed techniques for
mapping riparian vegetation. Computers and Electronics in Agriculture, 37, 113126.
http://doi.org/10.1016/S0168-1699(02)00108-4
Coops, N. C., Hilker, T., Wulder, M. A., St-Onge, B., Newnham, G., Siggins, A., & Trofymow, J. A.
(2007). Estimating canopy structure of Douglas-fir forest stands from discrete-return LiDAR. Trees,
21(3), 295310. http://doi.org/10.1007/s00468-006-0119-6
Coops, N. C., Wulder, M. A., & Iwanicka, D. (2009). An environmental domain classification of Canada
using earth observation data for biodiversity assessment. Ecological Informatics, 4(1), 822.
http://doi.org/10.1016/j.ecoinf.2008.09.005
Dalponte, M., Orka, H. O., Gobakken, T.; Gianelle, D., & Naesset, E. (2013). Tree Species Classification
in Boreal Forests With Hyperspectral Data. IEEE Transactions on Geoscience and Remote Sensing,
51(5), 2632-2645. doi: 10.1109/TGRS.2012.2216272
De Reu, J., Bourgeois, J., Bats, M., Zwertvaegher, A., Gelorini, V., De Smedt, P., … Crombé, P. (2013).
Application of the topographic position index to heterogeneous landscapes. Geomorphology, 186, 3949.
http://doi.org/10.1016/j.geomorph.2012.12.015
Dechka, J., Franklin, S., Watmough, M., Bennett, R., & Ingstrup, D. (2002). Classification of wetland
habitat and vegetation communities using multi-temporal Ikonos imagery in southern Saskatchewan.
Canadian Journal of Remote Sensing, 28(5), 679685. http://doi.org/10.5589/m02-064
Diggins, T. P., & Catterlin, R. G. (2014). Topographic patterns in forest composition and diversity on
slopes of Zoar Valley Canyon, western New York. Northeastern Naturalist, 21(3), 337350.
http://doi.org/http://dx.doi.org/10.1656/045.021.0301
Dobrowski, S. Z., Safford, H. D., Cheng, Y. Ben, & Ustin, S. L. (2008). Mapping mountain vegetation
using species distribution modeling, image-based texture analysis, and object-based classification.
Applied Vegetation Science, 11(4), 499508. http://doi.org/10.3170/2008-7-18560
Emili, L. (2003). Hydrochemical characteristics of hypermaritime dorest-peatland complexes, North
Coast British Columbia. University of Waterloo: PhD thesis.
Fekedulegn, D., Hicks, R. R., & Colbert, J. J. (2003). Influence of topographic aspect, precipitation and
drought on radial growth of four major tree species in an Appalachian watershed. Forest Ecology and
Management, 177(1-3), 409425. http://doi.org/10.1016/S0378-1127(02)00446-2
Fennessy, M. S. (2014). Wetland ecosystems and global change. In B. Freedman (Ed.), Global
Environmental Change (pp. 255261). Springer Netherlands. http://doi.org/10.1007/978-94-007-5784-4
Feret, J., & Asner, G. P. (2013). Tree Species Discrimination in Tropical Forests Using Airborne
Imaging Spectroscopy. IEEE Transactions on Geoscience and Remote Sensing, 51(1), 73-84. doi:
10.1109/TGRS.2012.2199323
Fitterer, J. L., Nelson, T. A., Coops, N. C., & Wulder, M. A. (2012). Modelling the ecosystem indicators
of British Columbia using Earth observation data and terrain indices. Ecological Indicators, 20, 151162.
http://doi.org/10.1016/j.ecolind.2012.02.024
Franklin, J. (1995). Predictive vegetation mapping: geographic modelling of biospatial patterns in relation
to environmental gradients. Progress in Physical Geography, 19(4), 474499.
http://doi.org/10.1177/030913339501900403
30
Franklin, J. (2013). Mapping vegetation from landscape to regional scales. In E. van der Maarel & J.
Franklin (Eds.), Vegetation Ecology (2nd editio). John Wiley & Sons, Ltd.
Fraser, R., McLennan, D., Ponomarenko, S., & Olthof, I. (2012). Image-based predictive ecosystem
mapping in Canadian arctic parks. International Journal of Applied Earth Observation and
Geoinformation, 14(1), 129138. http://doi.org/10.1016/j.jag.2011.08.013
Gillin, C. P., Bailey, S. W., McGuire, K. J., & Prisley, S. P. (2015). Evaluation of lidar-derived DEMs
through terrain analysis and field comparison. Photogrammetric Engineering & Remote Sensing, 81(5),
387-396. doi: 10.14358/PERS.81.5.387
Goward, S. N., Tucker, C. J., & Dye, D. G. (1985). North American vegetation patterns observed with the
NOAA-7 advanced very high resolution radiometer. Vegetatio, 64(1), 314.
http://doi.org/10.1007/BF00033449
Green, R. (2014). Reconnaissance level terrestrial ecosystem mapping of priority landscape units of the
coast EBM planning area: Phase 3. Prepared for BC Min. Forests, Lands & Natural Resource Ops., by
B.A. Blackwell & Associates, Vancouver, BC.
Green, R., & Klinka, K. (1994). A field guide to site identification and interpretation for the Vancouver
Forest Region. Victoria, BC: BC Ministry of Forests Research Program.
Handcock, R. N., & Csillag, F. (2002). Ecoregionalization assessment: Spatio-temporal analysis of net
primary production across Ontario. Ecoscience, 9(2), 219230.
Hargrove, W. W., & Hoffman, F. M. (2004). Potential of multivariate quantitative methods for
delineation and visualization of ecoregions. Environmental Management, 34 Suppl 1, S3960.
http://doi.org/10.1007/s00267-003-1084-0
Hogg, A. R., & Holland, J. (2008). An evaluation of DEMs derived from LiDAR and photogrammetry for
wetland mapping. Forestry Chronicle, 84(6), 840849. http://doi.org/10.5558/tfc84840-6
Holland, P. G., & Steyn, D. G. (1975). Vegetational responses to latitudinal variations in slope angle and
aspect. Journal of Biogeography, 2(3), 179183.
Holmgren, J. (2004). Prediction of tree height, basal area and stem volume in forest stands using airborne
laser scanning. Scandinavian Journal of Forest Research, 19(6), 543553.
http://doi.org/10.1080/02827580410019472
Hopkinson, C., & Chasmer, L. (2009). Testing LiDAR models of fractional cover across multiple forest
ecozones. Remote Sensing of Environment, 113(1), 275288. http://doi.org/10.1016/j.rse.2008.09.012
Johansen, K., Coops, N. C., Gergel, S. E., & Stange, Y. (2007). Application of high spatial resolution
satellite imagery for riparian and forest ecosystem classification. Remote Sensing of Environment, 110(1),
2944. http://doi.org/10.1016/j.rse.2007.02.014
Jones, T. G., Coops, N. C., & Sharma, T. (2010). Assessing the utility of airborne hyperspectral and
LiDAR data for species distribution mapping in the coastal Pacific Northwest, Canada. Remote Sensing of
Environment, 114(12), 28412852. http://doi.org/10.1016/j.rse.2010.07.002
Ke, Y. Quackenbush, L. J., & Im, J. (2010). Synergistic use of QuickBird multispectral imagery and
LIDAR data for object-based forest species classification. Remote Sensing of Environment, 114 (6), 1141-
1154. http://dx.doi.org/10.1016/j.rse.2010.01.002.
31
Kent, M., Gill, W., Weaver, R., & Armitage, R. (1997). Landscape and plant community boundaries in
biogeography. Progress in Physical Geography, 21(3), 315353.
http://doi.org/10.1177/030913339702100301
Knight, J. F., Tolcser, B. P., Corcoran, J. M., & Rampi, L. P. (2013). The effects of data selection and
thematic detail on the accuracy of high spatial resolution wetland classifications. Photogrammetric
Engineering & Remote Sensing, 79(7), 613623. http://doi.org/http://dx.doi.org/10.14358/PERS.79.7.613
Kupfer, J. A., Gao, P., & Guo, D. (2012). Regionalization of forest pattern metrics for the continental
United States using contiguity constrained clustering and partitioning. Ecological Informatics, 9, 1118.
http://doi.org/10.1016/j.ecoinf.2012.02.001
Leathwick, J. R., Overton, J. M., & McLeod, M. (2003). An environmental domain classification of New
Zealand and its use as a tool. Conservation Biology, 17(6), 16121623. http://doi.org/10.1111/j.1523-
1739.2003.00469.x
Lefsky, M. A., Cohen, W. B., Acker, S. A., Parker, G. G., Spies, T. ., & Harding, D. (1999). Lidar remote
sensing of the canopy structure and biophysical properties of Douglas-fir western hemlock forests.
Remote Sensing of Environment, 70(3), 339361. http://doi.org/10.1016/S0034-4257(99)00052-8
Lefsky, M. A., Cohen, W. B., Parker, G. G., & David, J. (2002). Lidar Remote Sensing for Ecosystem
Studies. BioScience, 52(1), 1930. http://doi.org/10.1641/0006-3568(2002)052[0019:LRSFES]2.0.CO;2
Lim, K., Treitz, P., Wulder, M. A., St-Onge, B., & Flood, M. (2003). LiDAR remote sensing of forest
structure. Progress in Physical Geography, 27(1), 88106. http://doi.org/10.1191/0309133303pp360ra
Long, J., Nelson, T. A., & Wulder, M. A. (2010). Regionalization of Landscape Pattern Indices Using
Multivariate Cluster Analysis. Environmental Management, 46(1), 134142.
http://doi.org/10.1007/s00267-010-9510-6
Loveland, T. R., & Merchant, J. M. (2004). Ecoregions and ecoregionalization: geographical and
ecological perspectives. Environmental Management, 34(S1), S1S13. http://doi.org/10.1007/s00267-
003-5181-x
MacKenzie, W., & Moran, J. (2004). Wetlands of British Columbia: A Guide to Identification. Victoria,
BC: BC Ministry of Forests Research Branch. Land Management Handbook 52.
Mackey, B. G., Berry, S. L., & Brown, T. (2007). Reconciling approaches to biogeographical
regionalization: a systematic and generic framework examined with a case study of the Australian
continent. Journal of Biogeography, 35(2), 213229. http://doi.org/10.1111/j.1365-2699.2007.01822.x
MacMillan, R. A., Moon, D. E., & Coupé, R. A. (2007). Automated predictive ecological mapping in a
forest region of B.C., Canada, 20012005. Geoderma, 140(4), 353373.
http://doi.org/10.1016/j.geoderma.2007.04.027
Magnussen, S., & Boudewyn, P. (1998). Derivations of stand heights from airborne laser scanner data
with canopy-based quantile estimators. Canadian Journal of Forest Research, 28(7), 10161031.
http://doi.org/10.1139/x98-078
Maxa, M., & Bolstad, P. (2009). Mapping northern wetlands with high resolution satellite images and
LiDAR. Wetlands, 29(1), 248260. http://doi.org/10.1672/08-91.1
32
McMahon, G., Wiken, E. B., & Gauthier, D. A. (2004). Toward a scientifically rigorous basis for
developing mapped ecological regions. Environmental Management, 34(S1), S111S124.
http://doi.org/10.1007/s00267-004-0170-2
Morgan, J. L., Gergel, S. E., & Coops, N. C. (2010). Aerial photography: a rapidly evolving tool for
ecological management. BioScience, 60(1), 4759. http://doi.org/10.1525/bio.2010.60.1.9
Næsset, E. (2002). Predicting forest stand characteristics with airborne scanning laser using a practical
two-stage procedure and field data. Remote Sensing of Environment, 80(1), 8899.
http://doi.org/10.1016/S0034-4257(01)00290-5
Niesterowicz, J., & Stepinski, T. F. (2013). Regionalization of multi-categorical landscapes using
machine vision methods. Applied Geography, 45, 250-258.
http://dx.doi.org/10.1016/j.apgeog.2013.09.023.
Olson, D. M., Dinerstein, E., Wikramanaya, E. D., Burgess, N. D., Powell, G. V. N., Underwood, E. C.,
… Kassem, K. R. (2001). Terrestrial ecoregions of the world: a new map of life on Earth. BioScience,
51(11), 933938. http://doi.org/http://dx.doi.org/10.1641/0006-3568(2001)051[0933:TEOTWA]2.0.CO;2
Olstad, T. A. (2012). Understanding the science and art of ecoregionalization. The Professional
Geographer, 64(2), 303308. http://doi.org/10.1080/00330124.2011.603656
Ozesmi, S. L., & Bauer, M. E. (2002). Satellite remote sensing of wetlands. Wetlands Ecology and
Management, 10, 381402. http://doi.org/10.1023/A:1020908432489
Pojar, J., Klinka, K., & Meidinger, D. (1987). Biogeoclimatic ecosystem classification in British
Columbia. Forest Ecology and Management, 22, 119154.
Powers, R. P., Coops, N. C., Morgan, J. L., Wulder, M. A., Nelson, T. A., Drever, C. R., & Cumming, S.
G. (2012). A remote sensing approach to biodiversity assessment and regionalization of the Canadian
boreal forest. Progress in Physical Geography, 37(1), 3662. http://doi.org/10.1177/0309133312457405
Resource Information Management Branch. (2005). Alberta Vegetation Inventory Interpretation
Standards. Version 2.1.1. In Chapter 3 Vegetation Inventory Standards and Data Model Documents.
Edmonton, AB: Alberta Sustainable Resource Development.
Resources Inventory Committee. (1998). Standard for Terrestrial Ecosystem Mapping in British
Columbia. Ecosystems Working Group, Terrestrial Ecosystems Task Force. Province of British
Columbia.
Roberts, D. W., & Cooper, S. V. (1989). Concepts and techniques of vegetation mapping. In Land
Classifications Based on Vegetation: Applications for Resource Management (pp. 9096). Ogden, UT:
USDA Forest Service GTR INT-257.
Rosenqvist, A., Finlayson, C., Lowry, J., & Taylor, D. (2007). The potential of long-wavelength satellite-
borne radar to support implementation of the Ramsar Wetlands Convention. Aquatic Conservation:
Marine and Freshwater Ecosystems, 17, 229244. http://doi.org/10.1002/aqc
Rouse, J. W. J., Haas, R. H., Schell, J. A., & Deering, D. W. (1974). Monitoring vegetation systems in the
great plains with ERTS. In Third Earth Resources Technology Satellite Symposium. NASA SP-351 (pp.
309317).
Sayre, R., Comer, P., Harumi, W., & Cress, J. (2009). A new map of standardized terrestrial ecosystems
of the conterminous United States. US Geological Survey Professional Paper 1768.
33
Schmidtlein, S., Tichý, L., Feilhauer, H., & Faude, U. (2010). A brute-force approach to vegetation
classification. Journal of Vegetation Science, 21(6), 11621171. http://doi.org/10.1111/j.1654-
1103.2010.01221.x
Tagil, S. and J.S. Jenness. 2008. GIS-based automated landform classification and topographic, landcover
and geologic attributes of landforms around the Yazoren Polje, Turkey. Journal of Applied Sciences.
8(6):910-921
Tarboton, G. (1997). A new method for the determination of flow directions and upslope areas in grid
digital elevation models. Water Resources Research, 33(2), 309319. http://doi.org/10.1029/96WR03137
Thomas, V., Treitz, P., Jelinski, D., Miller, J., Lafleur, P., & Mccaughey, J. H. (2002). Image
classification of a northern peatland complex using spectral and plant community data. Remote Sensing of
Environment, 84, 8399. http://doi.org/http://dx.doi.org/10.1016/S0034-4257(02)00099-8
Valeria, O., Laamrani, A., & Beaudoin, A. (2014). Monitoring the state of a large boreal forest region in
eastern Canada through the use of multitemporal classified satellite imagery. Canadian Journal of Remote
Sensing, 38(1), 91108. http://doi.org/10.5589/m12-014
Van Leeuwen, M., & Nieuwenhuis, M. (2010). Retrieval of forest structural parameters using LiDAR
remote sensing. European Journal of Forest Research, 129(4), 749770. http://doi.org/10.1007/s10342-
010-0381-4
Vierling, K. T., Vierling, L. A., Gould, W. A., Martinuzzi, S., & Clawges, R. M. (2008). Lidar: shedding
new light on habitat characterization and modeling. Frontiers in Ecology and the Environment, 6(2), 90
98. http://doi.org/10.1890/070001
Wang, T., Hamann, A., Spittlehouse, D. L., & Murdock, T. (2012). ClimateWNA - High-resolution
spatial climate data for western North America. Journal of Applied Meterology and Climatology, 51, 16
29. http://doi.org/http://dx.doi.org/10.1175/JAMC-D-11-043.1
Weiss, A., 2001.Topographic position and landforms analysis. Poster Presentation, ESRI User
Conference, San Diego, CA.
White, D. C., Lewis, M. M., Green, G., & Gotch, T. B. (2015). A generalizable NDVI-based wetland
delineation indicator for remote monitoring of groundwater flows in the Australian Great Artesian Basin.
Ecological Indicators, 112. http://doi.org/10.1016/j.ecolind.2015.01.032
Whittaker, R. (1967). Gradient Analysis of Vegetation. Biological Reviews, 49, 2077684.
Wulder, M. A., Skakun, R. S., Kurz, W. A., & White, J. C. (2004). Estimating time since forest harvest
using segmented Landsat ETM+ imagery. Remote Sensing of Environment, 93(1-2), 179187.
http://doi.org/10.1016/j.rse.2004.07.009
Wulder, M. A., White, J. C., Luther, J. E., Strickland, G., Remmel, T. K., & Mitchell, S. W. (2006). Use
of vector polygons for the accuracy assessment of pixel-based land cover maps. Canadian Journal of
Remote Sensing, 32(3), 268279. http://doi.org/10.5589/m06-023
Wulder, M. A., Han, T., White, J. C., Sweda, T., & Tsuzuki, H. (2007). Integrating profiling LIDAR with
Landsat data for regional boreal forest canopy attribute estimation and change characterization. Remote
Sensing of Environment, 110(1), 123-137. http://dx.doi.org/10.1016/j.rse.2007.02.002.
34
Xu, C. Sheng, S., Chi, T., Yang, X., An, S., & Liu, M. (2014). Developing a quantitative landscape
regionalization framework integrating driving factors and response attributes of landscapes.
Landscape and Ecological Engineering, 10(2), 295-307. 10.1007/s11355-013-0225-8
Zedler, J. B., & Kercher, S. (2005). Wetland resources: status, trends, ecosystem services, and
restorability. Annual Review of Environment and Resources, 30(1), 3974.
http://doi.org/10.1146/annurev.energy.30.050504.144248
35
36
Figure A.1. Composition of each cluster in terms of provincial map units (TEM site series). Rare
TEM site series and very small polygons were excluded
37
38
39
Figure A.2. Composition of each TEM site series in terms of the 18 clusters. Rare TEM site
series and very small polygons excluded.
... Aerial and satellite imagery have been successfully used since the early 2000s to estimate forested areas, with an 80% accuracy (Kunz et al., 2000;Haapanen et al., 2004;Wężyk and de Kok, 2005;Próchnicki et al., 2006;Wang et al., 2008;Pekkarinen et al., 2009;McRoberts, 2011McRoberts, , 2012Hościło et al., 2015;Kolecka et al., 2015;Thompson et al., 2016;Szostak et al., 2017). Using airborne scanner lasing (ALS) data makes it possible to obtain similar or better estimates of forest area (Castillo−Núńez et al., 2011;McRoberts et al., 2012;Pujar et al., 2014;Kolecka et al., 2015;Naesset et al., 2016;Thompson et al., 2016;Szostak et al., 2017). ...
... Aerial and satellite imagery have been successfully used since the early 2000s to estimate forested areas, with an 80% accuracy (Kunz et al., 2000;Haapanen et al., 2004;Wężyk and de Kok, 2005;Próchnicki et al., 2006;Wang et al., 2008;Pekkarinen et al., 2009;McRoberts, 2011McRoberts, , 2012Hościło et al., 2015;Kolecka et al., 2015;Thompson et al., 2016;Szostak et al., 2017). Using airborne scanner lasing (ALS) data makes it possible to obtain similar or better estimates of forest area (Castillo−Núńez et al., 2011;McRoberts et al., 2012;Pujar et al., 2014;Kolecka et al., 2015;Naesset et al., 2016;Thompson et al., 2016;Szostak et al., 2017). However, there is a lack of studies that detect potential forest areas, namely areas covered with trees which do not reach the required height (5 m) and canopy cover (10%) but are expected to, in 5 years or more, depending on the individual site conditions (according to the FAO/UN definition). ...
... Sophisticated methods for estimating forest area include object segmentation and super− vised classification (Pekkarinen et al., 2009;Szostak et al., 2017), but few of them consider the geometric characteristics of forest vegetation -height (Kolecka et al., 2015;Thompson et al., 2016) and canopy cover (Kolecka et al., 2015) User's accuracy -78.0 Table 5. Accuracy of the estimation of 'forests', 'potential forests' and 'non−forests' with an overall accuracy of 95%. However, their partial values (1 m, 2−20% canopy cover) do not match the values (5 m, 10%) given in the forest definitions formulated by the FAO/UN. ...
Article
Full-text available
The aim of this study is to estimate the area with forest vegetation that does not yet meet the criteria formulated in the FAO/UN definition (minimum height 5 m, minimum canopy cover 10%, minimum area 0.5 ha), but will potentially meet them in the future (5 years or more, depending on the individual site conditions), which means that (according to the definition) they also represent forest areas. The study was conducted in the Białowieża Glade. Tree species were classified individually and then divided into two groups: those that will reach a height of 5 m in the future and those that will not (grey willow, hawthorn). Hyperspectral (reduced with MNF transformation) and ALS−based features were used for classification with the SVM algorithm. Classification accuracy based on ALS data was better than that of hyperspectral data for indi− vidual species but similar for the two species groups-95.5% (Kappa 87.5%). Information about species and height was used to perform the classification of a fishnet layer into 'forests', 'potential forests' and 'non−forests', with an accuracy of 96% (Kappa 87.7%). A map of forests and potential forest vegetation was created in the form of a thematic map, taking into account height, canopy cover, area of the complex and land use. This study provides new solutions in the context of cli− mate change, deforestation and the need for reporting the forest area by individual countries (including Poland) to the FAO/UN.
... (above sea level) in the east of Calvert Island (Mount Buxton) and 495 m a.s.l. on central Hecate Island to relatively low gradient hummocky terrain in the west. The landscape is characterized by extensive wetlands, bog forests, and bog woodlands (Green, 2014;Thompson et al., 2016) on faulted and folded intrusive igneous bedrock (primarily quartz diorite; Roddick, 1996). The soils are thin (< 70 cm on average), rich in organics, and have formed in sandy colluvium and patchy morainal deposits (Oliver et al., 2017), resulting in a soil landscape with limited water storage potential. ...
... The soils are thin (< 70 cm on average), rich in organics, and have formed in sandy colluvium and patchy morainal deposits (Oliver et al., 2017), resulting in a soil landscape with limited water storage potential. Forest stands are generally short with open canopies, and the tree composition is dominated by western redcedar, yellow cedar, shore pine, and western hemlock (Thompson et al., 2016;Hoffman et al., 2021). Understory and wetland plants include bryophytes, salal, deer fern, Sphagnum mosses, and sedges. ...
... The KWO design captures the seven largest (3.0 to 12.8 km 2 ) watersheds draining into Kwakshua and Meay channels, signified as 626, 708, 703, and 693 (Calvert Island) and 1015, 819, and 844 (Hecate Island). The watersheds are on average 68 % forested, 24 % non-forested but vegetated (mainly wetlands with short vegetation), 4 % nonvegetated (mainly exposed bedrock), and 4 % covered by lakes (Thompson et al., 2016; Table 1). Contrasting watershed features are summarized by the topography, the presence of lakes, and land cover. ...
Article
Full-text available
Hydrometeorological observations of small watersheds of the northeast Pacific coastal temperate rainforest (NPCTR) of North America are important to understand land to ocean ecological connections and to provide the scientific basis for regional environmental management decisions. The Hakai Institute operates a densely networked and long-term hydrometeorological monitoring observatory that fills a spatial data gap in the remote and sparsely gauged outer coast of the NPCTR. Here we present the first 5 water years (October 2013–October 2019) of high-resolution streamflow and weather data from seven small (< 13 km2) coastal watersheds. Measuring rainfall and streamflow in remote and topographically complex rainforest environments is challenging; hence, advanced and novel automated measurement methods were used. These methods, specifically for streamflow measurement, allowed us to quantify uncertainty and identify key sources of error, which varied by gauging location. Average yearly rainfall was 3267 mm, resulting in 2317 mm of runoff and 0.1087 km3 of freshwater exports from all seven watersheds per year. However, rainfall and runoff were highly variable, depending on the location and elevation. The seven watersheds have rainfall-dominated (pluvial) streamflow regimes, streamflow responses are rapid, and most water exports are driven by high-intensity fall and winter storm events. The complete hourly and 5 min interval datasets can be accessed at 10.21966/J99C-9C14 (Korver et al., 2021), and accompanying watershed delineations with metrics can be found at 10.21966/1.15311 (Gonzalez Arriola et al., 2015).
... Different methods are used to implement these approaches (both within and between the approach classes). Most studies use the ''model-to-estimate'' approach to generate vegetation structure attributes to produce a time-series of vegetation metrics (e.g., Hudak et al., 2002;St-Onge et al., 2008;Sankey and Glenn, 2011;Varhola and Coops, 2013;Zald et al., 2016;Olsoy et al., 2017;Pitkänen and Käyhkö, 2017;Humagain et al., 2018;Matasci et al., 2018;Goldbergs et al., 2019;Hoffmann et al., 2019;Lymburner et al., 2020), while a few studies used the ''classify-to-monitor'' approach to understand vegetation dynamics (e.g., Sankey et al., 2010;Bolton et al., 2015;Thompson et al., 2016;Bolton et al., 2018;Guo et al., 2018;Campbell et al., 2020;Chirici et al., 2020;Francini et al., 2022). All of these studies have their advantages (e.g., suitable for watershed level areas and multiple environments, ability to detect vegetation changes, and suitable for heterogeneous environments) and disadvantages (e.g., not suitable for multiple environments, applied only to homogeneous landscapes, and used short time-series). ...
... LiDAR period Reflective satellite period Thompson et al. (2016) Calvert and Hecate Islands, 374.33 km 2 (British Columbia, Canada) / Cfb -Temperate oceanic climate. ...
... Aerial and satellite imagery has been successfully used since the early 2000s to estimate forest area with an accuracy of 80-99% for forest areas, 95% for forests developed during the secondary succession, and 75% for trees on agricultural land. The accuracies were calculated for the whole study sites and represent the percentage of area classified correctly according to the reference data (Kunz et al. 2000;Haapanen et al 2004;Wężyk and de Kok 2005;Próchnicki 2006;Wang et al 2007a;Pekkarinen et al 2009;McRoberts et al 2012;Pujar et al 2014;Kolecka et al 2015;Hościłło et al 2015;Thompson et al 2016;Szostak et al 2017). Unlike data from Airborne Laser Scanning (ALS, active remote sensing technology), passive remote sensing data do not contain information on elevation, which is important for the forest definition formulated for reporting purposes. ...
... Currently, ALS data are available in many countries of the world (in Poland within the ISOK project-IT system of Country Protection) and are increasingly used in remote sensing analyzes. The use of ALS data makes it possible to achieve or improve the results of forest area estimates (Castillo-Núñez et al. 2011;McRoberts et al. 2012;Kolecka et al. 2015; Thompson et al. 2016;Naesset et al. 2016;Szostak et al. 2017). Land use is an important factor responsible for differences in official forest area statistics (Seebach et al., 2011). ...
Article
Full-text available
• Key message The aim of the study was to distinguish orchards from other lands with forest vegetation based on the data from airborne laser scanning. The methods based on granulometry provided better results than the pattern analysis. The analysis based on the Forest Data Bank/Cadastre polygons provided better results than the analysis based on the segmentation polygons. Classification of orchards and other areas with forest vegetation is important in the context of reporting forest area to international organizations, forest management, and mitigating effects of climate change. • Context Agricultural lands with forest vegetation, e.g., orchards, do not constitute forests according to the forest definition formulated by the national and international definitions, but contrary to the one formulated in the Kyoto Protocol. It is a reason for the inconsistency in the forest area reported by individual countries. • Aims The aim of the study was to distinguish orchards from other lands with forest vegetation based on the data from airborne laser scanning. • Methods The study analyzed the usefulness of various laser scanning products and the various features of pattern and granulometric analysis in the Milicz forest district in Poland. • Results The methods based on granulometry provided better results than the pattern analysis. The analysis based on the Forest Data Bank/Cadastre polygons provided better results than the analysis based on the segmentation polygons. • Conclusion Granulometric analysis has proved to be a useful tool in the classification of orchards and other areas with forest vegetation. It is important in the context of reporting forest area to international organizations, forest management, and mitigating effects of climate change.
... 495 m a.s.l. on central Hecate Island to relatively low gradient hummocky terrain in the west. The landscape is characterized by extensive wetlands, bog forests, and bog woodlands (Green, 2014;Thompson et al., 2016), with shallow (typically < 1 m) but organic-rich soils (Oliver et al., 2017), on faulted and folded intrusive igneous bedrock 90 (primarily quartz diorite) (Roddick, 1996). Forest stands are generally short with open canopies and tree composition is dominated by western redcedar, yellow cedar, shore pine, and western hemlock (Thompson et al., 2016;Hoffman et al., 2021). ...
... The landscape is characterized by extensive wetlands, bog forests, and bog woodlands (Green, 2014;Thompson et al., 2016), with shallow (typically < 1 m) but organic-rich soils (Oliver et al., 2017), on faulted and folded intrusive igneous bedrock 90 (primarily quartz diorite) (Roddick, 1996). Forest stands are generally short with open canopies and tree composition is dominated by western redcedar, yellow cedar, shore pine, and western hemlock (Thompson et al., 2016;Hoffman et al., 2021). Understory and wetland plants include bryophytes, salal, deer fern, sphagnum mosses and sedges. ...
Preprint
Full-text available
Hydrometeorological observations of small watersheds of the northeast Pacific coastal temperate rainforest (NPCTR) of North America are important to understand land to ocean ecological connections and to provide the scientific basis for regional environmental management decisions. The Hakai Institute operates a densely networked and long-term hydrometeorological monitoring observatory, that fills a spatial data gap in the remote and sparsely gauged outer coast of the NPCTR. Here we present the first five water years (October 2013–October 2019) of hourly streamflow and weather data from seven small (
... accessed on 7 September 2020) ( Table 1) were acquired and mosaicked to cover the entire study area. These images have been widely used in cities, mainly to delimit urban forms of development, identify vegetation in urban environments, detect embedded ecosystems in urban areas, and monitor agricultural uses [30,[55][56][57][58], due to their high spatial resolution. The images had a level 3A, meaning they incorporate geometric correction and topographic correction based on 30 and 90 m DTM and projected in Universal Transverse Mercator (UTM), Zone 18 South, Datum WGS84. ...
Article
Full-text available
Coastal wetlands areas are heterogeneous, highly dynamic areas with complex interactions between terrestrial and marine ecosystems, making them essential for the biosphere and the development of human activities. Remote sensing offers a robust and cost-efficient mean to monitor coastal landscapes. In this paper, we evaluate the potential of using high resolution satellite imagery to classify land cover in a coastal area in Concepción, Chile, using a machine learning (ML) approach. Two machine learning algorithms, Support Vector Machine (SVM) and Random Forest (RF), were evaluated using four different scenarios: (I) using original spectral bands; (II) incorporating spectral indices; (III) adding texture metrics derived from the grey-level covariance co-occurrence matrix (GLCM); and (IV) including topographic variables derived from a digital terrain model. Both methods stand out for their excellent results, reaching an average overall accuracy of 88% for support vector machine and 90% for random forest. However, it is statistically shown that random forest performs better on this type of landscape. Furthermore, incorporating Digital Terrain Model (DTM)-derived metrics and texture measures was critical for the substantial improvement of SVM and RF. Although DTM did not increase the accuracy in SVM, this study makes a methodological contribution to the monitoring and mapping of water bodies’ landscapes in coastal cities with weak governance and data scarcity in coastal management.
... Wet and dry seasons are less pronounced than those found in the other catchments investigated in this study, but generally span from September through April and May through August, respectively (Oliver et al., 2017). Short statured bog forests and wetlands make up the main vegetation cover (Thompson et al., 2016). Western red cedar (Thuja plicata), yellow cedars (Chamaecyparis nootkatensis), western hemlock (Tsuga heterophylla), and shore pine (Pinus contorta) are the dominant trees. ...
Article
Full-text available
Plain Language Summary Periods of intense infiltration in the form of rainfall (storm events) or snowmelt push large volumes of dilute fluid through catchments, resulting in evacuation of dissolved solutes derived from weathering reactions. These events occur over short (daily or weekly) timescales, but these pulses of solutes released into small streams can represent a majority of total annual mass flux (from 40% to 70%). Despite their importance, storm events remain largely underrepresented in studies of weathering in watersheds due to challenges in obtaining high frequency measurements. Yet extreme events are expected to become more frequent in the future in response to a changing climate. In this study, we evaluate the dependence of weathering‐derived solute export on hydrology and biogeochemical cycling through measurements of dissolved silicon and silicon stable isotope ratios of stream waters collected during seven significant infiltration events in six small catchments spanning different bedrock lithology and climates. Each catchment was found to have unique silicon chemical and isotopic export signatures reflecting site‐specific subsurface water routing and combination of biogeochemical reactions. These signatures were found to be preserved despite the perturbation exerted by such enhanced infiltration and discharge. In this sense, catchments exhibit resilience, bending but not ceding to strong hydrologic forcing.
... Cluster analysis is an exploratory classification technique in which samples that are closer together are classified into one category by calculating the distance between them, and those that are farther apart belong to different categories (Kaufman and Rousseeuw 2009;Thompson et al. 2016). Without prior knowledge, that is, training samples, a set of forest health indicators (FHIs) is directly explored and constructed from UAV LiDAR point cloud data by an unsupervised clustering method. ...
Article
Full-text available
The Yellow River Delta (YRD) has China's largest artificial Robinia pseudoacacia forest, which was planted in the late 1970s and suffered extensive dieback in the 1990s. The health grade of the R.pseudoacacia forest (named canopy vigor grade, CVG) could be achieved by using high-resolution images and canopy vigor indicators (CVIs). However, a previous study showed that there was no significant correlation between CVG and the field-estimated aboveground biomass (AGB) of R.pseudoacacia forest. Therefore, this study aims to construct forest health indicators (FHIs) based on canopy spatial structure parameters extracted from LiDAR. The FHIs included Weibull_α (the scale parameter of the Weibull density function that reflects the shape of the tree canopy), VCI (vertical complexity index), sdCC (the standard deviation of canopy cover), H99 (the 99th percentile height) and cvLAD (the coefficient of variation of leaf area density), and could significantly distinguish three forest health grades (FHG) (p < 0.05). The FHG was positively correlated with forest AGB (rs = 0.51, p = 0.004), and the similarity value with CVG was 63.33%. The results of this study confirmed that the FHIs can reflect both canopy vigor and tree productivity, and distinguish forest health status without prior classification information.
... Calvert and Hecate Islands contain four different vegetative types that vary altitudinally and are classified based on dominant species and productivity (Banner et al., 2005;Thompson et al., 2016). All habitation and control sites were established in the zonal forest, the highly productive riparian interface between the coast and the bog forest further upland. ...
Article
Full-text available
Identifying how past human actions have influenced their environment is essential for understanding the ecological factors that structure contemporary ecosystems. Intertidal resource use by Indigenous Peoples for thousands of years has led to habitation sites containing vast shell midden deposits and facilitating long-term impacts on soil chemistry and drainage. Here we examine how these shell middens have impacted various forest metrics, such as species diversity, community composition, canopy height, and regeneration recruitment to determine if forests on habitation sites differ from the surrounding matrix. We surveyed known habitation sites with archeological evidence indicating past year-round human occupation, within the Hakai Lúxvbálís Conservancy on Calvert and Hecate Islands within the Great Bear Rainforest along British Columbia’s Central Coast. Our results demonstrate that habitation sites exhibit lower tree species richness, less relative species abundances, as such, displayed lower Shannon diversity and inverse Simpson values. The composition of tree communities on habitation sites was statistically different, with western hemlock and western redcedar densities increasing on non-habitation sites. Conversely, regeneration diversity at habitation sites was more even and exhibited elevated Shannon diversity and inverse Simpson values. The community composition of regeneration was more consistent among habitation and non-habitation sites; however, western redcedar, western hemlock and Sitka spruce were more abundant at habitation sites. For all tree species, maximum height was higher within the habitation sites; however, this trend was the most notable in western redcedar and Sitka spruce, which increased by an average of 4.8 m relative to non-habitation sites. Collectively, our findings suggest that long-term habitation alters forest community compositions. The landscape alterations within habitation sites promote conditions needed to support diverse, even, and abundant regeneration communities and consequently increase the height of the dominant coastal tree species. Thus, our results offer evidence that long-term influence by Indigenous communities have a persistent influence on coastal forests.
Article
Full-text available
Bioblitzes are a tool for the rapid appraisal of biodiversity and are particularly useful in remote and understudied regions and for understudied taxa. Lichens are an example of an often overlooked group, despite being widespread in virtually all terrestrial ecosystems and having many important ecological functions. We report the lichens and allied fungi collected during the 2018 terrestrial bioblitz conducted on Calvert Island on the Central Coast of British Columbia, Canada. We identified 449 specimens belonging to 189 species in 85 genera, increasing the total number of species known from Calvert Island to 194, and generated Internal Transcribed Spacer (ITS) sequences for 215 specimens from 121 species. Bryoria furcellata , Chaenothecopsis lecanactidis and C. nigripunctata were collected for the first time in British Columbia. We also found Pseudocyphellaria rainierensis , which is listed as Special Concern on the federal Species at Risk Act, and other rarely reported species in British Columbia including Opegrapha sphaerophoricola , Protomicarea limosa , Raesaenenia huuskonenii and Sarea difformis . We demonstrate that DNA barcoding improves the scope and accuracy of expert-led bioblitzes by facilitating the detection of cryptic species and allowing for consistent identification of chemically and morphologically overlapping taxa. Despite the spatial and temporal limitations of our study, the results highlight the value of intact forest ecosystems on the Central Coast of British Columbia for lichen biodiversity, education and conservation.
Article
Full-text available
A new map of standardized, mesoscale (tens to thousands of hectares) terrestrial ecosystems for the conterminous United States was developed by using a biophysical stratification approach. The ecosystems delineated in this top-down, deductive modeling effort are described in NatureServe's classification of terrestrial ecological systems of the United States. The ecosystems were mapped as physically distinct areas and were associated with known distributions of vegetation assemblages by using a standardized methodology first developed for South America. This approach follows the geoecosystems concept of R.J. Huggett and the ecosystem geography approach of R.G. Bailey.Unique physical environments were delineated through a geospatial combination of national data layers forbiogeog-raphy. bioclimate. surficial materials lithology. land surface forms, and topographic moisture potential. Combining these layers resulted in a comprehensive biophysical stratification of the conterminous United States, which produced 13.482 unique biophysical areas. These were considered as fundamental units of ecosystem structure and were aggregated into 419 potential terrestrial ecosystems.The ecosystems classification effort preceded the mapping effort and involved the independent development of diagnostic criteria, descriptions, and nomenclature for describing expert-derived ecological systems. The aggregation and labeling of the mapped ecosystem structure units into the ecological systems classification was accomplished in an iterative, expert-knowledge-based process using automated rulesets for identifying ecosystems on the basis of their biophysical and biogeographic attributes. The mapped ecosystems, at a 30-meter base resolution, represent an improvement in spatial and thematic (class) resolution over existing ecoregionaliza-tions and are useful for a variety of applications, including ecosystem services assessments, climate change impact studies, biodiversity conservation, and resource management.
Article
Full-text available
This chapter covers mapping vegetation from landscape to regional scales, with sections on scale, data, methods, examples of recent maps illustrating their uses, dynamic mapping, and the future of vegetation mapping research.
Article
Full-text available
Topographic analysis of watershed-scale soil and hydrological processes using digital elevation models (DEMs) is commonplace, but most studies have used DEMs of 10 m resolution or coarser. Availability of higher-resolution DEMs created from light detection and ranging (lidar) data is increasing but their suitability for such applications has received little critical evaluation. Two different 1 m DEMs were re-sampled to 3, 5, and 10 m resolutions and used with and without a low-pass smoothing filter to delineate catchment boundaries and calculate topographic metrics. Accuracy was assessed through comparison with field slope measurements and total station surveys. DEMs provided a good estimate of slope values when grid resolution reflected the field measurement scale. Intermediate scale DEMs were most consistent with land survey techniques in delineating catchment boundaries. Upslope accumulated area was most sensitive to grid resolution, with intermediate resolutions producing a range of UAA values useful in soil and groundwater analysis.
Article
Full-text available
A vegetation community map was produced for the Ozark National Scenic Riverways consistent with the association level of the National Vegetation Classification System. Vegetation communities were differentiated using a large array of variables derived from remote sensing and topographic data, which were fused into independent mathematical functions using a discriminant analysis classification approach. Remote sensing data provided variables that discriminated vegetation communities based on differences in color, spectral reflectance, greenness, brightness, and texture. Topographic data facilitated differentiation of vegetation communities based on indirect gradients (e.g., landform position, slope, aspect), which relate to variations in resource and disturbance gradients. Variables derived from these data sources represent both actual and potential vegetation community patterns on the landscape. A hybrid combination of statistical and photointerpretation methods was used to obtain an overall accuracy of 63 percent for a map with 49 vegetation community and land-cover classes, and 78 percent for a 33-class map of the study area.
Chapter
The effects of global climate change will have serious consequences for wetland ecosystems, which store a substantial portion of the world’s carbon and so act to buffer the increasing concentrations of atmospheric carbon dioxide. Warmer and drier conditions may lead to increases in the flux of carbon from wetlands to the atmosphere, the loss of both inland and coastal wetlands, their biodiversity, and the provision of ecosystem services.
Article
The radiant energy income of a slope influences its ambient temperatures and water movements, both of which are important controls on the growth behaviour, species composition and structure of its vegetation cover. Therefore, information about the radiation environments of topographically diverse areas should provide a basis for predicting the likelihood of local variations in vegetation composition and structure. From a simple model of the annual shortwave energy load of slopes of different angle and compass orientation we predict that aspect effects should be greatest at 45⚬ N/S and least in equatorial and polar regions. Other predictions concern likely physiological responses of plants to varying slope angle over the range of latitude. A literature review shows good agreement between these physically based predictions and observations of vegetation patterns in a geographically wide range of countries.
Article
The Ontario Ministry of Natural Resources (OMNR) and Ducks Unlimited Canada (DUC) have been engaged in developing an efficient and accurate methodology for inventorying wetlands. Their progress in this area has demonstrated that Digital Elevation Models (DEMs) are crucial input for wetland identification and boundary delineation. The provincial DEM, however, has known precision limitations in areas of minimal topographic relief that cause considerable mapping error. This study explored whether wetland mapping derived from bare-earth light detection and ranging (LiDAR) data would overcome the limitations of the provincial DEM. An automated wetland mapping approach was applied to the 2 elevation datasets and the results were compared using 2 methods of validation. One hundred aerial-photo-interpreted sample plots were used to quantitatively measure the ability of each source to separate upland from wetland. An overlay of wetland maps created from the 2 DEM sources was then qualitatively assessed to further clarify the magnitude of discrepancy between the 2 mapping sources. The study concluded that LiDAR showed a significant improvement at p = 0.05 over the provincial DEM for mapping wetlands, improving overall mapping accuracy from 76% to 84%. However, an overlay analysis and qualitative assessment showed the magnitude of this reported improvement is greater than was quantified by the accuracy assessment and that an assessment scheme with different sample units may further elucidate this discrepancy.