Open Access
How to translate text using browser tools
1 May 2007 Development of a Spatial Analysis Method Using Ground-Based Repeat Photography to Detect Changes in the Alpine Treeline Ecotone, Glacier National Park, Montana, U.S.A
W. Roush, J. S. Munroe, D. B. Fagre
Author Affiliations +
Abstract

Repeat photography is a powerful tool for detection of landscape change over decadal timescales. Here a novel method is presented that applies spatial analysis software to digital photo-pairs, allowing vegetation change to be categorized and quantified. This method is applied to 12 sites within the alpine treeline ecotone of Glacier National Park, Montana, and is used to examine vegetation changes over timescales ranging from 71 to 93 years. Tree cover at the treeline ecotone increased in 10 out of the 12 photo-pairs (mean increase of 60%). Establishment occurred at all sites, infilling occurred at 11 sites. To demonstrate the utility of this method, patterns of tree establishment at treeline are described and the possible causes of changes within the treeline ecotone are discussed. Local factors undoubtedly affect the magnitude and type of the observed changes, however the ubiquity of the increase in tree cover implies a common forcing mechanism. Mean minimum summer temperatures have increased by 1.5°C over the past century and, coupled with variations in the amount of early spring snow water equivalent, likely account for much of the increase in tree cover at the treeline ecotone. Lastly, shortcomings of this method are presented along with possible solutions and areas for future research.

Introduction

Repeat photography is the process of locating the position from which an existing photograph was taken, occupying that photo-point, and taking a new photograph to create a photo-pair of the same scene (Rogers et al., 1984). This method can effectively document landscape change and has been employed since 1888, initially for monitoring glacial movement but more recently in broader applications (Harper, 1934; Rogers et al., 1984; Klett, 2004). Over the past 25 years numerous books and studies have used repeat photography to provide compelling qualitative information about the magnitude and type of long-term ecological and geological processes that would not otherwise be possible and have produced photo-pairs that are easily interpreted by a broad audience (Rogers, 1982; Gruell, 1983; Vale and Vale, 1983, 1994; Klett et al., 1984; Baars et al., 1994; Webb, 1996; Fielder et al., 1999; Butler and DeChano, 2001; Klasner and Fagre, 2002; Rhemtulla et al., 2002; Munroe, 2003; Klett, 2004).

While ground-based repeat photography can effectively display qualitative landscape change, it has rarely been used to measure quantitative change, largely because an oblique perspective creates a continuously varying scale and prevents linking landscape features in the photograph to absolute spatial coordinates. Detailed quantitative measurements of landscape processes and changes can be made from airborne and space-borne platforms. However, systematic aerial photographic surveying only began ca. 1950, while satellite images only became available in the 1960s. Furthermore, both the methods used in, and results of, analysis of aerial photography and satellite images are often not easily understood by the general public. In contrast, ground-based photographic records extend as far back as the 1860s and are a universally familiar medium. The development of new methodologies capable of extracting meaningful, quantitative information from historical photographs would provide a technique for addressing questions of landscape change as well as a way to record and measure responses of landscapes on a longer time scale than those offered by satellites and other modern equipment.

Earlier efforts to extract quantitative information from ground-based photography include Hofgaard et al. (1991) and Kullman (1987), who repeated photographs of individual trees to measure tree fitness and growth during the intervening period. More recent studies have developed methodologies to quantitatively document landscape change. Notable among these are the use of on-screen sampling of vegetation in photo-pairs to determine percent change in vegetative cover in the Uinta Mountains, Utah (Munroe, 2003) and development of a GIS method that digitized polygons based on vegetation type and used a transition matrix to document vegetation changes in the montane zone of Jasper National Park, Canada (Rhemtulla et al., 2002).

The alpine treeline, defined here as the broad ecotone stretching from the end of full canopy forest to isolated krummholz patches, is an excellent landscape feature for use in further development and testing of these techniques for several reasons. First, abrupt and clear distinctions between plant communities makes changes in the treeline ecotone easier to detect visually than in other transitional environments. Second, the treeline ecotone is a dynamic environment responding to a variety of environmental factors (Cairns, 1990; Holtmeier, 2003), ensuring there will be landscape changes detectable by repeat photography. Finally, because significant ecological changes in alpine plant communities take place over longer time scales than most field studies, quantified repeat photography is one of the few ways to document change and increase our understanding of treeline ecotones.

Our study had two objectives: (1) to develop a method for quantitative analysis of repeated photographs, and (2) to apply this method to the treeline ecotone of Glacier National Park (GNP), Montana, to examine vegetation changes over multi-decadal timescales. Treeline change in GNP has been documented (Butler et al., 1994; Cairns and Malanson, 1997, 1998; Butler and DeChano, 2001; Cairns, 2001; Klasner and Fagre, 2002), but these previous studies could be augmented both temporally and spatially by using historical photographs taken during the last century. Here we describe a new method applying ArcGIS software (ESRI, 2002) to digital photo-pairs that allows vegetation change to be categorized and quantified. The patterns of tree establishment identified in GNP are described, and the possible causes of changes at the treeline ecotone are discussed. Finally, shortcomings of this method are presented along with possible solutions and areas for future research.

Study Area

Comprehensive photographic archives, extensive documentation of paleoclimates and past human interactions with the landscape, and considerable prior research on alpine and subalpine environments make GNP an ideal study area for repeat photography. GNP occupies 4075 km2 in northwestern Montana (Fig. 1), where elevations range from 948 to 3290 m a.s.l. (Shaw and On, 1979; Rockwell, 2002). Steep topography, climatic fluctuations, and past glaciations have created a treeline that is much more spatially variable than that occurring in the southern Rocky Mountains (Cairns and Malanson, 1997). Becwar and Burke (1982) used topographic maps to estimate that 80% of the forest-tundra transition spans an elevation range of 550 m in GNP, compared to 150 m in the southern Rocky Mountains.

Figure 1

Location of Glacier National Park. Latitude and longitude are given in Figure 2.

i1523-0430-39-2-297-f01.gif

Figure 2

Locations of the 12 sites which were rephotographed in Glacier National Park. See Table 2 for site details.

i1523-0430-39-2-297-f02.gif

The area that is now GNP was protected as a forest reserve prior to 1910, when the limited human impact in the area was further reduced by the establishment of the park. The area was never commercially logged, and grazing has been largely limited to horses used for pack trips (Shaw and On, 1979; Vogler, 1998). During the 1930s an outbreak of whitebark pine blister rust in high elevation sites infected whitebark pines (Pinus albicaulus Engem.) and by the 1950s had killed most of these trees (Kendall and Keane, 2001). Recent human impact in alpine areas of GNP is minimal and is largely due to trail use and maintenance. Probably the greatest direct human impact on vegetation in GNP has been a policy of fire suppression over the past 80 years that has contributed to increased forest cover and reduced early successional habitats. This policy now allows some fires to burn for wildland benefit. Fires occasionally burned into alpine areas in the past and may have affected treeline ecotone dynamics (Arno, 2001). Study sites for this project were located within the treeline ecotone both at the extreme upper limit of trees and in open subalpine meadows at slightly lower elevations. Topography at the sites varied from gentle meadows to steep rocky hillsides. The dominant treeline species at all sites is subalpine fir (Abies lasiocarpa [Hook.] Nutt), with isolated instances of Engelmann spruce (Picea engelmannii Parry ex Engelm.). Observed treelines were both orographic and altitudinal as defined by Holtmeier (2003).

Methods

Selection of Historical Photographs

The historical photographs replicated for this study were obtained from GNP's archive and were taken by three early 20th century photographers: George Grant, Tomer J. Hileman, and Fred Kiser. The collection of photographs from which selections were made was extensive and contained views from throughout the park, although particularly scenic or easily accessible views are more common. Approximately 2000 photographs were examined before final selections were made. Photographs were selected based on several criteria. First, the images needed to be of high quality and to clearly document the treeline ecotone. Second, photographs were selected with the goal of isolating climate as the primary factor that would cause a change in the characteristics of the treeline ecotone shown in the photograph. Third, photographs of a hillside of roughly constant slope that extended above a non-orographic treeline were preferred. Album prints of 17.5 × 26.5 cm (8 × 12 in.) were scanned at 273 dots per cm [600 dots per inch (DPI)], and saved as Tagged Image File Format (TIFF) files. Printed copies were used in the field to locate historic photo-points.

Locating Historical Photo-points

To locate the sites from which historical photographs were taken, information accompanying the photographs or prominent landmarks in the view were used to find the general area on a topographic map. To determine the precise location of photo-points in the field, the principle of parallax was used to find lines of equal perspective and proportionate length between features (e.g. background ridgeline intersections and permanent foreground objects, such as boulders) in the historical photograph and the present landscape (Rogers et al., 1984). Photo-points were located to within ~20 m using background features, then adjusted to within ~1 m of the original point using foreground features. In a few photo-pairs new vegetation obscured the view shown in the historical photograph, so the modern photograph was taken from a slightly different location (<2 m away). Figure 2 shows the locations of all photo-points included in this study.

A Nikon F-100 35 mm single lens reflex camera with a 24–200 mm f3.5 Nikon lens was used to shoot Kodak technical-pan black and white film at an ASA of 12. For each historical photograph site, 12 frames were taken with bracketed exposures; 6 of these used a Yellow 12 filter to increase contrast. The location of each photo-point was recorded using both a GPS unit and a topographic map. Camera height and bearing as well as aperture, shutter-speed, and focal length were also recorded.

Analysis of Photo-pairs

Of the 12 modern negatives per photo-pair, the image with exposure and contrast values which most closely matched the historical image was selected for comparative analysis. All modern negatives were converted to TIFF files by scanning at 1820 dots per cm (4000 DPI). To quantitatively analyze change between the modern and historic image files, an original method was developed using ArcGIS software (ESRI, 2002). First, the paired image files were imported into ArcMAP and ortho-referenced to each other using a minimum of five common points with the majority of these located in the vicinity of the area to be analyzed. This step ensured that the scales of the two images were equal and that they were overlain exactly in ArcMAP's spatial system. Second, an ESRI “fishnet” script (ESRI, 2002; Nicholas, 2003) was used to create a grid overlying the area of interest. Each cell of the grid is an independent polygon (Fig. 3), and information about what lies under each grid cell can then be used to populate a table corresponding to the grid in a spatially explicit manner. The grid was scaled to vegetation features depicted in each photo-pair, and because the scale of each photo-pair and corresponding grid varies depending on the distance of the subject from the photo-point, the amount of ground covered by each cell differs among photo-pairs. Thus, only relative changes can be measured between photo-pairs. Additionally, within a photo-pair, foreground space will occupy a larger area in the image than background space. To mitigate this issue, the area chosen for analysis in each photo-pair was held to a roughly constant distance from the photo-point.

Figure 3

Attribute table and associated grid overlain on study site Hileman 5053. Each cell has a unique identifier (FID). Information regarding historical and modern vegetation as well as the calculated type of vegetation change populates the table.

i1523-0430-39-2-297-f03.gif

Third, the grids scaled to each photo-pair were digitally overlain on the images, and cropped to the area from which cell-by-cell changes were to be measured. An attribute table was created for each grid file, with fields containing information regarding the vegetation visible in each cell (Fig. 3). Each cell received a value of 1, 2, or 3 indicating bare ground, partial tree cover, and complete tree cover, respectively. This information was then used to create mutually exclusive categories (no change, vegetation loss, infilling, and establishment) describing the type of tree cover change occurring within each grid cell. No change indicates no visible difference in the cell between the two photographs; vegetation loss is defined as a decrease in visible vegetative cover in the modern photograph; infilling is defined as increased or denser vegetation in areas which showed sparse vegetation in the historical photograph; and establishment is defined as new vegetation in areas which showed bare ground in the historical photograph (Table 1). It was not possible to determine the absolute change in tree cover documented in the photo-pairs because of the continuously varying scale created by the oblique perspective and the complex topography.

Table 1

Definitions of cell types showing categories of tree cover change.

i1523-0430-39-2-297-t01.gif

Finally, to evaluate the effect cell size had on the analysis, two photo-pairs were resampled with a new grid, the cell size of which was four times larger than used in the first analysis, and these results were compared to the original analysis.

Results and Discussion

Thirty-two historical photographs were replicated in July and August 2003. Of these, 11 (showing 12 sites) were suitable for analysis because they had good lighting, readily distinguishable vegetation features, and areas which appeared not to have burned. The time span represented by the photo-pairs ranged from 71 to 93 years (mean = ~75) (Table 2). The most notable trend illustrated by the photo-pairs is a change in percent tree cover, which increased in 10 out of 12 photo-pairs (Table 2), with a mean increase of 60% (Range: −4% to 366%; Standard Deviation (S.D.) = 103%). Excluding the two sites (Hileman 8042, and HDL 930a) in which tree cover increased most dramatically, the mean is 23% (S.D. = 23%).

Table 2

Relevant information for each site analyzed.

i1523-0430-39-2-297-t02.gif

Tree establishment was documented in all photo-pairs, including the two that showed an overall loss in tree cover (Table 3). Infilling occurred in all sites except Kiser 4714, though overall in lesser amounts. Establishment and infilling appear to be inversely related. The greatest infilling occurred at two sites (Kiser 11813 and Hileman 2006a) with low establishment values, and the highest establishment values occurred at sites (Hileman 8042 and HDL 930a) with low amounts of infilling (Table 3).

Table 3

Percent of cells showing establishment, infilling, vegetation loss, and no change in each photo-pair. Sites are ordered by percent establishment.

i1523-0430-39-2-297-t03.gif

The method developed for this project introduces quantitative rigor to the assessment of landscape change documented by photo-pairs and reduces subjective evaluations or a preferential focus on obvious changes in landscape elements. When applied to the alpine treeline ecotone of GNP, this method identified the magnitude and direction of vegetation change, the type of change (no change, vegetation loss, infilling, and establishment), and spatial distribution of change. In the following discussion, three separate photo-pairs that illustrate different spatial patterns of vegetation change are described, along with possible causes for these changes.

Photo-pair Hileman 8042

This photo-pair illustrated the most dramatic expansion of trees into a meadow environment, likely as a result of several factors that make this site ideal for tree growth and establishment (Fig. 4). First, stands of mature trees located just to the north (right) of the study site provide a substantial and proximal seed bank. Second, the physical environment at this site appears generally favorable for tree growth, as exemplified by the trees in the original Hileman 8042 photograph. Despite their location immediately below the upper treeline, these trees have a growth form more typical of those in lower elevation, mature forests. In contrast, other historical photographs show trees in krummholz forms or as isolated tree patches that are often located below cliffs and on more exposed mountainsides where winter desiccation and ice abrasion are greater impediments to upright tree growth (Holtmeier, 2003; Malanson et al., in press). Finally, the ground at this site is flat to gently sloping and includes many minor topographic highs, which Butler et al. (2003) observed to have a strong influence on tree position by providing well-drained soil that increases establishment.

Figure 4

Detail of photo-pair Hileman 8042 showing vegetation type superimposed on the historical photograph and the type of vegetation change superimposed on the modern photograph. Note the massive invasion of trees across the entire photograph.

i1523-0430-39-2-297-f04.jpg

Photo-pair Hileman 5053

The treeline in the photo-pair Hileman 5053, located at the base of a scree slope and near the upper elevation of tree growth, provides an example of two controls acting in unison on the treeline at a single location (Fig. 5). The photo-pair shows that the upper limit of tundra has not changed perceptibly in the intervening 72 years, suggesting that tundra vegetation is unable to colonize the talus. Upright trees however, have clearly expanded their range (approximately 10–20 vertical m), with some now occupying the upper limit of all vegetation at this site. It is possible that the functional treeline at this site may be depressed due to the presence of the talus, which is hypothesized to be the second largest control on treeline in GNP after climate (Cairns, 2001). This notion is supported by the presence of upright trees within the upper portion of the treeline ecotone at this site. Thus, the combined effect of tree establishment and infilling, and the inability of alpine meadow species to colonize the active talus, has caused a marked decrease in tundra area at this site while the density of trees has increased.

Figure 5

Detail of photo-pair Hileman 5053 showing vegetation type superimposed on the historical photograph and the type of vegetation change superimposed on the modern photograph. Note the disappearance of meadow below the talus and the preference of tree establishment to occur adjacent to historical vegetation in the three tongues of meadow which used to reach down into the forest.

i1523-0430-39-2-297-f05.jpg

The majority of the infilling and establishment in this photo-pair occurred along the edges of historical vegetation and may be a result of existing trees improving conditions for establishment in their immediate environment by altering soil characteristics and microclimates (Stevens and Fox, 1991; Holtmeier, 2003; Holtmeier and Broll, 2005). Specifically, establishment was concentrated in the three tongues of meadow that reach down into the trees in the historical photograph (Fig. 5). It is not possible to determine the direct cause of the observed vegetation change from the photo-pair alone. However, the distinct spatial pattern related to distance from both the talus and historical tree clumps suggests that protection from wind plays a role in the spatial variability of tree establishment within the treeline ecotone as was observed by Resler (2006) on Lee Ridge in GNP.

Photo-pair Kiser 11813

The change seen in photo-pair Kiser 11813 is notable because 61% of the grid cells record change (Fig. 6), but this occurred largely as infilling (41%), with a relatively low level of establishment (14%). Tree growth proceeded outward from the lower center of the site, where trees were densest in the historical photograph. The small amount of establishment occurred primarily on the rising slope at the left side of the site and represents an overall altitudinal increase of 5–10 m. Most notable is the disappearance of the ribbon forests present in the historical photograph. This forest type is thought to be caused by underlying bedrock topography (Butler et al., 2003) creating topographic highs and depressions in the landscape. The observed change indicates that the factors which caused the infilling of the ribbon forest are powerful enough to overcome geologic controls on tree location.

Figure 6

Detail of photo-pair Kiser 11813, showing vegetation type superimposed on both the historical and modern photographs. Note the disappearance of the ribbon forest environment present in the historical photograph and the slight upward movement of treeline on the far left side of the photograph.

i1523-0430-39-2-297-f06.jpg

General Trends and Interpretations

The most common trend documented by the photo-pairs is an increase in tree cover, both as infilling and establishment. Establishment occurred up to hundreds of meters away from historical vegetation, and half the sites experienced a 35% or greater increase in total tree cover. Even in the two sites where overall tree cover decreased, localized establishment and infilling occurred. Local factors certainly play a significant role in determining the magnitude of change that occurs at any given site (Holtmeier, 2003), yet the ubiquity of the observed changes indicates there is a more universal forcing mechanism that is amplified or attenuated by local conditions. Previous studies have documented an increase in tree establishment in subalpine forests in response to increased growing season temperature (Kullman, 1991; Rochefort et al., 1994; Cairns and Malanson, 1998; Kusnierczyk and Ettl, 2002; Lloyd et al., 2002), while Hofgaard et al. (1991) and Kullman (1991) in Scandinavia, and Rochefort et al. (1994) and Hessl and Baker (1997) in the American and Canadian Rockies, found a strong correlation between increased tree growth at treeline and rising summer temperatures. Data from the Kalispell WSO AP Station, part of the Historical Climate Network (Easterling et al., 1996) reveal that while the yearly mean temperature has increased only a small amount since 1900 (about 0.5°C), minimum summer temperatures have increased about 1.5°C in the past century (Fig. 7), lengthening the growing season by increasing the number of frost-free days. Watson et al. (in press) show a similar or greater increase in minimum spring and summer temperatures from 15 temperature stations along the continental divide. Butler et al. (1994) noted that increases in tetraterm temperatures (average temperature of the period June to September) observed in GNP from 1984 to 1991 could be linked to seedling establishment. And in a summary of studies examining seedling establishment in the alpine tundra, Alftine et al. (2003) found tundra invasion by forests of the Pacific Northwest to be associated with warmer, drier periods. Together these types of climatic changes would improve the carbon balance for existing trees and make it easier for seeds to germinate and prosper. While establishing direct causality for increased tree growth is not possible from the photo-pairs alone, changes in climate occur at a large enough scale to have influenced tree establishment at all sites in this study. Thus, the rise in mean summer minimum temperatures over the past several decades (Fig. 7) is a possible explanation for the observed changes, with differences in the magnitude and nature of the observed change demonstrating the importance of site specific conditions.

Figure 7

Annual spring snow pack (1 May Snow Water Equivalent) from Flattop Mountain and mean minimum summer (June–August) temperatures at the U.S. Historical Climatology Network's Kalispell WSO AP Station, 70 km from Glacier National Park. Black lines represent 10-year moving averages.

i1523-0430-39-2-297-f07.gif

Precipitation also affects tree growth and establishment but in more complex ways depending on whether water comes in the form of snow or rain (Holtmeier, 2003). Most of the precipitation at high elevations in GNP occurs as snow (Klasner and Fagre, 2002) and has more interannual variation than temperature (Fig. 7). Kusnierczyk and Ettl (2002) documented a positive correlation between non-growing season precipitation and tree growth, and given that cold soils and high winds are the largest limiting factors on tree establishment in continental climates, a deep snow pack would reduce these negative effects and aid tree establishment (Rochefort et al., 1994). The 10-year running mean of 1 May Snow Water Equivalent (SWE) from Flattop Mountain in GNP shows a marked increase during the period 1950–1975 (Fig. 7). This increase differs from the temperature record, which is fairly stable during this period and shows its most significant gains between 1915 and 1930 and again between 1970 and 1980 (Fig. 7). Selkowitz et al. (2002) demonstrated a link between longer term (20–30 yr) cycles of increased SWE within GNP and the negative phase of the Pacific Decadal Oscillation (PDO), and Alftine et al. (2003) found high levels of establishment occurring in the treeline ecotone of Lee Ridge in GNP during the negative phase (late 1940s to late 1970s) of the PDO. It is likely then that some of the establishment seen in the photo-pairs is a result of the increased snow levels during the period 1950–1975.

Consistent with other studies in GNP, photo-pairs from this study indicate that krummholz patches appear to have changed far less than upright trees at treeline or in the subalpine meadow environment. A rephotographic study by Butler et al. (1994) showed treeline position to have been quite stable throughout the park from 1972 to 1992, including krummholz patches that were exactly the same size in historical and modern photographs.

Photo-pair Kiser 11810, which includes two separate sites, underscores the complex interrelationships between altitude, growth form, precipitation, and exposure. Tree establishment was more than four times greater (Table 2) at the exposed mountain site where upright trees are present ~245 m above the sheltered, krummholz-dominated valley site. The valley site likely has a greater potential for excessive snow drifting, which has been negatively correlated with establishment by limiting growing season length and quality (Franklin et al., 1971) and by reducing the carbon balance of the trees (Cairns and Malanson, 1998). Other studies have shown that snow accumulation can benefit trees by protecting them from winter desiccation and injury (Walsh and Kelly, 1990; Cairns and Malanson, 1997; Hattenschwiler and Smith, 1999). In this case, it seems likely the valley site receives so much wind-deposited snow that the shortened growing season outweighs the sheltering effects of the winter snow pack and limits seedling establishment.

Finally, it should be noted that fire is not thought to be a major factor in causing the changes documented by the studied photo-pairs. Forest stand-age maps were consulted in selecting the historical photographs to eliminate the possibility that the changes documented in the photo-pairs were simply due to succession following a stand-clearing fire. These stand-age maps do not necessarily provide the level of detail necessary to identify less-extensive fires, but the effects of fire would be obvious in the photo-pairs. For instance, at one site—excluded from the final analysis—a fire occurred between the historical and modern photographs, resulting in a loss of tree cover in 73% of the grid cells and limiting tree cover increase to only 6% of the grid cells. Stand-age maps showed this site as having had no stand replacing fires in the past 200 years, yet evidence from the photo-pair was unequivocal. The thin bark of subalpine fir (the dominant treeline species in the photo-pairs) makes this species particularly vulnerable to high intensity fires (Alexander et al., 1990) and as a result, such fires leave considerable visual evidence in the landscape. Furthermore, while subalpine fir establishment does occur on burned areas, fire is not an essential part of the life cycle of this shade-tolerant tree (Alexander et al., 1990). Thus the observed establishment and infilling of subalpine fir does not require fire to create suitable habitat.

Shortcomings of the Applied Method

The method described in this paper improves our ability to detect landscape change using repeat photography, yet has limitations worth noting. The inability to determine scale within or among photo-pairs limits analysis to relative, rather than absolute, change and makes comparisons with other studies difficult. Within photo-pairs the continually varying scale between foreground and background areas requires that areas chosen for change analysis occupy a consistent distance from the photo-point. Relying on percentages to document change is also complicated by the size of the sampling area, as this will affect the values obtained. Similarly, if landscape features that will never be vegetated (e.g. lakes or cliff faces) are included in the study area, then the results will be biased toward a lack of landscape change.

Additionally, in an oblique view, prominent features may obscure parts of the landscape behind them. In cases where a tree branch covers bare ground it is necessary to decide whether to record the tree branch as vegetation or the ground behind it as bare ground. Thus, it is preferable for one person to conduct the analysis on all photo-pairs as was done in this study to maximize consistency. The scale of the grid cells also determines the scale at which questions can be asked and, therefore, the scale at which landscape change is evaluated. Requantification of vegetation changes in photo-pairs STI 932 and HDL 930b using a larger grid revealed that the overall trends in vegetation change and cover remain largely the same despite minor shifts in categories (Table 4). Larger grid cells result in a lower resolution and thus decrease the percent of “bare” and “completely” covered ground and increase the area of “partly” covered ground. Thus, grid cell size is an important consideration in designing a study of this type.

Table 4

Effect of grid cell size on results obtained from analysis of photo-pairs. Values are percent of grid cells for each photo-pair. Note that the trends in the type of vegetation change (light gray) are preserved regardless of cell size, while the percent of cells with partial vegetation cover (dark gray) increases with cell size, and the percent of cells with bare ground or complete vegetation cover (white) decreases with cell size.

i1523-0430-39-2-297-t04.gif

The best opportunity to resolve the inaccuracies and shortcomings of this technique lies in the integration of this method and those that convert oblique photography to a planar view by geo-referencing photographs to a digital elevation model (Aschenwald et al., 2001; Davis et al., 2002; Honda and Nagai, 2002; Corripio, 2004). Although it requires establishment of ground control points in the field, this approach would give spatial attributes to the photo-pairs and allow for the calculation of absolute change in the landscape. As a follow-up to the work described here, the authors successfully experimented with adapting the methodologies developed by Corripio (2004) to repeat photography.

Conclusion

Although the method developed for this project was applied to treeline in GNP, it could be utilized in mapping and measuring numerous other biotic and abiotic landscape changes. Despite its limitations, the quantification process presented in this paper allows for the extraction of considerable information from historical photographs in a medium easily understood by a wide audience and should be of interest to many researchers studying aspects of landscape change.

Acknowledgments

Field work was funded by a Mellon Foundation grant given through the Environmental Studies Program at Middlebury College. The archivists at Glacier National Park, particularly Somer Treat, provided excellent help in locating historical photographs. Blase Reardon assisted with field planning. The U.S. Geological Survey (USGS) Global Change Program and the USGS Glacier Field Station funded additional analysis. Middlebury College and its faculty, notably William Hegman and Dr. Andrea Lloyd, provided space, facilities, and much needed advice to the first author during development of the method and writing of the paper. The authors would also like to thank two anonymous reviewers for their detailed suggestions and comments.

References Cited

1.

R. R. Alexander, R. C. Shearer, and W. D. Shepperd . 1990. Subalpine fir. In R. M. Burns and B. H. Honkala , editors. eds. Silvics of North America Forest Service, U.S. Department of Agriculture, Agricultural Handbook 654. Google Scholar

2.

K. J. Alftine, G. P. Malanson, and D. B. Fagre . 2003. Feedback-driven response to multidecadal climatic variability at an alpine treeline. Physical Geography 24:520–535. Google Scholar

3.

S. F. Arno 2001. Community types and natural disturbance processes. In D. F. Tomback, S. F. Arno, and R. E. Keane , editors. eds. Whitebark pine communities Washington, D.C Island Press. 74–88. Google Scholar

4.

J. Aschenwald, K. Leichter, E. Tasser, and U. Tappeiner . 2001. Spatio-temporal landscape analysis in mountainous terrain by means of small format photography: a methodological approach. IEEE Transactions on Geoscience and Remote Sensing 39:4885–893. Google Scholar

5.

D. L. Baars, R. C. Buchanan, and J. R. Charlton . 1994. The canyon revisited: a rephotography of the Grand Canyon, 1923/1991 Salt Lake City University of Utah Press. 168. Google Scholar

6.

M. R. Becwar and M. J. Burke . 1982. Winter hardiness limitations and physiography of woody timberline flora. In P. H. Li and A. Sakai , editors. eds. Plant cold hardiness and freezing stress: mechanisms and crop implications New York Academic Press. Vol. 2. 307–323. Google Scholar

7.

D. R. Butler and L. M. DeChano . 2001. Environmental change in Glacier National Park, Montana: an assessment through repeat photography from fire lookouts. Physical Geography 22:291–304. Google Scholar

8.

D. R. Butler, G. P. Malanson, and D. M. Cairns . 1994. Stability of alpine treeline in Glacier National Park, Montana, U.S.A. Phytocoenologia 22:485–500. Google Scholar

9.

D. R. Butler, G. P. Malanson, M. F. Bekker, and L. M. Resler . 2003. Lithologic, structural, and geomorphic controls on ribbon forest patterns in a glaciated mountain environment. Geomorphology 55:203–217. Google Scholar

10.

D. M. Cairns 1990. Multi-scale analysis of soil nutrients at alpine treeline in Glacier National Park, Montana. Physical Geography 20:256–271. Google Scholar

11.

D. M. Cairns 2001. Patterns of winter desiccation in krummholz forms of Abies lasiocarpa at treeline sites in Glacier National Park, Montana, U.S.A. Geografiska Annaler 83A:157–168. Google Scholar

12.

D. M. Cairns and G. P. Malanson . 1997. Examination of the carbon balance hypothesis of alpine treeline location in Glacier National Park, Montana. Physical Geography 18:125–145. Google Scholar

13.

D. M. Cairns and G. P. Malanson . 1998. Environmental variables influencing the carbon balance at the alpine treeline: a modeling approach. Journal of Vegetation Science 9:679–692. Google Scholar

14.

J. G. Corripio 2004. Snow surface albedo estimation using terrestrial photography. International Journal of Remote Sensing 25:5705–5729. Google Scholar

15.

T. J. Davis, B. Klinkenberg, and P. C. Keller . 2002. Updating inventory using oblique videogrammetry and data fusion. Journal of Forestry 100:345–50. Google Scholar

16.

D. R. Easterling, T. R. Karl, E. H. Mason, P. Y. Hughes, D. P. Bowman, R. C. Daniels, and T. A. Boden , editors. 1996. United States historical climatology network (U.S. HCN) monthly temperature and precipitation data Oak Ridge, Tennessee Carbon Dioxide Information Analysis Center, Oak Ridge National Library. Google Scholar

17.

ESRI (Earth Sciences Research Institute) 2002. GIS and mapping software. Redlands, California. Google Scholar

18.

J. Fielder, W. H. Jackson, and E. Marston . 1999. Colorado 1870–2000 Englewood, Colorado Westcliffe Publishers. 224. Google Scholar

19.

J. Franklin, W. H. Moir, G. W. Douglas, and C. Wiberg . 1971. Invasion of subalpine meadows by trees in the Cascade Range, Washington and Oregon. Arctic and Alpine Research 3:215–224. Google Scholar

20.

G. E. Gruell 1983. Fire and vegetative trends in the Northern Rockies: interpretations from 1871–1982 photographs Ogden, Utah U.S. Department of Agriculture, Forest Service, Intermountain Forest and Range Experiment Station. Google Scholar

21.

A. Harper 1934. Glacial retreat. The New Zealand Alpine Journal 21:322–326. Google Scholar

22.

S. Hattenschwiler and W. K. Smith . 1999. Natural seedling occurrence in treeline conifers: a case study from the central Rocky Mountains, USA. Acta Oecologia 20:219–224. Google Scholar

23.

A. E. Hessl and W. L. Baker . 1997. Spruce and fir regeneration and climate in the forest-tundra ecotone of Rocky Mountain National Park, Colorado, U.S.A. Arctic and Alpine Research 29:173–183. Google Scholar

24.

A. Hofgaard, L. Kullman, and H. Alexandersson . 1991. Response of old-growth montane Picea abies (L.) karst forest to climatic variability in northern Sweden. New Phytologist 119:585–594. Google Scholar

25.

F. K. Holtmeier 2003. Mountain timberlines ecology, patchiness, and dynamics Dordrecht, Netherlands Kluwer Academic Publishers. 369. Google Scholar

26.

F. K. Holtmeier and G. Broll . 2005. Sensitivity and response of northern hemisphere altitudinal and polar treelines to environmental change at landscape and local scales. Global Ecology and Biogeography 14:395–410. Google Scholar

27.

K. Honda and M. Nagai . 2002. Real-time volcano activity mapping using ground-based digital imagery. ISPRS Journal of Photogrammetry and Remote Sensing 57:144–153. Google Scholar

28.

K. C. Kendall and R. E. Keane . 2001. Whitebark pine decline: infection, mortality, and population trends. In D. F. Tomback, S. F. Arno, and R. E. Keane , editors. eds. Whitebark pine communities: ecology and restoration Washington, DC Island Press. 221–240. Google Scholar

29.

F. L. Klasner and D. B. Fagre . 2002. A half century of change in alpine treeline patterns at Glacier National Park, Montana, U.S.A. Arctic and Alpine Research 34:49–56. Google Scholar

30.

M. Klett 2004. Third views, second sights: a rephotographic survey of the American West Santa Fe, NM Museum of New Mexico Press. 256. Google Scholar

31.

M. Klett, E. Manchester, J. Verburg, G. Bushaw, and R. Dingus . 1984. Second View: the Rephotographic Survey Project Albuquerque, NM University of New Mexico Press. 211. Google Scholar

32.

L. Kullman 1987. Tree-vigor monitoring by repeat photography in the forest alpine tundra ecotone. Ambio 16:160–162. Google Scholar

33.

L. Kullman 1991. Pattern and process of present tree-limits in the Tarna region, southern Swedish Lapland. Fennia 169:25–38. Google Scholar

34.

E. R. Kusnierczyk and G. J. Ettl . 2002. Growth response of Ponderosa pine (Pinus ponderosa) to climate in the eastern Cascades Mountains, Washington, U.S.A. Implications for climate change. Ecoscience 9:544–551. Google Scholar

35.

A. H. Lloyd, S. T. Rupp, C. L. Fastie, and A. M. Starfield . 2002. Patterns and dynamics of treeline advances on the Seward Peninsula, Alaska. Journal of Geophysical Research 108.D28161, doi:10.1029/2001JD000852. Google Scholar

36.

G. P. Malanson, D. G. Brown, D. R. Butler, D. M. Cairns, D. B. Fagre, and S. J. Walsh . in press. Ecotone dynamics: invasibility of alpine tundra by tree species from the subalpine forest. In D. R. Butler, G. P. Malanson, S. J. Walsh, and D. B. Fagre , editors. eds. The changing alpine treeline of Glacier National Park, Montana, USA Amsterdam, Netherlands Elsevier. Google Scholar

37.

J. S. Munroe 2003. Estimates of Little Ice Age climate inferred through historical rephotography, northern Uinta Mountains, U.S.A. Arctic, Antarctic, and Alpine Research 35:4489–498. Google Scholar

38.

R. Nicholas 2003. Create a grid polygon shapefile (FISHNET) ( http://arcscripts.esri.com/details.asp?dbid=128070). Last accessed 2007. Google Scholar

39.

L. M. Resler 2006. Geomorphic controls of spatial pattern and process at alpine treeline. The Professional Geographer 58:2124–138. Google Scholar

40.

J. M. Rhemtulla, R. J. Hall, E. S. Higgs, and S. E. Macdonald . 2002. Eighty years of change: vegetation in the montane ecoregion of Jasper National Park, Alberta, Canada. Canadian Journal of Forest Research 32:112010–2021. Google Scholar

41.

R. A. Rochefort, R. L. Little, A. Woodward, and D. L. Peterson . 1994. Changes in subalpine tree distribution in western North America: a review of climate and other factors. The Holocene 4:89–100. Google Scholar

42.

D. Rockwell 2002. Exploring Glacier National Park Guilford, CT Globe Pequot Press. 336. Google Scholar

43.

G. F. Rogers 1982. Then and now—A photographic history of vegetation change in the central Great Basin desert Salt Lake City University of Utah Press. 152. Google Scholar

44.

G. F. Rogers, H. E. Malde, and R. M. Turner . 1984. Bibliography of repeat photography for evaluating landscape change Salt Lake City University of Utah Press. 179. Google Scholar

45.

D. J. Selkowitz, D. B. Fagre, and B. A. Reardon . 2002. Interannual variations in snowpack in the Crown of the Continent Ecosystem. Hydrological Processes 16:3651–3665. Google Scholar

46.

R. J. Shaw and D. On . 1979. Plants of Waterton-Glacier National Parks and the Northern Rockies Missoula, Montana Mountain Press Publishing Company. 160. Google Scholar

47.

G. C. Stevens and J. F. Fox . 1991. The causes of treeline. Annual Revue of Ecological Systems 22:177–191. Google Scholar

48.

T. R. Vale and G. R. Vale . 1983. U.S. 40 today: thirty years of landscape change in America Madison University of Wisconsin Press. 198. Google Scholar

49.

T. R. Vale and G. R. Vale . 1994. Time and the Tuolumne landscape: continuity and change in the Yosemite high country Salt Lake City University of Utah Press. 212. Google Scholar

50.

J. B. Vogler 1998. An analysis of vegetation-environment relationships among vegetation communities of anthropogenically disturbed sites, Glacier National Park, Montana. MS thesis. University of North Carolina, Chapel Hill. 235. Google Scholar

51.

S. J. Walsh and N. M. Kelly . 1990. Treeline migration and terrain variability: integration of remote sensing digital enhancements and digital elevation models. Proceedings, Applied Geography Conference 13:24–32. Google Scholar

52.

E. Watson, G. T. Pederson, B. H. Luckman, and D. B. Fagre . in press. Glacier mass balance in the northern U.S. and Canadian Rockies: paleo-perspectives and 20th century change. In B. Orlove, E. Wiegandt, and B. Luckman , editors. eds. Darkening peaks Berkeley, CA University of California Press. Google Scholar

53.

R. H. Webb 1996. Grand Canyon a century of change: rephotography of the 1889–1890 Stanton Expedition Tucson, AZ University of Arizona Press. 290. Google Scholar
W. Roush, J. S. Munroe, and D. B. Fagre "Development of a Spatial Analysis Method Using Ground-Based Repeat Photography to Detect Changes in the Alpine Treeline Ecotone, Glacier National Park, Montana, U.S.A," Arctic, Antarctic, and Alpine Research 39(2), 297-308, (1 May 2007). https://doi.org/10.1657/1523-0430(2007)39[297:DOASAM]2.0.CO;2
Accepted: 1 October 2006; Published: 1 May 2007
Back to Top