A guide to grasslands of the mid south

Page 1

A G U I D E TO

GRASSLANDS OF THE MID-SOUTH A COOPERATIVE PROJECT OF: BOTANICAL RESEARCH INSTITUTE OF TEXAS AUSTIN PEAY STATE UNIVERSITY D. ESTES M. BROCK M. HOMOYA A. DATTILO S. FLEMING



PAGE 1 • A GUIDE TO GRASSLANDS OF THE MID-SOUTH

I N TRODUC TIO N Most people have heard the tale of vast forests that once blanketed Eastern North America, so dense a squirrel could travel through the treetops from the Atlantic Ocean to the Mississippi River without touching the ground. Accumulating evidence suggests such an enterprising squirrel must have taken a very circuitous route to accomplish such a feat, for the Mid-South was riddled with several million acres of naturally open grasslands at the time of European settlement. Broadly defined, grasslands include “any natural community or ecosystem in which the herbaceous layer is dominated by grasses, other graminoid (grass-like) plants, such as sedges and associated forbs (other herbaceous flowering plants)” (Noss 2012)*. This definition encompasses traditional concepts, such as prairies, barrens, savannas, and balds, but it can also include canebrakes, glades and open wetlands such as bogs, fens and meadows. Most of our grasslands have declined by greater than 90 percent of their original pre-settlement distribution. Today, barely discernible remnants dot the landscape, obscured by 230+ years of land use changes, fire suppression, succession to closed forests, infestation by invasive species and the loss of large grazers (bison) and browsers (elk). Most are a fraction of a hectare in size, tucked away in some lonely corner of a pasture, woodland edge or rural roadside. They are steadily slipping away, and with intensifying land-use threats, many will be lost forever. As they slip into oblivion, we will lose our ability to imagine these communities, making future conservation and restoration efforts difficult, if not impossible. If we are going to restore these grassland remnants, now is the time to act. If we wait 25 years, it will be too late. There is growing realization that grasslands hold the majority of Southeastern biodiversity and, as they have disappeared, the flora and fauna that call them home have also steadily disappeared. There is great need to conserve what is left and restore degraded remnants. We also must look to recreate if we want to prevent the extinction of grassland-dependent species. To do this, we must have a thorough understanding of what remains, where those remnants are found, what grows there and lives there. Until we have such a foundation, our conservation efforts will be encumbered by our lack of a deeper, broader understanding of what our grasslands were and are today. The intent in preparing this guide is to provide a very basic synthesis upon which further research and conservation can build. The specific objectives of this guide are to:

s PROVIDE AN UP TO DATE OVERVIEW OF THE GRASSLANDS OF THE -ID 3OUTH A REGION FOR THE PURPOSES OF THIS guide ) bound on the east by the Blue Ridge Mountains, the south and west by the Gulf Coastal Plain and Mississippi Embayment and the north by the glaciated plains of the Midwest.

s PRESENT DESCRIPTIONS FROM THE HISTORICAL RECORD THAT PINPOINT LOCATIONS OF SIGNIlCANT GRASSLANDS OBSERVED by early settlers and naturalists.

s PROVIDE THE lRST COMPREHENSIVE MAP SHOWING THE GENERAL DISTRIBUTION OF NEARLY ALL TYPES OF -ID 3OUTH grasslands

s PRODUCE A GUIDE THAT WILL BRING ATTENTION TO GRASSLANDS STIMULATE OTHERS TO lLL IN KNOWLEDGE GAPS AND AID efforts to restore grassland communities throughout the Mid-South.

s PROVIDE READERS WITH EXAMPLES OF SOME OF THE BEST REMAINING PUBLICLY ACCESSIBLE GRASSLAND REMNANTS SO they can visit, study and appreciate them. * Literature cited available as supplemental materials online at apsu.edu/herbarium/refrences


PAGE 2• A GUIDE TO GRASSLANDS OF THE MID-SOUTH

HOW DO WE KNOW WHERE GRASSLANDS WERE HISTORICALLY?

Reconstructing the historical distribution of MidSouth grasslands requires detective work—looking for existing remnants; sifting through early historical documents, maps, and land surveys; consulting studies of fire history, and determining the present distribution of plant and animal species that require open lands.

PHOTO CREDIT: D. ESTES

Existing Grassland Remnants Studying existing remnants helps us understand grasslands, but it is difficult to know how similar these remnants are to historic examples. They are leftovers from a bygone era, often just a fraction of a hectare in size. Some regions, such as the Cumberland Plateau and Highland Rim, contain many tiny remnants in powerlines and along rural roadsides. In other cases, such as the Catoosa Wildlife Management Area (WMA) near Crossville, Tennessee, remnants have reemerged after timber harvests released grasses and wildflowers from the seedbank or as plants re-sprouted from long-suppressed rootstocks. The absence of remnants in other areas cannot be taken as proof they never existed. There are areas we know once supported vast grasslands, but no remnants remain due to complete elimination of original habitat and deplettion of seedbanks. This is especially true in areas such as the Jackson Purchase of western Kentucky and Tennessee, a region dominated by agriculture, where almost all ground not amenable to farming is maintained by mowing and herbicide.

“The top of the mountain is...a vast upland prairie, covered with a most luxuriant growth of native grasses, pastured over as far as the eye could see, with numerous herds of deer, elk and buffalo, gamboling in playful security over these secluded plains….” (Ramsey 1853)


PAGE 3• A GUIDE TO GRASSLANDS OF THE MID-SOUTH

Historical Descriptions Unlike the Great Plains, the early grasslands of the Mid-South were neither photographed nor painted, and one has to rely on obscure references to locate former grasslands. Except for a few eloquent descriptions from select areas, most references to grasslands are frustratingly brief; sometimes a single phrase or word is all that is available among hundreds of pages of historical literature. Many areas (e.g., Nashville Basin) were not described in their pre-settlement condition. Compared to other parts of the East, there was a dearth of early naturalists who ventured into the Mid-South. Those who did, such as André and Francois Michaux, tended to take well-established travel routes and didn’t explore off the beaten path. Many who went off-trail, such as the longhunters (late 18th century hunters from NC and VA who went on “long” hunting expeditions to KY and TN), were illiterate, their observations mostly going undocumented. By the time trained naturalists such as Augustine Gattinger (1825-1903), Charles T. Mohr (1824-1901), James Safford (1822-1907), and Joseph B. Killebrew (1831-1906) began working, overgrazing, fire suppression, agriculture, and 20–50 years of forest succession had altered the landscape. Many of the best descriptions for each of the major regions of the Mid-South are included in the six maps accompanying this guide. Maps and Place names Early maps are key to reconstructing the former extent of Mid-South grasslands. TheFrench were the first to indicate extensive grasslands in this area. One French map from 1720 provides the inscription “Savana Land and Good Pasture Ground” over what is now central Kentucky and Tennessee. Grasslands appear on several early maps including Filson’s (1794), Munsell’s (1818), and Loughridge and Hoeing’s (1886) maps of Kentucky and Tanner’s (1795) map of Tennessee. Some of the largest grasslands were curiously not mapped, either because they were unknown or because cartographers chose not to exhaustively map all examples. Place names also lend important clues to historical vegetation. Many landmarks were named early in the region’s history, but their exact date of origin is often not known. Names such as Barren Plains, Barren Fork, Prairie Creek, Price’s Meadow, Pleasant View, Oak Grove, Strawberry Plains, Glade Branch, Pine Grove, Hazel Green, Meadow Creek, Crab Orchard, and Barren Fork refer to places we know were once open grasslands or closely connected to them. Munsell’s 1818 Map of Kentucky

Survey Records Some parts of the eastern U.S., such as Mississippi, Arkansas, Indiana, and Ohio, were mapped by Government Land Office (GLO) surveyors who divided the landscape into square-mile blocks of townships, ranges, and sections. Each of these areas was thoroughly surveyed, often with notes on vegetation. GLO survey records have been used to reconstruct the historic distribution of grasslands in the Black Belt of Mississippi and Alabama (Barone 2005). In Tennessee and Kentucky, surveyors used the “meets and bounds” system where individual land tracts were mapped, and the types of trees at property boundary corners were recorded and marked. References to marker or “witness” trees could be specific (e.g., “white walnut” for the single species Juglans cinerea) but more frequently were general (e.g. “hickory” which could be any one of several species in the region). Certain trees, such as post oak (Quercus stellata) and blackjack oak (Q. marilandica), were often indicative of savanna, grassy woodland, or prairie-like landscapes. When trees were absent, surveyors would use a wooden stake to mark the boundary. The relative abundance of stakes used in an area provides clues to the openness of the landscape. In areas too rocky to place a stake, such as the Nashville Basin, surveyors sometimes created rock piles as markers. Some of these piles still exist (Jack Masters, pers. comm.). In rare cases surveyors actually referred directly to grasslands, such as “stake in the barrens” or “hazel bush in the barrens.” Surveys of political boundaries also yield important information. In 1818, the State of Georgia commissioned surveys of the Georgia-Tennessee state line where trees were recorded at every mile (Camak 1818 in Coulter. 1951). In the Ridge and Valley section, surveyors recorded mostly pine (presumably shortleaf ) and post oak except for one notable gap of three miles wide just east of White Oak Mountain southeast of Chattanooga where no trees were recorded. These data suggest this part of the Ridge and Valley was likely open savanna at the time. Settlement Patterns A highly overlooked clue to the former distribution of grasslands is the settlement pattern in a given region. In reference to West Tennessee, Claybrooke (in Killebrew 1879), noted that thinly timbered lands, or lands barren of timber, were quickly settled by “the early emigration... because they quickly opened a farm that produced good cotton and corn….” Goodspeed (1887) wrote that many early settlers were “incited by the rich fields in West Tennessee.” But not all open lands were attractive to early settlers; the Pennyroyal Plain was settled later than other grasslands because its karst geology allowed for too few suitable water sources. The recent work by Drake et al. (2009) shows how there were almost no Revolutionary War land grants in the open grasslands of the Pennyroyal of Tennessee but they were instead concentrated along nearly all of the region’s stream valleys. More work is needed to evaluate settlement patterns and early land survey records available in state archives.


PAGE 4• A GUIDE TO GRASSLANDS OF THE MID-SOUTH

Landscape Clues The landscape of a region can provide clues to the historical vegetation. With the exception of floodplain forests, there were few flat landscapes in the Mid-South that were densely forested at the time of European settlement. Flat to rolling or slightly hilly landscapes tend to support prairie and savanna, respectively, because fire can move better than on more dissected terrains, which support forest and woodland under natural fire regimes where ravines and creek valleys serve as natural fire breaks. Plains and plateau surfaces with high fire frequency have large fire compartments (areas bounded by natural fire breaks). According to Dr. Cecil Frost (Univ. of North Carolina, Chapel Hill), fire compartments in the Pennyroyal Plain may have exceeded 940 km2 (362 mi2). These fires probably burned onto adjacent escarpment slopes such as the Dripping Springs Escarpment (Mammoth Cave Hills), an escarpment that today still supports numerous relict grassland species. In mountains where lightning may be frequent during thunderstorms, fires often burn relatively small compartments limited to exposed ridgelines and dry slopes. Dendroecological Studies Grasslands are fire-dependent communities. In general they require a fire every 1-3 (-5) years to suppress tree growth. Areas with longer fire-return intervals tend to favor woodlands and forests. Many once extensive grasslands have succeeded into forests with 100+ years of fire suppression. Dendroecology (tree-ring) studies are valuable to reconstructing changes in vegetation and fire regimes of an area from pre-settlement times to the present. Researchers from the Missouri Tree-Ring Laboratory, University of Missouri, have studied trees and fire scars from over 80 sites across the eastern U.S. in an effort to document and understand the historical ecology of grasslands and forests. For example, at Arnold Airforce Base, Coffee Co., Tennessee, they documented predominantly dormant-season fires (Sept-April) burned across the landscape on average once every 3.3 to 5.2 years in pre- and early-settlement times (Stambaugh et al. 2016) and that annual fires occurred during some periods. These short fire-return intervals corroborate other lines of evidence that suggest this area was a mosaic of open savanna and prairie prior to widespread fire suppression. Archaeological Evidence Keith (1980), citing the work of Dr. Patrick Munson (Indiana University), suggested that Paleo-Indian Clovis spear points have primarily been found concentrated in certain areas of the Southeast and Midwest, such as the Nashville Basin, the Pennyroyal Plain and the Mitchell Plain of Indiana. Since the Clovis hunters are thought to have primarily hunted in open lands, it suggests that areas with high concentrations of spear points may have been open landscapes 10,500-12,500 years ago. Specimens of Grassland Conservative Species One way that ecologists infer the original distribution of grasslands is to consult specimens found in museums and herbaria. For more than 200 years, naturalists have been collecting specimens and preserving them. These records are important for establishing the distribution and habitat preferences of species. At the New York Botanical Garden, New York City, one can examine the original specimen of American chaffseed (Schwalbea americana) collected by Ferdinand Rugel somewhere east of Jamestown, Tennessee in June 1842 where he described his collection in Latin as “In locis arenosis humidis regionis super. montium Cumberland inter Montgomery et Jamestown, Tennessee.” This translates to “In moist sandy places, over the country, Cumberland Mountains, between Montgomery [currently Wartburg, Morgan Co.] and Jamestown [Fentress Co].” This federally-endangered species has not been seen on the Tennessee Cumberland Plateau in 174 years. Without this specimen we would never know that this strictly fire-dependent pine savanna species ever occurred in the region. Less than 30 years ago there were also populations of the now federally-endangered red-cockaded woodpecker in the Cumberland Plateau of Tennessee and Kentucky. This species requires open, old-aged pine savannas. The last population on the Tennessee portion of the Plateau disappeared from Catoosa WMA in Cumberland Co. in 1984 (Clarence Coffey, Shortleaf Pine Initiave, pers. comm.). The presence of these two fire-dependent savanna species make it clear that parts of the Plateau were open, fire-maintained, pine savanna historically. The American chaffseed is an example of an ecologically conservative species. Conservative plants are those that require certain natural conditions to thrive, e.g., fire, flooding, deep shade, certain moisture conditions, or particular soil types. Some species cannot tolerate human disturbance, and their populations will decline if certain conditions aren’t met. Ecologists use the Floristic Quality Index (FQI) to assess site quality and the degree of conservativeness of the flora. Species are assigned a score of 0–10, where those with a 10 are extremely picky and are only found in sites that simulate high-quality natural conditions like those of pre-settlement times. Common and widespread species that are often dominant in some communities, such as white oak, would typically receive a 4–7. Weedy or non-native species receive low scores. When high numbers of conservative species disappear from a particular habitat type across a broad region this can be a sign of ecological collapse and should signal a need for urgent conservation measures to prevent further loss and/or restore degraded habitat. In the last 25 years, nearly half of the 40 rare savanna-obligate plant species known from the Tennessee Cumberland Plateau have become extirpated whereas several other species are on the brink.


PAGE 5 • A GUIDE TO GRASSLANDS OF THE MID-SOUTH

Heliophiles Heliophiles include species such as pine snake, Henslow’s sparrow, spiked blazingstar, and Eggert’s sunflower. These organisms are “sun lovers” and need naturally open conditions to survive. Plants requiring high light levels that become shaded by forest canopies will often fail to flower. When we find concentrations of highly conservative species that require naturally open conditions, such as goat’s rue (Tephrosia virginiana) or little bluestem (Schizachyrium scoparium) in the understory of a close-canopied forest, it is usually a strong indication that the site was historically open and that those species have persisted as the site closed-in over time. Of course there are many areas with high disturbance or lengthy forestation periods where the conservative light-loving species have disappeared completely. Grazers and Browsers Soon after the first European hunters began to hunt the Mid-South, buffalo and elk disappeared. Domesticated stock (cattle, sheep, and horses) brought by settlers took over the role of grazing. The grasslands proved to be very important for providing pasturage, but unfortunately many were overgrazed. Killebrew and Safford (1874) attest to the importance of these grasslands to Tennessee’s livestock industry in the 19th century, as evidenced by this quote pertaining to the Cumberland Plateau in Van Buren Co.:

“…for pasturage and meadows, [they] are scarcely surpassed. There is a course nutritious grass, well known in this part of the State as ‘mountain grass,’ which is indigenous to the soil, and afford rich and abundant pasturage to hundreds of cattle, sheep, and other animals.” Grasslands and Corridors With the loss of >90 percent of most grassland systems in the Mid-South, the majority of grassland-dependent plant and animal species are barely surviving, and many populations have been lost. Some persist in human-maintained openings such as powerline corridors, roadsides, and field edges where they survive because periodic mowing prevents woody plants from encroaching. As maintenance ceases, these corridors succeed to thickets, and the grassland flora disappears. Mowing mimics the fire and grazing that kept these sites open in pre-settlement times. However, increased herbicide use to maintain these corridors has destroyed many remnants in the past 20 years. The “Puryear Barren” in Henry Co., Tennessee, is one of the last known prairie remnants in west Tennessee and is a vestige of the former Jackson Purchase prairies that covered hundreds of thousands of square kilometers of western Kentucky and Tennessee prior to 1830. When visited by University of Tennessee ecologist Dr. Hal DeSelm in 1983 the site was an open prairie strip sandwiched between US Hwy 641 and the adjacent railroad. The cessation of mowing along this highway led to the rapid growth of trees and shrubs, which now form a dense thicket that has displaced the prairie. Without urgent action this remnant and its seed bank will disappear forever. The Tennessee Valley Authority (TVA), established in 1933 to address issues in energy, economic development, and environmental stewardship, is committed to clean air and a clean water supply for the Mid-South region, as well as protection of historical, cultural, and environmental resources across a seven state region. Significant grassland remnants occur on both TVA properties and transmission line rights-of-way (ROW). Grasslands on TVA property include Silurian barrens in west Tennessee, sandstone glades in northeastern Alabama, and limestone barrens in the Ridge and Valley of east Tennessee. Projects focused on reintroduction of fire and reduction of woody encroachment in these significant habitats are currently in the planning stages.


PAGE 6 • A GUIDE TO GRASSLANDS OF THE MID-SOUTH

In ROW situations, it is often the case that the only habitat for state and federally listed species occurs within the managed area under the transmission line and not in the surrounding non-ROW vegetation, which is usually forest. Imperiled plants that utilize ROW as primary habitat include the federally listed leafy prairie-clover (Dalea foliosa), fleshy-fruit gladecress (Leavenworthia crassa), Mohr’s Barabara’s buttons (Marshallia mohrii), white fringeless orchid (Platanthera integrilabia), and green pitcher plant (Sarracenia oreophila) as well as a host of other globally rare species indicative of open habitats. In all, over 300 occurrences of rare plant taxa have been identified on TVA ROW. Even in areas where there are no listed plant species, selective application of herbicide to woody plants as part of regular ROW maintenance can promote a rich herbaceous flora found almost nowhere else on the modern landscape. While conscientious ROW maintenance alone cannot always provide sufficient disturbance to support rich native plant habitats in all situations, ROW often support high quality habitat. The first highways were grasslands Long before wagons and automobiles, a succession of large mammals, from mammoths and rhinoceroses to elk and bison, roamed the Mid-South, moving from one large grazing area to another along networks of crisscrossing paths called traces. Later in the nation’s early history, many of the Mid-South’s grasslands served as major thoroughfares for westward expansion. The first longhunters, following some of these traces, entered the Mid-South via the grassy valleys of southwest Virginia and northeast Tennessee. After passing west through the forested mountains of the Cumberland region, they again found open grasslands in central Kentucky and Tennessee. Haywood (1823) described the first visit by the longhunters in 1779 to the upper Eastern Highland Rim:

“...came to an open country called barrens, to a place since called Price’s Meadow, in what is now called Wayne County [Kentucky]” and they hunted their way south to the Caney Fork River near modern-day McMinnville, Tennessee. “All the country through which these hunters passed was covered with high grass, which seemed inexhaustible….” Roadsides are now exceedingly important for grassland conservation because so many remnants are now sandwiched between roads an adjacent lands, often occupying thin strips less than 3 m (10 ft) or very rarely 9-15 m (30–50 ft) wide. These roadside grasslands are under constant threat from herbicide application, too frequent mowing, woody plant succession, and competition from invasive species. We need to identify ways to protect these remnants because they are all that is left in some regions. Fire in Mid-South Grasslands Most Mid-South grasslands need fire in some capacity, the major exception being flood-maintained riverscour grasslands. Prairies and savannas require frequent fire (1–3 year return) to retard tree growth and prevent succession to woodlands and forests. Glades and barrens are better able to withstand invasion by trees due to their shallow, drought-prone, rocky soils, but fire is needed to keep their margins open and grassy. Even some types of wetlands in grassland-dominated regions burned frequently. For example, Stambaugh et al. (2016) found that some swamps at Arnold Airforce Base, Coffee Co., Tennessee, burned yearly during certain periods. In the early literature there are numerous references to the importance of fire in maintaining the natural open landscapes of the Mid-South. One of the best is a description by Ross (ca. 1812) from the Pennyroyal Plain of northern Montgomery Co., Tennessee: According to several early written accounts from the turn of the 19th century, Indians were responsible for starting many of the fires that burned the Mid-South landscape. However, Frost (2014) studied the fire ecology of Mammoth Cave National Park, Kentucky, and calculated that, with an average of 4–8 strikes per km2/yr and an average of 478 strikes needed to produce one ignition, 8–15 ignitions would be expected per km2/yr. In Mammoth Cave, a fire compartment of 940 km2 would experience 23–47 ignitions per year, which allows for an average fire return interval of 1–3 years. Even in the absence of Indian burning this would be sufficient to maintain grasslands in the open, rolling Pennyroyal Plain. While there is no question that Indians burned the landscape, it also should not be assumed that they created grasslands. As emphasized by Noss (2012) in his Forgotten Grasslands of the South, Southern grasslands support thousands of species of plants and animals adapted to frequent fires that have been burning for millions of years, well before the first humans appeared in the Southeast less than 15,000 years ago.


PAGE 7• A GUIDE TO GRASSLANDS OF THE MID-SOUTH

“During this winter I first saw the tremendous fires caused by the burning of the dry grass. In many places, this grass was very thick and tall; and when perfectly dry, should it get on fire, the wind being high, the spectacle became truly sublime, especially at night. The flames ... would sometimes burn the leaves on trees twenty or thirty feet in height. No one who ever witnessed one of these great fires would ever afterward be at a loss to account for the scarcity of timber in the Barrens, as trees of all kinds, when small, were destroyed by them. Should a little twig or bush put up from the ground one season, it was sure to be burned the next. The Indians, in early times, used to set this grass on fire, when hunting, and killed great quantities of game as it fled before the flames.” (Ross undated, ca. 1812)

Photo Credit: Harold Gaston


PAGE 8• A GUIDE TO GRASSLANDS OF THE MID-SOUTH

TYPES OF MID-SOUTH GRASSLANDS There is a strong need for the adoption of a standardized classification of grasslands to facilitate effective communication, which itself is critical to successful conservation. It isn’t realistic to expect all people to agree to a standard classification, especially across broad geographic areas, but an attempt is made here to define and classify grasslands of the Mid-South in hopes that it will facilitate future conservation. Only time will tell whether the community is willing to adopt these concepts. In their typical form most of our grassland types are quite distinct, but historically they occurred in mosaics that were connected via transitional gradients and blurred ecotones. For example, the term “barren” means different things to different people. English-speaking settlers used the term barren to refer to naturally open, grassy, treeless areas or places with stunted or widely spaced trees with grassy understory, based on their assumption that they were “barren” of trees or failed to support forests due to infertile soils. DeSelm (1994) considered the barrens to be formerly extensive openings similar to Midwestern prairie, but noted “…the term has been used to include scrub forest, woodland and savanna with grass-dominated openings, and low density forest with grass understory” (DeSelm 1981). DeSelm (1988, 1990, 1992, 1993, 1994) used the term broadly to embrace deep-soil extensive grasslands, open grassy woodlands, rocky calcareous seepage fens, bedrock glades and rocky grassy slopes, and riparian cobble bars. The usage of “barrens” as an umbrella-term, which has been the norm for the past 30 years, minimizes the striking ecological differences among the myriad types that existed on the historic landscape. In this guide, a deliberate attempt is made to depart from the confusing historical usage of “barrens” as a catchall for all grassland types of the Mid-South. Instead the term is used in a restricted sense to denote rocky slope grasslands that are indeed “barren” due to infertile or shallow soils. “Grassland” is here used as the umbrella-term for all types. Defining characteristics are given below for the 10 major kinds of grasslands recognized in this guide.

Photo Credit: D. Estes

Prairies: (1) historically vast matrix communities ranging from dozens to thousands of square kilometers (with local patches smaller); (2) dominated by mid- to tall-statured (0.5–3 m, 1.6-10 ft) native perennial grasses and herbs; (3) trees and shrubs historically absent or sparse, often restricted to waterways or drainages or otherwise scattered and patchily distributed covering less than 10% of the landscape; (4) with deep, well-drained to hydric soils; (5) maintained historically by a combination of fire, drought, grazing by megafauna, and in some by the interaction of hydrology and a restrictive clay fragipan or shrink-swell clay soils; and (6) restricted to level to rolling landforms such as plains, plateaus, basins, and wide valleys. Photo Credit: A. Cressler


PAGE 9 • A GUIDE TO GRASSLANDS OF THE MID-SOUTH

Savannas: (1) historically large-patch or matrix communities ranging from one to dozens of square kilometers (with local patches smaller); (2) dominated by two vegetation layers consisting of a sparse tree layer with 10-30% canopy coverage and a dense grass/herb layer with scattered clumps of low shrubs (0.5–3 m, 1.6-10 ft); (4) with deep, well-drained to hydric soils or a clay fragipan; (5) maintained historically by fire; and (6) commonly associated with rolling to slightly hilly landforms of plains, plateau surfaces, broad ridges, foothills, basins, and wide valleys. Barrens: (1) small-patch communities typically less than 0.8 ha (2 ac) with rare examples larger; (2) dominated by short- to mid-statured (0.1–2 m, 0.3-6.5 ft) perennial grasses and forbs; (3) often with imbedded groves of tree/shrub islands; (4) shallow drought-prone rocky soils <25 cm (10 in) deep often with scattered outcroppings of bedrock; (5) association with gently to steeply sloping eroded landforms; (6) maintained historically by a combination of edaphic factors (drought, shrink-swell clay soils) and fire. Glades: (1) small patch communities mostly less than 0.8 ha (2 ac); (2) dominated by low- to mid-statured annual grasses, herbs, deep-rooted perennials, succulents, bryophytes, lichens, and cyanobacterial crusts (0.1–1 m, 0.3-3.2 ft); (3) bedrock exposed at or just below surface; (4) soils absent or thin, forming soil pockets or a thin veneer 5–12 cm (2-5 in) deep; (5) hydroxeric (saturated winter to spring and often with formation of pools, xeric summer to fall and desert-like); (6) on flat to slightly sloping sites of plateau surfaces, basins, wide valleys, and benches; (7) occur today as “islands” among juniper-oak-hickory-ash (limestone glades) or pine-oak-hickory (sandstone glades) woodlands but historically likely occurred in open savanna-woodland matrix; (8) the open glade proper maintained by intrinsic edaphic factors (margins fire-maintained). Balds: (1) small patch communities ranging from 2-100 ha (5-247 ac); (2) dominated by low- to mid-statured perennial grasses, sedges, and forbs (0.1–1 m, 0.3-3.2 ft), (3) periphery characterized by scattered dense shrub islands and thickets (0.5–3 m, 1.6-10 ft) between the open bald and nearby spruce-fir or northern hardwoods forests; soils moist, often with scattered outcrops and seeps; (4) located at high-elevations mostly above 1,676 m (5,500 ft) on ridgelines, summits, sometimes extending into high-elevation gaps; (5) relicts of alpine tundra that extended into the high-elevation Blue Ridge Mountains during the Pleistocene Ice Age; (6) maintained in the last 18,000 years since the close of the last Ice Age by a succession of megafauna and after European settlement by domesticated stock (Weigl and Knowles 2014). Riverscours: (1) linear small patch communities mostly less than 0.8 ha (2 ac); (2) dominated by low- to mid-statured perennial grasses and forbs, shrubs and small saplings (0.5–3 m, 1.6-10 ft) represented by a mix of upland, wetland, and riparian species; (3) substrate formed of unconsolidated cobbles or boulders with sandy interstices or of exposed bedrock; (4) soils mostly sand, often limited to interstices or forming deep accumulations near far edge of flood zone; (5) hydroxeric (saturated winter to spring and after rain events, xeric summer to fall); (6) located on perched alluvial bars in entrenched river gorges mostly along high-gradient streams. Meadows: (1) small patch or linear communities, original sizes uncertain but likely several to dozens of hectares; (2) dominated by mid- to tall-statured forbs, grasses, sedges, and shrub thickets 0.5–2 m tall (2-6.5 ft); (3) associated with floodplains of gently meandering small- to mid-sized streams of narrow valleys; (4) soils deep, consisting of gravelly or silty alluvium; (5) hydrology controlled by high water table from associated stream, groundwater seepage and surface runoff from adjacent slope bases, and periodic short-duration floods following flash flood events; (6) historically probably developed as part of the mosaic of beaver-created habitats that were likely common prior to French and Indian fur-trapping in the early 1700s and/or from Native American burning in valleys.

Photo Credit: Edward W. Chester


PAGE 10• A GUIDE TO GRASSLANDS OF THE MID-SOUTH

Photo Credit: J. Tuttle

Photo Credit: A. Cressler

Fens: (1) small patch communities mostly less than 0.1 ha (0.25 ac) but up to 0.8 (2 ac); (2) dominated by sedges, grasses, rushes, forbs, and shrubs, with Sphagnum absent to rare; vegetation <0.1–3 m tall (<0.3-10 ft); (3) soils varying from shallow gravels over bedrock to deep organic muck, calcareous to circumneutral, waterlogged; (4) underlying bedrock limestone or dolomite or mafic metamorphic rock (amphilbolite and serpentine); (5) associated with flat to gently or moderately sloping landforms of wide valley bottoms, narrow stream valleys, basins, or hillslopes; (6) occur as “islands” embedded in woodland matrix when on hillslopes or may have occurred as “islands” among grassland, savanna, woodland matrices in flatter terrains; (7) maintained historically by hydrology, slope creep (in shallow-soiled hillslope examples), and fire (likely important for margins). Montane Bogs: (1) small patch communities, historically from less than 0.5-20 ha (1–50 ac) though much larger examples may have occurred; (2) dominated by Sphagnum, sedges, forbs, and shrubs 0.1–3 m tall (0.3-10 ft); (3) soils acidic, permanently waterlogged peats to 2+ m (6.5 ft) deep; (4) often with small, shallow, braided rivulets; (5) associated with streamhead seepage swales of plateau surfaces, shallow ravine bottoms, or in bottom of montane basins and coves; (6) often occur today as open “islands” among forests or woodlands but historically may have occurred among matrix of open woodlands or savannas in some areas; (7) maintained historically by a combination of hydrology, fire, and probably beaver. Depression Ponds: (1) small patch communities <1–15 ha (<1-37.5 ac); (2) dominated by submerged and floating aquatics, emergent grasses, sedges, rushes, and forbs, and often with emergent shrubs, 0–2 m tall (0-6.5 ft); (3) associated mostly with limestone karst plains; (6) often occur today as open “islands” among forests or imbedded among agricultural fields or pastures but historically occurred among prairies, savannas, and flatwoods; (7) maintained historically by hydrology but fire important for keeping margins open.

Black Belt Prairie, c. 1775, Montgomery Co., Alabama, 2009, oil on canvas by Philip Juras


PAGE 11 • A GUIDE TO GRASSLANDS OF THE MID-SOUTH

PRAIRIES The term prairie (Latin pratarium = meadow-land) was first used by 17th-18th century French explorers in reference to areas in the Midwest characterized by vast, flat to rolling deep-soiled landscapes dominated by tall grasses with very few to no trees (Gleason 1909). In the Southeast, English-speaking settlers referred to such sites as barrens or meadows (DeSelm 1994). Note that some areas that bear the name of prairie, such as the Coosa Prairie of Georgia, are treated in this guide as a type of savanna. Major prairie systems of the Mid-South are discussed below. Pennyroyal Plain Prairie—The Pennyroyal or “Big Barrens” was the largest tallgrass prairie in the Southeast at 6,070 km2 (1.5 million ac) though some report as large as 1.5 million ha (3.7 million ac). It extended in a crescent from near Paducah, Kentucky, dipping southeast into northern Tennessee and then northward through Kentucky into south-central Indiana (Homoya 1994, Keith 1980). Early explorers such as Daniel Boone, early naturalists such as Andre Michaux, and Louis-Phillipe I, later the King of France (1830–1848), described this once majestic landscape. These prairies are associated with three karst plains: the Pennyroyal, Elizabethtown, and Mitchell plains. Early on, the area was avoided for settlement due to the scarcity of surface streams, but in the early 19th century the true fertility of the soils was realized, and the region blossomed into one of the most productive agricultural districts in the Mid-South. Fire suppression resulted in the swift loss of prairie, with numerous writers documenting the succession to oak-hickory woodlands from the 1830s–70s. Today’s best remnants are found at Fort Campbell Army Base where nearly 100 km2 (25,000 ac) exist—of which 80 km2 (20,000 ac) occur in a contiguous block in the Base’s impact zone; an area deemed off-limits by the Army due to unexploded ordinances. Jackson Plain Prairie—The loess plain of western Kentucky and Tennessee (the Jackson Purchase) supported ca. 4,050 km2 (1 million ac) of prairie at settlement. These prairies were among the first areas settled when the area opened in 1819. Unfortunately settlement occurred at least 30 years before the first trained naturalists explored the region. By the time these naturalists visited the area most prairies had become farms or succeeded into oak-hickory woodlands. Today, 99.99% of this system has been lost, and less than 40 ha (100 ac) are known to exist. The largest remnant is at Western Kentucky Wildlife Management Area near Paducah, Kentucky. Only two sites are known in Tennessee (Henry Co.) totaling less than 2 ha (5 ac). Black Belt Prairie—The crescent-shaped Black Belt Prairie extends from McNairy Co., Tennessee, south and east through Mississippi into south Alabama before terminating east of Montgomery. Barone (2005) mapped the prairies of Mississippi and Alabama. These prairies occupied undulating landscapes “full of gentle ascents” (Nairne 1988). The namesake fertile black soils were largely lost due to cotton farming in the mid-late 1800s. There are numerous written accounts of this prairie, which once covered 140,000 ha (>345,000 ac), the first dating to Hernando de Soto’s 1541 expedition. Barone (2005) provides extensive documentation of the prairies using many historical quotes from the 1700-1800s. Many prairies were “largely treeless, though surrounded by forest and often with small woodlands within them, usually on hilltops or ridges”(Barone 2005). The Black Belt was not a single large open prairie but rather was “a dense archipelago of prairie ‘islands’ imbedded in a matrix of woodland and forest (Barone 2005). Where the Black Belt tapers to a sliver in McNairy Co., Tennessee the prairies were probably very small and more savanna-like. Perkins (in Wright 1882) wrote that in 1825 the Saunders family found “A virgin soil of great fertility, landscapes of marvelous beauty made up of green savannas...” Today many remnants exist in powerlines and along roadsides in Alabama and Mississippi only. Important remnants in Mississippi are the 16th Section Prairie, Oktibbeha Co. and remnants in the Tombigbee National Forest, Chickasaw Co., Mississippi and the Old Cahawba Preserve in Dallas Co., Alabama. The single Tennessee remnant is limited to a few prairie species on a few square meters of a rural roadside banks, but nearby calcareous woodlands may be suitable for restoration with canopy thinning and fire. Highland Rim Prairie—The Eastern (EHR) and Western Highland Rims (WHR) of Tennessee, Alabama, and Kentucky once supported an estimated 4,050 km2 (1 million ac) of prairie. Much more is known about EHR prairies than those of WHR. By the time the WHR was visited by naturalists, it had already undergone decades of fire suppression, agriculture, and succession to woodlands. Early descriptions of these grasslands are very rare. Several EHR prairie remnants remain in Tennessee (Arnold Airforce Base and May Prairie State Natural Area, Coffee Co., TN) but only two or three are known in Kentucky (Hazeldell Meadow, Pulaski Co., KY). No remnants are documented in Alabama though they likely exist along rural roadsides and in irregularly mowed fields.


PAGE 12 • A GUIDE TO GRASSLANDS OF THE MID-SOUTH

S AVA N N A S

Photo Credit: C. Coffey

The first use of the term savanna for the Mid-South was by the Hernando de Soto expedition in 1540–41. They noted savannas in southeast Tennessee’s Ridge and Valley and the Black Belt region of northern Mississippi. Their usage of the term encompassed both treeless grasslands (prairie) and grasslands with scattered trees (savanna). Within the Mid-South, savannas occurred historically in rolling to gently hilly landscapes and wide basins and valleys. In some regions, such as on the Cumberland Plateau, savannas were the dominant grassland type. In other areas, like the Pennyroyal Plain prairies of Kentucky, there were several miles of savanna that separated the open treeless prairies (“the Barrens”) from the woodlands and forests of hilly, dissected landscapes. Within the Mid-South there were extensive areas of both oak and pine savannas, with the numerous types discussed below. East Gulf Coastal Plain—In 1795, Tanner provided the note “Light Soil Long Grass Little Timber Broken Ground on the Heads of these Rivers” for the upland hilly region of west Tennessee that forms the headwaters for the Obion, Forked Deer, and Hatchie rivers. These thinly timbered lands were settled quickly when west Tennessee was purchased from the Chickasaw Indians in 1819. These savannas likely graded into treeless prairies and occurred across multiple ecoregions. In the more calcareous regions such as the Black Belt and Flatwoods ecoregions, oak savannas would have predominated, but pines were likely more important in acidic, sandy ecoregions (Northern Hilly Gulf Coastal Plain, Pontotoc Ridge, Fall Line Hills, and Transition Hills), especially southward. University of Tennessee botanist Dr. A.J. Sharp collected several specimens from an “oak savanna northwest of Ramer” in McNairy Co., Tennessee, in the late 1940s, but that site has never been relocated (DeSelm 1989). No intact savanna remnants exist in this ecoregion but grassland vegetation is found on roadsides and in powerline corridors in places such as Chickasaw State Forest, Chester Co., Tennessee and Big Hill Pond State Park (McNairy Co., Tennessee). Pennyroyal Plain Savannas—Several early settlers and naturalists described the vast and nearly treeless Pennyroyal Prairies, but the bordering savannas received little attention. Francois Michaux (1805) described how, upon approaching the Tennessee state line, the open treeless prairies gave way to a belt of savanna about 8 km (5 mi) wide with scattered oaks before abruptly transitioning into woodlands as the rolling plains gave way to dissected country north of Nashville. In Indiana Edmond Dana (1819) noted “Small oaks, of a tolerable height, as thinly scattered as the apple trees in an orchard, usually commence at the termination of the barrens, and extend for a good distance, sometime for a space of two or three miles.” Limestone savannas also occurred along the northern edge of the Pennyroyal in the dissected Dripping Springs Escarpment region. Today there are many grassland remnants on these south-facing slopes, but most are found along roadsides or in powerline corridors. Some early maps of Kentucky show numerous places that bear the name “grove” (e.g., Oak Grove), which probably reflects settlements that were established in savanna where groves of trees would have provided welcome shade and a relief from the monotony of the rolling treeless plains. Pennyroyal Plain savannas only exist today at Fort Campbell Army Base in the impact zone. Highland Rim Savannas—Savannas also were found on the Highland Rims of Tennessee and probably also extended into north Alabama and south-central Kentucky. The Highland Rim savannas consisted of two basic types: dry and seasonally wet. Both were oak-dominated, but from the Alabama state line southward, loblolly pine probably was an important component of seasonally wet savannas and open flatwoods. From early writings it seems that much of the EHR was open prairie, no doubt with scattered inclusions of savanna and woodland, but like the Pennyroyal, savannas probably predominated where the flat to rolling surfaces of the HR surface began to break into more dissected terrain towards the Nashville Basin and Cumberland Plateau Escarpment. Savannas may have extended onto the lower slopes of the Plateau Escarpment where numerous grassland species still remain along trails, in powerline corridors, and in small open limestone barrens and glades. Nashville Basin—For decades, the general consensus among ecologists was that the Nashville Basin was primarily forested in pre-settlement times except for natural openings associated with limestone glades. This perception probably came from two sources: (1) early writers described the majestic forests of the Outer Nashville Basin dominated by tuliptree, basswood, maple, beech, and oak and (2) Safford’s (1869) description of large redcedars in the Inner Nashville Basin around glades. The hilly Outer Basin was probably mostly forested, but the large redcedars observed by Safford were likely an artifact of early 1800s fire suppression. Albrecht and Long (unpublished data) examined early land surveys from the 1780s–90s and found that junipers accounted for just 2 percent of the surveyed trees and that oaks and “stakes” were most abundant, suggesting that large areas of the Inner Nashville Basin consisted of an open post oak savanna matrix with imbedded limestone glades and barrens. Areas along streams, on hill-slopes, and sites with extreme surface karst and sinkholes likely would have supported woodlands and forests. For decades, botanists have puzzled over a botanical mystery. In river and stream floodplains of the Nashville Basin (and the Moulton Valley of Alabama, see below), there are small annual mustard plants known as bladderpods (Paysonia spp.). There are five species of bladderpods in central Tennessee and north Alabama. Most populations are restricted to regularly plowed agricultural fields. In fact these plants would be


PAGE 13 • A GUIDE TO GRASSLANDS OF THE MID-SOUTH

threatened if the fields, where these species occur were to be taken out of production. Strangely, two or three of the species also have one or two populations on limestone glades but agricultural fields seem to be preferred. This raises questions about what these stream valleys were like in pre-settlement times. Recent land survey records published by Drake et al. (2009) are providing insight into this question. In many of the larger stream valleys of Middle Tennessee, surveyors frequently used a “stake” to mark property boundaries, indicating areas devoid of trees. Today, with the exception of small floristically depauperate glades, these regions generally do not support conservative grassland species raising questions about what these areas were like. Given the high amount of disturbance, the Basin has undergone since the late 1700s, it is possible that all remnants of former Oute Basin deep-soiled grasslands have disappeared with the exception of the bladderpods. We will likely never know the true story. The concept of a Nashville Basin limestone savanna makes sense, given that numerous species of plants are restricted to this savanna type that do not occur in open pavement-type glades, including the federally-endangered Pyne’s groundplum (Astragalus bibullatus) and limestone clover (Trifolium calcaricum). Perhaps the thousands of old open-grown post oaks in the Inner Basin are another testament to the historic savannas but tree-ring studies will be needed to test this. Land managers are now realizing the need to restore this nearly extinct community. At Flatrock Cedar Glades and Barrens State Natural Area in Rutherford Co., Tennessee, efforts are underway by staff of the Tennessee Division of Natural Areas to remove thickets of redcedar to restore oak savannas. A high-quality savanna occurs at Couchville Glade and Barrens State Natural Area in Davidson Co., Tennessee. Additional efforts need to be initiated on other lands, including the Yanahli WMA, Maury Co. and Cedars of Lebanon State Park and Forest, Wilson Co. Moulton Valley—This region makes up portions of four north Alabama counties, but is a hotspot of biodiversity, with several endemic plant species. The Moulton Valley savannas likely existed as a mosaic of oak savannas, oak-loblolly pine flatwoods and limestone glades. Deeper soiled portions of the valley are used for cotton farming. Lands not kept open by cultivation or grazing have largely succeeded into forests. The Nature Conservancy’s Prairie Grove Glades Nature Preserve in Lawrence Co. is home to a number of rare species and remnant glade and savanna habitats and would be an ideal place to focus savanna restoration efforts in the future, similar to those underway in the Nashville Basin. Cumberland Plateau—Some of the most extensive savannas in the Mid-South occurred on the Cumberland Plateau of Tennessee, north Alabama, extreme northwest Georgia, and southeast Kentucky. Numerous historical descriptions describe the open lands that once existed, but most are from the northern plateau in Tennessee because this was one of the only places where roads (Avery Trace and Walton Road) crossed the Plateau in the late 1790s to early 1800s. Much of the rest of the Plateau remained unexplored by naturalists until after the Civil War, decades after an extensive turpentine and timber industry depleted the once extensive shortleaf pine savannas. Harper (1913) also noted extensive tracts of virgin longleaf pine (Pinus palustris) on the Plateau of northeasternWalker and southern Winston Cos., Alabama. Killebrew and Safford (1874) provided post-Civil War descriptions of the landscape and vegetation across numerous Tennessee Plateau counties. They used descriptions like “thinly wooded,” “unlimited pasturage,” “extensive flats, covered mainly with post oak and black jack,” “the native wild grass covering the entire face of the country,” and the abundance of “yellow pine” (=shortleaf pine) to describe the Plateau surface. Imbedded among the pine-oak savannas were prairie-like inclusions and scattered montane bogs. The flora and fauna of the Plateau hint at the former extent of savannas. Savanna endemics such as red-cockaded woodpecker and American chaffseed (Schwalbea americana) have now disappeared from the Plateau. Most conservation work on the Plateau has focused on preserving extensive forested areas associated with the numerous dissected gorges. However, little work has been focused on preserving and restoring the once extensive savannas of the Plateau surface. In fact, some 20th century ecologists have advocated that much of the Plateau surface should be managed as forest—a paradigm that if not abandoned will lead to continued widespread ecological collapse of now-rare habitats and the loss of sensitive species of plants and animals. In the last 25 years, nearly half of 40 rare savanna plant species have become extirpated completely from the Tennessee portion of the Plateau whereas several other species are on the brink. Since the 1990s, the Tennessee Wildlife Resources Agency (TWRA) has been spearheading efforts to restore shortleaf pine savanna at Catoosa WMA and the Bridgestone-Firestone Centennial Wilderness and parallel efforts are underway on the Daniel Boone National Forest in Kentucky. Several other public-land units in Tennessee present major opportunities for savanna restoration, which could reverse the current trend of species loss, including Savage Gulf State Natural Area (Grundy Co.), Fall Creek Falls State Park (Bledsoe and Van Buren cos.), Prentiss Cooper State Forest and Wildlife Management Area (Marion Co.). In Alabama, portions of Little River Canyon National Preserve (Cherokee, DeKalb cos.) would be ideal for savanna restoration. Opportunities to partner with foresters and the timber industry should be considered. Ridge and Valley Savannas—The Ridge and Valley was home to one of the most extensive and least understood grasslands in the Eastern U.S. Prior to European settlement they formed a natural “highway” that connected central Alabama to Pennsylvania. The earliest reference to Ridge and Valley savannas dates to the Hernando de Soto expedition of 1540 where they encountered savannas in southeast Tennessee, probably along the lower Hiwassee River. Missionaries Steiner and De Schweinitz (in Williams 1928) later described extensive savannas along the western base of the Great Smoky Mountains near Tellico Plains and Fort Loudon in the late 1700s. Francois Michaux briefly noted savannas south of present-day Oak Ridge in Roane Co., Tennessee, and the town of Strawberry Plains east of Knoxville was once an area of extensive grassland. Post oak and shortleaf pine were the dominant trees of Ridge and Valley savanna. Farther south, in the Coosa Valley of northeast Alabama and northwest Georgia, there were once extensive longleaf pine-pitcher plant savannas, which extended upslope into adjacent mountains where today patches of old-growth mountain longleaf pine can still be found, as at the Mountain Longleaf National Wildlife Refuge near Anniston, Alabama and the Berry College campus in Rome, Floyd Co., Georgia. Efforts are underway to restore savannas at the Coosa Prairie Preserve in Floyd Co., Georgia.


PAGE 14• A GUIDE TO GRASSLANDS OF THE MID-SOUTH

BARRENS As noted in the introduction to this guide, “barrens” is used here in a very specific way to refer to sloping rocky or gravelly grasslands dominated by perennial grasses and herbs. They develop primarily on slopes of limestone, dolomite and shale geologies. In southwestern North Carolina, barrens occur on serpentine, a very rare rock type for eastern North America. Lawless et al. (2006) advocated the use of the term “xeric limestone prairies” to refer to barrens over limestone in lieu of the commonly used term “cedar barren.” They also considered most limestone barrens to be of recent human origin caused by land-clearing and subsequent erosion, but this seems unlikely given the high number of conservative species barrens support across the region. Barrens are distinguished from savannas by their small patch size of one to few hectares, their occurrence in moderately to strongly sloping hilly landscapes, and their tendency to be imbedded in woodlands. Most eastern U.S. limestone barrens share much in common with Ozark “glades.” Western Valley Barrens—These barrens are found along the Western Valley of the Tennessee River separating West and Middle Tennessee. Foerste (1903) described “mound glades” along the Tennessee River that occur as conical hills “capped by whitish clays and soft limestones... visible for long distances and are often conspicuous landmarks.” Years earlier, Safford (1869) characterized them as “gravelly, and marly places, mostly naked, but presenting here and there patches of bushes, or shrubby cedars. They are sometimes several acres in extent, and occur usually upon hill-sides, but often entirely cover isolated and low knobs.” These barrens have been studied by DeSelm (1988). More than 150 individual barrens have been mapped from both sides of and within 1.6-8 km (1–5 mi) of the Tennessee River. Many examples have stunted, multi-stemmed, 200+ year-old redcedars that are interspersed in open gravelly barrens that are probably in sites naturally protected from frequent fire. The surrounding woodlands have become denser with fire suppression and much work is needed to thin trees and restore fire. Carol Cabin Barrens State Natural Area in Decatur Co. preserves several barrens but most remain in private ownership. The Western Valley is likely to see substantial development in future years due to the proximity to Kentucky Lake, a popular boating destination, and some of these barrens could be impacted by riverfront development. Cumberland Plateau Escarpment—Along the southern dissected escarpment of the Cumberland Plateau (including slopes in the Sequatchie Valley), barrens occur over limestone on dry lower to mid-slopes and mountainside benches. Approximately 30 individual barrens have been mapped, with the greatest concentrations in Franklin Co., Tennessee, and Madison Co., Alabama. These barrens exist as islands among rocky oak-hickory-smoketree woodlands. Historically, these lower mountain slopes were probably open and grassy with 30-75 percent tree canopy cover. The once extensive fires that maintained EHR prairies below the escarpment likely burned up the mountain slopes. With fire suppression, these woodlands have closed into forests with dense canopies and redcedar has increased, presumably in former openings. The barrens flora that was probably once widely distributed on dry slopes is now limited to the open barrens, roadside banks, trail edges and powerline corridors. One such corridor near the Cowan Tunnel, Franklin Co., Tennessee, supports large populations of grassland species such as purple coneflower (Echinacea purpurea), southern prairie dock (Silphium pinnatifidum), gray-headed coneflower (Ratibida pinnata) and white prairie clover (Dalea candida). A number of species are endemic to the rocky lower escarpment slopes, including the federally endangered Morefield’s leatherflower (Clematis morefieldii), the rare Cumberland rosinweed (Silphium brachiatum), and the newly described Kral’s beard-tongue (Penstemon kralii). Much work is needed to restore fire and openness to these communities. Several natural areas preserve these barrens remnants, including Bear Hollow Mountain WMA, Franklin Co., Tennessee and Wade Mountain Nature Preserve, Madison Co., Alabama. Several examples of what appear to be old-growth, perhaps virgin, chinkapin oak woodlands are found on the Plateau escarpment in the vicinity of Winchester, Franklin Co., Tennessee. Crawford-Mammoth Cave Uplands—The Crawford Upland (=Dripping Springs Escarpment) in west-central Kentucky is a maturely dissected escarpment that separates the gently undulating karst Pennyroyal Plain on the south from the sandstone, shale, and coal- dominated Shawnee Hills to the north. Side slopes of the hills are underlain by limestone, and in some areas grassy limestone barrens occur. Historically, fires that raced across the Pennyroyal of Kentucky and Mitchell Plain of Indiana annually likely burned up the slopes of these hills, resulting in a mosaic of open calcareous savanna and woodlands. Limestone barrens are associated with steep, somewhat eroded hillslopes. Aldrich and Bacone (1981) used the term “glade” for Indiana examples, but Homoya (1994) advocated the use of “barrens,” recognizing limestone and siltstone types. Baker Prairie Natural Area and Russellville Glade State Nature Preserve (SNP) in the town of Russellville, Logan Co., Kentucky, are examples of this rare community. In Indiana, Teeple Glade Nature Preserve, Harrison Co., is a protected example. Shawnee Hills—The Shawnee Hills of southeastern Illinois is a dissected landscape with a mixture of limestone barrens and sandstone glades. Limestone barrens occur on steep slopes over limestone bedrock. These have been studied by Kurz (1981). Similar examples occur in the Shawnee Hills Natural Region of Indiana (Aldrich and Bacone 1981, Homoya et al. 1985, and Homoya 1994).


PAGE 15• A GUIDE TO GRASSLANDS OF THE MID-SOUTH

Knobs and Bluegrass— The Bluegrass is a geologic basin located primarily in Kentucky but also extending into adjacent southeast Indiana and southwest Ohio. Although there is no evidence that the uplands of the Inner Bluegrass region supported naturally open grasslands in pre-settlement times (Campbell 1985), the Outer Bluegrass has areas of rocky limestone barrens on dry slopes and ridgetops, fringed with post and blackjack oak savanna and woodland. Although most examples are in small isolated patches, remnants can also be found in fencerows and rights-of-way, indicating a disturbance-based community. These barrens occur on Ordovician limestone or Silurian dolomite and harbor conservative species, such as juniper sedge (Carex juniperorum). Protected remnants are found at Versailles State Park, Ripley Co., Indiana (Homoya 1987) and Chaparral Prairie State Nature Preserve, Adams Co., Ohio, and Crooked Creek Barrens, Lewis Co., Kentucky. Some Indiana examples have been called “slump prairies” as they occur on steep slopes where periodic downward sliding of soil exists, often in areas bordering eroding creek banks. The Lynx Prairie in Adams Co., Ohio’s “Edge of Appalachia” is a complex of small limestone barrens sometimes referred to as “hanging prairies.” Chaparral Prairie SNP and Lynx Prairie, Adams Co., Ohio and Crooked Creek Barrens, Lewis Co., Kentucky. The Lynx Prairie in Ohio’s “Edge of Appalachia” is a complex of small limestone barrens sometimes referred to as “hanging prairies.” Other examples of barren or glade-like communities include tiny (few square meters), rocky openings amid oak-juniper woodlands along the Palisades (limestone cliffs) of the Kentucky River in the Inner Bluegrass. A protected example exists at Tom Dorman State Nature Preserve in Garrard Co., Kentucky. There are also areas of dry, rocky openings along the Licking River at Blue Licks State Park, Robertson, Nicholas and Fleming cos., Kentucky; these are home to the federally-listed Short’s goldenrod (Solidago shortii). Fringing the Bluegrass is a line of hills called “The Knobs”. These are cone-shaped erosional remnants of Devonian shale and Mississippian limestone. Small barrens occur on the slopes of these knobs, particularly on the escarpment of the Elizabethtown Plain. Protected areas include Jim Scudder SNP in Hardin Co., Kentucky and Thompson Creek Glades SNP in Larue Co., Kentucky. Western Allegheny Plateau—Along the western escarpment of the Allegheny Plateau, there are scattered isolated barrens that need more study. In Kentucky, these are primarily found in the vicinity of Clack Mountain, Rowan Co., near Carter Caves State Park, Carter Co., and in the vicinity of Irvine, Estill Co. The Carter and Rowan county sites harbor an undescribed species of Thaspium (D. Estes, unpubl. data). Ridge and Valley—Barrens are scattered throughout the Ridge and Valley and occur on limestone, dolomite and shale. Perhaps the most well-known barrens are the Ketona Dolomite Barrens found in Bibb Co., Alabama, which are home to a distinctive flora chocked with rare species, including about eight endemic species described to science in 2001 (Allison and Stevens 2001), and a couple described since, that are found nowhere else in the world. Dolomite barrens are also found in southwest Virginia and northeast Tennessee. Limestone barrens are found in The Cedars area of Lee Co., Virginia, throughout East Tennessee, particularly on the Oak Ridge National Laboratory Reservation (Roane Co.) and in Meigs Co. along the Tennessee River (DeSelm 1993, Ludwig 1999). Shale barrens are limited to two areas within the region: northwest Georgia on Taylor Ridge (Chattooga Co.) and northeast Tennessee and southwest Virginia. These shale barrens are not as well-developed and do not contain the high number of rare and endemic species characteristic of the Mid-Appalachian shale barrens farther north in western Virginia, West Virginia and Maryland. Blue Ridge—Barrens are very rare in the Blue Ridge Mountains, with the most notable type being the Buck Creek serpentine barrens of Clay Co., North Carolina within the Nantahala National Forest. These barrens are developed over serpentine/olivine rock, which has high concentrations of heavy metals in the soil that many plants can’t tolerate. They are positioned on moderate to steep mountain slopes at about 850-1000 m (2,800-3,280 ft) elevation. This is truly one of the South’s most unusual communities. The site of golden prairie grasses (e.g. prairie dropseed, Sporobolus heterolepis) growing on mountainsides among stunted pitch pine (Pinus rigida) is truly a site to behold. Numerous rare species are found in the Buck Creek barrens, including two recently described single-site endemic species found nowhere else in the world: Rhiannon’s aster (Symphyotrichum rhiannon) and the Buck Creek Ragwort (Packera serpenticola).

Level IV Ecoregions of Tennessee ! !

! ! !

! !

!

! ! !

!

! !

! !

!! ! ! ! !

!

! !

!

! !

! ! !! ! !! ! ! ! !! ! !! !!! !! !! !

!

!

!! !

! !

!

! !

!!

!

! !

!

!

!! ! ! !! ! !!! !! !! ! !! !! ! !! ! ! ! ! ! !! !!! ! !! ! ! ! ! ! ! ! !! ! ! ! ! !! ! ! ! !! ! ! ! ! !! ! ! ! ! !! ! !! ! ! ! ! ! ! ! !! ! ! !! ! ! ! ! ! ! ! !! !!! ! ! ! ! ! ! ! ! ! !! !! !! ! ! ! ! ! !! ! ! ! ! ! ! ! ! ! ! ! ! ! ! !! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! !! ! ! ! ! ! ! ! ! !! ! ! ! ! ! ! ! ! ! ! ! ! !! ! ! ! ! !! ! ! ! ! ! !! !!!! ! ! ! ! ! ! ! ! ! ! ! ! ! ! !!!!! !!! ! ! !! ! ! ! !! ! !! ! ! ! !! ! ! ! ! !! ! !!! ! ! ! ! ! !! ! ! ! ! ! ! ! ! ! ! ! ! !! !! ! ! ! ! ! ! ! ! ! ! !! ! ! ! ! ! ! ! !! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! !! ! ! ! ! !! ! ! ! ! ! !! ! ! ! ! ! ! ! ! ! !! ! ! !! ! ! !! ! ! ! !! ! ! ! !! ! ! ! ! ! ! ! !! !! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! !! ! ! ! ! ! !! ! ! ! ! ! ! ! ! ! ! ! !! ! ! ! ! !!! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! !!! !! ! ! !! ! ! ! !! ! ! !!!! ! ! ! !! ! ! !! ! !!!! ! ! ! ! ! ! ! ! ! !!! ! ! ! ! ! ! !! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! !! !! ! ! ! !!! ! ! ! ! ! ! ! !! !! ! !!!! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! !! ! ! ! ! ! ! ! ! ! !! !! !!!! ! ! ! ! ! !! ! !! !! ! ! ! ! ! ! ! !! ! ! ! ! ! ! !! ! ! ! ! ! ! ! ! ! ! ! ! !! !!! ! ! ! ! ! ! ! ! ! ! ! !! ! ! !!!! ! ! ! ! ! !!! ! ! !! ! !! !! ! ! !! !! ! !! !! ! ! ! !! ! ! ! !!! !! ! ! ! ! ! !! ! !! ! ! ! ! ! ! ! ! ! ! ! !!! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! !! ! ! ! ! ! ! !!! ! ! ! ! ! ! ! ! ! ! ! ! ! !! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! !! !! ! ! ! !! !! ! !! ! !! ! ! ! ! ! ! ! ! ! ! ! ! ! ! !! ! ! ! ! ! !!! !! !! !! ! ! !!! ! !! !! ! !! ! ! !! ! ! ! ! !! ! ! !! ! ! ! ! ! ! ! ! ! ! !! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! !!! !! ! ! ! ! ! ! ! ! ! ! ! ! ! !! ! ! ! ! ! ! ! !! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! !! ! ! ! ! ! ! ! ! ! ! ! ! ! !! ! ! ! !! ! !! ! ! !! !! ! ! ! !! ! ! !! ! ! ! ! !! ! ! ! ! !!! ! !!!!! ! ! ! !! !! ! !! ! !! ! ! ! !! ! ! ! !! !!!! ! !! !! ! ! ! ! ! ! ! ! ! ! ! !! ! ! ! ! !! ! !! !! !! !! !! ! ! !! ! !! ! ! ! !! ! ! ! !

! ! ! ! ! ! !

!

!

!! ! !! ! ! ! ! !!! !! !! !

!

! !! ! ! !!

!

Distribution of rare sun-loving plant species by ecoregion in Tennessee. Map Credit Sunny Fleming

Legend !

Strict Heliophytes

Bluff Hills

Level III Ecoregions

Loess Plain Meander Belt

PhysProvMerge2

Southeastern Plains

Interior Plateau Western Highland Rim

Undissected Eastern Highland Rim

Southwestern Appalachians

Undissected Western Highland Rim

Plateau Escarpment

Northern Hilly Gulf Coastal Plain

Outer Nashville Basin

Cumberland Plateau

Northern Holocene Meander Belts

Flat Woods

Nashville Basin Knobs

Sequatchie Valley

Northern Pleistocene Valley Trains

Pontotoc Ridge

Inner Nashville Basin

Crab Orchard

Northern Backswamps

Black Belt

Western Pennyroyal Karst Plain

Mississippi Aluvial Plain

Central Appalachians

Ridge & Valley Southern Limestone/Dolomite Valleys and Low Rolling Hills

Limestone Valleys and Coves Broad Basins

Southern Shale Valleys

Southern Crystalline Ridges and Mountains

Southern Sandstone Ridges

High Mountains

Southern Table Plateaus

Southern Metasedimentary Mountains

Southern Dissected Ridges and Knobs

Amphibolite Mountains

Blue Ridge


PAGE 16• A GUIDE TO GRASSLANDS OF THE MID-SOUTH

GLADES

Photo Credit: A. Cresler

Glades are sparsely-vegetated, rocky grasslands associated with flat-lying exposures of limestone or sandstone found in wide valley bottoms, bedrock benches on side slopes, along riversides, on plateau surfaces and clifftops. Unlike barrens, which are often strongly sloping, glades are generally level to slightly sloping. The thin soils or lack of soils only allow for the growth of annual grasses, small herbs, prickly-pear cactus, bryophytes and lichens. Shrubs and small stunted trees are restricted to deep-soiled cracks. In many glades, bedrock is exposed and soil is absent. In some types, only a thin mantle of soil covers the bedrock. In limestone glades, the bedrock mass is often broken at the surface into flat fragments (flagstones) and gravel residuum whereas in sandstone glades the bedrock weathers to sand. Glades are saturated in winter/spring, but become very dry and desert-like during summer and fall. Limestone glades are well known for their high number of endemic species (>25) and species disjunct from the Midwest and Ozarks. The major centers for limestone glades are the Inner Nashville Basin of Tennessee, the Moulton Valley of Alabama and the southern Ridge and Valley of northwest Georgia and southeast Tennessee. Compared to limestone glades, sandstone glades harbor fewer endemics, and several of the most restricted species occur both on sandstone and granite, the latter primarily of the Georgia and Alabama Piedmont. The vast majority of glades are imbedded in fire-suppressed woodlands where major thinning/burning efforts are needed to restore the once open margins of these imperiled communities. Nashville Basin Limestone Glades—The Nashville Basin is home to the largest concentration of limestone glades in Eastern North America. They were documented on early land surveys as early the 1780s. Unfortunately, the first trained naturalists did not describe the glades until about the Civil War. These descriptions depict open exposed bedrock openings (the glades proper) surrounded by dense forests of large redcedar (Safford 1869). Twentieth-century ecologists, such as Quarterman (1950), took Safford’s description as evidence that redcedar forests are a natural part of the glade ecosystem. In the last half of the 20th century, the role of fire has since been minimized or dismissed as an important factor in glade ecology. Instead, redcedar has been characterized as part of the climax community for woodlands surrounding glades and glades have primarily been managed as openings within closed forest. Biologists of the Missouri Botanical Garden’s Center for Plant Conservation (Albrecht et al. 2016), have shown that the federally endangered Pyne’s groundplum (Astragalus bibullatus) has likely declined due to encroachment of woody vegetation into its glade edge habitat. They examined early land survey records of Rutherford Co. and found evidence that redcedar has indeed “invaded” the Nashville Basin landscape. Their study reveals redcedar comprised just two percent of the trees in the 1790s whereas wooden stakes and oaks (likely post oak) were the most common survey markers. This evidence suggests parts of the Nashville Basin were more open and savanna-like. The abundance of old, open-grown post oaks throughout the Inner Nashville Basin is a testament to the openness of the historic landscape though dendroecological studies are needed to determine the region’s fire history. The concept of an extensive limestone savanna once covering much of the Inner Nashville Basin makes sense in light of this research. Urban and residential development poses the greatest threat to Nashville Basin glades given that the majority of glades are found within the Nashville-Murfreesboro-Lebanon area of central Tennessee, the so-called “Golden Triangle” and one of the fastest-growing regions of the country. Many glades were inundated by the impoundment of the Stones River to create the J. Percy Priest reservoir. Chinese privet (Ligustrum sinense) and bush honeysuckle (Lonicera maackii) are exotic shrubs aggressively invading glade margins. Overgrazing has destroyed the sensitive herbaceous flora of many glades. Fortunately, many large glade complexes are protected on state and federal lands. Notable preserves include Cedars of Lebanon State Park and Forest (Wilson Co.), Flatrock Glades and Barrens State Natural Area (Rutherford Co.), Couchville Glade and Barrens (Davidson Co.) and the Yanahli Wildlife Management Area (Maury Co.). Moulton Valley Limestone Glades—Outside the Nashville Basin, northwest Alabama’s Moulton Valley is the second-best region for limestone glades. Numerous glade species (Tennessee milkvetch, leafy prairie clover, Nashville breadroot) are shared between these two regions, but there are also a few species, which are endemic to the Moulton Valley, including Alabama glade cress (Leavenworthia alabamica), fleshyfruit glade cress (L. crassa), lyrate bladderpod (Paysonia lyrata) and glade blue-eyed grass (Sisyrinchium calciphilum). Most Moulton Valley glades are highly disturbed, and many have been lost to limestone quarrying and overgrazing. They were probably historically imbedded in a mosaic of calcareous oak savanna and open woodlands. The Prairie Grove Glades Preserve in Lawrence Co. protects several high-quality remnants.


PAGE 17 • A GUIDE TO GRASSLANDS OF THE MID-SOUTH

Pennyroyal Plain Limestone Glades—In south-central Kentucky, limestone glades are very rare and localized to a few areas in Christian, Simpson, Todd and Warren cos. Historically, these glades were likely imbedded in open calcareous oak savanna or open woodlands with grassy understory, which in turn graded into the vast Pennyroyal Prairie. Today, very few glades remain. Some are on private property in overgrazed cattle pastures where their distinctive glade flora has been displaced by common weeds. The best examples are found at Flatrock Glade SNP in Simpson Co. Unfortunately, these glades are surrounded by dense thickets and are in need of management by canopy thinning and prescribed fire. Trash dumping has also severely impacted some glades. Western Highland Rim Limestone Glade—A single example (0.1 ha, 0.2 ac) of this rare glade type exists in Lewis Co., Tennessee, at Dry Branch State Natural Area. It is located on a side slope in a small stream valley. Seeps flow along the margin of the glade, and one end contains a calcareous seepage fen. Several rare species are associated with this glade, including Tennessee yellow-eyed grass (Xyris tennesseensis) and large-flowered grass-of-Parnassus (Parnassia grandifolia). Outer Bluegrass—The Bluegrass is often regarded as the sister province to the Nashville Basin, but unlike the latter ecoregion, the Bluegrass lacks the abundance of limestone glades and savannas. Outer Bluegrass glades are primarily restricted to outcroppings of Silurian dolomite around Fort Knox (Jefferson and Bullitt counties). These glades support one endemic species, Kentucky Glade Cress (Leavenworthia exigua var. laciniata). Most sites have been destroyed or degraded by development and overgrazing. Ridge and Valley—Limestone glades of the Ridge and Valley are patchily distributed and small. The greatest concentration is found in northwest Georgia (Catoosa and Walker counties) near Chickamauga National Battlefield north into southeastern Tennessee (Bradley, Hamilton and Meigs cos). Many glades in private ownership are threatened by quarrying, urban and residential development, and impacts from overgrazing whereas protected glades are impacted by fire suppression. Near the Virginia-Tennessee border, there are a number of former glade areas, but no high-quality remnants are known and most are in cattle pastures. DeSelm (1993) studied many of the glades of the southern Ridge and Valley. Cumberland Plateau Sandstone Glades—Sandstone glades are scattered across the surface of the Cumberland Plateau. Today, these glades occur as islands among a sea of upland oak-pine forests, but in pre-settlement times they likely were imbedded in open shortleaf pine-oak savannas and open woodlands. They support a number of endemic species, including Lookout Mountain Tickseed (Coreopsis pulchra), flatrock fimbry (Fimbristylis brevivaginata) and Little River Canyon onion (Allium speculae). Several species are shared with the granite outcrops of the Georgia and Alabama Piedmont to the southeast. Northward into Tennessee and westward into northwest Alabama, sandstone glades are less well developed, more isolated and lack rare endemics. The best sandstone glades occur in northeastern Alabama near Little River Canyon National Preserve. The best examples in Tennessee are protected on the campus of the University of the South at Sewanee (Franklin Co.) and on private lands near the town of Jamestown, Fentress Co. Shawnee Hills—The Shawnee Hills of southern Illinois and western Kentucky contain scattered sandstone glades. These are mostly located along ridgetops in dissected landscapes and are sometimes associated with stunted woodlands. They are much less diverse and smaller in size than those of the Cumberland Plateau and lack the endemics and rare species. The Lusk Creek Wilderness Area in Pope Co., Illinois and numerous other sites on the Shawnee National Forest contain high-quality examples of this community. Mantle Rock Preserve (Livingston Co.) is a notable site in Kentucky.

Photo Credit: S. Fleming


PAGE 18 • A GUIDE TO GRASSLANDS OF THE MID-SOUTH

BALDS

Photo Credit: A. Cresler

Balds are literally the “coolest” grasslands of the Mid-South. They are found at high-elevations in the southern Blue Ridge Mountains above 1,676 m (5,500 ft) elevation on the summits and ridgelines. They are surrounded by dense thickets of Catawba rhododendron (Rhododendron catawbiense), flame azalea (R. calendulaceum) and blueberries which grade into forests of spruce-fir or birch-beech. The most extensive balds are found near Roan Mountain along the North Carolina/Tennessee state line, where they cover hundreds of acres (see cover photo of this guide). These and the balds of the Grayson Highlands near Mt. Rogers, Virginia, are the “northern” balds. They tend to be larger than the “southern” balds found near Great Smoky Mountains National Park, which are mostly just a few acres in size. All balds are threatened by invasion of trees and shrubs in the absence of grazing. According to Jamey Donaldson, botanist at East Tennessee State University and goat herder for the Baatany Goat Project on Roan Mountain, an estimated 75 percent of the grass balds on Roan have disappeared to invasion by trees in the last 100 years. Each summer, Donaldson tends a herd of Angora goats who help manage woody plant succession in the balds but most balds have not been grazed in decades. The need for grazing has led many to question how balds originated. A recent publication by Weigl and Knowles (2014) provides perhaps the most likely explanation. Their climate-herbivore hypothesis proposes that balds developed from alpine tundra that existed during the last Pleistocene Ice Age 18,000 years ago and have been maintained since by a long succession of native herbivores followed by European stock. Extinct species that once grazed/browsed the balds include mammoths, mastodons, horse, stag moose, fugitive deer, tapir, ground sloths and wooly rhinoceros. Other species have since retreated back to the north with warming post-glacial climates, including moose, musk ox, and caribou. At the time of European settlement, the balds were being maintained by bison, elk and deer. By the late 1700s their cattle, sheep and horses took over the role of grazing as these last native herbivores began to disappear. Since the early- to mid-1900s most balds have largely remained ungrazed as they were incorporated into the US Forest Service and National Park Service lands and are managed today by a combination of manual clearing, mowing and grazing. Balds near Roan Mountain support numerous rare species, including some that are found almost nowhere else outside New England, such as the green alder (Alnus viridis), Greenland sandwort (Mononeuria groenlandica) and New England ragwort (Packera schweinitziana). Endemics are also found on the balds including Gray’s lily (Lilium grayi) and Roan Mountain bluet (Houstonia montana). You haven’t lived until you’ve lain in the grass bald on Roan Mountain or watched the wild ponies frolic on Mt. Rogers. The balds are truly one of the most unique habitats in North America. The most similar communities to the Blue Ridge grass balds are the poloninas of the East Carpathian Mountains of Romania and the grass balds of the Oregon Coast Range (Weigl and Knowles 2014).

Heliophytes (sun-loving plants) of Southern Appalachian grass balds: Indian paint brush (Castilleja coccinea), Heller’s blazingstar (Liatris turgida), Gray’s lily (Lilium grayi).


PAGE 19 • A GUIDE TO GRASSLANDS OF THE MID-SOUTH

RIVERSCOURS Along entrenched, rugged, Appalachian streams there are flood-maintained grasslands (including glades) associated with riparian cobble and boulder bars or exposures of bedrock. These habitats are classified by ecologists as riverscour. NatureServe classifies these communities with the Cumberland Riverscour ecological system, primarily because they are found on sandstone substrates along the larger streams of the Cumberland Plateau (e.g., Daddy’s Creek, Obed River, Clear Creek, Big South Fork, Rockcastle River, Laurel River, Little River, Locust and Mulberry forks of the Black Warrior River). A few examples occur on limestone in the EHR of Tennessee (Caney Fork River) and Kentucky (Cumberland River and tributaries). Limestone riverscour is also present at the Falls of the Ohio River in the Bluegrass of Kentucky. In Indiana examples on limestone are most prevalent in the Shawnee Hills, Highland Rim, and Bluegrass natural regions. Dolomite riverscour occurs in the Ridge and Valley along the Little Cahaba River southwest of Birmingham. In the southern Blue Ridge Mountains, riverscour is well developed on the Hiwassee River (Polk Co., Tennessee) and a very rare “pine barren” type not studied since 1897 is reported from the French Broad River (Kearney 1897). Most Plateau riverscours are in deeply entrenched streams in gorges walled-in by sandstone cliffs and steep, forested, boulder-strewn slopes. In typical cases, the gorge bottom is choked with giant car-sized boulders. Floods that scour the gorges can exceed 4,247 m3/sec (150,000 ft3/sec) in rare cases. Unlike gravel- and sandbars, which are easily reworked by floods, the cobble and boulder bars are generally more stable over fairly long periods and thus are able to support conservative plant species. The scouring floods favor the development of shrubs, grasses and perennial herbs on the cobble bars, but on adjacent non-riparian slopes there are dense Appalachian forests of hemlock, white pine, magnolia, rhododendron and mountain laurel. The riverscour grasslands occur as small linear strips between the river channel and the forested slopes. They occur on elevated bars made of boulders, cobble and sand. These particles sort themselves into zones. Large, sparse boulders occur in and immediately adjacent to the channel. Moving away from the channel, particle sizes decrease such that medium to small boulders accumulate along the river’s edge forming elevated bars dominated by shrubs, ferns and herbs. Higher up and farther from the stream centerline are deposits of small boulders and cobble mixed with sand. Still higher upslope near the woodland margin at the back of the bar, there may be moderately deep deposits of sand which can support vegetation reminiscent of Coastal Plain sandhill habitats. The cobble and sandbar communities typically support grassland vegetation. Dominant species include important prairie grasses such as big bluestem (Andropogon gerardii), Indian grass (Sorghastrum nutans), little bluestem (Schizachyrium scoparium) and switchgrass (Panicum virgatum). These grasslands can be highly diverse with 89 species found during a single survey in one 100 m2 (1,076 ft2) plot (Rodgers in prep.). Several of the most common species on riverscours are endemics, including the Cumberland goldenrod (Solidago arenicola), Cumberland sandgrass (Calamovilfa arcuata), and the federally-threatened Cumberland rosemary (Conradina verticillata). They may contain seeps with Coastal Plain savanna species such as sundews and yellow-eyed grasses. Limestone glade scour contain Ozarkian disjuncts like the rare rock grape (Vitis rupestris), white prairie clover (Dalea candida), and maidenbush (Phyllanthopsis phyllanthoides). About six undescribed species exist on riverscour. Riverscour grasslands have been impacted by damming (e.g., Caney Fork River), but most are in relatively high-quality condition. One extinct species, Orbexilum stipulatum, is known to have inhabited the limestone scour at Falls of the Ohio, but has not seen since the 1830s. Invasive species, particularly Chinese bushclover, Autumn olive, multi-flowered rose, mimosa and Deutzia scabra, threaten some scours. Some of the best riverscour grasslands are found at the Big South Fork National River and Recreation Area (Scott Co., TN and McCreary Co., KY), Rockcastle River (Laurel and Pulaski cos., KY), Obed Wild and Scenic River (Morgan Co., TN), and Daddy’s Creek (Cumberland Co., TN). The best example of limestone riverscour occurs at Rock Island State Park, Warren and White Cos., Tennessee.


PAGE 20 • A GUIDE TO GRASSLANDS OF THE MID-SOUTH

HERBACEOUS WETLANDS The umbrella term “grassland” also may include numerous wetland types, including wet prairies and savannas, bogs, fens, meadows and emergent marshes, the last being associated with depression ponds imbedded in grassland. These wetlands are dominated by a combination of graminoids (grasses, sedges, rushes), herbs and scattered shrubs. Channelization and ditching have reduced wetland acreage in the region by 80–90%, but less well known is the adverse impact fire suppression has had on these communities. In the case of the ponds, fire may or may not have burned across them very frequently, but instead may have burned around and up to the very edge of the wettest parts of the wetland, thus keeping the pond shore open. Without fire, many of these smaller wetland types develop closed canopies, and their rare, sun-loving conservative plant and animal species decrease or disappear. Montane Bogs: Prior to the 1940s, the Cumberland Plateau was home to thousands of peat bogs—wetlands dominated by sedges with deep, peaty and strongly acidic soils. In pre-settlement times, these bogs occupied low flats and streamhead swales. Shortleaf pine-oak savannas likely dominated surrounding drier uplands. These bogs, a term widely used in the 1940s, were referred to by 19th century settlers as glades or meadows and many former bog sites can be inferred by names such as “Glade Branch” or “Meadow Creek.” Some modern ecologists prefer the term “poor fen” for such communities and argue that the term bog should be restricted to glaciated landscapes of the Northeast. Neel (1914) provided the best account of the Plateau bogs, describing what he called “natural meadows” dotting the Plateau surface and varying from less than one hectare to more than 20 hectares. These were “level as the prairie and practically free from tree growth.” He remarked that they were clear when the first settlers observed them and remained clear until the time of his observations. Killebrew and Safford (1874) called these areas “glades” similar to usage seen in West Virginia’s Cranberry Glades. They wrote: “In many places there are glades of greater or less extent, which are, in fact, small prairies, destitute of timber, and covered with coarse, rank grass. The superabundance of water in the soil and on the surface is the cause of the absence of timber.” Nearly all of these bogs were converted to farm ponds when mechanized machinery was introduced to the Plateau around World War II (Dr. Ed Clebsch, 2006 and 2016, pers. comm.). The bogs were dominated by sedges but were also home to numerous species which are now rare or extirpated from the Plateau, including the federally-endangered green pitcher plant (Sarracenia oreophila), rose pogonia (Pogonia ophioglossoides), grass pink orchids (Calopogon puchellus and C. oklahomensis), bog cotton grass (Eriophorum virginicum), whitehead bogbuttons (Lachnocaulon anceps) and white fringeless orchid (Platanthera integrilabia). Today, only a couple of sites of decent quality remain, and these are in desperate need of maintenance and restoration. Powerline corridors provide refuge for some remnants. Many examples have succeeded into forested wetlands, which are often classified as the Cumberland Acid Seepage Forest Ecological System by NatureServe. These “forests” contain species adapted for high-sunlight conditions and should be opened up and burned frequently. A number of public land units contain sites suitable for montane bog restoration, including Big South Fork National River and Recreation Area, Fall Creek Falls State Park, Savage Gulf State Natural Area and Little River Canyon National Park. In Alabama, much work has been focused on managing open habitats for the green pitcher plant (Sarracenia oreophila), but urgent efforts are needed on the Tennessee and Kentucky portions of the Plateau. There are no large bogs that remain. Some former bog sites have strong restoration potential but will require significant investment of resources to restore hydrology and bog vegetation. Fens: Fen is a term that has been mostly applied in glaciated regions of the Midwest and Northeast; however, a number of notable sites in limestone-dominated regions of the Mid-South would qualify, especially the Ridge and Valley and Highland Rim. Fens are small herbaceous wetlands fed by mineral-rich calcareous groundwater and typically dominated by sedges, grasses, rushes, and herbs with scattered shrubs. Soils are continuously waterlogged, and may be shallow and gravelly over bedrock or deep and mucky. In the Ridge and Valley of east Tennessee, muck fens were probably once common in valleys underlain by limestone or dolomite where they occurred around the many abundant valley springs. Degraded remnants exist near the town of Sweetwater, Monroe Co., Tennessee and at Red Clay State Historic Park, Bradley Co., Tennessee. Perhaps the best extant valley fens are those associated with the “Coosa Prairie Preserve” in Floyd Co., Georgia.

Photo Credit: A. Cresler

Photo Credit: A. Cresler


PAGE 21 • A GUIDE TO GRASSLANDS OF THE MID-SOUTH

Other fens are sometimes called “seepage fens” and are found on gentle to moderately sloping hillsides with shallow, gravelly soils. These hillslope fens are found in the Ridge and Valley of northeast Tennessee (Campbell, Claiborne, Union counties) and southwest Virginia, particularly along the Powell River. Another notable hotspot for seepage fens is in the WHR of Tennessee, southwest of Nashville in the vicinity of the Natchez Trace (Lewis, Williamson counties). The Tennessee Division of Natural Areas maintains a number of small limited-access natural areas that preserve these rare slope fens, including several sites in Lewis Co. Some of these examples have the federally-endangered Tennessee yellow-eyed grass (Xyris tennesseensis), a species that has been in decline in recent years. Dry oak-hickory forests surround these seepage fens, and the broad flat uplands found even farther upslope once supported extensive blackjack and post oak savannas. Fires likely once burned downslope to and around these seepage fens, maintaining them as open sites rich with species that require naturally open conditions. Much work is needed to evaluate the role of fire in these communities. It isn’t clear whether or not the fens themselves burned, but what is clear is that the adjacent upslope oak woodlands are fire dependent, contain conservative heliophytes, and are in need of thinning and burning. The limestone glades of the Nashville Basin and to a lesser extent other gladey regions support limestone glade fens. These are glades that are wet from winter through early May in most years and are supplied by mineral-rich groundwater. They are often dominated by obligate or facultative wetland species. Numerous rare species are associated with this community including the federally-endangered leafy prairie clover (Dalea foliosa), yellow sunnybells (Schoenolirion croceum), prairie Indian plantain (Arnoglossum plantagineum) and Quarterman’s hedge-hyssop (Gratiola quartermaniae). Depression Ponds: Historically, several of the large open grasslands of the Mid-South contained imbedded depression ponds, including the Pennyroyal Plain Prairie of Kentucky, Tennessee, and Indiana and the Highland Rim (especially the Eastern) prairie and savanna of Tennessee and Alabama. In the Ridge and Valley of northwest Georgia, these ponds are known as “sag ponds.” These grassland ponds contain a rich assemblage of species, many disjunct from the Coastal Plain pine savannas. In these systems, the margins are exceedingly important and are where the greatest diversity exists. With fire suppression, the grassy transition zone (ecotone) from open pond and emergent marsh to wet and then dry grassland disappears with the invasion of trees, leading eventually to a sharp pond margin and the disappearance of dozens of species that need open shorelines and sunlight. Goose Pond Registered National Natural Landmark in Coffee Co., Tennessee, is an excellent example of where the rich pond ecotone has been lost to forest succession, and as a result many of the rarest species, such as Carolina redroot (Lachnanthes caroliniana) and fringed yellow-eyed grass (Xyris fimbriata), have nearly disappeared from the periphery. Creeping St. John’swort (Hypericum adpressum) has disappeared entirely. In recent decades, the emergent grass maidencane (Panicum hemitomon) has displaced other rare pond species. This could be the result of fire suppression, which presumably would have kept the maidencane in check historically, but now this species forms nearly continuous stands across some ponds and are known as “maidencane ponds.” There is a small section of the Dripping Springs escarpment northwest of Bowling Green that has remained largely undissected. Flats around the town of Chalybeate were described as “sand swamps” by Sadie Price, a Kentucky botanist who worked in the late 1800’s. Based on the species historically documented from this area, it is likely that these plateaus had open acidic wetlands similar to what is found on the Highland Rim today. No modern remnants have been located on the flats, although wetland communities (such as Sloan’s Pond) have been found on the dissected ridges in Mammoth Cave National Park that suggest a similar, although less open, condition. More work is needed in this area. The Mitchell Karst Plain in Indiana is replete with sinkhole wetlands but most have been severely altered by artificial drainage and use as livestock ponds. Naturally vegetated remnants house an interesting assemblage of species, including Carex comosa, C. decomposita, Decodon verticillatus, Dulichium arundinaceum, Glyceria acutiflora and Torreyochloa pallida (Homoya and Hedge 1983). The Pennyroyal Plain was once home to thousands of sinkhole ponds that were imbedded in vast tallgrass prairie, essentially the South’s very own “prairie pothole” region similar to those of the northern Great Plains. Some are/were large ponds or lakes with nearly permanent open water supporting emergent marsh and shrub swamp vegetation. Remnant examples include Hundred Acre Pond (Hart Co., KY) and Chaney


PAGE 22 • A GUIDE TO GRASSLANDS OF THE MID-SOUTH

Lake (Warren Co., KY). Most depression wetlands of the Pennyroyal are/were ephemeral, holding water only in winter and early spring and for short periods following rain events in summer and fall. Today, nearly all examples of the ephemeral type are in private ownership and are imbedded in corn, wheat and soybean fields. This community is essentially extinct as a natural wetland type. Even though modern examples are devoid of natural vegetation and are surrounded by man-made environs, these ponds are a haven for waterfowl and wading birds including numerous species of ducks, Canada geese and sandhill cranes. It is easy to imagine how significant these thousands of ponds would have been to migrating waterfowl 200 years ago but it also leaves us to wonder just what we’ve lost that we’ll never realize. The only description of these ponds is one provided by C.W. Short (1836):

“Now, would burst upon the view a smooth sheet of water…whilst its surface was covered over with the large and floating leaves and splendid flowers…and then, in endless vistas, was stretched before the eye a waving sea of gigantic grasses.”

Meadows: Meadows may be the least known and most misunderstood grassland type within the Mid-South. They are small-patch communities that may have occurred in linear arrays on stream terraces in narrow stream valleys. Unlike most of the grasslands covered in this guide, which are fire-dependent communities of relatively undissected to poorly dissected landscapes, meadows occur in strongly dissected and otherwise densely forested landscapes. The hydrologic source may be subsurface seepage from adjacent underground aquifers, a locally high water table associated with an adjacent stream, and/or surface seepage from adjacent sedimentary rock layers. Overbank flooding may also contribute to community maintenance. Due to abundant water availability, high humidity and occurrence in generally fire-protected habitats, meadow communities were probably not maintained in pre-historic times by fire but were likely the result of perpetual occupation by beaver. Modern beaver (Castor canadensis) likely had a profound impact on floodplain landscapes for tens of thousands of years, if not longer. In some places in southeastern Canada where beaver are still abundant, dozens of densely spaced beaver dams occur for miles along stream systems. Presumably beavers had similar habitats in the Mid-South. Their activity possibly created repeating mosaics of beaver marsh, meadows, shrub thickets, and riparian and floodplain wetlands. The near-elimination of beavers in the Mid-South occurred early in our nation’s history as a result of the aggressive fur trade spear-headed by the French in the late 1600s–early 1700s. It is easy to imagine how the extirpation of beaver could transform Mid-South floodplains in short time. These transformations possibly involved succession to canebrakes and forested wetlands. Native Americans also kept some bottomlands open for villages, farming and hunting. The Mississippian Culture was strongly dependent on stream valleys where they built villages, mounds and had extensive agricultural fields. After the collapse of the Mississippian Culture in the mid-1400s, much of the Mid-South lacked permanent Indian settlements. From the 1500s to the early 1800s, the Cherokee, Chickasaw, Creek and Shawnee Indians, who primarily lived on the periphery of the Mid-South, used the region as a mutual hunting ground and no doubt still exerted much control over the landscape via burning. Archaeological research in East Tennessee along the Little Tennessee River demonstrates that Cherokee Indians and their predecessors intensively managed the stream valleys where they lived with fire. How much fire was used in stream valleys lacking permanent Indian habitations is less well known. William Bartram, during his first visit to the Cherokee country in western North Carolina in 1775, described how the stream valleys among the mountains were full of wild strawberries, and he used the word “lawns” to describe their appearance. Many Blue Ridge stream valleys are still rich in relict meadow communities. As Euro-Americans settlers began to settle the wilds of the Mid-South, many established farms centered on these stream valleys (Drake et al. 2009). They encountered extensive canebrakes, which was a favored food for bison. Presumably, these areas would have been among the most open landscapes available in these dissected lands and allowed for easy conversion to crops and pastures. Today, all of the stream valleys have been altered from their original pre-settlement condition via wetland drainage, channel alterations, planting and haying of cool-season forage grasses, and pond construction. Remnant sedge meadows are present and contain certain conservative indicator species that require open grass/herb-dominated wetlands, including New England aster (Symphyotrichum novae-angliae), purpleleaf willowherb (Epilobium coloratum), prairie loosestrife (Lysimachia quadriflora) and swamp lousewort (Pedicularis lanceolata). One can look to sites in the Missouri Ozarks, namely Grasshopper Hollow Fen in Reynolds County, to get a sense for what some of these meadow complexes may have been like, though Missouri examples have a higher representation of prairie species. Some high quality current examples occur in Land Between the Lakes National Recreation Area, Stewart Co., Tennessee and along Saline Creek in the impact zone at Fort Campbell Army Base, Trigg Co., Kentucky. One complexity that must be considered in future conservation and restoration efforts of meadow communities is that most are on private lands, and within a given stream valley there may be dozens of separate landowners. Such a fragmented ownership landscape can be a difficult hurdle for large-scale restoration. The general dislike for beaver, which may be important if not critical to restoration, could be an impediment in many areas, though mowing and perhaps fire could be used to manage some sites where beaver reintroduction isn’t possible.


PAGE 23 • A GUIDE TO GRASSLANDS OF THE MID-SOUTH

CONCLUS IO N S When most people think of the natural landscape of the Mid-South, the last thing that probably comes to mind is grasslands. This is because our grasslands were destroyed so long ago that they have escaped the collective memory of our modern society. They weren’t photographed or painted, and almost no stories of their former greatness have been passed down to current generations. Ecologists have had to struggle to reconstruct the grassland puzzle. As the pieces fall into place, we are beginning to see the picture we’ve been missing for the past half-century. The places we love, the parks we visit on a weekly basis, the scenery we drive by daily, was very different just 200 years ago. While we’ve struggled to find a way to fit the pieces together, the fragments of grasslands that managed to persist through the 1960s are mostly gone. The majority of those that held on a little longer vanished in the 1980s and 90s. The 21st century has been especially hard on grasslands and it isn’t getting any easier. As we push toward 2020 and the pieces continue to fall into place, our picture is nearly complete and it is telling us that we need a massive overhaul in current land management in the Mid-South. It won’t be popular and many, even many who read this now or attend the Mid-South Prairie Symposium, won’t accept it. Our society is in love with forests, but our desire to have a forested landscape is killing our native biodiversity and we have been in ecological collapse for decades. There are thousands of acres locked up in state forests, state parks, wildlife management areas, nature preserves and national parks and forests that are silently trying to tell us that they are suffering. They need their open lands restored with stand thinning and fire. This is the most parsimonious way to restore our native grasslands. The time for change is now.

Photo Credit: A. Dattilo


PAGE 24 • A GUIDE TO GRASSLANDS OF THE MID-SOUTH

Estimated Pre-Settlement Distribution of Grasslands and Related Communities of the Mid-South U.S.


PAGE 25 • A GUIDE TO GRASSLANDS OF THE MID-SOUTH

ACKNOWLEDGEMENTS MIKE STAMBAUGH BRYAN YAHN MARTINA HINES CLARENCE COFFEY JULIAN CAMPBELL MELISSA CASPARY BROOKE BEST SUGGESTED CITATION: Estes, D., M. Brock, M. Homoya, A. Dattilo. 2016. A Guide to Grasslands of the Mid-South. Published by the Natural Resources Conservation Service, Tennessee Valley Authority, Austin Peay State University and the Botanical Research Institute of Texas. PHOTOGRAPHY CREDITS: Mason Brock, Alan Cressler, Adam Dattilo, Marc Evans, Mike Homoya, Kentucky Nature Preserves Commission, Sunny Fleming, Clarence Coffey, Edward W. Chester, Mike Stanbaugh, Gary Fleming, Chris Ludwig, James Richardson, Harold Gaston, Philip Juras, Jovonn Hill, H.R. DeSelm, Julie Tuttle and Devin Rodgers.

Photo Credit: D. Rodgers


Austin Peay State University does not discriminate on the basis of race, color, religion, creed, national origin, sex, sexual orientation, gender identity/expression, disability, age, status as a protected veteran, genetic information, or any other legally protected class with respect to all employment, programs and activities sponsored by APSU. http://www.apsu.edu/files/policy/5002.pdf AP343/5-16/250


Turn static files into dynamic content formats.

Create a flipbook
Issuu converts static files into: digital portfolios, online yearbooks, online catalogs, digital photo albums and more. Sign up and create your flipbook.