Final Proceedings Copy4-4-07 CD

Transcription

Final Proceedings Copy4-4-07 CD
Proceedings of the Sixth Longleaf Alliance Regional Conference
November 13-16, 2006
Tifton Campus Conference Center
University of Georgia
Tifton, GA
Longleaf Alliance Report No. 10
March 2007
ii
Longleaf Pine: Seeing the Forest through the Trees
Proceedings of the Sixth Longleaf Alliance
Regional Conference
November 13-16, 2006
Tifton Campus Conference Center
University of Georgia
Tifton, GA
This conference would not be possible without the financial and logistical support of the following organizations:
USDA Forest Service
Georgia Forestry Commission
J. W. Jones Ecological Research Center
Simmons Tree Farm
International Forest Company
Stuewe & Sons
Meeks Farm & Nurseries, Inc.
Warnell School of Forestry & Natural Resources,
University of Georgia
School of Forestry and Wildlife Sciences, Auburn University
The Longleaf Alliance appreciates the generous support of these organizations.
Citation: Estes, Becky L. and Kush, John S., comps. 2007. Longleaf Pine: Seeing the Forest through the Trees,
Proceedings of the Sixth Longleaf Alliance Regional Conference; November 13-16, 2006, Tifton, GA. Longleaf
Alliance Report No. 10.
Longleaf Alliance Report No. 10
March 2007
iii
In changing with the times, The Longleaf Alliance is attempting to do its part to “green” the meeting. We have chosen to
distribute the proceedings for this year in a CD format. This will reduce the amount of paper used to print hardcopies and
allows each user to print only what is needed. The proceedings are in a PDF format that can be viewed using the Adobe
Free Reader, which can be downloaded at the following website: http://www.adobepack.com/. If you have any comments or suggestions, please direct them to us at the address listed below. Thank you.
Proceedings Order Form
If you would like to have another CD or a hardcopy of the proceedings sent to you, please complete and detach the form
below.
Name: ____________________________________
Address:___________________________________
City: _____________________________State: _________ Zip: ______________
Please check what format you would like:
____CD
_____ Hard Copy
Please mail to:
The Longleaf Alliance - Proceedings
3301 Forestry and Wildlife Sciences Bldg.
Auburn University, AL 36849-5418
phone 334.844.1012
fax 334.844.4873
e-mail [email protected]
iv
Foreward
Dean Gjerstad, The Longleaf Alliance
The goal of The Longleaf Alliance is to provide information
at our conferences on “all things longleaf”. This conference
met the goal. Topics presented in the concurrent sessions
included 1) wildlife management: quail, predators, birds,
herps 2) Herbicides: site preparation, herbaceous control 3)
Understory Restoration: ground cover, restoration, seed collection and handling 4) Fire: reintroduction, old growth
stands, frequency and seasonality, tree growth. More topics
were discussed in the poster session: herbicides, wiregrass
restoration, ecosystem management, climate effects, old
growth, cones, seedlings, conservation, growth and yield,
fire/fuel ladders, restoration, invasive/exotic species, redcockaded woodpecker, root pathogens, economics, forest
structure, fire councils, soils, planting spacing, restoration
planning, ecosystem management, and bird conservation.
There were 321attendees at the meeting. The invited
speakers discussed topics on local longleaf pine management, wildlife issues, herbicide use, understory restoration,
prescribed fire, and environmental history. The poster
session with 50 presentations expanded on the topics in the
sessions and generated fruitful discussion. There was substantial exchange of ideas through the speaker and poster
sessions and new and old friends were brought together
through education, good food, and entertainment.
With the threat of storms looming on the horizon, the field
trip began at the Tift/Brumby farm which is co-owned by
Mike Brumby, Jerry Tift, and Thomas Tift. A variety of
topics were covered on the property from pine straw harvesting, naval stores operations, prescribed burning, wildlife management, natural regeneration, and midstory control. After enjoying a relaxing lunch with a pleasant view
of the pond, the field trip continued to the Oakridge Farms
owned by E. Cody Laird, Jr. and Nancy Laird Croswell.
While here, the participants learned all about the intricacies of wiregrass restoration from seed harvest to planting.
Other topics covered were conservation easements, planting of longleaf pine seedlings, groundcover, and non-game
management. The field day was followed by a wonderful
dinner and picturesque setting complete with a warm fire
graciously hosted by Cody Laird. Music was provided by
Tanger who entertained us with Irish music, creating a beat
for energetic céilí dancing. Not even the threat of tornadoes and rain dampened our enthusiasm for longleaf pine.
South Georgia was the setting for the Sixth Regional Longleaf Alliance Conference. The Tifton area was an excellent
location for longleaf researchers, managers, land owners, and
enthusiasts to congregate as it is surrounded by a significant
portion of the remaining longleaf pine forests. The title for
the Sixth Regional Longleaf Alliance Conference was Longleaf Pine: Seeing the Forest through the Trees and it highlighted all aspects of the longleaf pine community: trees,
herbaceous and shrub layers, soils and wildlife. The general,
concurrent and poster sessions, provided landowners, private
and federal organizations, and other interested parties the
ability to develop partnerships and provided excellent examples of restoration of the understory, non-native invasive
control, silvicultural options, social and political challenges,
wildlife conservation, and uses of prescribed fire.
The regional conference would, of course, not be possible
without the dedication and hard work of many individuals
who help to develop and organize the meeting venue. To
everyone who had a hand in bringing the conference from
the beginning to the end many thanks are offered. I would
like to thank Karen McBrayer, Senior Event Coordinator
at the UGA Tifton Campus Conference Center, for serving
as our conference coordinator. We thank all of those who
assisted with the field trip during the conference. We extend our appreciation to the Georgia Forestry Commission
for undivided help throughout planning and during the
field trip. We also thank Ron Halstead for his guidance
and help in coordinating the field trip. Much thanks goes
to the J.W. Jones Ecological Research Center for providing
transportation and drivers at Oakridge Farms. The field
trip would of course not be possible without the access to
longleaf pine forests. For this, we extend a very special
thanks to Mike Brumby and Cody Laird for the use of the
Tift/Brumby and Oakridge farms. We also say thank you
to Cody Laird for graciously hosting dinner on the property. The fried peanut butter and jelly sanwiches added a
nice touch to the conference. And the barbeque and brews
were a perfect southern touch to a wonderful day.
The conference was held at Tifton Campus Conference Center which is located in the College of Agriculture and Environmental Sciences at the University of Georgia, Tifton
Campus. The staff and facilities at the conference center
proved to be perfect for the meeting, continuing the high
meeting standards from the past. Tifton is located in centralsouth Georgia in Tift County and is centered in the belt of
quail plantations that stretches from Albany to Thomasville,
Georgia and on to Tallahassee, Florida. These areas represent some of the best remaining intact longleaf pine habitats
in the southeast. The importance of quail plantations in
southern Georgia cannot be understated as they have done
enormous good for the conservation of longleaf pine forests
and associated wildlife species. There are also important
research facilities in the area. The J.W. Jones Ecological
Research Center at Ichauway is a 29,000 acre outdoor laboratory once owned by Robert W. Woodruff and now managed by the Woodruff Foundation and was the setting for the
post-conference field trip. Located south of Tifton is the Tall
Timbers Research Station whose mission is to conserve ecosystems.
v
vi
Table of Contents
Schedule of Events ...................................................................................................................................................... xii
Speaker Presentations..................................................................................................................................................2
Herbicides in Longleaf Pine: Guidelines for Selection and Use
Mark Atwater..................................................................................................................................................................2
Seed Cleaning and Germination Testing Procedures for the Restoration of Ground Layer Plants in
a Longleaf Pine Ecosystem
Jill Barbour .....................................................................................................................................................................3
Aspects of Predator Ecology and the Predation Process within a Longleaf Pine Forest
Mike Conner ...................................................................................................................................................................8
Current Studies on Selected Birds of the Longleaf Forest.
Jim Cox...........................................................................................................................................................................9
Grey Moss Plantation Overview
Jenny Crisp ...................................................................................................................................................................10
Restoring Longleaf Ground Layer Vegetation on Private Lands
E. David Dickens, Bryan C. McElvany and David J. Moorhead .................................................................................12
Opportunites for Restoring Longleaf Ground Layer Vegetation through the US Fish and
Wildlife Service.
Jeff Glitzenstein, Donna Streng, Jim Bates, Mark Hainds, Jill Barbour, and Joe Cockrell .........................................15
Reptiles and Amphibians of Longleaf Pine Forests: Update on Some Conservation Issues
Craig Guyer ..................................................................................................................................................................21
Benefits and Costs of Herbaceous Release Treatments in Longleaf Pine Establishment and
Management
M.J. Hainds...................................................................................................................................................................23
Frequency and Seasonality of Prescribed Fire Affect Understory Plants
James D. Haywood .......................................................................................................................................................25
How Does Longleaf Pine Native Groundcover Fit with Forest Management Goals?
Sharon M. Hermann ....................................................................................................................................................28
Fire History of a Georgia Montane Longleaf Pine (Pinus palustris) Community
Nathan Klaus ................................................................................................................................................................32
Fire Effects on Longleaf Pine Growth
John S. Kush .................................................................................................................................................................33
New Findings for Site Preparation with Chopper Herbicide.
Dwight K. Lauer and Harold E. Quicke .......................................................................................................................37
Reintroduction of Fire to Fire Suppressed Longleaf Pine Stands: An Overview of the Problem
John McGuire ...............................................................................................................................................................38
vii
Physiological Effects of Organic Soil Consumption on Mature Longleaf Pines (Pinus palustris)
Joseph O'Brien, J. Kevin Hiers, Kathryn Mordecai and Doria Gordon .......................................................................39
Pineywood's Cattle Breed: History and New Uses for Small Acreages
Chuck Simon ................................................................................................................................................................41
Bobwhite Quail Issues and Research Efforts in the Longleaf Region: An Overview
Lee Stribling .................................................................................................................................................................42
The Georgia Coastal Flatwoods Upland Game Project: Launching a War on Wiregrass
Chris Trowell................................................................................................................................................................45
Lessons Learned About Ground-layer Restoration: What we think we know and what we don’t
Joan Walker and Lin Roth ............................................................................................................................................46
Poster Presentations ...................................................................................................................................................49
Use of Herbicide Site Preparation Treatments to Promote Longleaf Seedling Growth and to
Enhance Fuels Structure for Longer Term Fire Management
Robert N. Addington, Thomas A. Greene, Catherine E. Prior, Wade C. Harrison ......................................................49
Longleaf Pine Plant Community Restoration at the Savannah River Site: Design and Preliminary
Results
Todd A. Aschenbach, Bryan L. Foster, and Don W. Imm ...........................................................................................54
Longleaf pine ecosystem management at Eglin Air Force Base, Florida
Chadwick Avery ...........................................................................................................................................................58
Establishment and Management of Longleaf Pine (Pinus palustris Mill.) Seed Production Areas
Jill Barbour ...................................................................................................................................................................59
The Dendrochronology of Pinus palustris in Virginia
Arvind A. R. Bhuta, Lisa M. Kennedy, Carolyn A. Copenheaver and Philip M. Sheridan .........................................60
Old-Growth Longleaf Pine on Horn Mountain, AL (Talladega National Forest)
David Borland, Art Henderson, John S. Kush, and John McGuire..............................................................................61
The Longleaf Pine Cone Crop Story
Elizabeth Bowersock, William D. Boyer and John S. Kush ........................................................................................63
Restoring and Maintaining Ecological Integrity of Special Communities Embedded within
Longleaf Pinelands
Joyce Marie Brown and Johnny P. Stowe ....................................................................................................................65
Longleaf Pine Ecosystem Restoration Project: Lessons Learned from LPER
Shan Cammack .............................................................................................................................................................68
Longleaf Pine Seedling Survivorship and Growth on Poorly Drained Soils
Susan Cohen and Joan Walker......................................................................................................................................71
viii
Restoring and Managing Longleaf Pine Ecosystems in the Southern United States: Southern
Research Station Research Work Unit 4158 – Auburn, AL: Clemson, SC; Pineville, LA
K.F. Connor, D.G. Brockway, J.D. Haywood, J.C.G. Goelz, M.A. Sword-Sayer, S-J.S. Sung, and
J.L. Walker ...................................................................................................................................................................72
South Carolina Lowcountry Forest Conservation Project
W. Conner, T. Williams, G. Kessler, R. Franklin, P. Layton, G. Wang, T. Straka B. Humphries,
C. LeShack, K. McIntyre, R. Mitchell, S. Jack W. Haynie, A. Nygaard, L. Hay D. Beach, J. Lareau ,
J. Johnson, and M. Robertson, M. Prevost, M. Nespeca ..............................................................................................74
An Investigation of Old-field Longleaf Growth, Yield, Diameter Distributions, Product Class
Distributions, Pine Straw Production, and Economics of Management Intensities in Georgia
E. David Dickens, Bryan C. McElvany and David J. Moorhead .................................................................................75
Old Resinous Turpentine Stumps as an Indicator of the Range of Longleaf Pine in Southeastern
Virginia
Thomas L. Eberhardt, Philip M. Sheridan, Jolie M. Mahfouz, and Chi-Leung So ......................................................79
Spatial Patterns of Fuels and Fire Intensity in Longleaf Pine Forests
B.L. Estes, D.H. Gjerstad, and D.G. Brockway ...........................................................................................................83
Evaluating Forest Development and Longleaf Pine Regeneration at Mountain Longleaf National
Wildlife Refuge
Bill Garland, John S. Kush, and John C. Gilbert,.........................................................................................................87
Wiregrass – Overrated
John C. Gilbert, John S. Kush, and John McGuire.......................................................................................................89
Longleaf Pine Re-Discovered at Horseshoe Bend National Military Park
John C. Gilbert, Sharon M. Hermann, John S. Kush, Lisa McInnis and James Cahill ................................................93
A Container-Grown Seedling Quality DVD
Mark J. Hainds, Elizabeth Bowersock, and Dean Gjerstad..........................................................................................95
Longleaf Pine Forest Restoration at Horseshoe Bend National Military Park: Evaluation of
Residual Stands and Re-Introduction of Fire
Sharon M. Hermann, John C. Gilbert, John S. Kush, Caroline Noble and Herbert “Pete” Jerkins .............................97
What Happens to Top-Killed Seedlings?
Rhett Johnson and Mark J. Hainds .............................................................................................................................100
Effects of Two Native Invasive Trees on the Breeding Bird Community of Upland Pine Forests
Nathan Klaus and Tim Keyes .....................................................................................................................................101
The Regional Longleaf Pine Growth Study – 40 years old
John S. Kush and Don Tomczak ................................................................................................................................102
Chopper® Herbicide Site Prep Improves Quality of Weed Control
Dwight K. Lauer and Harold E. Quicke .....................................................................................................................104
ix
Red-Cockaded Woodpecker Recovery and Longleaf Pine Ecosystem Conservation: Sharing and
Selling the Success through the Eyes of the Advocates
Jon Marshall, Ralph Costa, John Maxwell and Dave Case ........................................................................................107
Pathogenicity of Leptographium serpens to Longleaf Pine
George Matusick, Lori Eckhardt and Scott Enebak ...................................................................................................108
An Economic Model for Multiple-Value Management of Longleaf Pine
B.B. McCall, R. K. McIntyre, S. B. Jack, and R. J. Mitchell.....................................................................................109
Spatio-Temporal Patterns of Forest Structure and Understory Species Composition in Longleaf
Pine Flatwoods along Florida's Gulf Coast
George L. McCaskill and Shibu Jose .........................................................................................................................110
Tale of Two Forests: Light Environments in Slash and Longleaf Pine Forests and Their Impact on
Seedling Responses.
J.D. McGee,R.J. Mitchell, S.D. Pecot, J.J. O'Brien, L.K. Kirkman, and M.J. Kaeser ...............................................110
Linking State Prescribed Fire Councils as a Coalition: A Proposal to Promote Media and Public
Understanding of Rx Fire, and to Nationally Address Key Management, Policy, and Regulatory
Issues
Mark A. Melvin, Johnny Stowe, Frank Cole, Lane Green, Scott Wallinger, and Lindsay Boring ............................112
Longleaf Pine Genetics Research at the Harrison Experimental Forest
C.D. Nelson, L.H. Lott, J.H. Roberds, T. L. Kubisiak and M. Stine..........................................................................114
Long Term Research on the Effects of Fire Regime on Upland Longleaf Pine Forests
Thomas E. Ostertag and Kevin M. Robertson............................................................................................................116
Spatial and age structure of old-growth mountain longleaf pine, (Pinus palustris), stands in the
Talladega National Forest of northeastern Alabama
Brett Rushing, Kevin Jenne, and Robert Carter .........................................................................................................118
Spacing recommendations for longleaf pine
David B. South ...........................................................................................................................................................121
Ichauway’s prescribed fire management program 1994-2006: A balanced approach
Jonathan M. Stober and Steven B. Jack......................................................................................................................122
Allatoona Lake Longleaf Pine Ecosystem Restoration Project
Terrell Stoves..............................................................................................................................................................125
The Role of Ritual and Ceremony in Wildlands Conservation: Reestablishing Primal Connections
Johnny Stowe..............................................................................................................................................................126
Private Property Rights vis-à-vis Establishing and Maintaining Invasive Exotic Plant Species:
Legal and Ethical Ramifications of the “Right to Plant” versus Other’s “Right to Maintain
Landscape Integrity and Property Values”
Johnny Stowe..............................................................................................................................................................129
A Framework for Restoration: Increasing the Success of Longleaf Pine Restoration Projects
Rob Sutter, Brett Williams, Alison McGee and Michelle Creech..............................................................................133
x
Repeated Fire Effects on Soil Physical Properties in Two Young Longleaf Pine Stands on the
West Gulf Coastal Plain
Mary Anne Sword-Sayer ............................................................................................................................................134
Preliminary Density Management Diagram for Naturally Regenerated Longleaf Pine
Curtis L. VanderSchaaf, Ralph S. Meldahl, and John S. Kush ..................................................................................137
The East Gulf Coastal Plain Joint Venture: A Regional, Landscape-Scale Approach to All-Bird
Conservation
Allison Vogt ...............................................................................................................................................................141
A Continued Pinus palustris Burn Study Comparing Frequency and Season of Fire to Basal Area
Growth Loss
Ben Whitaker, William D. Boyer, and John S. Kush .................................................................................................142
Landscape Scale Ecosystem Classification in Longleaf Pine Forest of the Talladega Mountains,
Alabama
Brent Womack and Robert Carter ..............................................................................................................................145
Fuel Loads, Tree Community Structure, and Carbon Storage in Mountain Longleaf Pine Stands
Undergoing Restoration
Rebecca Worley and Martin Cipollini........................................................................................................................148
xi
Sixth Longleaf Alliance Regional Conference
Schedule of Events
UGA-Tifton Campus Conference Center
Tifton, GA
Longleaf Pine: Seeing the Forest through the Trees
Monday, November 13, 2006
5:00 – 8:00 PM
Registration
Social, poster session, vendors & photographers with featured displays
Tuesday, November 14, 2006
7:00 – 8:30 AM
Registration
8:30
Welcome & Introductions (John Johnson - Deputy Administrator for Farm Programs, Farm
Service Agency; Rick Hatten - Management Chief, Georgia Forestry Commission; Richard
Brinker - Dean, School of Forestry & Wildlife Sciences, Auburn University; Bob Izlar - Director for Forest Business, Warnell School of Forestry & Natural Resources, University of
Georgia)
9:00
Local Efforts with Longleaf Management, Restoration and Research
Session Moderator: Peter Stangel, NFWF
Peter Stangel - National Fish & Wildlife Foundation
Rick Hatten - Georgia Forestry Commission
Alison McGee - The Nature Conservancy
Robert Brooks - U.S. Fish & Wildlife Service
Mark Whitney - Georgia Department of Natural Resources
9:45
BREAK
10:15
Local Efforts with Longleaf Management, Restoration and Research
Jim Cox - Tall Timbers Research Station
Susan Gibson - U.S. Army
Kevin McIntyre - J.W. Jones Ecological Research Center
Amy Carter - National Environmentally Sound Production Agriculture Laboratory
11:00
Regional Longleaf Recovery Plan – Dave Case, DJ Case & Associates
11:30
The State of the Longleaf Alliance – Rhett Johnson & Dean Gjerstad, Co-Directors
12:00
LUNCH (provided)
1:30 PM
Concurrent Session I - Wildlife Issues in the Longleaf Ecosystem
Session Moderator: Eric Darracq, Georgia Department of Natural Resources
xii
1:30
Bobwhite Quail in the Southeast 1930 - 2006 - A story of increase, decline, research, outreach,
and recovery. Lee Stribling, School of Forestry and Wildlife Sciences, Auburn University,
AL
1:55
Aspects of predator ecology within the longleaf pine forest. Mike Connor, J.W. Jones Ecologi
cal Research Center, Newton, GA
2:20
Current studies on selected birds of the longleaf forest. Jim Cox, Tall Timbers Research Sta
tion and Land Conservancy, Thomasville, GA
2:45
Some important amphibians and reptiles of the longleaf pine forests. Craig Guyer, Department
of Zoology, Auburn University, AL
1:30 PM
Concurrent Session II - Herbicide Use in Longleaf Pine Forests
Session Moderator: Mark Hainds, The Longleaf Alliance, Andalusia, AL
1:30
New findings for site preparation with chopper herbicide. Dwight
K. Lauer, Silvics Analytic, Ridgeway, VA
1:55
Herbicides in longleaf pine:Guidelines for selection and use. Mark Atwater, Weed Control
Unlimited, Inc., Donalsonville, GA
2:20
Post-plant herbaceous weed control timing considerations for longleaf pine. E. David Dick
ens, Bryan C. McElvany, and David J. Moorhead, Warnell School of Forestry & Natural Re
sources, The University of Georgia
2:45
Benefits and costs of herbaceous weed control: Three to five years post-application. Mark
Hainds, The Longleaf Alliance, Andalusia, AL
3:00 PM
BREAK
3:30
Concurrent Session III - Understory Restoration
Session Moderator: Jim Bates, U.S. Fish and Wildlife Services
3:30
Longleaf groundcover: What good is it and what do you need to know to make informed dec
sions for forest management? Sharon M. Hermann, Department of Biological Sciences, Au
burn University, AL
3:55
Seed cleaning and germination testing procedures for the restoration of ground layer plants in a
longleaf pine ecosystem. Jill Barbour, USDA Forest Service, National Seed Laboratory, Dry
Branch, GA
4:20
Opportunities for restoring longleaf ground layer vegetation through the USFWS Partners for
Fish and Wildlife Program. Jeff Glitzenstein, Tall Timbers Research Station, Newton, GA
4:45
Lessons learned about ground-layer restoration: What we think we know and what we don’t.
Joan Walker, USDA Forest Service Research Plant Ecologist and Lin Roth, Department of
Forestry and Natural Resources and Belle W. Baruch Institute of Coastal Ecology and Forest
Research, Clemson University, SC
xiii
3:30 PM
Concurrent Session IV – Good Fire/Bad Fire
Session Moderator: Kevin Hiers, J.W. Jones Ecological Research Center, Newton, GA
3:30
Reintroduction of fire-to-fire suppressed longleaf pine stands: An overview of the problem.
John McGuire, The Longleaf Alliance, Auburn University, AL
3:55
Why mature longleaf pine trees die following organic soil (duff) consumption. Joseph
O'Brien, USDA Forest Service, Southern Research Station, Athens, GA
4:20
Frequency and seasonality of prescribed fire effects on understory plant community development. James D. Haywood, USDA Forest Service, Pineville, LA
4:45
Fire effects on longleaf pine growth. John Kush, Longleaf Pine Stand Dynamics Lab, Auburn
University, AL
5:05
Adjourn
5:30 – 8:00 PM
Social/Poster Session
Wednesday, November 15, 2006
8:00 AM
Depart from UGA-Tifton Campus Conference Center to Field Day Location #1
8:30
Tift/Brumby Farm
Mike Brumby, Jerry Tift and Thomas Tift, owners
Field Stations
Product Utilization - Mike Harrison, Consultant
Pine Straw/Fertilization - David Moorhead, University of Georgia
“Enviro-grid” Stream Crossing - Chad David and Bert Earley, Georgia Forestry Commission;
Cal Callahan Associates
Natural Regeneration - John Kush, Auburn University
Prescribed Burning - Ron Halstead, Consultant
Bobwhite Quail Management - Eric Staller, Tall Timbers Research Station and Jimmy Atkin
son, J.W. Jones Ecological Research Station
Naval Stores History and Demo - Grady Williams, Local Expert
Midstory Control with Herbicides - Dwight Lauer, Silvics Analytic
Midstory Control with Mulcher and Demonstration - Jerry Marchant, Vendor
Lunch @ Tift Farm
1:00 PM
Oakridge Farms
E. Cody Laird, Jr. and Nancy Laird Crosswell, owners
Focal Area 1
Wiregrass Restoration - Kevin McIntyre, J.W. Jones Ecological Research Center
Hand-Planted Wiregrass Plugs
Small Wiregrass Plugs Planted with Tobacco Planter
Wiregrass Seed Planted with Grasslander (equipment demo)
Wiregrass Plugs Planted with Whitfield Planter (equipment demo) - Terry Whitfield
Focal Area 2
Conservation Easements - Alison McGee, The Nature Conservancy
xiv
Focal Area 2
Conservation Easements - Alison McGee, The Nature Conservancy
Underplanting Longleaf Pine Under Mature Slash Pine - Mark Hainds, The Longleaf Alliance
and Bob Franklin, Clemson Extension
Old-Field Groundcover versus Intact Groundcover - Kay Kirkman and Melanie Kaeser, J.W.
Jones Ecological Research Center
Nongame Management - Todd Engstrom, Jimmy & Sierra Stiles and Wilson Baker
Safe-Harbor Agreement - Ralph Costa, U.S. Fish and Wildlife Service and Phil
Spivey, Georgia Department of Natural Resources
Wednesday Evening, November 15, 2006
4:30 – 8:30 PM
Dinner & Social – Oakridge Farms
Entertainment provided by: Tanager (Irish Band)
Thursday, November 16, 2006
8:00 AM
Longleaf Embryogenesis. John Pait, VP of Business Development CellFor, Inc.
8:30
Environmental History
Moderator: John McGuire, Outreach Coordinator, The Longleaf Alliance
8:30
ral
Fire history on Pine Mountain, GA. Nathan Klaus, Senior Wildlife Biologist, Nongame NatuHeritage Section, Georgia Department of Natural Resources
9:00
Pineywood's cattle breed, history and new uses for small acreages. Chuck Simon, County Ex
tension Agent-Coordinator, Covington County, Alabama Cooperative Extension System
9:30
The Georgia Coastal Flatwoods Upland Game Project: Launching a war on wiregrass. Chris
Trowell, Emeritus Professor of Social Science, South Georgia College, Douglas
10:00
BREAK
10:30
Landowner Panel:
John Norman - Quail Ridge Plantation
Bill Moody - South Carolina landowner
Jenny Crisp - Grey Moss Plantation
Mayo Livingston - Cyrene Turpentine Company
11:30
Wrap-up & Adjourn - Rhett Johnson & Dean Gjerstad,
Co-Directors, The Longleaf Alliance
Friday, November 17, 2006
TBA
Post-conference Tour
J.W. Jones Ecological Research Center
xv
Speaker Presentations 1
Speaker Presentations
Herbicides in Longleaf Pine: Guidelines for Selection and Use
Mark Atwater1
1
Weed Control Unlimited, Inc., Donalsonville, Georgia, 39845, USA
Abstract
Timely, effective use of herbicides in longleaf pine systems is dependent on recognizing site specific goals, features and limitations including: reforestation vs. restoration, time, economics, manpower, presence or absence of
invasive plant pests, sensitive non-target organisms and
others. Effective recognition of these and other criteria
combined with a well thought out management plan are
essential for success in using herbicides to achieve the site
objective.
Speaker Presentations 2
Seed Cleaning and Germination Testing Procedures for the Restoration of
Ground Layer Plants in a Longleaf Pine Ecosystem.
Jill Barbour1
1
USDA Forest Service, National Seed Laboratory Dry Branch, Georgia, 31020, USA
Introduction
Upland habitats in southeastern Coastal Plain, USA,
were dominated historically by open woodlands of
longleaf pine with a very rich herbaceous ground layer
(Peet and Allard, 1993). Presently intact stands of this
vegetation type are very much reduced, probably to less
than 2% of the original extent (Frost 1993). Restoration
of the longleaf pine ecosystem is currently an important
priority of several federal and state agencies (Brockway
et al., 2005). In addition, programs are in place to aid
private landowners who wish to accomplish such restorations (http://ecos.fws.gov/partners/). Reinstating the
rich herbaceous ground layer is a critical step in overall
ecosystem restoration.
Large-scale restoration of longleaf ground layer vegetation depends ultimately on a ready source of viable seed
for many ground layer species. Unfortunately, this resource is not yet available for most of the species. One
bottleneck is lack of basic information on seed cleaning,
seed germination, long-term seed storage, and efficient
procedures for nursery propagation (see Pfaff and Gonter 1996, Glitzenstein et al. 2001, Pfaff et al. 2002, Coffey and Kirkman 2006, for preliminary results).
This study was implemented with three main objectives:
(1) Develop seed cleaning procedures for longleaf
ground layer plants. Cleaned seed should be pure, i.e.
free of trash, and with high germination given suitable
germination conditions. (2) Determine optimal germination protocols, including the need for moist cold stratification and scarification. (3) Determine if laboratory
germination successes can be duplicated in a working
nursery environment.
Materials and Methods
Winter 2005/2006 collections
Seeds from 34 species of ground layer plants (9 grass
species, 25 forb species) were collected by hand in 2005
and 2006 from 6 sites within Alabama, Georgia, and
South Carolina (Fort Benning, GA, Aiken Gopher Heritage Preserve, SC, three private landholdings in Russell
County, AL, one private landholding in Stewart County,
GA). Seeds were allowed to dry naturally in paper
bags, than taken to the USDA Forest Service National
Seed Laboratory for cleaning in spring 2006. Some
collections were too small to clean with equipment, so the
seeds were extracted by hand and germinated in laboratory
dishes.
Seed Conditioning
The general process for seed conditioning was as follows:
Step 1 Remove seeds from inflorescence with Westrup
laboratory brush machine
Step 2 Use an aspirator or blower to remove very small
trash
Step 3 Remove sticks and large trash with hand held
screens or Ideal indent cylinder screens
Step 4 Remove lighter weight material with blowers
(Stultz or General Blower) or Oliver 30 specific gravity
table.
Step 5 Hand pick out debris for small lots
Step 6 Prepare sample for seed testing and storage
The Forsberg scarifier was used instead of the brush machine on Baptisia lanceolata, and Tetragonotheca helianthoides. Tephrosia virginiana was cleaned without the
sandpaper inside the scarifier. Polygonella americana was
cleaned with the brush machine and the Forsberg scarifier.
Coreopsis major seeds were run over the Oliver 30 specific
gravity table with a linen deck. Seeds fell through 3 chutes
at the end of the deck and were labeled upper, middle and
lower designating deck position. Seeds from the middle
chute were rerun and divided into middle upper and middle
lower. X-rays revealed that the upper and middle upper
seed portions had full seed, but the middle lower and lower
portions contained mostly empty seeds. These portions
were combined with the trash. The upper and middle upper still contained a significant amount of debris, so water
floatation was tried. Seeds sunk in the water, but debris
and seeds floated too, resulting in further cleaning with the
General Blower.
Water floatation was tried with Lespedeza hirta. The cotyledons on some of the floating seeds swelled open and had
to be removed from the lot.
Speaker Presentations 3
Table 1. Germination data for the winter 2005/2006 seed collections.
Species
Nursery Germ
%
Amsonia ciliata
Aristida condensata
Lab germ
%
Comments
12
20
4
Aristida stricta
38, 24
Aster tortifolius
0
Baptisia lanceolata
2
Chamaecrista fasciculata
767,512
27
20
6
11
48
No prechill; 28 day
prechill
Brown;
Green
1
36
40% after cleaning
Coreopsis major
9
Desmodium
12
32 upper
33
22
16
4
65% dormant
Brown
Green
No prechill
7 day prechill
4
40% dormant
Eriogonum tomentosum
0
Eupatorium hyssopifolium
648,000
61% dormant
73% dormant
93295
762,353
27% after cleaning
528,671
71% dormant
65% dormant
1,109,046
1,537,637
35% after cleaning
Galactia macri
Lespedeza capitata
Lespedeza hirta
Liatris elegans
3
Covered seed
Brown
Green
Brown, no float
Brown float
Green no float
Green float
11,554
5
7% after cleaning
177,742
327,272
62
17% after cleaning
30
7% after cleaning
Manfreda virginica
3
23
4,086,486
47
22
49
57
39
26
20
33
Liatris tenufolia
Paspalum bifidum
927,607
None dormant
16
Eupatorium rotundifolium
Mimosa quadrivalvis
Sd/lb
after
cleaning
355 dormant after test
Chrysopsis gossypina
Liatris secunda
Sd/lb before
cleaning
0
Aristida purpurascens
Eupatorium album
Comments
527,442
14
0
35% dormant
Pityopsis graminifolia
2
Saccharum alopecuroides
6
53
16
Caryopsis
Whole seed
Schizachyrium scoparium
5
54
49
2
19
Caryopsis
Whole seed
No prechill
28 day prechill
4
2
No cleaning
After cleaning
24
12
Caryopsis
424,925
305,572
Caryopsis
No prechill
7 day prechill
After cleaning
905,389
343,896
28
18
0
1
25
0
93% dormant
394,434
3
After cleaning
207,692
Silphium compositum
Solidago odora
Sorghastrum nutans
Sorghastrum secundum
Sporobolus juneus
Tephrosia virginiana
Tetragonotheca helianthoides
Vernonia angustifolia
2
855,849
Speaker Presentations 4
180,501
caryopsis
247,463
whole
13,095
11% after cleaning
462,857
978,314
Germination
hand and seeds were discarded that had been damaged by
insect predation or that had clearly failed to develop (i.e.
seed was unfilled). Collection dates for the two species
were November 23, 2006 and December 22, 2006, respectively. Ionactis trays were put out in the nursery January
15-January 30, 2007 and Pityopsis was put out February 1015, 2007. Sample sizes were n=945 seeds for Ionactis and
n=900 seeds for Pityopsis. Germination results were tallied
on February 28, 2007.
Standard unstratified germination tests were conducted on
seeds from each collection. Small germination dishes
were utilized. Kimpak and blue blotters were the media,
dampening them with 47.5 ml of water. Germination
temperature was 20° C (68° F) for 16 hours of darkness
and 30° C (86 ° F) with 8 hours of light. For Schizachryium, Sorghastrum, and Saccharum genera, the caryopsis
and whole seeds were germinated separately to determine
germination differences. Green and brown seeds were
tested separately in the legume genera, Lespedeza, Baptisia, and Desmodium; the green seeds germinated slightly
better than the brown seeds. Legume seeds that did not
germinate were examined by cutting them open. If the
embryos were not dead, they were classified as dormant.
Stratified germination tests were performed on seeds of
Aster tortifolius, Silphium compositum, Eriogonum tomentosum, Sporobolus junceus, and Vernonia angustifolia.
Results
2005/2006 collections
The Westrup laboratory brush machine satisfactorily removed seeds from the inflorescence. Usually seeds were
expelled from the opening at the chute end and the trash fell
through the opening underneath the machine. Small seeds
from some forb species came out both openings and got
mixed with the trash. The brush machine created a large
amount of trash. A variety of methods and a large amount
of time was required to separate the debris from the seeds.
Nursery propagation
Seeds from 15 species were hand planted in August 2130, 2006 in hard plastic propagation trays at the American
Tree Seedling Nursery in Bainbridge, Georgia. Germination for all species was checked on October 4, 2006.
Nursery and laboratory germination for the winter
2005/2006 collections are listed in table 1 and illustrated
in figure 1.
Lespedeza was run through the brush machine several
times. To keep from injuring the seedcoat, the seeds were
screened out of the debris between each run. Three species
of Liatris were damaged in the brush machine resulting in
higher germination before cleaning than after cleaning. A
softer mantle, with less cutting action, in the brush machine
may alleviate this problem. The Asteraceae family inflorescences were difficult to clean, because so much debris was
created in the brush machine, further requiring the use of
additional equipment to remove the debris from the seeds.
Seeds of grass species were easily cleaned in the brush machine. The caryopsis separated from the whole seed yielded
higher germination in Saccharum, Schizachyrium, and Sorghastrum Table 1.
Winter 2006/2007 collections
% Ger mi nat i on
Considerable additional seed was collected during winter
2006/2007 from the private landholding in Stewart
County, GA, and from Francis Marion NF (FMNF) near
Charleston, SC. Much of this seed is still in the initial
stages of processing and testing but nursery propagation
data on seed of two FMNF species of Asteraceae, Ionactis
(Aster) linariifolius and Pityopsis graminifolia is already
available. These two species were cleaned carefully by
65
60
55
50
45
40
35
30
25
20
15
10
5
0
62
20
12
32 33
27
0
53 54
23
6
2
9
1
Nu rse ry g e rm
12
17
4
0
3
3
28
24 25
23
14
2
0
L a b g e rm
Speaker Presentations 5
6
5
12
Figure 1
4
2
The Forsberg scarifier without sandpaper removed the seeds
from the legumes without breaking the seed coats. Tephrosia virginiana seeds were easily cleaned using this method.
The laboratory germination was higher than the nursery
germination except for Mimosa quadrivalvis, and Sorghastrum nutans Figure 1. The ambient temperature in August
was probably too high for germination of most species’
seeds. Tephrosia virginiana nursery germination was 3
percent higher than laboratory germination.
2006/2007 collections
Nursery germination data from the two FMNF fallflowering Asteraceae were as follows: (1) Ionactis linariifolius: 450 germinated out of 945 seeds put out, germination rate 47.6%, (2) Pityopsis graminifolia: 574 germinated
out of 900 seeds put out, germination rate 63.8%. These
numbers are underestimates since germination was still proceeding on the tally date.
Discussion
In a previous exploratory study Glitzenstein et al. (2001)
demonstrated moderate success in germinating seed of
many of these same species. Of 42 species, 32 had germination rates in excess of 20% in at least one test. Methods
used in that study were to place seed outside in the nursery
shortly after collection; thus, seeds were exposed to natural
germination cues. Seeds were cleaned by hand and damaged seeds were discarded. Using similar techniques very
respectable germination percentages were obtained for two
species of Asteraceae collected in fall 2006 from FMNF
(2006/2007 collections discussed above). These generally
positive results contrast with the very low germination percentages in the present study of nursery propagated Asteraceae put out in August 2006 and to a lesser extent with
the laboratory results for this plant family.
Two explanations can be suggested for this discrepancy.
First, it is likely that high August temperatures are outside
the natural range of temperatures under which these species
will germinate (Pfaff and Gonter 1996, Pfaff et al. 2001).
Most of the Asteraceae examined in both studies fruit in
late fall and will germinate in the field during winter and
early spring when soil moisture levels tend to peak and temperatures are cool (Glitzenstein et al. 2001, Coffey and
Kirkman 2006). When these conditions are replicated in the
nursery high germination percentages are obtained. The
laboratory tests which simulated late spring conditions may
have likewise been suboptimal for some Asteraceae.
When damaged seeds were eliminated in the 2006/2007
collections higher germination rates were found. This
suggests that insect-damaged or predated seeds were not
satisfactorily screened out during the machine cleaning
process and that new protocols will need to be developed to achieve that goal.
Results for Fabaceae and Poaceae were similar to the
earlier (Glitzenstein et al. 2001) study. Satisfactory (>
20%) germination percentages were obtained for most
of the major species in these two families. However,
cleaning techniques employed in the current study did
not appreciably increase the germination percentages in
comparison to the uncleaned seed as tested herein or in
the earlier study. Further iterations of seed separation
equipment (i.e. the aspirator, the blower) may be needed
to accomplish the objective. Most of the grass seeds
have not been cold stratified, a technique that may appreciably increase germination for some species. Use
of the Forsberg scarifier without sandpaper was a convenient method for extracting seeds from legumes while
at the same time ensuring sufficient scarification for
good germination. An exception was Baptisia which
had low germination and high residual dormancy. More
intense scarification would appear to be indicated to
break dormancy in this genus.
Conclusions
We are still learning how to handle, clean and germinate understory species’ seeds in the longleaf pine ecosystem. More collection sites need to be added to examine seed characteristics over the species’ range. We
need to collect seed over several years to determine the
limits of fecundity, seed set, and viability. The conditioning process needs to be fine tuned by collecting
more seeds from fewer plants to increase effectiveness
of the cleaning equipment. Additional equipment needs
to be tried to determine the optimum cleaning process.
Different germination temperatures need to be examined to determine optimum germination. Seed pretreatments such as stratification need to be further explored to maximize germination in the laboratory and
nursery. Little is known or written about the nursery
propagation of these species: sowing dates, nutrient
regime, growing habit, and preparation for out-planting.
Available information indicates that nursery propagated
plugs will survive well in the field over the short term
(Glitzenstein et al. 2001), but long-term survival and
demography information is not available for most species.
Further laboratory experiments are planned to explore the
effects of temperature regimes on germination rates for the
Asteraceae. A second explanation for low germination
rates of the 2005/2006 Asteraceae collections may have
been high rates of seed damage due to insect predation.
Speaker Presentations 6
Acknowledgements
We would like to recognize the staff at the USDA Forest
Service National Seed Laboratory for their assistance on
the germination and cleaning of the seeds in this project
and Chuck Whittaker from American Tree Seedling
Nursery for donating the containers and media for growing the plants.
Literature Cited
Brockway, D.G. , K.W. Outcalt, D.J. Tomczak, and E.E.
Johnson. 2005. Restoration of longleaf pine ecosystems. Gen. Tech. Rep. SRS-83. Asheville, NC: U.S.
Department of Agriculture, Forest Service, Southern
Research Station. 34 pp.
Coffey, K.L. and L.K. Kirkman. 2006. Seed germination
strategies of species with restoration potential in a
fire maintained pine savanna. Natural Areas Journal
26: 289-299.
Frost, C.C. 1993. Four centuries of changing landscape
patterns in the longleaf pine ecosystem. Proceedings
Tall Timbers Fire Ecology Conference 18: 17-44.
Glitzenstein, J.S., D.R. Streng, D.D. Wade and J. Brubaker.
2001. Starting new populations of longleaf pine ground
layer plants in the outer Coastal Plain of South Carolina, USA. Natural Areas Journal 21: 89-110.
Peet, R.K. and D.J. Allard. 1993. Longleaf pine vegetation
of the Southern Atlantic and Eastern Gulf Coast regions: a preliminary classification. Proceedings Tall
Timbers Fire Ecology Conference 18: 45-82.
Pfaff, S. and M.A. Gonter. 1996. Florida native plant collection, production and direct seeding techniques: interim report. USDA, NRCS, Plant Materials Center,
Brooksville, FL, USA, 76 pp.
Pfaff, S., M.A. Gonter, and C. Maura. 2002. Florida native
seed production manual. USDA, NRCS, Plant Materials Center, Brooksville, FL, USA, 76 pp.
Speaker Presentations 7
Aspects of Predator Ecology and the Predation Process within a Longleaf Pine Forest
Mike Conner1
1
J.W. Jones Ecological Research Center, Newton, Georgia, 39870, USA
Abstract
Predators are an important natural component of forested
ecosystems. However, in the southeastern United States,
habitat loss and fragmentation have resulted in the loss of
many top carnivores, such as the red wolf and Florida
panther. Because large carnivores are largely absent in
the southeast, remaining predator communities are
dominated by smaller predators. The role of these smaller
predators is poorly understood, and their ecology within
longleaf pine-dominated systems has received little research attention. Here, the basic ecology of a variety of
predator species is discussed. Finally, the concept that
habitat serves as a template for the predation process is
introduced as a potential tool for better understanding and
managing predation.
Speaker Presentations 8
Current Studies on Selected Birds of the Longleaf Forest
Jim Cox1
1
Tall Timbers Research Station and Land Conservancy, Thomasville, Georgia, 31799, USA
Abstract
The red-cockaded woodpecker is certainly one longleaf
specialist that draws a lot of attention because of its rarity,
but there are many other declining birds associated with
longleaf systems that need attention. For example, both
brown-headed nuthatch and Bachman's sparrows disappeared from
some longleaf areas before the red-cockaded woodpecker,
and both species are considered very rare in parts of their
range. Today, Jim Cox will talk about research underway at
Tall Timbers Research Station that addresses some of the
key management questions associated with all three of these
rare and declining pineland birds.
Speaker Presentations 9
Grey Moss Plantation Overview
Jenny Crisp1
1
Grey Moss Plantation, Lee County, Alabama, 36849, USA
My name is Jenny Crisp. The name of our property is Grey
Moss Plantation. It is located in North Lee County and has
been in our family for three generations.
Our property is only 2000 acres, which is small in size as
far as plantations are concerned, but our property is diverse.
We have three natural water sources and two ponds. Due to
the abundance of water, we have natural hardwood areas,
cypress bottoms, and pine uplands. This makes our property a haven for upland and lowland wildlife. We have
deer, turkey, quail and countless other non-game species
including gopher tortoises.
We decided years ago to manage not only for timber and
wildlife, but also for aesthetic beauty. As you probably
realize from the mention of quail, deer and turkey not all of
these species have the same requirements for maximizing
their abundance. We are instead focused on being the best
blend possible for all aspects of our desired goals. This
requires knowing your property so you can apply the best
silvicultural and wildlife management practices to each different area of your property.
As landowners yourselves, you know property taxes are in
an upward spiral and there are long dry spells of money
flow between timber harvests. Even with the help of
CUVA (Conservation Use Value Assessment) property tax
season is no Christmas present. As I’m sure you already
realize, the best way to bridge the gap between paychecks is
with hunting leases.
Recreation is big business. Hunting leases can be quite
lucrative. So can renting a relaxing get-a-way in the country to city dwellers. You should keep all of these options in
mind when designing your property management plan.
Only by setting clear goals can you achieve a desirable outcome.
Now let’s focus on timber.
Although we have loblolly, slash and longleaf pines on our
property we are most proud of the longleaf. Our property is
in a natural longleaf area so longleaf trees have been on the
land for hundreds of years. I can remember in the not so
distant past when my father and I were concerned about the
survival of the longleaf species and lamented that we felt
the timber industry was over-looking the benefits of growing longleaf.
Thankfully the government instituted the 15-year CRP
contract for longleaf pines. Although the program got off
to a shaky start due to human error and the learning curve,
we are now on the way to a brighter forestry future. As
landowners, foresters, and contractors learned which sites
lent them to the longleaf; that containerized seedlings
gave better overall survival; and exactly how deep to plant
the soil plug; survival of the seedlings increased and cost
of planting as well as frustration decreased. These experiences are a milestone in themselves, but we have miles
left to travel.
I personally would like to see the CRP program modified
to allow landowners to plant greater numbers of trees per
acre to account for seedling mortality, disease and cankers, genetic inferiority, cotton rat attacks, deer damage,
lightening strikes, wind damage, and insect and beetle
infestation. Currently the program does not allow enough
seedlings to be planted to cover all the potential losses and
still produce a productive, genetically superior stand after
the initial thinning. Additionally, the sparse spacing creates lower octopus-like branches, which not only make
the trees ugly, they make them impossible to mow between. Burning can help make these trees self-prune but
you have to have enough fuel to carry a beneficial fire
without turning into a killing fire. Remember longleaf are
fire resistant, not fireproof and burning during candling is
a no-no.
In our experience of burning field planted CRP longleaf
you get hit and miss results. You only have a short window of opportunity. For us this is after hunting season
and before candling. Also our fields have varied types of
grasses, shrubs and brush. Some carry fire well, others
don’t. Remember not all your pines will candle at the
exact same time, so some of the earliest will be killed if
care is not taken.
Now I would like to address additional concerns associated with the longleaf pine. These are uneven release
from the grass stage within even aged stands and genetic
inferiority of available seedlings. Due to the fact that
landowner interest has only in recent years focused on the
longleaf species we are behind in the effort to provide
genetically superior seedlings to the market. This is a
disadvantage to those persons interested in growing longleaf timber.
Speaker Presentations 10
If you personally agree with the above statements please
pass these concerns on to your nursery stock providers
and governmental agencies. Remember the squeaky
wheel gets the grease. Advances in the industry take time,
research and money. We must push for all three to be a
viable part of today’s market.
Let’s talk about different treatments, which can be applied
to old-field CRP to affect natural pruning. First and easiest application; do nothing. Trees will continue to have
many octopus-like lower limbs for years. This will lower
the timber quality and therefore revenues. Second and
slightly more expensive; burn the stand. This method will
help to start the natural pruning process by killing the
lowest limbs. This will increase timber quality without
adding much cost. Third and most costly; trim or de-limb
the trees. With this method you send a crew of men
through the stand with loppers, pruning saws, and pole
saws to reach the higher limbs. They should leave only
the upper third of the tree untrimmed. This is expensive
and time consuming.
We use a combination of burning and trimming. I like
this method best because you can burn some of the lowest
limbs without having to mechanically trim them. This
saves time and money. Since we employ one full time
man and one part time man year round we use this work
as filler to keep them busy.
Finally, I want to address natural regeneration of longleaf
stands. In my experience there are two ways to achieve
this goal. First, you must cut the timber to a seed tree
stand leaving your best trees on an even spacing to disperse a good seed crop and wait for that 1 in 10 years
when the seeds are plentiful. Then time the cutting while
the seeds are falling. You want to get the seed trees off
the property before the seeds start to sprout. The timbering operation will disturb the soil to get a good seed bed
for the regeneration.
Next pray for enough rain to carry the tender seedlings
through the first season. After that you are on your way.
Remember you are at Mother Nature’s mercy so some
places may have too many seedlings survive and some
won’t have enough but, you will have a natural stand with
which to work. You can thin an over populated stand and
you can hand plant into an under populated stand. One
other thing to keep in mind is the seed crop rules the play
of the game, not the timber market. So you might sell
your seed trees for less than you would like to. You can
look at this as an offset to planting costs. The only hitch
is if your seedlings fail to survive.
where sunlight can hit the ground. Once you get a seedling catch, you withhold fire until the trees are big enough
to withstand burning. If the stand is too dense you can
burn to kill some of the smaller trees or you can mechanically thin the stand. This method will give you a mixture
of different aged trees in the stand hence the name uneven
pine stand management.
There are times when the circumstances of a naturally
regenerated longleaf stand require more relief than fire
can supply for overpopulation. This situation can be
caused by too little fuel to carry the fire, or spotty areas of
regeneration leading to too few trees in an area and too
many trees in an adjoining area. This causes hot fires
where you have too few trees and lots of grasses and very
little fire where you have too many trees and no grasses.
Using a Bobcat with a mulching head to cut rows through
the stand in thick areas and avoid the thin spots can solve
this potential problem. Although this is an expensive
process there are government programs to help defray the
costs. We thinned an area using this method with the
assistance of the Southern Pine Beetle Cost Share program. This is a relativity new process and, although I was
dubious about using it, our results have been good so far.
To conclude this talk I would like to focus on the good
traits of the longleaf pine. They have a high resin content; which makes them more insect resistant, increased
tonnage harvested per acre, produces more lightwood
stumps for potential profit and allows for turpentine harvesting when feasible. Longleaf is fairly ice tolerant considering their long needles. Their needles also bring a
premium as baled pine straw. They are a fire tolerant
pine species, which makes them the natural pine species
for the wiregrass region. They produce a higher percentage of poles per acre as compared to other pine species.
In my opinion, they are the most attractive pine species
grown in the Southeast. If pines and wildlife are your
interests, you should consider the longleaf pine. If you
like the old Deep South plantation look, you should consider the longleaf pine. If you like natural and graceful
beauty, you should consider the longleaf pine.
The longleaf pine is the complete package. If they were
women they would have beauty, brains and big dowries.
Certainly the longleaf pine has its place in history. It is
now making a comeback. With continued effort it can
survive and hopefully thrive into the future.
Another method of natural regeneration is a shelter wood
stand that allows for holes to be created in the stand
Speaker Presentations 11
Post-Plant Herbaceous Weed Control Timing Considerations for Longleaf Pine.
E. David Dickens1, Bryan C. McElvany1 and David J. Moorhead1
1
School of Forestry and Natural Resources, The University of Georgia, Tifton, Georgia, 31793 USA
Introduction
year survival ranged from 90% with the April 7th
Oust+Velpar L treatment to 40-65% with the May 9th herbicide treatments (Figure 1). All dead seedlings were
replanted (Dec 2000) in the May 9th treatments. The April
Oust+Velpar L treatment had significantly greater percent
trees out of the grass stage and significantly greater
heights at the end of the third growing season than the
100
90
a
80
70
Survival (%)
b
b
60
b
b
50
b
b
b
b
b
b
40
30
20
10
C
May A4P1.2
May A6P1.2
May A8
May A6
May A4
May A4O2
May OS13
May A4V24
May
A4OS6.5
0
April O2V24
Proper timing of herbicide applications is critical to obtain effective herbaceous weed control. Herbaceous weed
control (HWC) applications can be made pre-emergent
before weeds germinate or are still at the 1-2 leaf stage, or
post emergent following weed germination. Annual
variations in spring to early summer rainfall patterns influence herbicide efficacy within these application windows. Soil moisture is one of the most critical factors in
determining the effectiveness of a weed control treatment
as well as seedling survival and early growth. Examination of long-term precipitation records for the Coastal
Plain of Georgia can be useful in predicting treatment
application windows. Soil drainage class knowledge can
be tied in with historic rainfall patterns for a given area to
best estimate optimal a HCW treatment window. A study
area in Emanuel County, Georgia was installed to discern
the effectiveness of various herbicides and timing over
newly planted (Dec 1999) old-field longleaf pine (Pinus
palustris) on a moderately well to well drained Tifton
soil.
Treatment
Figure 1. First year survival at the Emanuel County
Georgia old-field longleaf site (Tifton soil).
Study Design
The experimental design was randomized complete block
with 4 replications of each treatment. There were a total
of 11 treatments, but only 7 were followed through six
growing seasons: control, a 7 April 2000 Oust @ 2 oz/ac
+ Velpar L @ 24 oz/ac, then a 9 May 2000 application of
Arsenal @ 4 oz/ac, Arsenal @ 6 oz/ac, Arsenal @ 4 oz/ac
+ Oust @ 2 oz/ac, Arsenal @ 4 oz/ac + Oustar @ 6.5 oz/
ac, and Oustar @ 13 oz/ac. The herbicide applications
were over the top banded in a 4 foot swath. Four rows of
ten living seedlings were wire flagged and numbered in
each plot (initially 1760 seedlings) with approximately
20” of buffer between plots. Each plot was revisited 6-10
weeks, 21-25 weeks and at the end of the first, second and
third growing season to determine % survival, % seedlings out of the grass stage, and mean height of those
seedlings out of the grass stage. Statistical analyses were
run on treatment means using SAS and one way analysis
of variance. Least squares means were compared using
Duncan’s Multiple Range Test (5% alpha level).
First Three Year Results
The early (April 7, 2000) Oust+Velpar L herbicide treatment gave greater initial survival, % trees out of the grass
stage, and height growth compared to nine later herbicide
treatments (May 9, 2000) or an untreated control. First
control and the May herbicide treatments. We did the
same 9 May HWC treatments as described above over
newly planted slash pine (mechanical site prep; shear; pile
and no bedding; set up for eventual pine straw raking) on
a poorly drained Pelham soil 2 mile north of the longleaf
site. In that case, the slash survival was acceptable ranging from 80 to 90% at the end of the first growing season.
Soil moisture on this poorly drained, non-bedded site did
not appear to be as critical into April and May as it was on
the better drained Tifton soil.
Six-Year Results
We returned to the site in March 2006 and re-established
the seven aforementioned treatment plots (April 2000
treatment, best 5 May 2000 treatments, and control), retagging only the interior two rows of trees in replications
2, 3, and 4. After six growing seasons, mean diameters,
heights (Figure 2), and green weight per acre were significantly greater with the 7 April application than the control
and 9 May treatments. Sixth year survival ranged from
78% in the 7 April application to 61% for the May 9th
applications to 41% for the control. In the May 9th treatments, originally planted trees had significantly larger
Speaker Presentations 12
17
a
16
height (ft)
15
Table 1. A guideline for HWC application timing based
on soil drainage class and herbicide type.
ab
14
b
b
b
b
b
13
Soil Drainage
Class
12
11
10
control
May A4
May A6
May A4O2
May OS13
May A4OS6
April OV
Timing - Treatment
Figure 2. Sixth year mean height at the Emanuel
County Georgia old-field longleaf site (Tifton soil).
heights and diameters than the replanted trees. Original
planted trees had an average diameter of 3.4 inches and
height of 16 feet. Replanted trees had an average diameter of 1.9 inches and height of 9 feet. No differences exist
between the May 9th treatments. During the spring of
2000, rainfall patterns were 25% of normal (1.79 versus
7.06 inches, Figure 3). It appears that the April 7th herbicide treatment allowed for the seedlings to survive this
critical dry period. These results indicate that substantial
establishment costs can be saved with an earlier herbicide
application under severe spring drought conditions.
Rainfall at this Site During the First Growing Season
and the Local Weather Station
Rainfall at the Emanuel County site (old-field, moderately
to well drained Tifton soil) during April and May 2000
was well below normal; 1.58 inches in April and 0.21
inches in May 2000 compared to the 50 year running average of 3.30 inches in April and 3.76 inches in May.
Over 80% of the first year mortality in the 9 May treated
Somewhat to
excessively well
drained
Moderately well
to well drained
Poorly to somewhat poorly
drained
Pre- to early
post emergence
herbicide application timeframe
Late Feb to
mid-March
Early post to
post emergence
herbicide application timeframe
March
March
Mid-March early
April
April to early
May
Mid-April to
mid-May
longleaf seedling plots was noted 6 weeks after treatment
(21 June 2000 first survival count) compared to the 7 April
Oust+Velpar L treatment mortality being 10% on 21 June
2000. Historically, April and May have been the two
months with the lowest rainfall for the first half of the year
for all the interior Georgia Coastal Plain (CP) weather stations. To confound the early growing season soil moisture
dilemma, the Plains, Tifton (Figure 4), and Statesboro
weather stations have had April and May rainfall amounts
less than to much less than the historical averages in 5 to 7
years in the period of 2000 through 2006. March rainfall
has historically been variable with record low rainfall
amounts in 2004 and 2006 (the 2nd and 3rd worst on record)
at some of the Georgia CP stations. For example, the Tifton
weather station had the following rainfall amounts during
March: 4.72” in 2000, 9.95” in 2001, 5.16” in 2002, 8.22”
in 2003, 0.42” in 2004, 6.53” in 2005, and 0.29” in 2006.
7
6
7
2000
5
2001
Rainfall (in)
6
Rainfall (in)
5
4
2002
4
2003
2004
3
2005
2006
50 Yr Avg
2000
3
2
50 Yr Avg
1
2
0
1
April
May
Month
0
Jan
Feb
Mar
Apr
May
Jun
Jul
Aug
Sep
Oct
Nov
Dec
Month
Figure 3. Vidalia, Georgia weather station rainfall for
the year 2000 and the fifty year average.
Figure 4. Tifton, Georgia weather station rainfall during
April and May 2000 through 2006 and the 50 year average.
Speaker Presentations 13
Timing Considerations for Post-Plant HWC over Longleaf (and other southern pines) for the Georgia Coastal
Plain
From these findings and recent findings by Yeiser and Ezell
(2006 unpublished), HWC treatment timing can be critical
in (1) controlling competing vegetation effectively
(generally herbaceous plants under stress are harder to control with herbicides then when these competitors have new
growth and are vigorously growing), (2) maximizing herbicide efficacy with early post to post emergence herbicides
(i.e., they work better on plants in the 1-2 leaf stage; once
plants get larger than the 1-2 leaf stage declines), and (3)
optimizing survival and early growth of planted pine
seedlings. Since April and May have been the two driest
months of the first ½ year and that the last 5 to 7 of the
seven years have been below the historical average for
these two months, early HWC applications would appear
to be more effective than later HWC applications. Table
1 may be used as a guideline based on soil drainage class
and type of herbicide used and Table 2 lists common
Georgia Coastal Plain soils by drainage class.
Table 2. A list of common Georgia Coastal Plains Soil Series by drainage class and subsoil type.
Drainage
Surface Depth
Very Poorly
0-10
Poorly to somewhat poorly
Somewhat to
excessively well
drained
None
Loamy
Clayey
Rutledge
Torhunta
Surrency
Bayboro
Chipley
Osier
Scranton
Rains
Lynchburg
Bladen
Coxville
Brady
20-40
Pelham
40-80
Albany
Plummer
Spodic—
Argilic Not
Present
Murville
Wesconnet
Rigdon
Ridgeland
Resota
Pactolus
Ortega
Goldsboro
Tifton
Dothan
Hurricane
Pottsburg
Faceville
Nankin
Greenville
10-20
Onslow
Seagate
20-40
Lucy
Fuquay
Stilson
40-80
Bonifay
40-80
Spodic—
Argilic Present
Mascotte
Sapelo
10-20
0-10
Moderately well
to well drained
Subsoil Type
Lakeland
Kershaw
Troup
Speaker Presentations 14
Baymeade
Echaw
Rimini
Kureb
Restoring Longleaf Ground Layer Vegetation on Private Lands
Jeff Glitzenstein1, Donna Streng1, and Jim Bates2 Mark Hainds3, Jill Barbour4, Joe Cockrell5
1
2
Tall Timbers Research Station, Tallahassee, Florida, 32312, USA
US Fish and Wildlife Service, West Georgia Ecological Services, Ft. Benning, Georgia, 31995, USA
3
Longleaf Alliance, Solon Dixon Forestry Education Center, Andalusia, Alabama, 36420, USA
4
USDA Forest Service National Seed Laboratory, Dry Branch, Georgia, 31020, USA
5
USDI Fish and Wildlife Service, Charleston, South Carolina, 29407, USA
Introduction
Restoration Approach and Philosophy
Longleaf ground layer vegetation is among the most
floristically rich in North America and is worthy of
preservation and restoration in its own right (Peet and
Allard 1993). In addition, it provides the trophic base
for a rich and unique arthropod community (Hermann et
al. 1998). Rare and threatened vertebrate populations
including flatwoods salamander (Ambystoma cingulatum), eastern indigo snake (Drymarchon corais couperi), gopher tortoise (Gopherus polyphemus), redcockaded woodpecker (Picoides borealis) and many
others may depend on the rich ground cover. Popular
game species including northern bob-white quail
(Colinus virginianus), turkey (Meleagris gallapavo) and
white-tail deer (Odocoileus virginianus) also thrive in
this cover type (Brockway et al. 2005). For these reasons restoration of diverse ground layer vegetation in
longleaf pine stands is an important objective of
USFWS, other state and federal agencies, and some
private landowners in southeastern USA (Brockway et
al., 2005, Roth et al, 2006).
The basis of longleaf ecosystem restoration is reestablishment of appropriate open woodland structure accompanied by, or followed by, re-initiation of prescribed
fire (Brockway et al. 2005). Mechanical or chemical treatments may often be useful if used cautiously (Brockway et
al. 2005). Nursery propagation and out-planting of longleaf
pine seedlings is required when this defining canopy tree is
not already established on the site (Brockway et al. 2005).
These basic steps are well recognized and accepted. Partners funds are available to help landowners with these activities.
A few longleaf understory restoration projects have
been implemented in recent years, mostly on federal
lands or Nature Conservancy Preserves (Roth et al,
2006). Substantial funding and expertise were involved
that would not be available to most private landowners.
USFWS Partners for Fish and Wildlife is a program of
USDI Fish and Wildlife Service that works with private
landowners to enhance native habitats for wildlife generally and rare species in particular (http://ecos.fws.gov/
partners/). An important goal of the program in southeast USA is to make available expertise and funding to
allow private landowners to carry out longleaf understory restoration projects. In cases where the project is
not funded through Partners we can still provide expertise and contractor contacts should the landowner
choose to provide the entirety of the funding. This article summarizes the approach and current status of longleaf ground cover restoration on a number of Partners
co-funded projects in AL, GA, and SC and on one privately funded project near Charleston, SC.
Less well understood is the need for active restoration of
ground layer plant communities (Roth et al, 2006). By active restoration we mean planting plugs or direct seeding to
put back species that have been lost from a site. It is often
assumed that such efforts will not be needed because the
majority of plant species will re-establish from buried seed
or disperse into the site from nearby areas (Brockway et al.,
2005). Observations of longleaf ground layer plants reappearing after disturbances, e.g. plantation establishment,
have been cited in support of this point of view (Walker
1998, Hedman et al. 2000, Brockway et al. 2005). However, such observations are more to likely indicate persistence of surviving plants or re-sprouts from buried rhizomes
than recruitment from seed. Recent data by Coffey and
Kirkman (2006) tend to contradict the assumed importance
of the seed bank in restoration. In this study the investigators buried seeds of characteristic perennial grass and forb
species of longleaf groundcover and then followed the fate
of the seeds over a four year period. Buried legume (plants
in the pea family) seeds did indeed persist in the soil, and
legume seedlings emerged from the seed bank during the
entire period (Glitzenstein et al. (2001), demonstrated similar, though shorter term, results). In contrast seed of major
non-weedy Asteraceae and Poaceae (plants in the sunflower
and grass families) did not have a long-lived seed bank.
Thus Coffey and Kirkman’s (2006) study strongly indicated
the need for active re-establishment of dominant perennial
Asteraceae and Poaceae in sites where these species had
disappeared due to past disturbances.
Speaker Presentations 15
Another point is that the seed bank itself may be disrupted
by intense disturbances that disrupt the soil profile. This
may explain why even within Fabaceae (i.e. legumes) there
may be meaningful variation in re-establishment potential.
Certain legumes, e.g. Lespedeza spp and Desmodium spp.,
do tend to reappear even after severe soil disturbances or
long-term fire suppression. Other legumes, e.g. Tephrosia
virginiana (Goat’s Rue), have a much lower capacity for
spontaneous recovery. Indeed, like wiregrass, a stand of
abundant Goat’s Rue is a good indicator for a history of
undisturbed soils and relatively pristine stand conditions.
Chigger Run Farm, one of our project sites in south GA
(Table 2), was a good example. The restoration site was a
young loblolly plantation with a history of frequent fire and
decent remnant ground layer diversity. Goat’s Rue itself
was, however, conspicuously absent from the restoration
stand despite its abundance in adjacent undisturbed mature
longleaf. The main goal of this project, or at least the Partners contribution, was to propagate and plant Goat’s Rue
back into the plantation area. This was accomplished in
January 2007 when, following canopy removal and site
preparation, we planted out ~ 1700 Tephrosia virginiana
plugs. Among other benefits, Tephrosia virginiana is a
long-lived and prolific seed producer, and thus an important
food of bobwhite quail.
As indicated by Coffey and Kirkman (2006) and the Goat’s
Rue example, we do believe that there is a need for active
ground layer restoration on many fire suppressed or disturbed sites. The Chigger Run example also emphasizes the
point that many “good” (i.e. characteristic perennials of
non-soil disturbed fire maintained pinelands) species may
sometimes survive in soil disturbed or even fire suppressed
stands. Thus it is extremely important to do a careful examination of any site before embarking on a restoration.
Remnant or residual populations are like money in the bank,
and taking care to protect or increase such populations
through use of appropriate site preparation (including burn
only if that is called for) will greatly simplify the ultimate
task of re-establishing diverse high quality understory.
The two main approaches to active restoration are: (1)
Grow plugs (tubelings, containerized seedlings) in the nursery and plant them out. (2) Direct seed, generally using notill drills adapted to handling fluffy seed. The first approach is more certain but is much more labor intensive.
The labor issue is mitigated presently by the ready availability of relatively low-cost planting crews. These crews
typically plant containerized pine seedlings but will also
plant perennial herbaceous species. We have attempted
direct seeding but the results are not yet certain. While this
method is often recommended as cheaper and easier to implement, the available evaluations seem based almost entirely on cover of a few dominant grasses (e.g. wiregrass,
little bluestem, Indiangrass) with little analysis of community level objectives (Roth et al. 2006). Consequently we
have relied mainly on the plug planting approach which
is nearly always successful if plugs are outplanted in fall/
early winter when conditions are cool and soil moisture
tends to be highest on dry upland sites (Glitzenstein et al.
2001, Roth et al. 2006). The scale, numbers and densities
of plugs planted out depends on the project budget and
landowner objectives. For low budget (i.e. < $10,000)
projects the goal is generally to establish a rather open
grid of dominant (“matrix”) grasses with the expectation
that the interstices will gradually fill in via rhizome expansion or seedling establishment. The target for forbs is
20-30 species with sufficient numbers to initiate new
populations that ought to increase over time given appropriate management. For bigger budget projects higher
numbers and densities are possible.
The issue of seed to use in restoration is at present somewhat controversial. From the standpoint of local adaptation and protection of local gene pools the safest approach
is to use seed locally collected from the nearest available
donor site (Hufford and Mazer 2003). In selecting a donor site it is also important to match ecological characteristics as closely as possible. We have tried to adhere to
this approach and have thus far collected all of our own
seed from nearby sites, although some plugs have been
moved up to 120 miles from the seed source. We have
thus far not utilized commercial seed and will not do so
even if marketed for southeastern USA unless the source
of the seed is explicitly identified by the producer and can
be determined to be local (i.e. within state and climate
zone) and appropriate for the project in question. Our
plugs are grown at several nurseries in FL, GA and SC
(Table 1). We are cooperating with the nurseries on
strategies to improve germination and efficiency of plug
production. Research to improve seed cleaning and testing is also being carried out at the USDA Forest Service
Seed Laboratory in Macon, GA (Barbour et al. 2007).
Table 1. Cooperating nurseries. If your nursery
would like to be included please contact Jim Bates,
Jeff Glitzenstein or Mark Hainds.
Nursery
Ownership
Location
ATS Partners
GA
Private
Bainbridge,
Andrews Nursery
FL
FL Division of Forestry
State
Chiefland,
Blanton Container Nursery Private
FL
Madison,
Deep South Nursery
GA
Private
Douglas,
Taylor Nursery
SC Forestry Commission
State
Taylor, SC
Speaker Presentations 16
Summary of Available Services
In sum, we, or our associated contractors, will provide the
following services to facilitate active ground layer restoration: (1) Evaluate potential donor sites and restoration site
from the standpoint of existing vegetation structure, composition and diversity. (2) Choose donor site. (3) Recommend site prep decisions to help conserve plant biodiversity and “climax” native grasses. (4) Collect seed, including machine harvesting as appropriate. (5) Direct seed if
situation warrants. (6) Grow and outplant plugs. (7)
Monitor plug survival. (8) If needed, help with follow-up
burns or other control activities.
Accomplishments to Date
Status and accomplishments for nine projects with meaningful progress are summarized in Table 2. Included are
four projects in AL, two in GA and three in SC. Ninety
one perennial plant species have thus far been utilized in
the projects, either via seed collection, propagation and/or
out-planting. Species are listed and utilization thus far
summarized in Table 3.
Acknowledgements
A number of folks were influential in each of the various
projects. These individuals and their affiliations are acknowledged in Table 2 in association with the project
name.
Literature Cited
Hedman, C.W., S.L. Grace and S.E. King. 2000. Vegetation
composition and structure of southern Coastal Plain
pinelands, an ecological comparison. Forest Ecology
and Management 134: 233-247.
Hermann, S. M., T. Van Hook, R. W. Flowers, L. A. Brennan, J. S. Glitzenstein, D. R. Streng, J. L. Walker, and
R. L. Myers. 1998. Fire and biodiversity: studies of
vegetation and arthropods. Transactions of the 63rd
North American Wildland and Natural Resource Conference:384-401.
Hufford, K.M. and S.J. Mazer. 2003. Plant ecotypes: genetic differentiation in the age of ecological restoration.
Trends in Ecology and Evolution 18(3): 147-155.
Peet, R.K. and D.J. Allard. 1993. Longleaf pine vegetation
of the Southern Atlantic and Eastern Gulf Coast regions: a preliminary classification. Proceedings Tall
Timbers Fire Ecology Conference 18: 45-82.
Roth, L. (editor), M.E. Barnwell, B. Beck, N.J. Bisset, J.K.
Hiers, L.K. Kirkman, C.S. Matson, G. Seamon, J.
Walker (major contributing authors). 2006 (final draft,
J.L. Walker, personal communication). Restoration of
ground layer vegetation in dry site longleaf pine communities: a working guide to practical methods.
Walker, J.L. 1998. Ground layer vegetation in longleaf pine
forests, an overview of restoration and management. In
Kush, J.S., comp. Ecological restoration and regional
conservation strategies. Longleaf Alliance Rep. 3.
Andalusia, AL: Solon Dixon Forestry Education Center: 2-13.
Barbour, J., V. Vankus, J. Glitzenstein, D. Streng, J.
Bates. 2007. Seed cleaning and germination testing
for the restoration of ground layer plants in a Pinus
palustris, Longleaf Pine, ecosystem. In Estes, B.L.
and Kush, J.S Proceedings of the Sixth Longleaf Alliance Regional Conference; November 13-16, 2006,
Tifton, GA. Longleaf Alliance Report No. 10.
Brockway, D.G. , K.W. Outcalt, D.J. Tomczak, and E.E.
Johnson. 2005. Restoration of longleaf pine ecosystems. Gen. Tech. Rep. SRS-83. Asheville, NC: U.S.
Department of Agriculture, Forest Service, Southern
Research Station. 34 pp.
Coffey K.L. and L.K. Kirkman. 2006. Seed germination
strategies of species with restoration potential in a
fire-maintained pine savanna. Natural Areas Journal
26:289-299.
Glitzenstein, J.S.; D.R. Streng, D.D. Wade and, J.
Brubaker. 2001. Starting new populations of longleaf pine ground-layer plants in the outer Coastal
Plain of South Carolina, USA. Natural Areas Journal
21: 89-110.
Speaker Presentations 17
Table 2. Private lands restorations for longleaf ground cover. All except Yeaman’s Hall project funded by USFWS Partners for Fish and Wildlife Program. Some of the property names are altered to protect landowner identity. Important
collaborators in each project are listed parenthetically beneath the property name. SDC = Solon Dixon Center, Andalusia, AL.
Landholding
Location
Acres
Status of Project
Russell County, AL
20
Kudzu dominated field at start of project. Seeds for restoration collected fall 2005/2006 onsite and in Tuskegee NF. Chemically treated
20 acres in 2005 and again in 2006 for removal of kudzu. Active
restoration not yet initiated.
Coffee County, AL
20
Clearcut and burned prior to project initiation. Seed collected from
Blackwater River State Forest donor site autumn 2002. Plugs planted
summer/fall 2003.
20
On-site donor site identified. Planted Saccharum spp (plumegrasses)
and Tridens ambiguus in wet disturbed site. Plans are to chemically
treat food plots infested with alien Lespedeza bicolor and plant native Lespedeza plugs currently in production at Andrews Nursery.
Alabama
Creek Stand Plantation
(Beau Dudley, Dudley Asset Management)
“Elba” Farm
Kowikee Creek Plantation
(Beau Dudley, Dudley Asset Management)
Pinecone Plantation
Russell County, AL
Intensive mechanical and chemical site preparation and prescribed
fire prior to project initiation. Seed collections from Blackwater State
Forest, west FL, 2004/2005 and SDC. Longleaf and groundcover
plugs planted for upland restoration during spring 2005-winter 2007.
Ground cover plugs planted for wet savanna/seep restoration winter
2007. Prescribed burns on sections of restoration areas March 2006
and January 2007.
Coffee County, AL
Georgia
Chigger Run Farm
20
Loblolly plantation with history of soil disturbance but decent burn
history. Pines clearcut, site prepared via Gyrotrac treatment fall
2006. Tephrosia virginiana seed collected June 2006, donor site
adjacent on same property. Approximately 1700 plugs Tephrosia
virginiana planted out January 2007.
20
Seven fields dominated by broomsedge (Andropogon virginicus) and
other native early successionals but with patches of potentially troublesome invasive aliens including Coastal Bermudagrass. Chemically treated with Roundup (glyphosate) during fall 2005 and subsequently burned. Donor sites included good quality mesic pineland on
same property and SDC. Seed collected fall 2005/2006. Plugs
planted December 2005 through February 2007.
Bamberg County, SC
50
Acquired by SCNPS from TNC in 2004. At that time the site consisted of high quality wetland depression with federally endangered
Oxypolis canbyi (Canby’s Dropwort) surrounded by fire suppressed
loblolly pine plantation. The latter is being restored to longleaf,
wiregrass and associates, in part to restore the natural fire regime for
the benefit of the dropwort population. Plantation loblolly logged in
autumn 2005. Pescribed burns in February and November 2006.
Plugs of longleaf pine, ~40,000 wiregrass and ~10,000 other species
planted in November 2006.
Clarendon County, SC
50
Dr. Porcher is directing his own restoration with Partners funding.
Glitzenstein is cooperating in growing and planting out federally
endangered Schwalbea americana.
Berkeley County, SC
50
Loblolly plantation logged fall 2006, Most of site chemically treated
with Chopper (Imazapyr) fall 2006. Prescribed burned December
2006. Seeds collected fall 2006, plug propagation target 50,000
plugs, ~30 ground layer species. Privately funded.
Thomas County, GA
“Five Chimneys” Plantation
(Beau Dudley, Dudley Asset Management)
Stewart County, GA
(Smoky Bartlett, Land Manager)
South Carolina
Lisa Mathews Memorial Bay, South
Carolina Native Plant Society
(John Brubaker, SCNPS; Guy San
Fratello, SanBar Forestry and Wildlife
Services).
Pocotaligo Plantation
(Richard Porcher, owner, retired professor of botany, the Citadel)
Yeaman’s Hall Club
(Lisa Lord, YHC naturalist/ecologist)
Speaker Presentations 18
Table 3. Native perennial herbaceous species of longleaf pine ground layer utilized in the various restoration projects.
Codes: C = seed collected, T= seed cleaned and germination tested, P = plugs propagated, O = plugs out-planted. (R
following species name indicates globally rare and endangered). Nomenclature follows Kartesz (1994). Species with
low numbers (<20) of out-plants were, for the most part, collected as part of mixed seed collections. “In propagation”
refers to plugs presently in propagation trays maturing or waiting to be out-planted.
Species
Common Name
Activity
Plugs planted
(regular text) or
in propagation (italics)
Ageratina aromatica
lesser snakeroot
CP
450
Agrimonia incisa (R)
incised groovebur
CP
50
Agrimonia microcarpa
smallfruit agrimony
C
Amsonia ciliata
fringed bluestar
C
Andropogon gerardii
big bluestem
CPO
8
Andropogon gyrans
Elliott’s bluestem
CPO
300
Anthaenantia rufa
purple silkyscale
CPO
3
Aristida beyrichiana
wiregrass
CTPO
70000
Aristida condensata
piedmont threeawn
CPO
180
Aristida lanosa
woolysheath threeawn
CPO
100
Aristida purpurascens
arrowleaf threeawn
CPO
200
Aster adnatus
scaleleaf aster
CPO
9
Aster concolor
Eastern silver aster
CPO
50
1250, 250
Aster tortifolius
Southern aster
CTPO
Baptisia albescens
spiked wild indigo
C
Baptisia lanceolata
gopherweed
CTPO
80
Baptisia perfoliata
catbells
CPO
106
Baptisia bracteata
creamy wild indigo
CPO
10
Brickellia eupatorioides
false boneset
C
Carphephorus bellidifolius
bluntleaf deerstongue
CPO
Carphephorus odoratissimus
vanilla leaf
CPO
80
Centrosema virginianum
spurred butterfly pea
CPO
212
Chasmanthium sessiliflorum
longleaf dpikegrass
CP
1500
Chrysopsis gossypina
cottony goldenaster
CTPO
30
Chrysopsis mariana
Maryland goldenaster
CP
1500
600
Coreopsis linifolia
Texas tickseed
CPO
30
Coreopsis major
greater tickseed
CTPO
120
Crotolaria rotundifolia
rabbitbells
CPO
170
Ctenium aromaticum
toothache grass
CPO
106
300
Dalea pinnata
summer farewell
CPO
Desmodium ciliare
hairy small-leaf ticktrefoil
CPO
130
Desmodium spp.
beggar’s lice
CTPO
1700
Elephantopus elatus
tall elephant’s foot
CPO
50
Eragrostis spectabilis
purple lovegrass
CPO
20
Eriocaulon decangulare
Tenangle pipewort
CPO
2
Eriogonum tomentosum
dogtongue wild buckwheat
C
Eryngium yuccifolium
rattlesnake master
CPO
5
Eupatorium album
white thoroughwort
C
5
Eupatorium coelestinum
mistflower
C
Eupatorium hyssopifolium
hyssop leaf thoroughwort
CT
Eupatorium pilosum
rough boneset
CPO
Speaker Presentations 19
1
Table 3. (cont.)
Eupatorium rotundifolium
roundleaf thoroughwort
CP
Galactia macreei
downy milkpea
CPO
50
Helianthus atrorubens
Appalachian sunflower
CP
1350
Helianthus divaricatus
spreading sunflower
C
Helianthus occidentalis (R)
naked-stem sunflower
CPO
200
Helianthus radula
roundleaf sunflower
CPO
500
Helianthus resinosus
woodland sunflower
C
Ionactis linariifolius
flaxleaf aster
COP
Lespedeza capitata
roundhead lespedeza
CTP
3000
Lespedeza hirta
hairy lespedeza
CTPO
30, 1200
50, 1400
Liatris elegans
pinkscale gayfeather
C
Liatris gracilis
slender gayfeather
CPO
800
Liatris graminifolia
grassleaf gayfeather
CP
100
Liatris secunda
piedmont gayfeather
CT
Liatris spicata
dense gayfeather
CPO
20
Liatris tenuefolia
shortleaf gayfeather
CPO
1000
Mimosa quadrivalvis var. angustata
sensitive brier
CTPO
500
Panicum anceps var. rhizomatum
beaked panicum
CPO
4200
Parnassia caroliniana (R)
grass of parnassus
CPO
1, 20
Paspalum bifidum
pitchfork crowngrass
C
Paspalum floridanum
Florida paspalum
C
Paspalum setaceum
thin paspalum
CPO
1
Pityopsis graminifolia
grassleaf goldenaster
CTPO
2500, 2000
Rhexia alifanus
common meadowbeauty
CPO
40
Rhynchospora latifolia
white-top sedge
CPO
60
Saccharum alopecuroides
silver plumegrass
CPO
3000
600
Saccharum giganteum
giant plumegrass
CPO
Sarracenia leucophylla
white-topped pitcher plant
CPO
60
Schizachyrium scoparium
little bluestem
CPO
2700
Schwalbea americana (R)
American chaffseed
CPO
100
Silphium compositum
kidneyleaf rosinweed
CPO
14
Silphium asteriscus
starry rosinweed
C
Solidago odora
anise-scented goldenrod
CPO
Sorghastrum elliottii
slender Indiangrass
CPO
3350, 500
4200
Sorghastrum nutans
Indiangrass
CPO
1640
320
Sorghastrum secundum
lopsided indian grass
CPO
Strophostyles umbellata
pink fuzzybean
C
Tephrosia hispidula
sprawling hoarypea
C
Tephrosia spicata
spiked hoarypea
C
Tephrosia virginiana
goat’s rue
CPO
1800
Tridens ambiguus
pinebarren fluffgrass
CPO
500
Tridens carolinianus (R)
Carolina fluffgrass
C
C
Tridens flavus
Purpletop
Verbasina aristata
coastal plain crownbeard
C
Vernonia angustifolia
tall ironweed
CTPO
100
Xyris ambigua
coastal yelloweyed grass
CPO
25
Xyris scabrifolia (R)
Harper’s yelloweyed grass
CPO
1
Speaker Presentations 20
Reptiles and Amphibians of Longleaf Pine Forests: Update on Some Conservation Issues
Craig Guyer1
1
Department of Zoology, Auburn University, Alabama, 36849, USA
Abstract
Amphibians and reptiles are abundant, diverse, and
therefore, important components of longleaf pine forests. In places where maintenance of this diversity is a
management objective an examination of recent information suggests key components of the ecosystem that
provide specific characteristics necessary to support the
herpetofauna. In this paper, I review features of amphibians and reptiles that lead to their diversity in longleaf pine forests and describe the size and qualities of
areas needed to maintain this diversity.
Introduction
The longleaf pine ecosystem harbors a remarkable diversity of amphibians and reptiles. Many of these are
restricted in their distributions to this forest type and
appear to have evolved in it. One of these, the gopher
tortoise, is thought to be a keystone species of longleaf
because of the effect that the burrows created by these
terrestrial turtles have on maintenance of plant and animal diversity in the forest. Because the region has experienced intense land use activities by humans, the
ancestral landscape has changed dramatically for the
native amphibians and reptiles, resulting in a growing
list of species that receive federal protection under the
Endangered Species Act, are candidate species under
that legislation, or are listed by state law as being of
conservation concern. In this paper, I review selected
upland and wetland species of amphibians and reptiles
that should be of interest to land owners in the longleaf
region and review habitat management methods that can
be used to retain these organisms in the landscape. Finally, I review recent developments in conservation of
gopher tortoises (Gopherus polyphemus) and suggest
management objectives that might reduce the need to
list this species throughout its range.
Some portions of the longleaf pine ecosystem are created by deep sand ridges that are characterized by extremely low soil moisture and fertility but in which animals can dig easily. Because amphibians and reptiles
have low resting metabolism (because they are ectothermic), they are able to survive and diversify in such habitats in ways that vertebrates with high resting metabolism (endothermic birds and mammals) cannot. For this
reason there are six upland reptile species that deserve
special consideration by those who list maintenance of
the ancestral fauna of the region as a management goal.
These species are the gopher tortoise, indigo snake
(Drymarchon corais couperi), Florida pine snake (Pituophis
melanoleucus migitus), Louisiana pine snake (Pituophis
ruthveni), southern hognosed snake (Heterodon simus), and
eastern diamondback rattlesnake (Crotalus adamanteus).
The first two species receive federal protection, the next
three are candidate species for federal protection, and the
last species is widely recognized as deserving protection,
except for the fact that it is venomous. These species all
require a consistent set of landscape features to maintain
their populations.
Summary of Some Conservation Issues
Primary among important features of the landscape is retention of an open canopy, allowing penetration of light to the
understory. The ground cover vegetation must be dominated by grasses and forbs to provide a food base for herbivorous species, like the gopher tortoise, and to enhance
insect abundance that serves as the prey sources for other
vertebrate species that form the primary diet of predators
like the five species of snakes. Indigo snakes consume
other snakes, the two species of pine snakes and the rattlesnake largely consume rodents, and the hognosed snake
principally consumes toads. Maintenance of an open canopy is the single feature that will create the broadest impact
on retaining the forage base of these reptiles that are endemic to the xeric ridges of longleaf pine. A second important landscape feature is the presence of abundant stump
holes and downed logs. These provide refuges from predators, hibernacula for overwintering, and nest sites for reproduction. Gopher tortoises create their own refuges by digging burrows that can be used by all four species of snakes
during their yearly activities. However, stump holes are
frequently used by pines snakes and rattlesnakes as refuges
and these are at least as important as gopher tortoise burrows in the lives of these snakes. Management strategies
that fail to implement thinning and prescribed fire or a surrogate for fire (judicious use of herbicides; mowing of thick
shrub) and that disturb the understory vegetation (bedding;
creation of wind rows) typically are deleterious to the target
reptiles. A second vital feature of the longleaf landscape is
fish-free wetlands, principally because these habitats provide reproductive sites for amphibians. The flatwoods salamander (Ambystoma cingulatum), dusky gopher frog (Rana
sevosa), pine barrens treefrog (Hyla andersoni),
Speaker Presentations 21
striped newt (Notophthalmus perstriatus), mimic glass lizard (Ophisaurus mimicus) are five key taxa associated with
these wetland sites and that are of conservation concern.
The first two are federally listed under the Endangered Species Act, the third was recently de-listed but is regulated by
the states of Alabama and Florida, the fourth is protected by
state laws in Georgia and Florida, and the fifth is a rare,
recently described species that deserves conservation attention. These species all require wetlands that lack of fish,
have undisturbed drainage patterns, and connect to upland
habitats of high quality (see above). Sinkhole ponds (dusky
gopher frog, striped newt), Carolina bays (striped newt),
pitcher plant bogs (pine barrens treefrog), and flatwoods
(flatwoods salamander), and other such seasonally-flooded
sites are the primary reproductive sites of the four amphibians. The single most important habitat feature for retaining
the amphibians is exclusion of fishes from the reproductive
sites because fish can consume the entire production of eggs
and larvae. These wetlands all dry frequently enough that
fish cannot colonize them without the aide of humans who
frequently stock fish to expand opportunities for sport fishing. If the reproductive sites can be maintained, then all
five species require appropriate habitat structure. In general, this means use of prescribed fire to maintain open
habitat in the wetlands as well as the uplands surrounding
the wetlands. Stand thinning is an important tool, but this
must be done in a fashion that does not alter the drainage
patterns that accumulate waters. Mechanical site preparation must be avoided within the wetlands themselves if the
habitat is to retain the five target species.
These will require as little as 50 ha and as many as 1500
ha, depending on the quality of the forest structure.
Within such sites, tortoise introductions must include all
age categories to ensure population viability. The spatial
distribution of public lands upon which tortoises might be
maintained is already sufficient to cover ancestral distribution of gopher tortoises and these lands likely will be
the primary focus on tortoise conservation. Regardless of
whether tortoise conservation occurs on public or private
lands, maintenance of habitat quality through thinning,
use of frequent low-intensity fires during the growing
season, and constant monitoring of response of the tortoise populations will be required.
Gopher tortoises are a crucial feature of longleaf pine forests that are managed for maintenance of the ancestral
fauna. The species currently receives federal protection in
all areas west of the Mobile-Tombigbee drainage. However, a proposal to list the species throughout its geographic
range is receiving serious consideration by Fish and Wildlife Service. This is because of loss of upland habitat, restriction or improper use of fire as a management tool, fragmentation due to urbanization, disease, and human predation. In order to avoid the need for listing, the US Department of Defense recently implemented a memorandum of
understanding (MOU) of understanding among important
stakeholders within the range of the gopher tortoise. The
intent is to begin a process of action that will create areas
that will be managed to the benefit of tortoises so that viable populations will be retained across the current geographic range of the species.
Associated with this effort is generation of data that will
allow establishment of reserve areas where viable populations of gopher tortoises can be established by moving them
from places where tortoise densities are too low to indicate
population viability. Recent data collected from my lab
suggest that such reserves should target 100-150 animals as
constituting a minimum viable population.
Speaker Presentations 22
Benefits and Costs of Herbaceous Release Treatments in Longleaf
Pine Establishment and Management
M.J. Hainds1
1
The Longleaf Alliance, Solon Dixon Center, Andalusia, Alabama, 36420, USA
Table 3. 2002- Monroe Herbicide Screening Trial
Abstract
Longleaf survival rates one year post planting were
negatively affected by the majority of labeled herbicides
applied “over-the-top” of newly planted seedlings. The
primary benefit of herbaceous release treatments appears to be increased growth of newly planted seedlings, with the majority of labeled release treatments
leading to increased growth of planted longleaf pine
seedlings. Survival and growth of rates of longleaf are
examined three-to-five years post-application.
The Longleaf Alliance has installed numerous herbicide
screening trials across the Southeastern, US. In these
studies, The Longleaf Alliance has tested virtually
every herbicide labeled for herbaceous release over
longleaf pine seedlings at different timings and at varied
rates. Some examples of herbicide applications and
studies are:
Table 1. 1997 Herbicide Screening Trial (Bareroot)
Product (oz/acre)
Date Applied
Herbicide
Oz. of Prod.
Timing
Oust XP
3
3/28/2002
Oust Alt Form
3
3/28/2002
Oustar
10
3/28/2002
Oustar
13
3/28/2002
Oust Alt Form/Velpar DF
2.0 & 16.0
3/28/2002
Oust Alt Form/Velpar DF
3.0 & 8.0
3/28/2002
Velpar DF
10.67
3/28/2002
Velpar DF/Oust XP
10.67 & 2.0
3/28/2002
Velpar DF/Oust Alt Form
10.67 & 2.0
3/28/2002
Untreated Check
Oust XP/Velpar DF & Arse- 2.0 & 10.67 & 5.0 3/28/02 & 4/23/02
nal
Oust XP
2
3/28/2002
Arsenal
4
3/28/2002
Arsenal
6
3/28/2002
Arsenal
8
3/28/2002
Arsenal/Pendulum
4.0 & 3.33 lb
3/28/2002
Arsenal/Pendulum
4.0 & 3.33 lb
3/28/2002
Arsenal/Pendulum
8.0 & 3.33 lb
3/28/2002
Arsenal/Oust
5.0 & 2.0
3/28/2002
Arsenal/Oust
5.0 & 2.0
4/23/2002
#1
Check (No Chemical)
#2
Accord 16 oz
20-May-97
#3
Accord 20 oz
20-May-97
#4
16 oz Accord 2 oz Oust
20-May-97
#5
5 oz Arsenal 2 oz Oust
20-May-97
#6
6 oz Arsenal 2 oz Oust
22-Apr-97
#7
2 OZ Oust
22-Apr-97
#8
4 OZ Oust
22-Apr-97
Product oz/acre
Timing-Application
#9
22-Apr-97
Check
N/A
22-Apr-97
Vel. DF 10.7 / Oust 2
4/7/1999
#11
10.72 oz Velpar
1.5 Oust+ 10.72 oz
Velpar
2.5 Arsenal
20-May-97
#12
5 oz Arsenal
20-May-97
#13
7.5 oz Arsenal
20-May-97
#14
Escort .5 oz
20-May-97
#15
Escort 2 oz
20-May-97
Oust 2 & Ars. 4
4/7/99 & 5/12/99
#16
Escort 1 oz
20-May-97
Fusilade 24
4/7/99 & 5/12/99
Velpar DF 21.34
5/12/1999
Velpar DF 10.67
4/7/1999
#10
Table 2. 1999 Herbaceous Releases –Old Pecan
Orchard (Container)
Oust 2 & Ars. 4
4/7/1999
Arsenal 4 / Oust2
4/7/1999
Arsenal 4 / Oust 2
5/12/1999
Atrazine 64
4/7/1999
Atrazine 64 / Oust 2
4/7/1999
Speaker Presentations 23
May 1st. Typically this is applied with 4-5 ounces of Arsenal® and 2 ounces of Oust®.
Additional Herbicide Research in 2003
Denton, Georgia (Old Field)
Milledgeville, Georgia (Abandoned Ag Site)
Lexington, South Carolina (Cutover)
Andalusia, Alabama (Cutover)
From these studies and trials, The Longleaf Alliance
found two treatment regimes that consistently rate among
our best herbaceous release treatments on agricultural
sites. These two treatments are:
#1 The Split Treatment, which is two ounces of Oust®
applied between March 15 – April 15, followed by 4-6 oz
Arsenal after May 15. For example, the Split Treatment
ranked or yielded:
Best of 11 treatments in 1999 trial
5-7% mortality on Samson Site
#2 Alternatively, the Arsenal® Oust® Tankmix is a good
single application which we recommend be applied after
Table 4. 2002- Samson Herbicide Screening Trial
(Container)
Chemical
Oz. of Prod.
Timing
Oust XP
3
3/27/2002
Oust Alt Form
3
3/27/2002
Oustar
10
3/27/2002
Oustar
13
3/27/2002
Oust Alt Form/Velpar DF
2.0 & 16.0
3/27/2002
Oust Alt Form/Velpar DF
3.0 & 8.0
3/27/2002
Velpar DF
10.67
3/27/2002
Velpar DF Oust XP
10.672. & 2.0
3/27/2002
Velpar DF/Oust Alt Form
10.67 & 2.0
3/27/2002
Untreated Check
Oust XP/Velpar DF & Arsenal 2.0 & 10.67 & 5.0
3/27/02 & 4/22/02
Oust XP & Envoy
2.0 & 34.0
3/27/02 & 4/22/02
Arsenal
4
3/27/2002
Arsenal
6
3/27/2002
Arsenal
8
3/27/2002
Oust XP & Arsenal
2.0 & 4.0
3/27/02 & 4/22/02
Oust XP & Arsenal
2.0 & 6.0
3/27/02 & 4/22/02
Oust XP & Arsenal
2.0 & 8.0
3/27/02 & 4/22/02
Arsenal/Oust XP
5.0 & 2.0
3/27/2002
Arsenal/Oust XP
5.0 & 2.0
4/22/2002
The Arsenal/Oust Tankmix (Post-emergent) ranked or
yielded:
2nd best in 1997 (out of 16 treatments)
3rd best in 1999 (out of 11 treatments)
5% mortality in 2002 on Samson Site
Tied for #1 in 2002 on Monroe Site (out of 20 treatments)
Not releasing seedlings is always an option one should
consider. Not releasing seedlings sometimes yields the
best survival at age one. We have found that it is very
important to check for root growth before applying soil
active herbicides. If little or no new root growth is present, consider postponing or skipping a herbaceous release
treatment. In terms of survival at year one, the “Check”
or “No Herbaceous Release Treatment” ranked or yielded:
Best of 16 treatments in 1997 bareroot study
1999 Site Prep and Herbicide Study
Scalping Site Prep =8th out of 11 treatments
Chemical Site Prep = Worst out of 11 treatments
Check (Rip Only) Site Prep = 4th Worst out of 11 treatments
Tied for 1st out of 20 treatments in 2002 Monroeville
Study.
Tied for 1st out of 20 treatments in 2002 Samson Study
Overall, we have found that herbaceous release treatments
rarely increase survival. Examining dozens of different
treatments, on average, we reduced longleaf survival 3 out
of 4 times (74%) by applying a herbaceous release treatment. Even after we removed treatments that were off
label (Atrazine/Oust, Escort), we still reduced survival
70% of the time. The main benefit of herbaceous release
treatments appears to be increased growth. On average,
we increased height growth by applying a herbaceous
release treatment with 4 out of 5 treatments (80%).
In one of the first herbaceous release studies installed by
The Longleaf Alliance, bareroot longleaf pine seedlings
were planted on a cutover site and released with a Velpar/
Oust tank mixture in 1995. 4 ½ growing seasons later in
the untreated (not released) plots, seedlings averaged 5.4’
in height, 67% survival, and 11% of surviving seedlings
were still in the grass stage. Bareroot seedlings that had
been released one time in the first growing season with
the Velpar/Oust tank mix averaged 9’ height, 63% survival, and only 4% in the grass stage.
These results were consistent with many of our other studies, where we fairly consistently increased heights of
longleaf seedlings with herbaceous release during the first
growing season. However, we rarely increase survival
with herbaceous release treatments, and we believe that
survival rates can be improved through good site preparation treatments prior to planting.
Speaker Presentations 24
Frequency and Seasonality of Prescribed Fire Affect Understory Plants
James D. Haywood1
1
USDA Forest Service, Southern Research Station, Pineville, Louisiana, 71360, USA
Prescribed fire applied repeatedly over a number of
years can profoundly change forest structure and the
productivity of the understory. Both the frequency at
which fires are applied, whether annually, biennially, or
triennially, and the season of burning, whether in
March, May, or July, are believed to be important.
However, proving these differences requires the installation and monitoring of field studies over many years.
Herein, I present the results from four long-term studies
that support the belief that the frequency and season of
burning affect herbaceous plant productivity and the
basal area and richness of woody vegetation.
In the first study, prescribed fire was applied over 20
years in a direct seeded stand of longleaf pine (Pinus
palustris P. Mill.) (Haywood and Grelen 2000). Evidently, not applying fire or any other vegetation management treatment allowed natural loblolly pine (P.
taeda L.) and hardwoods to occupy the unburned plots
to the detriment of the longleaf pine regeneration and
the understory vegetation (Table 1). Conversely, the
understory production on the biennial March-burn plots
is the most productive and the overstory basal area is
the lowest. The triennially burned plots have less understory productivity and more hardwood basal area
than the biennially burned plots regardless of the season
of burning. Season of burning also affected productivity; both the biennial and triennial May-burn plots have
productive understories despite high overstory basal
areas. A sparse hardwood midstory on the four burning
treatments was doubtlessly a factor in maintaining productive herbage layers.
The second study started in a four-year-old slash pine
(Pinus elliottii Engelm.) plantation (Grelen 1983). Although the prescribed fires kept brush suppressed over
the next 8 years, on the unburned plots, fire-intolerant
species such as eastern baccharis (Baccharis halimifolia
L.), flowering dogwood (Cornus florida L.), American
holly (Ilex opaca Ait.), and sassafras (Sassafras albidum (Nutt.) Nees) flourished. Blackberry (Rubus
spp.) grew into impenetrable thickets in places, and
natural loblolly pines grew as fast as the planted slash
pines. Although the pine basal area on the unburned
plots was comparable to most of the burned treatments
at stand age 12 years, the brush that developed suppressed herbaceous plant productivity on the unburned
plots (Table 2). The annual March-burn treatment had
the most herbaceous plant productivity and the lowest
pine basal area. As the interval between prescribed
Table 1. Stand characteristics after 20 years of prescribed
burning; initially, the site was directed seeded in November 1968, prescribe burned 16 months later in 1970, and
the overstory trees were felled in early 1973 to form an
even-aged stand of longleaf pine regeneration.
Treatments
Unburned after
1970
Biennial March
burns
Triennial March
burns
Basal areas
All Un- Longleaf
Loblolly
Hardderstory
pine
pine
woods
vegetation
(lb/acre) (ft2/acre) (ft2/acre) (ft2/acre)
<1
6
143
13
1,810
40
0
<1
219
78
0
2
Biennial May
burns
491
93
8
1
Triennial May
burns
303
106
1
3
fires increased, herbaceous productivity decreased. The
May-burn plots maintained high herbaceous productivity
and high pine basal areas.
Longleaf pine trees in the third study originated from natural regeneration. In 1962, all pine and hardwood trees and
shrubs above one-ft tall were severed and removed to help
create uniform cover conditions over the entire area
(Haywood et al. 2001). However, scattered longleaf and
loblolly pines outside the study area continued to be seed
sources. Prescribed fire ceased on the unburned plots in
1961. The plots were mowed and raked in 1962 and 1963
as part of a simulated grazing study, but no further treatments were applied after 1963. The remaining plots were
burned from 1962 through 1998. After 37 years, the herbaceous plant community was nearly eliminated on the unburned plots due in large part to a well-developed hardwood
midstory, a large number of hardwood trees and shrubs in
the understory, and accumulated litter that smothered the
herbage (Table 3). Conversely, the July-burn plots had the
highest herbaceous plant productivity, the lowest pine basal
area, no hardwood midstory, and a sparse hardwood understory. The March-burn treatments had the lowest
Speaker Presentations 25
herbaceous productivity among the three burning treatments, but it also had a high pine basal area and the most
understory trees and shrubs of greatest stature among the
three burning treatments. The May-burn plots maintained a
productive herbaceous community, the highest pine basal
area, and a sparse hardwood understory. The lack of a
hardwood midstory on the burning treatments was doubtlessly a factor in maintaining productive herbage layers.
Table 2. Stand characteristics after 8 years of prescribed
burning; initially, the site was a 4-year-old slash pine
plantation when prescribed burning began.
Herbaceous
Pine
Treatments
vegetation
basal area
Unburned
Annual March burns
(lb/acre)
183
1,002
(ft2/acre)
94
75
Biennial March burns
730
84
Triennial March burns
346
87
Annual May burns
581
99
Biennial May burns
513
97
Triennial May burns
413
93
In the third study, the hardwood community in the understory on the March-burn plots was different from the one
on the May- and July-burn plots (Haywood et al. 2001).
On the March-burn plots, blueberries (Vaccinium spp.),
waxmyrtle (Morella cerifera (L.) Small), shining sumac
(Rhus copallinum L.), blackberry, and sassafras were
common, but were sparse on the other two prescribed
burned treatments. Southern red oak (Quercus falcata
Michx.) was common on the May-burn plots, but was
sparse on the March- and July-burn plots. On-the-otherhand, blackjack oak (Q. marilandica (L.) Muenchh.) was
found on all treatments. From the thirteenth through
thirty-seventh growing season, American beautyberry
(Callicarpa americana L.), flowering dogwood, white
oak (Q. alba L.), and New Jersey tea (Ceanothus americanus L.) were eliminated from all three prescribe burned
treatments (Grelen 1975, Haywood et al. 2001). Prescribed fire was clearly having an effect on the understory
woody vegetation in terms of number of stems, plant stature, and species richness (Table 3).
In the fourth study, prescribed fires were applied biennially in a longleaf pine plantation in March, May, or July
beginning in the seventh growing season (Haywood
2002). When the study began, the understory was dominated by grasses with low scattered brush due to weeding
treatments applied over the entire area in the fifth and
sixth growing seasons. In addition, I installed an untreated control (no more treatments after the fifth growing
season) and a biennial chemical weeding treatment. By
the fourteenth growing season, the herbaceous plant
Table 3. Stand characteristics after 37 years of prescribed burning, prescribed fire ceased on the unburned
plots in 1961, but the other plots continued to be burned
from 1962 through 1998.
Herbaceous
vegetation
Pine
Trees &
Hardwoods
AverHeig
age
ht
shrubs
(lb/acre)
(ft2/acre)
(ft2/acre)
(stems/
(ft.)
acre)
Unburned
after
1961
11
80
36
8,000
Biennial
March
burns
839
97
…
15,350 2.1
Biennial
May
burns
907
132
…
2,950
1.1
Biennial
July
burns
1,232
66
…
4,400
1.2
Treatments
3
community had collapsed on the untreated and chemically
weeded plots (Table 4). An accumulation of litter in the
absence of burning was the main reason for the collapse in
herbaceous cover, although a greater longleaf pine basal
area on the untreated and chemically weeded plots than on
the three prescribed burned treatments was a contributing
factor. In addition, percentage of tree and shrub cover in
the midstory and understory of the untreated plots had a
further adverse effect on grass productivity. Woody vine
cover was greater on the two unburned treatments than on
the three burning treatments.
Fire intensities in the May burns were lower on average
than in the March and July burns; the latter two averaged
similar intensities (Table 4). This difference in intensities
may help explain treatment differences. However, it
should be noted that the fire intensity experienced in the
fourth study were four times greater than the threshold of
50 BTU/sec/ft recommended for low intensity fires, and
such high intensities do not always result from prescribed
burning.
July burning was associated with higher grass and forb
cover than March or May burning. Once again, May
burning was associated with a higher pine basal area than
either March or July burning but cover of grasses and
forbs was still high on the May-burn plots. March-burn
plots had the lowest basal area of the three burning treatments but the cover of grasses and forbs was no better
than for the May-burn treatment.
Speaker Presentations 26
In conclusion, it is difficult to prove treatment differences
in herbaceous productivity and percentage of cover in
prescribed fire studies. However, across a series of studies, I was able to demonstrate that basal area of the overstory and understory woody vegetation is influenced by
the fire regime, and basal area in turn influences herbaceous plant productivity and cover, especially if a hardwood midstory develops. Frequency of burning is inversely related to herbaceous productivity. Burning in
May rather than March or July resulted in better longleaf
pine growth while maintaining a productive herbaceous
community. March burning was associated with more
woody understory plants and greater richness than May or
July burning, which is counter productive if herbaceous
vegetation is the primary concern. Therefore, spring
would be a better time to prescribe burn than late winter
or summer in order to have acceptable overstory growth
while maintaining herbaceous vegetation.
Chemical treatments can control woody vegetation, but
without fire, vine cover will increase and herbaceous plant
cover will decrease because accumulating litter smothers
the herbaceous plants. Still, chemical treatments as a supplement to prescribed burning may allow for a longer frequency between prescribed burns without sacrificing
woody plant control. Regardless, prescribed fire is necessary to remove litter and maintain longleaf pine grassland
communities.
Literature Cited
Grelen, Harold E. 1975. Vegetative response to twelve
years of seasonal burning on a Louisiana longleaf
pine site. Res. Note SO-192. New Orleans, LA: U.S.
Department of Agriculture, Forest Service, Southern
Forest Experiment Station. 4 p.
Grelen, H.E. 1983. Comparison of seasons and frequencies of burning in a young slash pine plantation. Res.
Pap. SO-185. New Orleans, LA: U.S. Department of
Agriculture, Forest Service, Southern Forest Experiment Station. 5 p.
Haywood, J.D. 2002. Delayed prescribed burning in a
seedling and sapling longleaf pine plantation in Louisiana. P. 103-108 in Outcalt, Kenneth W., ed. Proceedings of the eleventh biennial southern silvicultural research conference. Gen. Tech. Rep. SRS-48.
Asheville, NC: U.S. Department of Agriculture, Forest Service, Southern Research Station. 622 p.
Haywood, J.D., and H.E. Grelen. 2000. Twenty years of
prescribed burning influence the development of
direct-seeded longleaf pine on a wet pine site in Louisiana. Southern Journal of Applied Forestry. 24
(2):86-92.
Haywood, J. D., F.L. Harris, H.E. Grelen, and H.A. Pearson. 2001. Vegetative response to 37 years of seasonal burning on a Louisiana longleaf pine site.
Southern Journal of Applied Forestry. 25(3):122-130.
Table 4. Stand characteristics after four biennial prescribed fires or biennially applied chemical weeding
treatments in a longleaf pine plantation; treatments began in the sixth growing season and ended in the
twelfth growing season.
Treatments
Average fire
intensity
Longleaf pine
basal area
Percent
cover
Grasses
Forbs
Woody
vines
Trees &
shrubs
(BTU/sec/ft)
(ft2/acre)
(%)
(%)
(%)
(%)
Untreated
…
105
2
1
13
53
Biennial weeding
…
95
4
1
11
5
214
48
35
3
2
17
185
63
32
3
1
10
209
58
44
9
1
8
Biennial March
burns
Biennial May
burns
Biennial July
burns
Speaker Presentations 27
How Does Longleaf Pine Native Groundcover Fit with Forest Management Goals?
Sharon M. Hermann1
1
Department of Biological Sciences, Auburn University, Alabama, 36849, USA
Abstract
When the ecology and biodiversity of longleaf pine forests are discussed, the large number of plant species in the
ground layer is always mentioned. How groundcover
interacts with various forest management goals of a landowner is less frequently considered. Maintaining and/or
restoring components of groundcover are not always
among the primary objectives for a stand. However, management of this vegetation layer often enhance primary
objectives in cost-effective ways. In this paper I review
the interrelationship between groundcover and traditional
forest management interests, including timber, hunting,
application of prescribed fire, and pine straw raking. In
addition, I outline questions to ask in order to determine
when and how a landowner might benefit from managing
existing groundcover or undertaking some level of restoration of this vegetation layer.
Introduction
In the last decade there has been increased attention paid
to the ground layer of longleaf pine forests. This layer is
dominated by grasses and forbs with scattered stems of
hardwoods (in some cases also palms) plus longleaf seedlings and juveniles. The native ground layer flourishes
under open canopies with frequent fire and grows on a
wide-range of soils that have experienced little anthropogenic disturbance, that have had no prolonged period of
fire exclusion and that have not been invaded by exotic
species. An impressive number of plant species (200300) have been recorded stands, and in some mesic areas
with high site index there may be 30-40 species in a
square yard. A high percentage of species in the ground
layer are grasses, legumes (peas and beans), and composites (asters and daisies). There is no region-wide estimate
of the acreage of longleaf pine that supports this component of the forest. However estimates of remaining longleaf lack information on the ground layer, and so area
supporting groundcover must be well less than the approximately three million acres that currently supports the
trees.
There are two general categories of groundcover, distinguished by types of wiregrass (Aristida stricta and A.
beyrichiana) or bluestem grasses (Andropogon spp.).
These two categories of groundcover once occupied approximately equal acreage of the longleaf range (Frost
1993). Wiregrass is common in Florida and the lower
Coastal Plain from east Alabama east though Georgia and
in North Carolina. Bluestem is found from central Alabama to Texas and dominates the understory of much of
South Carolina north of the lower Coastal Plain. Although prior to European settlement bluestem species
might have been the most prevalent grasses at some sites
in the range of wiregrass, information is lacking and this
issue is often discussed if ground layer restoration is
planned.
Value of Groundcover in Fire Management and Natural
Regeneration
Although there is high biodiversity in the native ground
layer of longleaf pine forests, for some landowners this
may not be a sufficient reason to maintain it. However, in
addition to ecological value, there is a practical advantage
to retaining native groundcover; it has value for attaining
some types of management goals, especially those related
to application of fire and/or facilitating natural regeneration of longleaf pine. High-quality groundcover is a useful commodity because it may enhance effectiveness of
prescribed burning. The likelihood that fire will carry
through a stand and top-kill hardwood stems is, in part,
related to the availability of fine fuel. The components of
fine fuel are pine needles, grasses, forbs, and litter (dead
vegetation). Grasses and forbs enhance the effectiveness
of prescribed burns and are especially important in areas
that lack pine needles and, therefore, may develop into
hardwood thickets or bramble patches. The useful fine
fuel associated with native groundcover reduces competitors of regenerating longleaf trees and decreases the incidence of the fungal needle pathogen, brownspot. In addition, fire effects created by native groundcover and pine
needles results in exposure of mineral soil, thus improving
the seed bed conditions required for longleaf seedling
establishment.
The presence of native groundcover may also expand the
window of opportunity for the application of fire. Widespread fine fuel permits application of fire over a broader
range of seasons, especially spring and summer burns, and
allows more frequent use of fire. Both season and frequency of burns are important components of fire regimes. Frequent fire has been demonstrated to be especially important in maintaining an open forest structure in
ground layers dominated by wiregrass (Glitzenstein et al.
2003) or by bluestem (Glizenstein et al. 2003, Hermann
1995).
Speaker Presentations 28
Groundcover Value to Animals
In addition to its importance as a component of fine fuel,
groundcover has value for animals in longleaf pine forests. The native ground layer is the natural equivalent of
early successional habitat and so is important to bobwhite
quail and other bird species. A ground layer of grasses
and forbs is an important component of quail management
because it provides protective cover and nest sites for the
birds. The interaction of native groundcover and fire
helps maintain the necessary open vegetation structure
and important food plants of quail, including many native
species of lespedeza and other legumes that are useful
alternatives to exotic food plants that may invade longleaf
stands and alter fire patterns.
Van Lear et al. (2005) reviewed the habitat needs of a
suite of vertebrate species associated with longleaf pine.
These animals are of conservation concern and Van Lear
et al. (2005) determined that open forest structure was an
important habitat feature to them. Because fine fuel may
be important in facilitating frequent fire, the animals of
conservation concern may indirectly benefit from the
presence of native groundcover. This may be especially
true for red-cockaded woodpeckers (Cox this publication)
and gopher tortoises (Guyer this publication). There are
some studies that suggest that populations of some bird
species (Bachman’s sparrow, Pine warbler and Eastern
wood-pewee) may display a positive response to the effect
that growing season burns have on vegetation. There is
increasing evidence that growing-season fires do not
negatively affect ground-nesting birds (cf. Tucker et al
2004). In addition, the forb component of the ground
layer has value to many invertebrate species, especially
butterflies and bees (Hermann et al. 1998). Increasingly
native pollinators are of concern to conservation.
Requirements of Native Groundcover
To maintain native groundcover it is important to understand the basic requirements of the grasses and forbs and
their tolerances to disturbance. Many groundcover species (especially grasses) do not tolerate heavy soil disturbance, especially over large areas. Once soil is tilled it is
difficult for native bunchgrasses to recover however light
roller chopping is not likely to result in long-term damage.
In addition, many ground layer species, especially grasses,
lack a soil seed bank and dispersal distance is short (just a
few feet from the parent plant) so re-colonization is impossible for some plants. However, there are some species that do persist as seeds in the soil (Cohen et al 2004)
These species may be useful in groundcover restoration
efforts for (see below).
The high biodiversity of the ground layer might appear to
require many species-specific management approaches.
However, in general this is not the case because most understory plant species share common growth requirements
with seedling and grass stage longleaf. These requirements
include:
• High sunlight (little/no tolerance of shading)
• Fire to remove litter & hardwood encroachment
• High burn frequency
• Periodic growing-season fire is useful but probably not
mandatory for many species.
• Little mechanical disturbance once established
Land Management Goals and Native Groundcover
When land management goals include promotion of natural
regeneration of longleaf pine, opportunities emerge to enhance native groundcover with little compromise to the
primary management interests. This is because when natural regeneration is employed, goals include:
• Maintaining an open canopy
• Minimizing mechanical activity
• Applying frequent fire.
• Under these activities, a landowner can engage in tree
production, especially for saw-timber, and/or quailhunting opportunities.
Other land management goals may create challenges for
maintaining native groundcover. If a landowner is interested in planting a longleaf pine plantation, site preparation
and subsequent closed canopy may eliminate many important species in the ground layer. However at some sites, a
proportion of the native ground cover may remain and/or
re-colonize when the canopy of the plantation becomes
more open (Smith et al. 2002). Some points to consider
include:
• Some types of site preparation are likely to depress
existing groundcover species but usually do not eliminate them
• Densely planted plantations go through a closedcanopy phase that creates low light
• Retaining and enhancing remnants of groundcover that
exist at the time of planting require low planting density; studies are currently under way by Dr. Joan
Walker and others. Some workers have suggested that
tree density may need to be 500/ac or less; this is an
area that needs additional research.
• To maintain existing groundcover species, site preparation for tree planting should be relatively light, for example fire, hardwood-specific herbicide, and/or a single pass with a roller chopper.
• Coupling longleaf plantations and managing for
groundcover is most appropriate when the tree canopy
remains sparse and open.
• New CRP opportunities are developing that will facilitate sowing seed of native warm-season grasses at the
same time longleaf is planted on former agriculture
land.
Speaker Presentations 29
These considerations can help evaluate site potential and
guide future actions.
If a landowner is interested in raking pinestraw:
•
Often tree density is relatively high and there is low
light at ground level
Fire frequency is usually low
Raking almost always is associated with mechanical
disturbance
Consequently frequent raking is not very compatible
with maintaining or restoring components of native
groundcover
•
•
•
Considerations for Restoration of Native Groundcover
Restoration covers a range of activities and goals. It spans a
continuum (Walker 1999) from planting trees, to re-creation
of open forest structure, to establishment of some species
(dominant or those of special interest), to re-creating the
entire ecosystem. Each step along the restoration continuum has potential to provide benefits to forest ecology and/
or management. Recent publications, such as Brockway et
al. (2005), Walker and Sillette (2006), provide guidance for
longleaf groundcover restoration. Restoration of key elements of the ground layer is possible (Barbour and Glitzenstein this publication, Glitzenstein and Streng this publication, Glitzenstein et al. 2001, Walker and Roth this publication), however improved technologies and additional seed
sources are needed. Maintenance and enhancement of existing groundcover is important to consider before undertaking a management action that would degrade an example of
this natural resource.
Before embarking on a project to restore groundcover to a
longleaf forest, it is useful to consider issues suggested by
Johnson and Gjerstad (2006) for restoring trees to a site.
Similar logic is appropriate for the ground layer and is reviewed at the Longleaf Alliance website (http://
www.auburn.edu/academic/forestry_wildlife/
longleafalliance/).
•
•
•
•
•
•
•
•
•
•
•
Determine landowner goals
Understand requirements of species slated for restoration
Evaluate current condition and past use of site, including
Pasture (grazing)
Row Crop (traditional agriculture)
Soil compaction, alteration of soil quality
Planted Pine Plantation
Presence of exotic species
Fire Exclusion
Other potential incompatible histories (phosphate
mining, etc)
Possibility of residual native ground layer adults and/
or soil seed bank.
Conclusions
In addition to its significance to biodiversity, native
groundcover has value to management. Longleaf groundcover plants are important as resources for animals and
enhancing fire effects and maintaining open forest structure as well as being important for promoting natural regeneration. Maintenance of native groundcover is compatible with saw timber production and uneven-aged management that includes providing quail-hunting opportunities. Basic information required to restore the ground
layer exists and new technologies are in development,
although currently there is a shortage of local seed
sources. One of the most far-reaching choices a landowner can make is to eliminate existing native groundcover. It is more challenging and more costly to restore
this resource that it is to maintain it. If current forest condition and future management goals are compatible with
promoting a ground layer of native species, landowners
are encouraged to consider maintaining this important part
of the region’s natural heritage.
Literature Cited
Brockway, D.G, K.W. Outcalt, D.J. Tomczak, and E.
Johnson. 2005. Restoration of longleaf pine ecosystems.
General Technical Report SRS-83. Asheville, NC. U.S.
Department of Agriculture, Forest Service, Southern Research Station.
Cohen, S., R. Braham, and F. Sanchez. 2004. Seed bank
viability in disturbed longleaf pine sites. Restoration
Ecology 12(4):503-515.
Frost, C. 1993. Four centuries of changing landscape
patterns in the longleaf pine ecosystem. Pages 17-43. In
(S. Hermann, ed.) The Longleaf Pine Ecosystem: Ecology, Restoration, and Management. Proceedings Tall
Timbers Fire Ecology Conference 18.
Glitzenstein, J.S.; D.R. Streng, D.D. Wade and, J.
Brubaker. 2001. Starting new populations of longleaf
pine ground-layer plants in the outer Coastal Plain of
South Carolina, USA. Natural Areas Journal 21: 89-110.
Glizenstein, J.S., D.R. Streng, and D.D. Wade. 2003.
Fire frequency effects on longleaf pine (Pinus palustris, P.
Miller) vegetation in South Carolina and northeast Florida, USA. Natural Areas Journal 23:22-37.
Hermann, S.M. 1995. Stoddard fire plots: lessons for
land management thirty-five years later. Pages 13-30.
Proceedings of the Tall Timbers Game Bird Seminar.
Speaker Presentations 30
Hermann, S.M., T. Van Hook, R.W. Flowers, L.A. Brennan, J.S. Glitzenstein, D.R. Streng, J.L. Walker and, R.L.
Myers. 1998. Fire and biodiversity: studies of vegetation
and arthropods. Transactions of the North American Wildlife and Natural Resources Conference 63:384-401.
Johnson, R. and D. Gjerstad. 2006. Restoring the overstory
of longleaf pine ecosystems. Pages 271-295. In (S. Jose, E.
Jokela, and D. Miller, eds.) The Longleaf Pine Ecosystem:
Ecology, Silviculture, and Restoration. Springer Science,
New York.
Smith, G.P., V.B. Shelburne, and J.L. Walker. 2002.
Structure and composition of vegetation of longleaf pine
plantations compared to natural stands occurring along an
environmental gradient at the Savannah River Site. Pages
481-486. General Technical Report SRS-48. U.S. Department of Agriculture, Forest Service, Southern Research
Station, Asheville, NC.
Tucker, J.W.; W.D. Robinson, and J.B. Grand. 2004. Influence of fire on Bachman's sparrow, an endemic North
American songbird. Journal of Wildlife Management 68
(4):1114-1123.
Van Lear, D.H.; W.D. Carroll, P.R. Kapeluck, and R. Johnson. 2005. History and restoration of the longleaf pinegrassland ecosystem: Implications for species at risk. Forest Ecology and Management 211(1-2):150-165, Sp. Iss. SI
Walker, J.L. 1999. Longleaf pine ecosystem restoration on
small and mid-sized tracts. Pages 19-22.
Walker, J.L. and A. Silletti. 2006. Restoring the ground
layer of longleaf pine ecosystems. Pages 297-325. In (S.
Jose, E. Jokela, and D. Miller, eds.) The Longleaf Pine Ecosystem: Ecology, Silviculture, and Restoration. Springer
Science, New York.
Speaker Presentations 31
Fire History of a Georgia Montane Longleaf Pine (Pinus palustris) Community
Nathan Klaus1
1
Georgia Department of Natural Resources Non-game Endangered Wildlife Program, Forsyth, Georgia 31029, USA
Abstract
The montane longleaf pine forests are endangered firedependent ecosystems of the Piedmont and Ridge and
Valley and Cumberland Plateau of Georgia and Alabama. Little is known about the historic fire regimes of
mountain longleaf forests or how to apply prescribed
fire to achieve restoration and conservation goals. I
used two lines of investigation to investigate historic
fire regimes: 1) a dendrochronological study of fire
scars on Sprewell Bluff Natural Area and 2) calculations of the average fire tolerance of tree species recorded on 1820s land lottery maps and 2005 surveys.
Three distinct periods of fire history were revealed: pre1840 with an average fire interval of 2.6 years, 18401915 with an average fire interval of 1.2 years, and
1915-present with an average fire interval of 11.4 years.
Season of fire differed between periods with
all seasons of fire common prior to 1840, mostly winter
fires from 1840-1915 and mostly spring fires from 1915present. Land lottery data indicated that montane longleaf
forests of the 1820s were most similar in fire tolerance to
areas of longleaf wiregrass compared to several other historic Georgia forest types. Modern forests had much lower
scores of fire tolerance. Differences in species composition
accounted for these changes in scores, historic mountain
longleaf forests had larger components of pine (Pinus spp.),
post oak (Quercus stellata), and blackjack oak (Q. marilandica) while modern forests have higher densities of
chestnut oak (Q. prinus). Our results suggest a fire return
interval of two to three years is needed to stop the continued
loss of the montane longleaf pine ecosystem.
Speaker Presentations 32
Fire Effects on Longleaf Pine Growth
John S. Kush1
Abstract
Results: Biennial Burns
Two long-term, on-going studies located on the Escambia Experimental Forest south of Brewton, AL will be
used to discuss the effects of prescribed fire on longleaf
pine growth. The first study was established in 1973 to
examine understory succession and overstory growth in
longleaf pine small pole stands following biennial
burns, mechanical, and chemical treatments. In response to this project, a second study was established in
1984 to determine the comparative impact of both winter and spring prescribed fires at intervals of 2, 3, and 5
years on the growth of a longleaf pine overstory and
development of hardwood competition. This presentation will follow the growth of longleaf pine (diameter,
height, basal area and total volume) over time.
The results presented in this paper are from a subset of the
data that has been collected. Prior to 1995, this study was
re-measured every 3 years. By 1982, 10 years after the initiation of the study, the diameter at breast height (DBH) for
trees on the no burn plots was significantly greater than for
all of the burn plots. There is no significant difference
among the burn treatments. This is a trend which continues
up to now (Table 1). If you look at the growth increases
from 1989 to the end of the 2004 growing season, there was
a significant difference in diameter growth between the
summer burn and the other treatments. The summer burn
trees grew 1.5” compared with 2.2” for the other 3 treatments.
Introduction: Biennial Burns
A study was established on the Escambia Experimental
Forest (EEF) in 1973 to determine the effects of various
understory hardwood control treatments on the growth
of the longleaf pine overstory. This study included fire,
mechanical and chemical treatments (see Boyer 1983,
1987, 1991, 1993, 1994, 1995). This paper will focus
on the impacts of fire on longleaf pine growth.
Methods: Biennial Burns
This study is comprised of 3 blocks on the EEF. The
predominant soil series on all 3 blocks is a Troup that is
a deep loamy sand with an A-horizon down to 51inches. At the time of study establishment, all areas
supported well-stocked young longleaf pine stands.
These stands were naturally regenerated from the 1958
seed crop.
Each treatment plot is 0.4 acres with a square 0.1 acre
measurement plot located in the center. All plots were
thinned to 500 dominant trees/acre in 1973. Four burn
treatments were conducted with prescribed fire at twoyear intervals in winter, spring, summer and an unburned check. The winter burns occur in January or
February, spring in April or May and summer in July or
August. All plots, but the check plots were burned with
a cool winter fire in 1974 to precondition them for future burning by removal of excessive fuel loads as the
previous prescribed fire occurred in 1962.
And likewise for total height, by 1982, the total height for
trees on the no burn plots was significantly greater than for
all of the burn plots (Table 2). There is no significant difference among the burn treatments. And this is a trend
which continues up to now.
Table 1. Average tree DBH (inches) for biennial prescribed fire treatments on the Escambia Experimental
Forest. Underlined numbers are statistically different than
the other numbers in the column at the 0.05% level.
Treatment
1973
1989
1990
1992
1995
2004
Winter
3.2
5.9
7
7.5
8
9.2
Spring
3.3
5.9
7.1
7.6
8.1
9.3
Summer
3.1
5.9
7.2
7.7
8.2
8.7
No burn
3.2
6.2
7.7
8.2
8.8
10
Table 2. Average total height (ft) for biennial prescribed
fire treatments on the Escambia Experimental Forest.
Underlined numbers are statistically different than the
other numbers in the column at the 0.05% level.
Treatment
1973
1989
1990
1992
1995
2004
Winter
22
54
60
64
68
75
Spring
23
55
60
65
68
75
Summer
22
55
58
65
68
73
No burn
22
59
64
69
73
78
Speaker Presentations 33
And basal area showed the same trend (Table 3). By 1982,
the basal area for the no burn plots was significantly greater
than for all of the burn plots. There is no significant difference among the burn treatments. In 1989, with the basal
area at 110 square feet and the burn plots not having
reached 100 square feet, the plots were thinned, mainly
from below, to an after-cut basal area of 70 square feet/acre.
Since the thinning, there has not been a significant difference among the treatments which is probably due to the
fewer trees on the no burn treatment plots than on the burn
plots.
The same holds true for volume, by 1982, total volume
(inside bark) was significantly different (Table 4). The total
volume is calculated from local volume equations that were
developed from work on the EEF.
There was a difference in DBH growth between the no burn
and summer burn plots. There were no differences in
height and basal area growth but a significant difference in
average yearly volume growth, where the no burn grew at
131 ft3/year, the spring and winter treatments at 123 ft3/year
and the summer 111 ft3/year.
Table 3. Average total basal (feet2/acre) for biennial prescribed fire treatments on the Escambia Experimental
Forest. Underlined numbers are statistically different
than the other numbers in the column at the 0.05% level.
Treatment
1973
1989
1990
1992
1995
2004
Winter
30
95
70
81
91
116
Spring
30
96
68
78
87
114
Summer
28
89
70
80
89
112
No burn
31
110
71
81
93
119
Table 4. Average total volume (feet3/acre - inside bark)
for biennial prescribed fire treatments on the Escambia
Experimental Forest. Underlined numbers are statistically different than the other numbers in the column at the
0.05% level.
Treatment
1973
1989
1990
1992
1995
2004
Winter
296
2265
1732
2144
2544
3594
Spring
314
2298
1683
2079
2479
3533
Summer
262
2109
1742
2139
2455
3410
No burn
317
2798
1884
2317
2769
3854
Pine mortality over the first 16 years was 8.8%, or 2.8
trees/acre/year. Mortality was significantly greater with
summer burning (14.2%) than with the other treatments,
8.1% for both winter and spring and 4.9% for the unburned treatment. In the following 15 years, mortality
averaged 2.2 trees/acre/year. Mortality was highest with
summer burning 2.9 trees/acre/year, but it was not significantly different from the other treatments. The majority
of mortality was due to lightning strikes and insects which
came afterwards in all treatments except for the summer
treatment where the majority was related to natural mortality of suppressed trees.
Introduction: 2-, 3-, and 5-Year Fire Intervals
Boyer (1987) summarized the first 10 years of the biennial burn study. As a result of these findings, a study was
initiated in 1984 to determine if prescribed fire at intervals
of 3 or 5 years would reduce the impact on pine growth
and still be reasonably effective for hardwood control.
Methods: 2-, 3-, and 5-Year Fire Intervals
The study was established in young longleaf pine stands
regenerated by the shelterwood system. Treatments include both winter and spring burns repeated at intervals of
2, 3, or 5 years plus an unburned check. Despite the overwhelming evidence regarding volume losses with biennial
burns, 2-year spring and winter burn treatments were installed to compare with data from the first study.
The trees originated from the 1973 seed crop and were
released from the parent overstory during the 1976 winter.
Like the previous study, there are 3 blocks but for this
study, the 0.4 acre plots were thinned to leave 400 dominant/co-dominant trees per acre. Spacing between trees
was made as uniform as possible. Before initiation of the
study, all stands were last burned during the 1979 spring
for competition control.
The burning treatments were initiated in the winter and
spring of 1985. To the extent possible, winter fires are
completed in January or February and spring fires in April
or May. Fires are prescribed and executed to minimize
crown scorch on pines. Normally flank or strip head fires
are used. Fires follow soaking rains as soon as conditions
of fine fuel moisture of 7-10%, relative humidity of 3555% and reasonably steady winds of 3-10 mph.
Results: 2-, 3-, and 5-Year Fire Intervals
While there have been no significant differences in DBH
and total height (Tables 5 and 6), there has been a significant difference in basal area (Table 7). By 1994, the no
Speaker Presentations 34
burn and 5-year spring burn were different from the other
burn treatments. By 1999, the 3-year spring burn joined
the no burn and 5-year spring burn as being significant
from the other treatments. However, by 2004, the 3-year
spring burn grew the least of all treatments and was again
different from the no burn and 5-year spring burn. Of
note was the 2-year spring burn which has been significantly different from the other treatments since 1999.
Table 5. Average tree DBH (inches) for different season
and frequency of prescribed fire treatments on the Escambia Experimental Forest. Underlined numbers are
statistically different than the other numbers in the column at the 0.05% level.
Treatment
Winter2
Spring2
Winter3
Spring3
Winter5
Spring5
No burn
1984
1987
1990
1994
1999
2004
2.2
3.3
4.2
5.2
6.4
7.4
2.2
3.2
4.1
5.2
6.4
7.3
2.2
3.4
4.3
5.4
6.4
7.4
2.2
3.4
4.4
5.6
6.6
7.5
2.3
3.3
4.2
5.4
6.4
7.3
2.3
3.4
4.4
5.6
6.7
7.6
2.2
3.5
4.5
5.8
7
7.8
Table 6. Average total tree height (feet) for different
season and frequency of prescribed fire treatments on
the Escambia Experimental Forest. Bold numbers are
statistically different than the other numbers in the column at the 0.05% level.
The response for total volume is similar to that for basal
area; only the significant difference did not show up until
the 1999 measurement (Table 8). In 1999, the no burn and
5-year spring burn were different from the other burn treatments. This trend continued through the 2004 measurements. Likewise, the 2-year spring burn was significantly
different from the other treatments.
After 19 years, mortality has been minimal with survival
running from 91% on the spring 2 & 3 year burns to 99%
on the no burn, winter burns and spring 5-year burn treatment.
There has been a 12% reduction on volume between the no
burn and the average for the 3- and 5-year burns with very
little difference in volume between the no burn and the 5year spring burn.
The 2-year winter burns and 3-year spring burns grew, on
average, 158 feet3/year. The 5-year spring grew at a rate of
176 and the no burn grew at a rate of 182 feet3/year. For
further information on this study, please see the paper by
Whitaker et al. (this issue).
Conclusions
Do you want a forest stand that is thick with competing
vegetation because you do not burn or, in other words, what
are you management objectives? If it is growing fiber, no
matter what the fiber is, then this may be what you want
and you do not need to worry about burning.
However, if you want a forest stand favorable for wildlife,
recreation, hunting, pine straw, wildflowers, or natural regeneration, then fire is going to need to be a part of your
management prescription. There is some loss in volume
when using prescribed fire, whether to the species you are
managing or unwanted/competing species.
Prescribed Burning is Defined As
Treatment
1984
1987
1990
1994
1999
2004
Winter-2
15
24
32
44
53
63
Spring-2
15
23
32
44
53
62
Winter-3
15
24
34
46
55
65
Spring-3
15
23
33
45
54
64
Winter-5
16
24
34
46
54
64
Spring-5
16
24
34
47
55
64
No burn
15
24
35
48
58
67
Fire applied in a knowledgeable manner to forest fuels on a
specific land area under selected weather conditions to accomplish predetermined, well-defined management objectives. What I have reported on here are the results from a
coastal plain site in south Alabama on moderately productive sites. What happens to the stands you burn will depend
on how you burn. There is a loss due to fire but you have to
ask yourself what the cost will be if I don’t burn.
Acknowledgements
To George Ward, Bill Thompson and Ron Tucker who have
collected most of these data and the T.R. Miller Mill Co. for
allowing the U.S. Forest Service to conduct all these years
of research into longleaf pine management.
Speaker Presentations 35
Literature Cited
Table 7. Average total basal area (feet2/acre) for different
season and frequency of prescribed fire treatments on the
Escambia Experimental Forest. Bold numbers are statistically different than the other numbers in the column at the
0.05% level.
Treatment
Winter2
Spring2
Winter3
Spring3
Winter5
Spring5
No
burn
1984
1987
1990
1994
1999
2004
11
25
39
62
86
111
11
23
38
61
80
104
11
26
41
63
87
112
12
27
44
68
94
114
21
25
40
65
89
116
13
26
42
71
97
127
11
27
45
74
102
127
Table 8. Average total volume (feet3/acre) for different
season and frequency of prescribed fire treatments on the
Escambia Experimental Forest. Bold numbers are statistically different than the other numbers in the column at the
0.05% level.
Treatment
Winter2
Spring2
Winter3
Spring3
Winter5
Spring5
No
burn
1984
1987
1990
1994
1999
2004
74
256
546
1155
1869
2840
76
236
524
1125
1750
2639
78
275
595
1208
1969
2966
80
268
612
1300
2096
2987
86
262
573
1230
1969
3022
94
284
608
1366
2199
3340
72
274
638
1468
2390
3460
Boyer, W.D. 1983. Growth of young longleaf pine as affected by biennial burns plus chemical or mechanical
treatments for competition control. In Proceedings of
the second biennial southern silvicultural research
conference, 4-5 Nov. 1982, Atlanta, GA. Edited by
Earle P. Jones, Jr. USDA For. Serv. Gen. Tech. Rep.
SE-24, pp. 62-65.
Boyer, W.D. 1987. Volume growth loss: a hidden cost of
periodic burning in longleaf pine? South. J. Appl.
For. 11:154-157.
Boyer, W.D. 1991. Effects of a single chemical treatment
on long-term hardwood development in a young pine
stand. In Proceedings of the sixth biennial southern
silvicultural research conference, 30 Oct. – 1 Nov.
1990, Memphis, TN. Compiled by Sandra S. Coleman
and Daniel G. Neary. USDA For. Serv. Gen. Tech.
Rep. SE-70, pp. 599-606.
Boyer, W.D. 1993. Season of burn and hardwood development in young longleaf pine stands. In Proceedings
of the seventh biennial southern silvicultural research
conference, 17-19 Nov. 1992, Mobile, AL. Edited by
John C. Brissette. USDA For. Serv. Gen. Tech. Rep.
SO-93, pp. 511-515
Boyer, W.D. 1994. Eighteen years of seasonal burning in
longleaf pine: effects on overstory growth. In Proceedings of the 12th international conference on fire
and forest meteorology, 26-28 Oct. 1993, Jekyll Island, GA. Soc. Am. For. Pp. 602-610.
Boyer, W.D. 1995. Responses of groundcover under longleaf pine to biennial seasonal burning and hardwood
control. In Proceedings of the eighth biennial southern silvicultural research conference, 1-3 Nov. 1994,
Auburn, AL. Edited by M. Boyd Edwards. USDA
For. Serv. Gen. Tech. Rep. SRS-1, pp. 512-516.
Speaker Presentations 36
New Findings for Site Preparation with Chopper Herbicide
Dwight K. Lauer1 and Harold E. Quicke2
1
Silvics Analytic, Ridgeway, Virginia, 24148, USA
BASF Corporation, Raleigh, North Carolina, USA
2
Abstract
Loblolly pine on cutover sites is often re-established using
at least two herbicide treatments. The first is a site preparation treatment designed to provide long-term control of
perennials such as trees, shrubs, vines and grasses. The
second treatment is called herbaceous weed control
(HWC) and is applied after the pines are planted. This
treatment targets mainly re-colonizing forbs and grasses.
Vegetation control studies often focus on either site preparation or HWC. However, these treatments are not necessarily independent. For example, when a high rate of a
soil active herbicide is applied late in the year, herbaceous
weed development after planting may be delayed and this
may influence the optimal timing for herbaceous weed
control. These studies look at integrated systems of bedding, Chopper® herbicide site preparation and Arsenal®
AC herbaceous weed control.
In the Upper Coastal Plain studies, different rates of
Chopper® were applied at two different times of the year.
Each Chopper® rate and timing treatment was followed
by HWC treatments applied at different times of the year.
Results based on three years of pine growth indicated that:
1) Chopper site preparation earlier in the year (July) resulted in better pine growth than site preparation later in
the year (October); 2) There was a synergistic response
between Chopper site preparation and Arsenal AC +
Oust® herbaceous weed control with pine growth from
the combined treatments far exceeding the growth from
either treatment applied alone; 3) Optimal herbaceous
weed control timing on these Upper Coastal Plain sites
was not influenced by timing of Chopper site preparation;
4) On sites where planted pines and herbaceous weeds
developed quickly in the first year, herbaceous weed control early in the first pine year resulted in the best pine
response; 5) On a site with slow initial pine and herbaceous development, herbaceous weed control in June of
the first pine year resulted in the best pine response; 6) A
second year of herbaceous weed control increased pine
growth by up to 23%; 7) Modern silvicultural techniques
resulted in pine growth that equaled or exceeded the
growth in complete vegetation control studies where pines
were kept weed free for up to five years.
Mid-season bedding occurred between May and July and
late-season bedding between September and November.
Results indicate that many of the historical timing limitations placed on Chopper®, applied in an oil emulsion carrier, are not necessary. Treatments can occur as early as
February and as late as November, up to the day before
bedding and immediately after bedding. Seasonal timing of
application was also found to have a major impact on pine
growth. For those treatments that provided good vegetation
control, earlier season treatments resulted in the best pine
growth. Optimal timings for Chopper® applications were
identified as: 1) June through September at least 3 weeks
after mid-season bedding, 2) February to the day before
mid-season bedding, and 3) February through July followed
by late-season bedding. An exception is that if deciduous
woody species such as hardwood or blackberry are targeted,
early season applications (February-April) should not occur
until these species have leafed out. The optimal timing
windows are wide, allowing forest managers plenty of flexibility in scheduling operations. Use of optimal timing windows has the potential to substantially increase productivity
without increasing costs.
On Lower Coastal Plain sites, vegetation control following different timings of Chopper® herbicide application
relative to bedding was examined. Two bedding regimes,
mid-season and late-season, were examined at each location.
Speaker Presentations 37
Reintroduction of Fire to Fire Suppressed Longleaf Pine Stands: An Overview of the Problem
John McGuire1
1
The Longleaf Alliance, Auburn University, Alabama, 36849, USA
Abstract
The use of fire in the development and maintenance of
longleaf pine forests has long been recognized by forest
practitioners and scientists. However, the reintroduction of
prescribed fire-to-fire suppressed longleaf pine forests has
proven to be problematic. Burning prescriptions that are
appropriate to maintain longleaf pine stands are often
unsuitable for older stands that have a history of firesuppression. Across the range of longleaf pine, the misapplication of fire to restore older stands has resulted in
significant mortality to mature longleaf pine trees that
should otherwise be retained. The intent of this presentation is to highlight the degree of the problem of fire reintroduction and present some remedies to the problem.
Speaker Presentations 38
Physiological Effects of Organic Soil Consumption on Mature Longleaf Pines (Pinus palustris)
Joseph O'Brien1, J. Kevin Hiers2, Kathryn Mordecai1 and Doria Gordon3
1
USDA Forest Service, Southern Research Station, Athens, Georgia, 30602, USA
2
J.W. Jones Ecological Research Center, Newton, Georgia, 39870, USA
3
The Nature Conservancy, Gainesville, Florida, 32601, USA
Introduction
Longleaf pine ecosystems depend on frequent fires to
maintain both the overstory pines and a high diversity
understory plant community. At one time these forests
dominated the coastal plain of the southeastern US, but
currently longleaf stands occupy less than 3% of their
historical extent, usually as isolated fragments (Gilliam
and Platt 2006). One consequence of this fragmentation
has been the reduction of fire frequency with some stands
remaining unburned for decades. A major effect of this
reduction in fire frequency is the development of an organic soil horizon. In frequently burned stands, fire consumes litter and the mineral soil surface remains mostly
exposed. In unburned stands, low litter decomposition
rates, especially in xeric sites, result in the formation of a
deep forest floor (Hendricks et al. 2002). Mature trees
colonize a well developed O-horizon with numerous fine
roots. These roots are lost after the organic soil is consumed by fire. Root consumption could cause both acute
and chronic stress to a tree. Loss of fine roots could immediately lower nutrient and water uptake rates, reduce
carbohydrate pools, and require resources to be allocated
for root replacement and repair, perhaps at the expense of
some other process. The impact of stored carbon losses
and inhibited mineral nutrient uptake would also likely
cause chronic effects such as reduced leaf area, changes in
leaf nitrogen balance, lowered tissue repair rates, and inhibition of chemical defenses. As a first step to understanding the physiological consequences of fine root loss, we
measured whole tree transpiration and chlorophyll concentrations in mature longleaf pines growing in an area
with a well developed O-horizon exposed to varying degrees of duff consumption and crown scorch.
treatments. No direct observations were made of the wildfire, but post fire measurements indicated that both forest
floor consumption and crown scorch were highly variable
and extensive. Two study trees escaped the fire, and the
remaining 18 received varying degrees of scorch and root
consumption. An additional three unburned trees were
added to the study after the fire.
Following the fire, we measured the amount of duff consumed in a 4 m diameter circle centered on the tree bole.
Consumption was estimated using both duff pins that were
in place prior to the fire and by visual estimates. A linear
correlation between visual estimates of consumption and
the duff pins measurements showed the effectiveness of the
visual estimates (R2=0.92, p<0.0001). Percent crown
scorch was estimated visually in 10% increments. One
month prior to the fire, sap flow sensors were inserted into
each tree. The wildfire destroyed the originals and replacements were deployed approximately one month following
the fire. The first post-fire sap flux measurements occurred
in mid-August 2005 after damaged crowns had reflushed
new leaves. The outputs from the probes were averaged
and transformed into sap flux estimates after Granier
(1987). Chlorophyll content was measured in foliage samples collected immediately prior to the fire, then at 3
months following the fire. Post-fire chlorophyll content
was standardized and presented as an index by subtracting
the post-fire content from pre-fire content, then dividing by
the pre-fire content. Consumption effects on sap flux were
analyzed using random effects ANOVA and the chlorophyll
content was analyzed using linear regression.
Results and Discussion
Methods
We conducted experimental burns in a stand of long unburned longleaf pine forest at Fort Gordon Military Base,
Augusta, GA, USA. The site had not been burned in at
least 50 years and had a well developed O-horizon, with
an average depth of 16.1 cm (± 3.1 S.D.). Tree sampling
was stratified by soil type (Arenic Frangiudults) and topography (ridge tops). Within the site, 20 trees with a
DBH of approximately 35 cm and similar stature were
chosen. The 20 selected trees had a mean DBH of 37.1
cm (± 2.7 S.D.) and a height of 19.6 m (± 2.4 S.D.). Initially, the plan was to experimentally manipulate fire
damage, but on June 25, 2005 a wildfire passed through
the study area three days after the first root consumption
All fire damaged study trees had at least 20% O-horizon
consumption (mean = 50.2, ± 35.6 S.D.). Scorch rates averaged 10% (±19.2 S.D.), though 11 of the 18 trees with Ohorizon consumption had undamaged crowns. Mean sap
flux rates did not vary among the study trees prior to the fire
(F1, 19=0.653, p=0.861). Sap flux rates were estimated simultaneously in all 20 trees for a total of nine days postfire. The mean daily sap flux over the nine day period was
0.488 kg dm-1 hr-1 (± 0.35 S.D.). The results of the random
effects ANOVA indicated that only O-horizon consumption
had a significant impact on mean post-fire sap flux rates and
there was no interaction (Table 1).
Speaker Presentations 39
Table 1. Results of the random effects ANOVA. The multiple R2 for the model was 0.44 (p=0.001).
Intercept
SS
3.77
D.F.
1
MS
3.767
F
47.67
Scorch
0.01
1
0.007
0.09
Consumption
1.08
1
1.078
13.65
Scorch*
Consumption
0.05
Error
1.50
p
>0.00
01
0.77
0.001
1
0.050
0.63
0.436
19
0.079
Sap flux was negatively correlated with consumption as
indicated by the beta coefficient of -0.70 (t=-3.69, p=0.002).
Figure 1 displays the results of a linear regression showing
the negative relationship between amount of crown scorch
and standardized chlorophyll content (beta= -0.81, R2
=0.60, p=0.009).
Summary and Conclusions
Post-duff fire damage in longleaf pines seems to cause a
cascade of chronic stressors with mortality often occurring
several months after the fire. However, the loss of fine
Chlorophyll content index
-0.2
-0.3
-0.4
-0.5
-0.6
-0.7
-0.8
-0.9
-5
0
5
10
15
20
25
30
35
40
roots after O-horizon consumption can also create acute
stress due to the reduction in water uptake and transpiration rates. Recent research suggests that longleaf pines
maintain a constant allocation ratio between roots and
aboveground tissue (Hendricks et al. 2006). Observations
of study trees post-fire suggest that leaf area did decrease
following O-horizon consumption. This decrease in photosynthetic capacity coupled with the loss of stored carbohydrates appears to be catastrophic to the trees; three of
the 16 study trees with root consumption died within one
year of the study and two more appeared near death with
sparse crowns. The immediate loss in the ability to supply leaves with water could lead to chronic carbon limitations through a compensatory reduction in leaf area.
Crown scorch is an additional acute stressor that may or
may not occur simultaneously with root consumption.
Since mineral nutrients are required to replace scorched
leaves, root loss has the potential to further limit photosynthetic capacity. The loss of stored carbon and diminished photosynthetic capacity could lead to a cascade of
indirect effects such as inhibited chemical defenses and
reduced overall vigor. It appears that the inhibition of
herbivore defenses might be critical as all three dead
study trees succumbed to insect attack. These results suggest that the lowering of water and nutrient uptake rates
might be the ultimate cause of mortality after duff fire.
Literature Cited
Gilliam, F.S. and W.J. Platt. 2006. Conservation and restoration of the Pinus palustris ecosystem. Journal of
Applied Vegetation Science 9:7-10.
Granier A. 1987. Evaluation of transpiration in a Douglas
fir stand by means of sap flow measurements. Tree
Physiology 3, 309–320
Hendricks, J.J., R.L. Hendrick, C.A. Wilson, R.J.
Mitchell, S.D. Pecot and D. Guo 2006. Assessing the
patterns and controls of fine root dynamics: an empirical test and methodological review. Journal of
Ecology 94:40-57.
Hendricks, J. J., C. A. Wilson, and L. R. Boring. 2002.
Foliar litter position and decomposition in a firemaintained longleaf pine-wiregrass ecosystem. Canadian Journal of Forest Research 32:928-941.
45
Percent scorch
Figure 1. Linear regression of chlorophyll content and
crown scorch. All trees had at least 20% duff consumption.
It appears that resource limitation may have lead to a reduction in the ability to adequately provision new leaves
with chlorophyll.
Speaker Presentations 40
Pineywood's Cattle Breed: History and New Uses for Small Acreages
Chuck Simon1
1
County Extension Agent-Coordinator, Covington County, Alabama Cooperative Extension System, Andalusia, Alabama,
36420, USA
Abstract
The intent of this discussion is to introduce the audience
to a new concept of using an antique (or minor) breed of
livestock to help control vegetation on small land holdings. We will review what pineywood cattle are, the
breed history
and historical use and end with their uses today. Likewise,
the discussion will focus on how to save this breed from
extinction. Mr. Simon has raised pineywood cattle since
1995.
Speaker Presentations 41
Bobwhite Quail Issues and Research Efforts in the Longleaf Region: An Overview
Lee Stribling1
1
School of Forestry and Wildlife Sciences, Auburn University, Alabama, 36849, USA
Abstract
Bobwhite quail have existed in what is today the southeastern United States for thousands of years. Bobwhite populations flourished in an ideal habitat created by European settlers to the region. They opened the vast unbroken oldgrowth forests to build homes, farms, and towns separated
by patches of cut-over and frequently burned forest land.
Further increase of quail numbers was caused by the hunting practices of that era which reduced quail predator populations. Quail populations grew and remained high until the
1960's. For a relatively short time in the history of the
Southeast a set of circumstances had produced a landscape
and environment conducive to high quail populations. In
response to socioeconomic changes across most of the rural
south, a slow decline of quail numbers began. This decline
continued and accelerated through the rest of the 20th century. A growing human population, new farming practices,
exclusion of woodland fires, improved pasture, intensive
pine plantations, and high predator populations all contributed to the drop in quail numbers. It is ironic that the same
forces of human settlement, responsible for the explosion of
quail that began 200 years ago, are now the same socioeconomic agents that have driven them to desperately low levels in modern times. It seems that most things have to get
really bad before they attract enough attention to get better.
This was the case for bobwhite quail and all other species
associated with the early successional, forest-savannah-type
ecosystem of the Southeast. Starting in the early 1990's a
renewed interest and emphasis on quail biology and management began. Research, management, and outreach programs by southern Land Grant Universities, Wildlife Agencies, and Research Organizations all combined to restore
bobwhite quail to areas where they once were abundant.
While some portions of the bobwhite’s traditional range are
still at the “work in progress” stage, other areas have maximized conditions for quail and are producing record breaking populations.
contributions were made toward recovery of this historically, economically, and ecologically important species.
The Albany Quail Project
The tried and true extension approach of identifying and
working with a key landowner who is watched and respected by others in the area has been very successful
getting new programs started. If new approaches and
techniques are used by that key individual and prove to be
successful other landowners quickly adopt the practice as
well. The landowner was the R.K. Mellon Family and
their 20,000 acre property, Pineland Plantation, located
just south of Albany, GA had been intensively managed
for bobwhite quail since the early part of the 20th century.
Even though they were using the most intense quail management techniques available, they were experiencing
significant declines in their quail populations. The Mellon Family agreed to provide funding to determine the
cause of the quail decline and develop measures to halt
and reverse it (research). Additionally, they indicated that
all information from the study would be made immediately available not only to them, but to anyone else who
may be interested. This resulted in project personnel
spending a great amount of time and funds on visits to
other properties and tours of the study area (extension/
outreach). In this way everyone could decide how best to
use information from the work and positive results might
be produced more quickly.
This program became known as the Albany Area Quail
Project. Even though it was named after the area of Georgia where it was started, it was destined to have far reaching impacts on bobwhite quail populations and quail management across their entire geographical range.
Research Directed Management
Introduction
During the late 1980's bobwhite quail populations were
experiencing an alarming decline across their geographic
range. Quail biologists were aware of the problem but had
no solution. Some were even predicting possible threatened
or endangered status for the species if the situation was not
soon reversed. However, with the right approach and a
combined research and extension program, significant
In 1992, we began trapping wild quail and tagging them
with very small radio-transmitters which allowed our research associates, students, and technicians to locate the
birds at any time. The locations told us what habitat components were needed or not needed for survival and reproduction. Knowing what was required versus its availability showed what types of habitat needed to be added and
where. By providing this information and working with
the property land manager, techniques were tested to create and structure missing habitat components.
Speaker Presentations 42
After only 2 years, the research had provided sufficient
information to change management protocols and the
population of quail on the property began to rapidly respond. By 1994, quail populations had improved 40%
compared to when the study began. As predicted the
neighboring quail plantations became very interested in
this success and wanted to modify quail management
techniques on their land. A group of 9 owners and managers met with us and indicated that they were using the
new information had other quail management and hunting
problems that they needed help solving. They were willing to fund additional studies like the one which was
proving to be so successful on their neighbor’s property.
The project began to expand in scope, study sites, personnel, and success. In 2003 quail populations have tripled
(300% increase) on the original study site. Hunting success and quail abundance is almost twice as high as anytime listed in the nearly 100 years of hunt records on
Pineland Plantation. A great many other quail properties
following the “Albany Program” are having similar successes. Over the last 14 years we have conducted quail
studies on 10 different study sites (7 in Georgia & 3 in
Alabama). We are engaged in cooperative work with Tall
Timbers Research Station in Florida, the University of
Georgia, and Mississippi State. Information from our
work is made available through newsletters, magazine
articles, other multi-media outlets, field days, seminars,
group presentations, and individual landowner to project
personnel interactions. Publication in peer-reviewed and
other professional outlets were slower to production than
the outreach types chiefly due to the manner in which the
project was structured because the priority given to the
land owners and managers. During the last few years the
number of scientific publications has greatly increased
and will continue to do so in the future.
One of the most important values of this project comes
from its length and large sample size. The project has
existed for 14 years and is showing no signs of slowing
down. We have attached over 8,000 radio-transmitters to
quail during that time and have learned a great deal more
about quail than I ever expected. A project of this scope
provides a perspective on how an organism fits into the
human altered landscape. Quail declined because the entire landscape of its range changed and the techniques
developed by Stoddard in the 1930's were still in use.
These techniques, although good for 30-40 years after
they were developed, were designed for a world that does
not exist anymore for the bobwhite. Agriculture and forestry management has changed greatly as well as rural
socio-economic patterns. Quail had not changed their
requirements since Herbert Stoddard’s work. What
changed is the manner in which we had to go about providing them.
Impact Program
The success of this quail project caused a ripple effect
across quail country in the Southeast and Midwest. There
was now hope that bobwhite quail were not going to be a
fond memory of the past as some biologist had predicted.
Using the new management methods quail populations improved on intensively managed quail plantations across
Georgia and the region. In 1999 the success of these research projects and resulting management techniques
prompted a group of Georgia Biologists and State Legislators to visit the Albany Quail Project. Following this visit
the State of Georgia developed the Bobwhite Quail Initiative One. Millions of dollars were provided for education,
landowner assistance, and cost-share programs to increase
quail populations and quail hunting in areas of the state
where it once had been commonplace. In 2002 a similar
plan was developed by the Southeastern Quail Study Group
to address the same issue across the entire southeastern US.
The plan was called the Northern Bobwhite Conservation
Initiative Two. In August 2004, President Bush announced
a new part of the Conservation Reserve Program (CRP)
called the Northern Bobwhite Quail Habitat Initiative
Three . The goal of this $125 million (3 year) program is to
promote establishment of 250,000 acres of early successional buffers along agricultural field borders, a practice
developed, recommended, and proven by the Albany Quail
Project. This increase in nesting and brood-rearing cover is
hoped to increase bobwhite quail numbers by 750,000 birds
annually in the Southeast and Midwest.
Over the past decade the Albany Quail Project has been
directly involved with improving quail populations on over
500,000 acres, in Georgia, Alabama, South Carolina, Florida, Tennessee, Mississippi, Virginia, North Carolina, and
Kentucky. Not only have more quail resulted from the
habitat modifications and management approaches we recommended, but many other species dependant on this type
of early successional habitat also have benefited. In addition, land values of good quail hunting properties are much
higher than those where populations are low. A recent survey of realtors who broker quail management properties
indicated that the properties which are managed using our
program usually have very good quail populations are in
high demand and sell for about $1,000 more per acre than
similar non-managed hunting properties. This gives a total
added value to properties under the Albany Quail Project
management scheme of over $500 million.
Speaker Presentations 43
The Albany Quail Project began in 1993 with a single gift
of $80,000 from the R.K. Mellon Family Foundation to
fund the first year of the first project. In addition to continuing The Albany Quail Project’s work, we began another
quail research/ extension project in 2001, this time located
on Alabama. We employed the same model as used to start
the Albany Project, a single landowner funding a single
project, then getting others involved based on his successes.
This project is called the Alabama Quail Project and finetunes quail management to soil and other differences found
on quail lands in Alabama. All of the approximately
$400,000 to $500,000 per year in total operating funds for
both projects come from unsolicited, private donations.
Web Links:
1 Bobwhite Quail Initiative
http://georgiawildlife.dnr.state.ga.us/content/
displaycontent.asp?txtDocument=108
2 Northern Bobwhite Conservation Initiative
http://www.qu.org/seqsg/nbci/nbci.cfm
3 Northern Bobwhite Quail Habitat Initiative
http://www.alfafarmers.org/headlines/headline.phtml?
id=4467
Speaker Presentations 44
The Georgia Coastal Flatwoods Upland Game Project: Launching a War on Wiregrass
Chris Trowell1
1
Emeritus Professor of Social Science, South Georgia College, Douglas, Georgia, 31533, USA
Abstract
The piney woods of south Georgia have long been perceived as an impoverished region. These lands were
viewed by many during the 18th and 19th centuries as
land useful only as livestock range and even the pasturage
was believed to be impoverished, especially the wiregrass
on which the open-range wiregrass cattle foraged.
As the land was penetrated by railroads between 1860 and
1920 the unbroken longleaf pine-wiregrass range was
transformed into a forest of stumps. The lands in and
around the Okefenokee Swamp were burned over by wildfires during the droughts of 1931-1932. Much of the land
and many of the people were hopeless and abandoned in a
post-logging environment. Local communities sought
hope and relief through collective projects. The Waycross
Lions Club initiated several projects to improve the cattle
by improving pasturage. Cattle was believed to be the
core of the local economy, especially since timber was no
longer the economic base of the area.
Federal title to the property was transferred to the State of
Georgia in 1955. The area was renamed the Waycross State
Forest. Part of the 30,000 acre tract was developed as
Laura S. Walker State Park and a small tract on Cowhouse
Island was leased to become Okefenokee Swamp Park.
Over the years most of the wiregrass was removed in the
process of site preparation for slash pine plantations. Wiregrass was completely removed or covered with imported
sand in the fairways and greens of “The Lakes” golf course.
The Waycross State Forest was renamed the Dixon Memorial State Forest in 1974. The district offices of the Georgia
Forestry Commission are located on the site of the old
Georgia Coastal Flatwoods Uplands Game Project headquarters.
Many of the federal New Deal programs during Great
Depression of the 1930s focused on the alleviation of poverty and on relief of the desperate conditions of the demoralized people and the denuded land. Congressman
Braswell Deen of the Eighth District of Georgia promoted
a federal scheme to improve scrub cattle by replacing
wiregrass with carpet grass. By the time the Georgia
Coastal Flatwoods Upland Game Project was launched in
1935, the plan encompassed a 30,000 acre comprehensive
multiuse land use area near Waycross as well as an employment program.
Between 1935 and 1938, the U.S. Department of Agriculture funded reforestation, wildlife and range improvement, and recreation and tourism projects. The proposal
to replace the wiregrass range with carpet grass did not
receive a high priority. The Works Progress Administration employed about 500 unemployed workers. In 1938
management of the projects was leased to the State of
Georgia. Some of the projects were continued until 1944,
but the establishment of the Okefenokee National Wildlife
Refuge in 1937 and the Second World War restricted their
priorities and funding. Two decades of neglect and fire
suppression resulted in a buildup of fuel in the replanted
forest. In 1955, most of the pines were killed by a wildfire
that swept across much of south Georgia.
Speaker Presentations 45
Lessons Learned About Ground-Layer Restoration
Joan Walker1 and Lin Roth2
1
2
Department of Forestry and Natural Resources, Clemson University, South Carolina, 29634, USA
Belle W. Baruch Institute of Coastal Ecology and Forest Research, Clemson University, South Carolina, 29634, USA
Introduction
Methods
Ground layer restoration in longleaf pine communities is an
area of active investigation, through adaptive management
projects and formal research. While there is no comprehensive “Restoration Manual” for the longleaf pine community,
restoration practitioners develop their action plans based on
an ecological reference model and project goals, and
achieve their objectives using conventional natural resources management and horticultural methods. We recognized the growing base of practice-based knowledge, and in
2003 proposed to capture that knowledge by visiting restoration projects, interviewing those who developed and managed the projects, and organizing that information into “A
Practical Guide to Restoring the Ground Layer of Upland
Longleaf Pine Communities.” This project was initially
funded by the natural resources program at Fort Gordon.
This Department of Defense installation is located on the
fall-line in Georgia and has an abundance of upland, sandy
soils where restoring the ground layer was planned. For
this reason, the project focuses on upland, dry to intermediate moisture sites. In this paper, we summarize general
lessons learned from restoration practitioners. The Guide
has been drafted, but a date for its publication is not determined; we will communicate that information through the
Longleaf Alliance when it is available. Additional information on longleaf pine ground layer restoration is given in
Walker and Silletti (2006).
In 2003, we searched for “study sites” and contacted possible contributors. During 2003 and 2004 sites were visited, initial interviews about projects were conducted, and
results were assembled. As the format for the guide solidified, we communicated with contributors to fill in
missing data. For well-developed projects we prepared
detailed Case Studies that described aspects of restoration
in detail, including the reference model, starting conditions, prior land use, goals, methods, costs, results. For
smaller projects we prepared less complete descriptions,
but highlighted some special features or learning opportunities.
Our preliminary review of restoration projects showed that
they are being conducted on a relatively narrow subset of
possible longleaf pine habitats. Significant projects we
knew of when we began were concentrated in the Atlantic
and Gulf Coastal Plains; mesic savannas and flatwoods,
loamy upland sites, and xeric to sub-xeric sites in the fallline sandhills were and are represented. We note the absence of projects in the middle Atlantic Coastal Plain where
few examples of remnant vegetation remain, in the mountain longleaf pine communities of the Blue Ridge and Cumberland Plateau, and in the longleaf pine-bluestem communities. Information presented in this brief overview draws
on projects conducted by researchers and restoration practitioners in Florida, South Carolina, and Georgia.
Results
The individual projects varied widely in starting site conditions, restoration objectives, approaches, methods, and
resources available for work. The project yielded much
valuable specific information about how to do restoration,
and we attempted to organize the details so that some
wanting to do a restoration project could learn from these
case studies. Here we present general lessons that were
learned by examining a variety of restorations.
We discuss eight general lessons learned from reviewing
restoration projects and developing Case Studies (Table
1).
Table 1. Eight lessons learned from restoration case studies.
1.
2.
3.
4.
5.
6.
7.
8.
Restoration is site specific; know the site conditions.
Historical land use contributed to current conditions that
drive restoration protocols.
Protocols generally included control of undesirable species, adding desirable species, and burning.
“Undesirable species” are not created equal.
·
Exotic pasture grasses are not off-site pines—that is,
you may choose to keep some of them for awhile.
Many different native species can be established from
seed.
·
Start with good seed.
·
Control competition.
·
Don’t bother with irrigation or fertilization.
A range of seed collection and sowing methods are used
successfully.
Plan for drought and deluge; weather happens.
Restoration is a long-term commitment.
Speaker Presentations 46
1. Restoration is site specific; it is critical to know
and understand the current and previous site conditions.
The physical conditions, such as soil type or slope, at a
restoration site influence which species can grow there
and how effective management actions will be. Further,
the longleaf ecosystem wide area, but most of the species in the system have much smaller ranges. It is important to understand the historical ranges of species in
order to make good choices about what species to reintroduce. In the studies we reviewed, practitioners used
reference information to help identify ecological goals
for restoration projects.
Intact remnant patches of the target ecosystem, such as
nearby “natural areas”, were sometimes identified as
reference sites. Generally, reference sites are selected to
match the restoration project site with respect to geography and physical environment and are believed to represent the historic or contemporary potential conditions.
Besides reference sites, other kinds of reference information include a current site assessment of the project
area along with accurate historical information about
the same site. Desirable historical information includes
historical photographs, written descriptions, plant and
animal species lists, how often the area burned and under what conditions, and/or reports of significant disturbances or past land uses.
Practitioners sometimes used historical or contemporary
information from other sites, or from less specific geographic areas. Though such information may be useful,
it is important to remember that information about
places is generally place- and time- specific. The more
distant or more general the information source, the less
likely it will accurately represent a specific project site
and the less useful it will be for setting feasible objectives.
2. Historical land use contributed to current conditions that drive restoration protocols.
Altered fire regimes, forest management, and agriculture (animal husbandry and row crops) plantation establishment, and conversion of forest lands to agriculture
have resulted in loss of the ground cover diversity
throughout the longleaf pine. Mining was a more localized use, but one that produced profound changes. Ongoing and completed restoration projects that we reviewed all fall into one of these recent land use history
classes. The recent land use strongly determined the
protocols for restoration. For example, sites changed
primarily by an altered fire regime often retained valuable trees, but had accumulated dense litter and duff
layers and lost much of the herb layer. In contrast, pastures had no trees and were dominated by exotic
perennial grasses, very difficult to eliminate and control.
Pine plantations were highly variable with conditions affected by intensity of site preparation, fire management,
intervening treatments, and even the age of the stand. They
generally had high pine densities compared to reference
sites and low herbaceous cover; the pine species may be
longleaf or another species; hardwoods may be increased or
decreased.
3. Protocols generally included (1) control of undesirable
species, (2) adding desirable species, and (3) burning.
As suggested in the previous section, the current conditions
created by land recent uses determined the relative need for
any of these 3 activities; however nearly all sites required
some sort of site preparation to control competition. Mechanical treatments, chemical treatments (herbicides), and
prescribed burning were applied in the projects were reviewed. In the case of pasture grasses, multiple chemical
treatments through 1 or 2 growing seasons were applied
prior to planting desired native grasses and forbs. Prescribed fire was always recommended for maintaining restoration sites.
4. “Undesirable species” are not created equal.
Species that may be considered “undesirable” include native weedy species that are better competitors than desired
reintroductions, exotic species (aggressive or not), and
pines other than longleaf pine. Immediate elimination is not
necessarily the only choice in a restoration. Although it
was always considered necessary to eliminate perennial
pasture grasses, in some cases it was helpful to retain a canopy of off-site pines to provide habitat value for wildlife or
to prove fine fuels as the ground layer developed. Off site
pines may be restored gradually to longleaf.
5. Many different native species can be established from
seed or grown into plugs and out-planted. Start with
good seed. Control competition. Don’t irrigate or fertilize.
More information about seeds, seed cleaning, and viability
are provided by Glitzenstein, et al. in this Proceedings.
Many seeds do not need special treatments to stimulate germination, though cold stratification may benefit composites
and heat treatments or scarification may facilitate germination for legume species.
Factors that can affects successful establishment include
suitable temperatures, adequate moisture during early seedling development, presence of surface litter, and abundance
of competitors. As noted above, site preparation to control
competition is usually beneficial, but cover or nurse crops,
fertilization, and mulch are not beneficial. Assuring good
seed to soil contact is beneficial and can be achieved by
rolling the surface after planting. Viable seed can be sown
directly into prepared sites, or grown into nursery stock for
Speaker Presentations 47
out-planting. Both approaches have advantages (Table 2).
8. Restoration is a long-term commitment.
6. A range of technologies are available to accomplish
tasks associated with species additions.
The final strong lesson is that restoration is a multi-year
process. Even after all the undesirable species are removed and the desired species added, it is necessary to
continue some treatments, e.g. prescribed burning, and to
monitor (observed systematically to detect changes as
they occur) to make sure the restored site continues to
develop as expected.
Seed collection can be done by hand or with a variety of
machines including handheld models based on “weedeaters” and large (12 foot wide) tractor mounted rotating
brushes. The costs of mechanical seed harvesters may be
prohibitive for individuals, but they are becoming more
common in the region, and arrangements for sharing may be
possible. Collecting seed at the time when it would naturally ripen is recommended. Like collection, seeding may
be accomplished by hand (inexpensive and a good opportunity to involve community volunteers) or mechanically with
hay blowers (widely available, but not efficient, or specialized seed drills. Plugs are readily planted manually, and
have been successfully planted with a tree planted. Planting
just ahead of the rainy season is advised. In projects we
reviewed, post-planting management varied from handweeding to herbicide application, and prescribed burning.
7. Plan for drought and deluge.
Nearly every project manager interviewed reported stories
of “setbacks” associated with extreme weather events, from
drought to heavy rains. Weather happens. These restoration practitioners learned to plan for it, thereby minimizing
despair when the inevitable occurred.
Status of the Practical Guide project
A complete draft of the Guide has been completed, but we
do not know when it will be published and ready for circulation. We will share this information through the
Longleaf Alliance as soon as it is available.
Literature Cited
Walker, J.L. and A.M. Silletti. 2006. Restoring the ground
layer of longleaf pine communities. In: S. Jose, Jokela, and
Miller, eds. Longleaf pine Ecosystems: Ecology, Management,
and Restoration. Springer-Verlag.
Acknowledgments
Thanks to Fort Gordon for funding this project and for opportunities to learn about gound layer restoration in practice. A special thanks to all those who welcomed Lin to their field sites,
shared their knowledge, and responded to our relentless requests
for more information.
Table 2. Comparison of direct seeding and out-planting to reintroduce species into restoration sites.
DIRECT SEEDING
OUT-PLANTING PLUGS
Advantages
Advantages
Can choose individual target species
Economical ($ 3K/acre)
Simultaneously introduce many species
No need to disrupt existing conditions
Can create custom seed mixes by varying
No special planting tools
Can be done on slopes where seeding equiptiming and methods of collection
Can be done with site preparation
ment cannot be used safely
Can be done in winter when competition for
Conducive to volunteer assistants
Good success for many species
labor is lower
Can treat large areas
Reduced early drought susceptibility
Genetically diverse seeds can be used so
Appropriate for rare species
Few seeds are needed
that site conditions “select” most suitStock can be propagated any time
able individuals
Shorter period of competition control
Disadvantages
Disadvantages
Expensive (up to $ 10K /acre)
Large seed supplies are required
Not as useful for rare species
Introduce only one species at a time
Special care needed to create seed mixes
Available stock may be limited by seed
Seeding rates difficult to determine to enavailability, nursery size
Germination and establishment in greensure outcome
Competition control essential
house conditions may favor genotypes
less suitable for field conditions
Speaker Presentations 48
Poster Presentations 49
Poster Presentations
Use of Herbicide Site Preparation Treatments to Promote Longleaf Seedling Growth and to
Enhance Fuels Structure for Longer Term Fire Management
Robert N. Addington1, Thomas A. Greene2, Catherine E. Prior1, Wade C. Harrison1
1
The Nature Conservancy, Fort Benning Field Office, Georgia, 31905, USA
2
The Nature Conservancy, Fort Hood Field Office, Texas, 76545, USA
Abstract
Introduction
The positive influence of site preparation treatments on
planted longleaf pine (Pinus palustris Mill.) seedling
growth has been documented by numerous studies. Yet,
in addition to growing trees faster, site preparation treatments may also be aimed at enhancing fuel cover and
composition so that longer term management and ecosystem restoration goals can be achieved with fire alone. We
describe results from a field study on Fort Benning, GA,
whereby herbicide site preparation treatments were applied with the goal of reducing woody plant competition
without negatively impacting native warm-season perennial grass fuels. Two herbicide treatments commonly
used on Fort Benning were compared to one another, and
to an untreated control. Herbicide treatments included (1)
imazapyr/glyphosate and (2) hexazinone. Two years after
planting, treatment effects on longleaf seedling growth
were obvious – root collar diameter was an average 40%
higher on treated plots compared to control plots while
height growth was two-fold greater. Effects of herbicides
on woody stem density were variable, but by 2006 hexazinone treated plots had significantly fewer woody stems
compared to imazapyr/glyphosate and control plots, and
both herbicide treatments significantly reduced the density
of hardwood tree species, such as sweetgum
(Liquidambar styraciflua). No significant impact on
warm-season perennial grasses was detected for either
herbicide. All sites were burned two years after planting.
Fire effects measurements indicated more desirable fire
intensity on treated plots, hexazinone plots in particular.
Overall, hexazinone plots appeared better poised for
longer term fire management compared to imazapyr/
glyphosate plots. Release of woody shrubs and vines was
lower on hexazinone plots and bluestem grasses also responded exceptionally well. Our results suggest that herbicide site preparation treatments are important not only
in promoting seedling growth, but are also effective in
enhancing fuels structure and may be an important early
step in establishing a desirable longleaf pine restoration
trajectory.
Fort Benning has pursued an aggressive longleaf pine
restoration strategy for its uplands for the past decade,
primarily for the purpose of restoring habitat for the federally endangered red cockaded woodpecker (Picoides borealis) (USAIC 2001). In total, Fort Benning’s restoration
project covers some 90,000 acres, most of which is believed to have been occupied by longleaf pine but is presently dominated by mixed pine-hardwood communities.
A variety of techniques are typically employed by Fort
Benning land managers to promote longleaf pine, including selective harvesting of non-longleaf timber and prescribed fire. Additionally, roughly 1000-1500 acres of
upland forest, typically plantation loblolly, are clearcut
each year and replanted to longleaf pine. These areas are
site-prepared, planted, and then included in the installation's prescribed fire program, which has a 2-3 year fire
return interval. Site preparation methods that minimize
soil surface disturbance are generally favored. For this
reason, herbicide treatments followed by burning are often
the chosen method of site preparation. The objective of
this study was to compare the effectiveness of two herbicides commonly used on Fort Benning in reducing woody
plant competition and promoting longleaf pine seedling
growth. Additionally, we were interested in evaluating
the impact of herbicides on herbaceous fuels – warmseason perennial grasses in particular – as continued management of young longleaf plantations on Fort Benning
will rely primarily on prescribed fire alone.
Methods
Fort Benning is an 182,000 acre U.S. Army training installation located in west-central Georgia and eastern Alabama (32.4°N latitude, 84.8°W longitude). The area selected for study was a patchwork of upland loblolly pine
plantations located at the southern tip of the installation
that were clearcut harvested in 2002. Study sites were
established in May-June 2003 using a randomized complete block design with six replications (sites) and three
treatments (imazapyr/glyphosate, hexazinone, and untreated control) for a total of
Poster Presentations 50
18 plots. Sites were located across a range of soils, from
loamy sands to sandy loams, representative of upland
soils found on Fort Benning. Each plot was 2 ´ 3 chains
(20 ´ 40 m) in dimension. Herbicide treatments were applied at recommended rates during August 2003 as follows:
(1) imazapyr/glyphosate:
16 oz. imazapyr
(Arsenal®) product/acre (0.28 kg/ha active ingredient) per 4 quarts glyphosate (Accord®) product/
acre (4.5 kg/ha active ingredient) plus 0.5% nonionic surfactant
(2) hexazinone: 4 lb. (Velpar® ULW) product/
acre (3.36 kg/ha active ingredient)
Imazapyr/glyphosate was applied in liquid form using a
backpack sprayer and hexazinone was in granular form
applied by an operational contractor using a skidder
mounted applicator. All sites were burned following initial brownout. Nursery-grown, containerized longleaf
pine seedlings were then planted during November 2003
at a spacing of 6 ´ 12 ft to achieve a density of 605 seedlings per acre. Seedlings came from a regional supplier
(International Forest Company; Moultrie, GA) and were
8-9 months old at the time of planting. No further herbicide treatments were applied.
Field measurements were conducted on all plots at the
time of plot establishment in May-June 2003, prior to the
herbicide treatments, and again in June 2005, the second
growing season after planting. All sites were burned
again in January 2006 and post-burn data were collected
within six weeks of burning. Plots were then inventoried
again in June 2006. All measurements were made along
diagonal transects bisecting each plot using a 1-m2 PVC
frame (subplot). A total of ten 1-m2 subplots were measured per plot. Within each subplot, percent cover of perennial grasses was estimated visually to the nearest
percent. Perennial grasses included: Andropogon virginicus, Andropogon/Schizachyrium spp., Aristida purpurascens, Chasmanthium sessiliflorum, Dichanthelium spp.,
Panicum spp., Paspalum spp., Paspalum urvillei, Saccharum spp., Setaria spp. Several of these genera also contain
annual species, but effort was made to restrict measurements to perennial species only. Emphasis was placed on
perennial grasses over other herbaceous groups (like forbs)
because perennial grasses tend to be the most important
herbaceous group for carrying fire during the dormant season. All woody trees, shrubs and vines rooted within each
subplot were tallied by species to yield woody plant density
(#/m2). Planted longleaf pine seedlings were tallied within
a 4 m wide belt transect within each plot following their
second growing season. Root collar diameter and terminal
bud height were also measured on each seedling. Post-burn
assessments were made within each 1-m2 subplot using a
qualitative scoring system described by the National Park
System (USDI NPS 2003). Scores ranged from 1 to 5, with
1 being heavily burned and 5 being unburned. Treatment
effects
were evaluated using Analysis of Variance
(ANOVA) with multiple comparison procedures in SAS
(SAS v. 9.1.3; SAS Institute, Cary, NC, USA). Where multiple years of data were present, repeated measures
ANOVA was conducted with pretreatment (2003) data used
as a covariate. Slope homogeneity tests and treatment ´
year interactions were evaluated to determine how plots
changed relative to one another over time. Data were transformed as necessary to meet assumptions for parametric
tests. Both arcsine (for percentage data) and log transformations were used where needed. Statistical significance
was interpreted using α = 0.05.
Preliminary Results and Discussion
Longleaf seedling root collar diameter was significantly
greater for both imazapyr/glyphosate and hexazinone
Figure 1. Longleaf seedling (A) root collar diameter and (B) height measured at the beginning of
their third growing season (2006); averaged by treatment (n = 6 reps per treatment; means ±1 SE).
Means with different letters are significantly different at α = 0.05.
Poster Presentations 51
Figure 2. (A) Woody stem density averaged by treatment for each of the measurement years (n = 6 reps
per treatment, except 2005 n = 5 for control plots due to missing data; means ±1 SE). Significant differences among means at α = 0.05 are denoted by different letters. (B) Mean change in woody stem density
from 2003 to 2006 for trees, shrubs and vines.
treated plots relative to control plots (Figure 1A). For both
treatments, seedlings were typically at or above the 25 mm
approximate diameter threshold at which height growth
generally initiates (Boyer 1990). Seedlings were significantly taller on hexazinone plots compared to imazapyr/
glyphosate and control plots (Figure 1B). There was no
difference in seedling density among treatments (data not
shown), indicating similar seedling survival. Effects of
herbicide treatments on total woody stem density were variable, but by 2006 hexazinone treated plots had significantly
fewer woody stems present compared to imazapyr/
glyphosate and control plots (Figure 2A). The density of
hardwood trees species such as sweetgum (Liquidambar
styraciflua), red maple (Acer rubrum), and water oak
(Quercus nigra) was significantly lower for both imazapyr/
glyphosate and hexazinone plots compared to control plots
(Figure 2B). This was the one trend observed to date that
could potentially explain the increase in seedling growth on
treated plots. Total woody stem density on imazapyr/
glyphosate plots was no different from control plots in 2006
because of the prodigious release of shrubs such as sand
blackberry (Rubus cuneifolius) that occurred on these plots
(Figure 2A, B).There was no adverse effect of either herbicide treatment on warm-season perennial grass cover
(Figure 3). In fact, the percent increase in cover from 2003
to 2005 and from 2003 to 2006 was slightly though not significantly greater on both imazapyr/glyphosate and hexazinone treated plots compared to control plots. Bluestem
grasses (Andropogon spp. and Schizachyrium scoparium)
responded particularly well to hexazinone (data not shown).
Percent cover of these species in 2006 after the prescribed
burn was sixteen times greater than that measured in 2003
on hexazinone plots. Percent cover of these same species
over this time period was seven times greater on control
plots and only about two times greater on imazapyr/
glyphosate plots. An increase in graminoid cover following
hexazinone treatments in particular has been documented
by other studies as well (Brockway and Outcalt 2000, HaySmith and Tanner 2002).
Post-burn measurements also showed more favorable fire
intensity on treated plots compared to control plots (Table
1). Again, statistically significant differences were observed only for hexazinone plots. Within-plot variation in
burn severity scores was also evaluated to assess burn
patchiness or continuity. No significant differences
emerged from this analysis, however, indicating similar
burn continuity among treated and control plots.
Figure 3. Percent cover of warm season perennial grasses
averaged by treatment (n = 6 reps per treatment, except
2005 n = 5 for control plots due to missing data; means ±1
SE). No significant differences between treatments were
found and there was no interaction between treatment and
year.
Poster Presentations 52
Table 1. Mean burn severity code as defined by the
National Park Service (USDI NPS 2003) averaged by
treatment (n = 6 reps per treatment; means ±1 SE).
Measurements were made following a prescribed fire in
2006. Codes are from 1 to 5, with 5 = unburned and 1
= heavily burned. Means with different letters are significantly different from one another at α = 0.05.
Acknowledgements
We thank Fort Benning’s Environmental Management Division (EMD), specifically John Brent, Bob Larimore, and
Pete Swiderek, for supporting this project. We also thank
Mark Byrd for help with site selection and herbicide application, Ricky Caldwell and Brant Slay for help with fieldwork, and Matt Nespeca and Don Imm for reviewing earlier
versions of this report. Funding for this project was provided by the Department of Defense.
Treatment
Mean Burn
Severity Code
Standard
Error
Statistical
Significance
(α = 0.05)
Imaz/Glyph
3.72
0.301
ab
Literature Cited
Hexazinone
3.59
0.225
b
Control
4.30
0.092
a
Boyer, W.D. 1990. Pinus palustris., Mill. Longleaf Pine.
In: Burns, R.M. and B.H. Honkala (Eds.), Silvics of
North America, Vol. 1, Conifers. USDA Forest Service
Handbook 654. pp. 405-412.
Brockway, D.G. and K.W. Outcalt. 2000. Restoring longleaf pine wiregrass ecosystems: Hexazinone application
enhances effects of prescribed fire. For. Ecol. Manage.
137:121-138.
Hay-Smith, L. and G.W. Tanner. 2002. Restoring longleaf
pine sandhill communities with an herbicide. University of Florida Extension, WEC131. 6 p.
USAIC (U.S. Army Infantry Center). 2001. Integrated
Natural Resources Management Plan, Fort Benning
Army Installation 2001-2005.
USDI NPS (U.S. Department of the Interior National Park
Service). 2003. Fire Monitoring Handbook. Boise (ID):
Fire Management Program Center, National Interagency Fire Center. 274p.
Conclusions
Both herbicide treatments had favorable effects on longleaf seedling growth and fuels structure compared to control plots, though of the two herbicides, hexazinone
treated plots appeared better poised for longer term management with fire alone compared to imazapyr/glyphosate
plots. Woody plant density was significantly lower on
these plots, bluestem species responded exceptionally
well, and burn severity measures indicated more favorable
fire intensity likely necessary for continued control of
woody competition. We emphasize herbicide site preparation for creating or enhancing desirable fuels structure,
under the assumption that other desirable ecosystem attributes such as groundcover species richness and diversity will come with time following repeated fire application. Future work, however, should evaluate the impact of
herbicide treatments and any associated legacy effect on
other ecosystem attributes. Results from this study,
though, suggest that herbicide site preparation treatments,
hexazinone treatment in particular, are effective in promoting longleaf seedling growth and are likely an important early step in establishing a desirable longleaf pine
ecosystem restoration trajectory by enhancing fuels structure.
Poster Presentations 53
Longleaf Pine Plant Community Restoration at the Savannah River Site:
Design and Preliminary Results
Todd A. Aschenbach1, Bryan L. Foster1, and Don W. Imm2,3
1
Department of Ecology and Evolutionary Biology, University of Kansas, Lawrence, Kansas, 66045, USA
2
Strategic Environmental Research and Development Program, Ft. Benning, Georgia, 31995, USA
3
Savannah River Ecology Laboratory, University of Georgia, Aiken, South Carolina 29803, USA
Introduction
Losses of longleaf pine (Pinus palustris) ecosystems in the
southeastern United States have reduced the original extent
of longleaf pine ecosystems by 97% (Frost 1993, Outcalt
and Sheffield 1996). As a result, there is considerable interest in attempting to restore elements of the historic longleaf
pine ecosystem. Here we introduce a long-term experimental study established at the Department of Energy Savannah
River Site (SRS) in South Carolina to evaluate the potential
for restoring long-leaf savannah at local and landscapescales.
Restoration efforts at the landscape scale are need in order
for restoration to be an effective tool for biodiversity and
ecosystem conservation (Bell et al. 1997, Radeloff et al.
2000). However, before developing landscape-level restoration strategies, one must understand the constraints to reestablishment, persistence and dispersal of species within a
landscape. Constraints can be viewed from regional to local
levels of organization (Figure 1). Regional species pools
can place strong constraints/filters on plant colonization,
species composition and diversity at the local scale. The
diversity and composition of regional species pools may
vary depending on large-scale patterns of migration and
speciation, and the cumulative effect of anthropogenic
disturbance and subsequent loss of regional species
pools (Zobel 1997, Foster 2001). Seed addition studies
have shown that species diversity can be limited by
dispersal (Tilman 1997, Foster and Tilman 2003, Foster
et al. 2004). Therefore, dispersal constraints from the
regional species pool can be circumvented by the addition propagules. Local abiotic and biotic conditions act
as filters to invasibility, population persistence and thus
community assembly (Zobel 1997, Rejmánek 1996).
Abiotic resistance to establishment comprises the impacts of local edaphic conditions, microclimatic conditions, etc., and determines which subset of species from
the regional pool is able to survive under the local
abiotic conditions. Biotic resistance comprises the influence of species interactions on establishment and
persistence of available, abiotically-suitable species.
The disturbance regime also acts as a major selective
factor in local community assembly through its mediation of both abiotic and biotic conditions. In the context
of restoration, the re-establishment of fire alters both
Figure 1. Conceptual model showing the multi-scaled constraints on community assembly at the local scale.
Native Species Pool
Dispersal Filter
Intervention:addition
of seeds, transplants.
Intervention:re-establish
fire, species-site matching.
Local Abiotic Filter
Intervention:re-establish
fire, stand thinning, weed
control.
Local Biotic Filter
Local Extant Community
Poster Presentations 54
abiotic and biotic conditions and aid in the reestablishment of conditions suitable for the persistence of
native species.
At the Department of Energy’s Savannah River Site
(SRS), a long-term and landscape-level approach to longleaf pine savannah restoration was initiated in 2001 in
order to investigate the relative importance of dispersal
limitation, abiotic and biotic resistance in constraining
colonization and community assembly. Reduced regional
species pools and local degradation of existing savannah
vegetation (Duncan et al. 1996) necessitates intense restoration efforts including the addition of native plant
propagules. Intensely restored native plant communities
located throughout the landscape will serve as local
propagule pools for the remaining landscape, including
degraded longleaf pine forests located outside the SRS
boundaries.
Figure 2. Illustration of the 2 x 2 factorial experimental
design used at each of six separate study sites to examine
interactions between wiregrass (Aristida beyrichiana) and
other non-matrix species. +/- ARI = presence or absence
of A. beyrichiana; +/- NM = presence or absence of nonmatrix species.
In this experiment, biotic resistance is investigated
through the interactions between wiregrass (Aristida beyrichiana) and other planted species. Wiregrass helps to
regulate the floristic composition of longleaf pine ecosystems primarily due to its ability to act as an important fuel
source for prescribed and natural fires and is therefore
considered a keystone component of the longleaf pine
ecosystem (Clewell 1989, Noss 1989, Platt et al. 1989).
This facilitation of fire helps to reduce the extent of invasive species that are poorly adapted to fire (Ahlgren 1979,
Provencher et al. 2001, Reinhart and Menges 2004). Although wiregrass can help facilitate the establishment of
other fire-adapted herbaceous species, it may also act as a
strong competitor with other herbaceous species. As a
result, we have incorporated the planting of wiregrass to
investigate the potential impacts of this grass component
on the performance of the other transplanted species.
Methods
The objectives of this project are to a) reestablish native
understory savannah vegetation, b) examine the interplay
between local abiotic and biotic factors in constraining
species establishment, persistence, and coexistence, and
c) explore the potential of intensively restored founder
populations to disperse to areas throughout the adjoining
landscape. Here we describe how the experimental design of the project and present preliminary data on survivorship of the restoration plantings in light of differing
site conditions and biotic interactions among wiregrass
and other planted species.
This study was conducted at six separate sites within the
80,125 ha (310 mi2) U.S. Department of Energy Savannah River Site (SRS) in Barnwell and Aiken Counties,
South Carolina. All study sites were in agricultural production prior to the establishment of SRS in 1950.
Within 25 years after the establishment of SRS, all study
sites were planted to longleaf pine plantations.
Table 1. A biotic characteristics of restoration sites at SRS including total soil nitrogen (TN), total
soil carbon (TC), total phosphorus (TP), and total organic matter (TOM). Year of most recent burn
and mapped soil type area also given. Values within each measurement category are significantly
different (P<0.050) between sites that exhibit the same letters with different subscripts.
TN (%)
TC (%)
TP (%)
TOM (%)
Recent burn Soil Type
Site 1
0.024 a0
0.86
10.6 a0b0
1.4
1997
FuB Fuquay Sand (FuB)
Site 2
0.037
1.30
5.1 a0b1
2.2
2001
Troup Sand (TrB)
Site 3
0.029
0.88
9.6 a0b0
1.6
1993
Troup Sand (TrB)
Site 4
0.029 a0
1.06
12.3 a0b0
2.0
2005
Fuquay Sand (FuA)
Site 5
0.024 a0
0.90
10.6 a0b0
1.6
2000
Dothan Sand (DoB)
Site 6
0.045 a1
1.46
3.6 a1b0
2.3
2005
FuB Fuquay Sand (FuB)
Poster Presentations 55
Figure 4. Overall species survivorship (+ SE) among restoration sites. Overall survivorship is
significantly different (P<0.050) between sites that exhibit the same letters with different subscripts.
Figure 5. Comparison of non-matrix species survival (+ SE) with and without wiregrass (Aristida
beyrichiana) among restoration sites. An asterisk (*) indicates a significant difference at P<0.050
in non-matrix species survival.
Site preparation prior to restoration plantings consisted of
thinning the 30-50 year old longleaf pine stands to densities of 20-59 trees/ha in March 2001 followed by an
herbicide application in June 2002. Container-grown
transplants of wiregrass (Aristida beyrichiana), an important matrix grass of the longleaf pine savannah understory, and 29 other herbaceous “non-matrix” species native to longleaf pine savannahs were planted in July 2002
and June 2003 at a rate of 1-12 individuals per plot.
Plantings were irrigated for one month after planting with
approximately 8.4-12.6 L water/m2 only if a precipitation
event of 1.27 cm (0.5 in) did not occur within one week.
Plantings were not fertilized at any time during the
study.A 2 X 2 factorial design was used to elucidate biotic
interactions between Aristida beyrichiana and other nonmatrix species. Treatments were replicated 25 times and
randomly assigned to study plots at each site. Each plot
was 3 x 3 m in size resulting in a total study area of 900
m2 at each site (Figure 2).
Average total soil nitrogen, carbon, phosphorous, and
organic matter were determined from 12 42.75 cm3 soil
samples collected from the top 15 cm at each site in September 2005. Site characteristics, including recent prescribed burn history, are given in Table 1. All plots were
surveyed in June 2004 for vegetative cover and individual
number of all planted species.
Little bluestem
(Schizachyrium scoparium) and slender little bluestem (S.
tenerum) are combined for data analyses. Scheirer-RayHare, Kruskal-Wallace and Mann-Whitney U tests were
used for statistical analyses.
Poster Presentations 56
Results and Discussion
Overall species survival was greatest at site 4 at 52%
which was significantly greater (P<0.050) than species
survival at sites 2 and 5. Species survival at site 6 was
significantly lower than at all other restoration sites
(Figure 4). Site 6 exhibits the greatest soil nutrient levels
compared to all other sites (Table 1). Therefore, the lower
survival of planted species at site 6 may be due to competitive exclusion by non-planted species.
Non-matrix species survival is positively affected by the
presence of wiregrass (Aristida beyrichiana), but this relationship depends on site conditions. At sites 4 and 6, nonmatrix species survival was significantly greater
(P<0.050) in plots with wiregrass than in plots without
wiregrass (Figure 5). However, this increase in survival
for non-matrix species cannot be the result of facilitation
of fire by wiregrass since no sites had been burned between the time of planting and the collection of these data
(2004). This increase in survival of non-matrix species
may still reflect a type of abiotic facilitation by wiregrass,
however, the specific mechanism for this facilitation has
yet to be determined. The pine savannah restoration project at SRS provides for the opportunity to examine how
abiotic and biotic factors affect species establishment and
coexistence at the local level. Specifically, the planting
design employed allows for the examination of the biotic
interactions between wiregrass and other planted species.
Continuing research will examine these interactions as a
function of differing site conditions. In addition to elucidating the biotic interactions between wiregrass and other
species, this experiment will also help to determine the
limits to establishment of other understory species. Furthermore, this landscape-level project provides for the
opportunity to gauge the potential of intensively restored
founder populations to disperse to areas throughout the
adjoining landscape. Tracking the dispersal and colonization of planted species over time will help to elucidate the
advantages and shortfalls of this approach to landscape
restoration.
Literature Cited
Ahlgren, C.E. 1979. Emergent seedlings on soil from
burned and unburned red pine forest. Minnesota Forest Research Notes 273. 4 pp.
Bell, S.S., M. S. Fonseca, and L.B. Motten. 1993. Linking
restoration and landscape ecology. Restoration Ecology 5:318-325.
Clewell, A.F. 1989. Natural history of wiregrass (Aristida
stricta Michx., Gramineae). Natural Areas Journal
9:223-233.
Duncan, R. P., and R. K. Peet. 1996. A template for reconstructing the natural, fire-dependent vegetation of
the fall-line sandhills, south-eastern United States.
(Report 96-23-R).
Foster, B.L. 2001. Constraints on colonization and species
richness along a grassland productivity gradient: the
role of propagule availability. Ecology Letters 4: 530535.
Foster, B.L. and Tilman, D. 2003. Seed limitation and the
regulation of community structure in oak savannah
grassland. Journal of Ecology 91: 999-1007.
Foster, B.L., T.L. Dickson, C.A. Murphy, I.S. Karel, and
V.H. Smith. 2004. Propagule pools mediate community
assembly and diversity-ecosystem regulation along a
grassland productivity gradient. Journal of Ecology 92:
435-449.
Frost, C.C. 1993. Four centuries of changing landscape pattern in the longleaf pine ecosystem. Pages 17-43 inS.M.
Hermann, ed., Proceedings of the 18th Tall Timbers
Fire Ecology Conference, The longleaf pine ecosystem:
Ecology, restoration and management. Tall Timbers
Research, Inc., Tallahassee, Florida.
Noss, R.F. 1989. Longleaf pine and wiregrass: Keystone
components of an endangered ecosystem. Natural Areas Journal 9:211-213.
Outcalt, K.W. and R.M. Sheffield. 1996. The longleaf pine
forest: Trends and current conditions. Resource Bulletin SRS-9. Asheville, N.C.: U.S. Dept. of Agriculture,
Forest Service, Southern Research Station. 23 p.
Platt, W.J., J.S. Glitzenstein, and D.R. Streng. 1989. Evaluating pyrogenicity and its effects on vegetation in longleaf pine savannahs. Pages 143-163 in S.M. Hermann,
ed., Proceedings of the 17th Tall Timbers Fire Ecology
Conference, High intensity fire in wildlands: Management challenges and options. Tall Timbers Research
Station, Tallahassee, FL.
Provencher, L., B.J. Herring, D.R. Gordon, H. L. Rodgers,
K.E.M Galley, G.W. Tanner, J.L. Hardesty, and L.A.
Brennan. 2001. Effects of hardwood reduction techniques on longleaf pine sandhill vegetation in northwest Florida. Restoration Ecology 9:13-27.
Radeloff, V.C., D.J. Mladenoff, and M.S. Boyce. 2000. A
historical perspective and future outlook on landscape
scale restoration in the northwest Wisconsin Pine Barrens. Restoration Ecology 8:119-124.
Reinhart, K.O. and E.S. Menges. 2004. Effects of reintroducing fire to a central Florida sandhill community.Applied Vegetation Science 7:141-150.
Tilman, D. 1997. Community invasibility, recruitment limitation, and grassland biodiversity. Ecology 78: 81-92.
Zobel, M. 1997. The relative role of species pools in determining plant species richness: an alternative explanation of species coexistence. Trends in Ecology and
Evolution 12: 266-269.
Poster Presentations 57
Longleaf pine ecosystem management at Eglin Air Force Base, Florida
Chadwick Avery1
1
Eglin Air Force Base, Florida, 32542, USA
Abstract
Eglin AFB, totaling approximately 464,000 acres (724
square miles), is the largest Air Force base in the U.S.
Located in the panhandle of northwest Florida, over
362,000 acres of this area consist of the fire-dependent
longleaf pine ecosystem in the form of longleaf pine sandhills, flatwoods, and uplands. Eglin's longleaf pine forest is of great conservation significance as it is the largest
contiguous tract of the world's remaining old growth
longleaf pine and is home to 77 state and federally listed
species, including gopher tortoises, eastern indigo snakes,
and the 4th largest population of red-cockaded woodpeckers. Eglin's forest is managed by the Natural Resource
Section within the Environmental Management Division's
Stewardship Branch. The Natural Resource Section is
comprised of the Forestry, Wildland Fire, and Wildlife
Elements, each of which is responsible for multiple programs that support Eglin's test and training mission and
Integrated Natural Resource Management Plan (INRMP).
The mission of Eglin's Natural Resources Section is to
support the Air Force through responsible stewardship of
the installation's natural resources. This is accomplished
by integrating natural resource management and using an
adaptive ecosystem management approach, which maintains ecosystem viability and conserves biodiversity while
providing compatible multiple uses. Overall, ecosystem
management and the military mission have been compatible and highly successful at Eglin.
Poster Presentations 58
Establishment and Management of Longleaf Pine ( Pinus palustris Mill.) Seed Production Areas
Jill Barbour1
1
USDA Forest Service, National Seed Laboratory Dry Branch, Georgia, 31020, USA
Abstract
Interest in planting longleaf pine (Pinus palustris Mill.)
in the southeastern United States is growing. Seedling
demand is fueling the demand for longleaf pine seed,
but due to the species’ periodic nature in cone production, seed shortages can occur; thereby, limiting the
number of seedlings that can be produced. Therefore,
more longleaf pine seed production areas are needed to
meet the demand for seed.
The reproductive biology of longleaf pine is more unpredictable than other southern pines. More care is
needed in the handling of its reproductive structures,
making the species difficult to propagate. Longleaf
pine’s vegetative grass stage creates conditions that
favor managing natural stands for seed production
rather than creating grafted seed orchards. Provenance
and progeny tests, which select for early emergence
from the grass stage, can be converted to seedling seed
orchards an and managed for cone production too.
Much of what is known about longleaf pine cone production was published over 30 years ago, and the practical experience by foresters has been lost. Not much information
has been published on managing longleaf pine seed production areas; therefore, inferences are made from the management of other southern pines species’ seed production areas.
This poster is a compilation of knowledge and practical
experience on how to establish and manage mature longleaf
pine areas for cone production.
Poster Presentations 59
The Dendrochronology of Pinus palustris in Virginia
Arvind A. R. Bhuta1, Lisa M. Kennedy1, Carolyn A. Copenheaver1 and Philip M. Sheridan2
1
Department of Geography, Virginia Tech, Blacksburg, Virginia, 24060, USA
Meadowview Biological Research Station, Woodford, Virginia, 22580, USA
2
Abstract
Dendrochronological research of Pinus palustris has been
limited to its central and southern range. For this study,
we investigated how climate and disturbance affects ring
width growth at two sites that support naturally regenerated Pinus palustris at its northernmost range in Virginia.
For both sites, we measured height and diameter of all
Pinus palustris and cored a total of 71 trees (Seacock
Swamp, n = 32; Everwoods, n = 39) greater than 10 cm in
diameter at breast height. All cores were cross-dated and
measured and cross-dating was verified using
COFECHA. Our initial within site correlations were r2 =
0.373 for Seacock Swamp and r2 = 0.377 for Everwoods.
The low correlations indicate a strong competition signal
within the tree-ring record and are probably due to the
high density of Pinus taeda in both the understory and
overstory within these stands. We calculated release and
suppression events based on boundary-line radial growth
patterns and found long periods of suppression punctuated
by release events experienced at the tree level rather than
the stand level. To account for the effects of climate on
annual ring width growth, we adjusted our within site
correlations so both sites would better correlate with climate records. This yielded within site correlations of r2 =
0.509 for Seacock Swamp and r2 = 0.494 for Everwoods.
We are presently examining the effects of monthly and
seasonal precipitation, temperature, PDSI, and PHDI on
annual ring growth of Pinus palustris from both sites.
Poster Presentations 60
Old-Growth Longleaf Pine on Horn Mountain, AL (Talladega National Forest)
David Borland1, Art Henderson2, John S. Kush3, and John McGuire3
1
2
The Nature Conservancy Birmingham, Alabama, 35210, USA
USDA Forest Service, Talladega National Forest, Talladega District, Talladega, Alabama, 35160, USA
3
Auburn University School of Forestry & Wildlife Sciences, Auburn, Alabama, 36849, USA
Abstract
Objectives
A truly unique opportunity exists to increase the knowledge base about the dynamics of longleaf pine in the
montane region. Horn Mountain is located on the Talladega Ranger District of the Talladega National Forest.
It is located about 5 miles southeast of Talladega, AL at
an elevation of 1,500 feet. As a result of an Auburn
University School of Forestry & Wildlife Sciences senior project, several observations were made that make
this area truly unique. It may be the largest intact oldgrowth longleaf stand and may have the highest density
and basal area of any old-growth stand in the montane
region.
This study will 1) describe the age and stand structure, 2)
evaluate the size, age, and variability, and 3) shed light on
past disturbance and replacement patterns of old-growth
mountain longleaf pine stands.
The US Forest Service, The Nature Conservancy and
the Auburn University School of Forestry & Wildlife
Sciences are collaborating to document conditions and
to develop a restoration strategy for this and other montane old-growth longleaf pine stands.
Introduction
Approach
We have measured and stem-mapped all living longleaf
pines > 1.0 inch DBH (100% sample) for DBH (to nearest
0.1 inch), and sub-sampled all living hardwoods and other
pines > 4.0 DBH, and distance and direction from permanent plot centers. All snags will be measured for DBH,
total height, and distance and direction from permanent plot
centers. Notes will be made of trees and snags that have
fire scars and those trees and snags that had been turpentined. All living pines > 4 inches DBH will be cored at 4
feet above the root collar. The Talladega Ranger District
has indicated they will be able to cut samples from firescarred trees which will be used to determine what some of
the fire history may have been for the montane region.
Although vestiges of the pre-European forest occur in
pockets of the present day Ridge and Valley landscape
of northern Alabama, the structure of the forest today
bears little resemblance to that which was there prior to
significant Euro-American disturbances. A few remaining untrammeled ridges in northern Alabama counties
with intact forests provide insight into the structure
(species composition) and function of the native ecotypes in the region.
Preliminary Findings
Longleaf pine is typically thought of as a coastal plain
species. Many are surprised to learn that Alabama is
unique from the other states where longleaf pine is
found in that forests once stretched into the mountains
of northern Alabama. Within the Blue Ridge and Ridge
and Valley physiographic landscape, longleaf pine was
the dominant forest cover on south and southwest facing slopes up to about 2,000 feet in elevation. Although
still a common tree, longleaf pine (and its associated
pyrogenic ground cover) often faded out in damp bottomland valleys and north facing slopes where fire frequency and intensity was greatly reduced and allowed
for the establishment of non-fire tolerant related species. However, countless microsite differences in elevation, slope, fuels, etc. allowed fire to either move into or
Several field surveys conducted by David Borland have
found the following number of species: Forbs – 37, Shrubs/
small trees – 22, Trees – 17, Grasses and sedges – 11, Vines
– 9, and Ferns – 4.
The figure above is an aerial photograph of the old-growth
longleaf pine stand being studied on Horn Mountain. The
green points indicate the location of all longleaf pine > 4.0
inches DBH (diameter at breast height).
Figures 1-3 preset the preliminary data collected on Horn
Mountain.
Work on this project will continue in early 2007.
Poster Presentations 61
60
Tre e s /acre
50
40
30
20
10
0
0
5
10
15
20
25
30
40
35
DBH(inches)
Figure 1. Diameter distribution for longleaf on Horn Mountain. This figure
indicates that fire has been lacking because of the low numbers of small diameter longleaf pine.
9
Num ber of Trees
8
7
6
5
4
3
2
1
95
10
5
11
5
12
5
13
5
14
5
15
5
16
5
17
5
18
5
19
5
20
5
21
5
22
5
23
5
24
5
25
5
85
75
65
55
45
35
15
25
0
Age Class (yrs.)
80
70
60
50
hi
te
oa
k
oa
k
w
bl
ac
k
ch
er
ry
m
on
bl
ac
k
oa
k
pe
rs
im
re
d
oa
k
re
d
no
rt
h.
hi
ck
or
y
so
ut
h.
ap
le
re
d
m
oa
k
ac
k
oo
d
bl
ac
kj
so
ur
w
bl
ac
k
ch
es
nu
t
gu
m
40
30
20
10
0
oa
k
Trees/acre
Figure 2. Age class distribution for longleaf pine on Horn Mountain. The figure indicates that longleaf pine has had frequent cone crops for regeneration.
Hardwood species
Figure 3. Species density for hardwoods on Horn Mountain. Most the species listed
would be expected in fire-maintained longleaf pine ecosystems in the montane region
of its range except for red maple and black cherry. These species provide an indication
that fire has been infrequent in the stand.
Poster Presentations 62
The Longleaf Pine Cone Crop Story
Elizabeth Bowersock1, William D. Boyer2 and John S. Kush1
1
Auburn University School of Forestry & Wildlife Sciences, Auburn, Alabama, 36849, USA
2
USDA Forest Service, Auburn Alabama, 36849, USA
Abstract
For successful natural regeneration, the minimum size of a
cone crop is considered to be 750 cones/acre or roughly 30
cones per tree. In the past 30 years, 5 of the 8 cone crops
considered adequate for natural regeneration have occurred
since 1990. The 1996 seed crop occurred throughout the
longleaf pine range.
In any given locality, longleaf pine bears irregular cone
crops. Adequate cone crops for natural regeneration of
longleaf pine typically occur every 5-7 years, much to
the frustration of forest managers. The three-year duration over which the seeds develop may be the cause of
infrequent production. A time-line will be displayed
that illustrates this three-year process. Stand characteristics which may optimize the amount of seed available
for natural regeneration will be presented.
Reason for the Problem
The visual development of longleaf pine seed extends into 3
calendar years. The following is an abbreviated guideline
for the longleaf pine seed development process.
Longleaf pine cone crops have been monitored at several locations, at various times, across the Southeast.
The longest sequence from the Escambia Experimental
Forest in south Alabama will be shown, along with one
for southwest Georgia and an overall average for 5-9
sites located across the Southeast from Louisiana to
North Carolina. Forecasts will be presented for 2006
and 2007 longleaf pine cone crops across the Southeast.
Months prior to seedfall and what happens:
27 months 22 months -
Problem
19 months One of the major concerns in longleaf pine restoration,
regeneration, and management is its relatively sporadic
seed production. Compared to the other southern pines,
longleaf is a sporadic seed producer where good seed
crops may occur every 5 to 7 years. Whether the interest is natural or artificial regeneration, it is important to
know when to expect a bountiful seed crop.
5 months -
Differentiation between male and female
flowers occur; usually July, 2-years prior
to seedfall.
Male flowers appear, usually December,
2-years prior to seedfall.
Female flowers appear and pollination oc
curs, usually February to April, 1-year
prior to seedfall.
Fertilization occurs, usually May to June
of seedfall year.
Seed ripen and fall between late September and early November.
140
C ones/Tree
120
100
80
60
40
20
0
66
68
70
72
74
76
78
80
82
84
86
88
90
92
94
96
98
00
02
04
Year
Figure 1. Average of 5 – 9 locations across the Southeast. The bold line at 30 cones per
tree represents the desired number that may be required for successful regeneration.
Poster Presentations 63
6
2006 and 2007 Cone Crop Data
Female Conelet
Male Catkin
Initiation
Male Catkins Appear
Cone
Opens
Male Catkin
Female Conelet
Buds Appear
Cone Ripens
Pollination
Rapid Cone Growth
Conelet - Early
Fertilization
Conelet - Late
Figure 2. Development cycle of longleaf pine seed
(adapted from Croker and Boyer 1975; Boyer 1990).
Long-term records of longleaf pine cone production were
obtained from natural regeneration trials conducted at the
EEF with binocular counts using the method described
by Croker (1971).
Notes
1. The 2006 cone crop estimate is above the longterm average, with bumper (100 or more cones/
tree) at one site, good at three sites, fair at two,
poor at three and failure at one (Table 1).
2. Regional outlook for 2007 longleaf cone crop,
based on flower counts, is at the failure level overall.
Failures are indicated at seven sites, poor crops at
the remaining three sites. While cone crop estimates
from flower counts are unreliable due to highly variable flower losses during first year, the lack of flowers is a reliable indication of failure.
3. The 40-year regional cone production average = 28
cones/tree. The heavy 1996 cone crop averaged 125
cones/tree; the 1987 cone crop (2nd best) averaged
65 cones/tree.
4. The normal minimum needed for successful natural
regeneration is 750 cones/acre (30 cones/tree with
25 seed trees/ acre).
Literature Cited
Table 1. Longleaf pine cone crop prospects –
2006/2007, from springtime binocular counts (spring
2006 survey). Counts made on selected mature trees
(average 15 – 17” dbh) within low-density (shelterwood)
stands.
LOCATION
ESTIMATED CONES/TREE
Escambia, AL
Santa Rosa, FL
Okaloosa, FL
Leon, FL
Baker, GA
Chesterfield, SC
Bladen, NC
Grant, LA
from conelets
2006
22
30
14
15
109
93
63
45
from flowers
2007
13
0
6
4
5
21
19
6
Lee, AL
5
5
Thomas, GA
REGIONAL AVERAGE
57
16
45.3
9.5
Boyer, W. D. 1998. Long-term changes in flowering and
one production by longleaf pine. In: Proceedings of
the Ninth Biennial Southern Silvicultural Research
Conference, T.A. Waldrop (ed.), USDA Forest Service, Southern Research Station, Gen. Tech. Rep.
20, pages 92-98.
Croker, T.C., Jr. 1971. Binocular counts of longleaf pine
strobili. USDA Forest Service, Southern Forest Experiment Station, Research Note SO-127. 3 p.
Croker, T.C., Jr. and W.D. Boyer. 1975. Regenerating
longleaf pine naturally. U.S. Department of Agriculture, Forest Service, Research Paper SO-105.
Maki, T.E. 1952. Local longleaf seed years. Journal of
Forestry. 50(4):321-322.
Wahlenberg, W.G. 1946. Longleaf pine: Its use, ecology,
regeneration, protection, growth, and management.
Charles Lathrop Pack Forestry Foundation in USDA
Forest Service. 429 pp.
Poster Presentations 64
Restoring and Maintaining Ecological Integrity of Special Communities
Embedded within Longleaf Pinelands
Joyce Marie Brown1 and Johnny P. Stowe2
1
2
University of Central Florida, Orlando, Florida 32816, USA
South Carolina Department of Natural Resources, Columbia, South Carolina 29202, USA
At the Longleaf Alliance Regional Conference in Alexandria, LA in 2000, eminent longleaf pine researcher
Dr. Bill Boyer of the USFS encouraged the group to not
-- in its zeal to restore the longleaf pine ecosystem (the
trees, and famed groundcover and fauna) -- neglect to
pay similar attention to and concomitantly restore, the
Special Communities Embedded Within the Longleaf
Pine Landscape (SCELPL). As a land manager interested in restoring native ecosystems, I (co-author
Stowe) heard this and never forgot it, recognizing it as a
path suggested by a wise man. Since then I pay special
attention to this subject. I peruse the popular and scientific literature for developments, and make special note
of SCELPLs as I manage longleaf pinelands. Especially salient and fascinating to me are Atlantic whitecedar wetlands and canebrakes. I seek to follow Dr.
Boyer's advice by bringing attention to them.
As a scientist interested in restoration and management,
I (co-author Brown) recognize the importance of a hierarchical approach to biodiversity conservation. Although restoration of longleaf habitat has been successful, I fear that SCELPLs are often neglected by a singlelevel focus on landscape or species. Because SCELPLs
interact with longleaf pine habitats through fire dynamics, hydrology, energy cycling and nutrient cycling, and
contribute to the natural heritage value of the forest, we
must preserve, maintain and restore them to protect the
ecological integrity -- i.e. the composition, structure and
function -- of longleaf pine ecosystems. The core of my
research is the study and restoration of seasonally
ponded, isolated wetlands embedded in the longleaf
pine landscape.
Seasonally Ponded Isolated Wetlands
Characteristics:
•
•
•
•
•
•
•
•
•
Dry completely periodically
No regular surface flow in or out
Habitat patches for metapopulations
May interact with groundwater
Form in shallow topographic depressions
Affect and are affected by fire dynamics
Exchange energy and nutrients with uplands
Overstory of cypress, gum or pond pine
Understory structure dependent on fire and hydrologic regime
• High diversity and productivity, especially am-phibian
Threats:
•
•
•
•
•
•
ATVs and other ORVs
Exotic/invasive species
Fertilizers, pesticides (e.g. herbicides, insecticides)
Fire suppression and firebreaks in/around wetlands
Fish reduce amphibian diversity and abundance
Disturbance associated with agriculture, silvicul t u r e
and ‘development’ (e.g. ditching, draining, filling)
Management Recommendations:
•
•
•
•
•
•
•
•
Restrict ORVs to designated trails
Do not alter natural hydrology
Restore hydrology if possible (e.g. plug ditches)
Minimize soil disturbance in/around wetlands
Do not allow fishes in fishless isolated wetlands
Remove exotics/invasives
Use only wetland-approved chemicals
Prescribed fire
Canebrakes
Characteristics:
• Arundinaria is the only bamboo native to US
• Historically were vast, dense stands of cane
(Arundinaria)
• Largest stands are/were found in alluvial floodplains
• Possible artifact of abandoned American Indian a g r i cultural lands
• Cane is an understory plant in other habitats
• Cane tolerates inundation but not prolonged sub m e r gence
• Disturbance dependent/ sensitive
• Burns readily, culms killed by fire but resprout quickly
• Too frequent fire favors fire-resistant trees and shrubs
• Fire exclusion also leads to woody plant dominance
Threats:
• Overgrazing/browsing (palatable and high quality livestock forage)
• Low survival of transplants used in restoration ef forts
• Land clearing for agriculture silviculture dev.
Poster Presentations 65
encroachment, ATV use, hydrologic alterations, flood control and pollution. The unique wetlands embedded in longleaf pinelands were once extensive throughout the Southeastern US, but because so many has been lost, they now
require protection, maintenance and restoration.
• Altered flood regime
• Altered fire regime
Management Recommendations:
•
•
•
•
•
•
•
Restore flood regime/hydrology
Restore natural fire regime
Remove overstory
Exclude or limit grazing/browsing
Fertilize transplants
Mulch around transplants
Transplant culms to restoration sites using hand
tools
• Excavate, transport and plant cane clumps with
roots and soil intact
• Use only native Arundinaria for restoration
Atlantic White-Cedar Wetlands
Characteristics:
• Peat accumulation is often significant
• Often has a well developed shrub layer
• Dominant or co-dominant tree is Atlantic WhiteCedar (AWC)
• Communities often include gums, bays, red maples,
cypress, oaks and pines
• Variety of wetland types distributed over AWC
range, from ME to FL and MS
• Organic acids, high cation concentration, low nutrient availability, low pH, low DO
Threats:
•
•
•
•
•
Harvest/silviculture
Urbanization
Deer browsing
Altered fire and hydrologic regime
Clearing for agriculture/silviculture/development
Management Recommendations:
•
•
•
•
•
•
•
Restore hydrology
Remove slash after timber harvest
Harvest by clearcutting in strips
Control competing hardwoods
Use prescribed fire precisely
Plant seedlings to re-establish AWC stands
Shade and use peat to bolster seedling growth
Wetlands are among the most threatened SCELPLs.
Between the 1780’s and 1980’s 53% of wetlands in the
conterminous US were lost, and 47k ha of wetlands per
year continue to be lost. These figures do not take into
account conversion from one wetland type to another,
or degradation due to introduction of exotic/invasive
species, ditching, draining, fire suppression, urban
It is not safe to assume that successful management of longleaf pine uplands will automatically benefit SCELPLs.
These examples demonstrate that SCELPLs require special
attention, especially to hydrologic and fire regimes. In
some cases planting native vegetation or removal of invasive species may be necessary to restore the composition,
structure and function of special communities. Due to special soil and disturbance sensitivities of SCELPLs, harvesting and other silviculture practices may need to be modified
to avoid negatively impacting special communities.
Management plans should take SCELPLs into consideration. A hierarchical approach will ensure that ecological
integrity is conserved at every level, the landscape, the
communities, the populations of species and the gene pool.
Because composition, structure and function at these multiple levels are intimately connected, only in this way can we
assure the long-term persistence of longleaf pinelands.
Bibliography and Suggested Reading
Bailey, M.A., J.N. Holmes, K.A. Buhlmann, and J. C.
Mitchell. 2006. Habitat management guidelines for
amphibians and reptiles of the southeastern United
States. Partners in Amphibian and Reptile Conservation
Technical Publication HMG-2.
Brantley, C.G. and S.G. Platt. 2001. Canebrake conservation in the southeastern United States.Wildlife Society
Bulletin. 29:4.
Dattilo, A.J. and C.C. Rhoades. 2005. Establishment of the
woody grass Arundinaria gigantean for riparian restoration. Restoration Ecology. 13:4.
De Steven, D. and M.M. Toner. 2004. Vegetation of upper
coastal plain depression wetlands:environmental templates and wetland dynamics within a landscape framework. Wetlands. 24:1.
Eason, G.W. and J.E. Fauth. 2001. Ecological correlates of
anuran species richness intemporary pools: a field
study in South Carolina, USA.. Israel Journal of Zoology. 47:347-365.
Ehrenfeld, J.G. 1995. Microtopography and vegetation in
Atlantic white cedar swamps: the effects of natural
disturbances. Canadian Journal of Botany. 73:474-848.
Ehrenfeld, J.G. and J.P. Schneider. 1991. Chamaecyparis
thyoides wetlands and suburbanization:effects on hydrology, water quality and plant community composition. The Journal of Applied Ecology. 28:2.
Poster Presentations 66
Frost, C.C., J. Walker and R.K. Peet. 1986. Fire dependent savannas and prairies of theSoutheast: Original
extent, preservation status and management problems. Pages 348–357 in Wilderness and Natural Areas in the Eastern United States: A Management
Challenge. (D. L. Kulhavy and R. N. Conner, Eds.).
Center for Applied Studies, School of Forestry, Steven F. Austin State University, Nocogdoches, Texas.
Gibbons, J.W., C.T. Winne, D.E. Scot, J.D. Willson, X.
Glaudas, K.M. Andrews, B.D. Todd, L.A. Fedewa, L.
Wilkinson, R.N. Tsallagos, S.J. Harper, J.L. Green,
T.D. Tuberville, B.S. Metts, M.E. Dorcas, J.P.
Nestor, C.A. Young, T. akre, R.N. Reed, K.A.
Buhlmann, J. Norman, D.A. Croshaw, C. Hagen, and
B.B. Rothermel. 2006. Remarkable amphibian biomass and abundance in an isolated wetland: implications for wetland conservation. Conservation Biology. 20:5.
Hughes, R.H. 1951. Observations of cane (Arundinaria)
flowers, seed, and seedlings in the North Carolina
coastal plain. Bulletin of the Torrey Botanical Club.
78:2.
Kirkman, L.K. 1994. Vegetation disturbance and maintenance of diversity in intermittently flooded Carolina
bays in South Carolina. Ecological Applications. 4:1.
Kirkman, L.K. 1995. Impacts of fire and hydrological
regimes on vegetation in depression
wetlands of southeastern USA.. Pages 10-20 in Susan I.
Cerulean and R. Todd Engstrom, eds. Fire in wetlands: a management perspective. Proceedings of the
Tall Timbers Fire Ecology Conference, No. 19. Tall
Timbers Research Station, Tallahassee, FL.
Kirkman, L.K. 1999. Biodiversity in southeastern, seasonally ponded, isolated wetlands:management and policy perspectives for research and conservation. Journal of the North American Benthological Society.
18:4.
Kirkman, L.K., M.B. Drew, L.T. West, and E.R. Blood.
1998. Ecotone characterization between upland longleaf pine/wiregrass stands and seasonally-ponded
isolated wetlands. Wetlands. 18:3.
Kirkman, L.K., P. C. Goebel, L. West, M.B. Drew, and
B.J. Palik. 2000. Depressional Wetland vegetation
types: a question of plant community development.
Wetlands. 20:2.
Laderman, AD. 1989. The ecology of Atlantic white cedar
wetlands: a community profile.U.S. Fish and Wildlife
Service Biological Report. 85:7.21.
McKinley, C.E. and Frank P. Day, Jr. 1979. Herbaceous
production in cut-burned, uncut-burned, and control
areas of a Chamaecyparis thyoides (L.) BSP
(Cupressaceae) stand in the Great Dismal Swamp. Bulletin of the Torrey Botanical Club. 106:1.
Mylecraine, K.A. and G.L. Zimmerman. 2000. Atlantic
white-cedar ecology and best management practices
manual. Department of Environmental Protection, Divisioin of Parks and Forestry. New Jersey Forest Service.
Noss, R.F. 1983. A regional landscape approach to maintain
diversity. Bioscience. 33:700-706.
Noss, R.F. 1990a. Can we maintain biological and ecological integrity? Conservation Biology. 4:3.
Noss, R.F. 1990b. Indicators for monitoring biodiversity: a
hierarchical approach. Conservation Biology. 4:4.
Noss, R.F. 1999. A citizen’s guide to ecosystem management. Boulder, CO: Biodiversity Legal Foundation.
Noss, R.F. 2002. Context matters: considerations for large
scale conservation. Conservation in Practice. 3:3.
Pechmann, J.H. K., D.E. Scott, J.W. Gibbons, and R.D.
Semlitsch. 1989. Influence of wetland hydroperiod on
diversity and abundance of metamorphosing juvenile
amphibians. Wetlands Ecology and Mangement. 1:1.
Platt, S.G. and C.G. Brantley. 1997. Canebrakes: an ecological and historical perspective. Castanea. 62:1.
Schurbon, J.M. and J.E. Fauth. 2003. Effects of prescribed
burning on amphibian diversity in a southeastern U.S.
national forest. Conservation Biology. 17:5.
Semlitsch, R.D. and J.R. Bodie. 1998. Are small, isolated
wetlands expendable? Conservation Biology. 12:5.
Wicker, M. and E. Hinesley. 1998. Restoring an Atlantic
white cedar bog. USFWS EndangeredSpecies Bulletin.
23:5.
Poster Presentations 67
Longleaf Pine Ecosystem Restoration Project: Lessons Learned from LPER
Shan Cammack1
1
The Georgia Department of Natural Resources Division, Wildlife Resources Division, Nongame Wildlife & Natural
Heritage Program, Forsyth, Georgia 31029, USA
Project Overview
Doerun Pitcher Plant Bog Natural Area
The main objective of LPER is to restore the off-site pine
plantations to a healthy longleaf pine ecosystem. Goals
of this project included: 1) gradually converting the canopy from planted pine to longleaf pine through thinning
and the creation of small gaps, 2) reducing hardwood
competition and exotic species with fire and chemical
treatments, 2) utilizing frequent prescribed fire to restore
native groundcover species, and 3) acquiring wiregrass
seed from private lands through private landowner incentive payments to be planted on state lands. Monitoring is
underway to determine the effectiveness of these different
management techniques.
This 650 acre natural area in Colquitt County boasts a
healthy longleaf pine/wiregrass community with an
imbedded pitcher plant bog matrix. About 100 acres are
pine plantation enrolled in the Conservation Reserve Program. No timber thinning was allowed in that contract.
Emphasis here was on prescribed burning, underplanting
of wiregrass seed and plugs, and restoration of a 20 acre
agricultural field.
Management accomplishments include impacting about
1,100 acres on state lands with timber management, burning, and hardwood control. An estimated 54,500 longleaf
pine seedlings, 70,000 wiregrass plugs, and abundant
wiregrass seed have already been planted in small or medium-sized gaps or underplanted in thinned stands. Specialized equipment, including the Woodward flail vac,
Grasslander, and Whitfield tree planter have been purchased and are being utilized across the state. Equally
important, the lessons learned are helping drive adaptive
management and will be used to influence management
on other state and private lands.
Fire was applied several years in a row. Growing season
burns in natural stands have increased species diversity
and to helped develop our own wiregrass seed source.
The fire brought bog species back to areas that did not
even appear bog-like. Pitcher plants have been observed
sprouting in a drain where privet has recently been mechanically removed.
With monitoring help from the Jones Center, wiregrass
seeds and plugs were underplanted in the CRP pines.
Methodology included no pre-treatment, mowing, and
disking, as well as seed versus mechanically planted
plugs. In hopes to make groundcover restoration affordable to private landowners, techniques were developed to
use a tree planter to plant plugs. Our manual planter was
able to plant about 2,500 plugs per acre.
Growing season burns have helped reduce hardwood competition and are crucial in preparing
areas for future wiregrass harvest.
Different growers produced very different plugs. Some were predominantly wiregrass. Seed
must be carefully cleaned to achieve this. Unclean seed produces more diversity of species but
also more competition for the wiregrass.
While applying Chopper to Bermuda grass reduced its extent and vigor, it did not control it
enough to reseed the field with wiregrass
Harvesting on private lands produces high quality wiregrass seed that can be used to promote
ground-cover on depauperate state lands. We’ve learned lessons on harvesting (decrease flail
speed and increase ground speed), storing (can store up to a year in humidity-controlled room
without loss of viability), and growing plugs (cleaning seed is the key).
An 8 month old plug mechanically planted. The small mound made from the planter falls out it
time and does not appear to be a problem.
Poster Presentations 68
The field restoration proved to be the most challenging.
Bermuda grass has slowly encroached since retirement of
the field from peanuts. Wiregrass plots that were installed
inside the field were completely taken over by the aggressive exotic species. Three years of heavy hitting with
herbicide (Chopper) proved only mildly successful. To
prevent loss of precious wiregrass seed, we ended up
planting the whole field with a mixture of slash and longleaf pine in hopes of shading out the bermuda grass. The
slash will be thinned out in time, leaving a longleaf forest.
More planting of longleaf will result in an uneven-aged
stand. Wiregrass planting will be held off until the bermuda grass is “under control.”
Part of this project has focused on the economics of using
seed versus plugs. Plugs are a lot more expensive, with
growing costs of at least 10 cents and hand-planting costs
of at least 10 cents. Finding management techniques that
are affordable to private landowners is key to the success
of this project. Using a tree planter has proven to be an
important part of finding cost-effective techniques.
Moody Forest Natural Area
This 4,300 acre Natural Area in Appling County lies
along the Altamaha River. Habitats range from bottomland floodplain communities, longleaf pine/wiregrass,
pine-oak woodlands, loblolly flats, and pine plantation.
Redcockcaded woodpeckers inhabit the mature longleaf
pine. Due to weather, timber harvesting in the plantations
was delayed. Focus has been on carefully bringing fire
back to this fire-suppressed site. The prescribed burning
has been successful, with almost 1,300 acres burned last
year. Careful prescriptions in the mature stands with duff
issues and diligent work in the turpentined areas have
proven effective. We have found that if you let the turpentine trees burn for several minutes then extinguish
them with a backpack pump, the wounds tend to heal
over. Future burning does not ignite the trees as violently
or at all. Little mortality has been documented in areas
using this technique. Little mortality has been observed in
the old growth areas where fire is carefully applied after
rains with high soil moisture and high winds.
Big Dukes Pond Natural Area
This 1,692 acre Natural Area in Jenkins County Big Dukes
Pond hosts a large Carolina bay which supports significant
natural communities and rare species populations. In the
uplands, there is a 155 acre loblolly pine plantation interlaced with isolated wetlands. Restoration began in 2003 to
slowly convert this to a longleaf pine ecosystem.
Timber thinning occurred between April 2003 and February
2004. The whole stand was thinned to 60 basal area using
fourth row operator select and 23 quarter acre gaps were
randomly placed in the stand. An important lesson learned
is that onsite supervision of the harvesting crew is important
in proper gap creation. In August 2004, the entire stand
was sprayed with Chopper to kill competing hardwood
vegetation. Prescribed burning was conducted on this stand
in December 2004. Fire effects were moderate. Fuels in
several of the gaps and logging decks were sparse, resulting
in a patchy burn. Observations on the herbicide treatment
looked favorable. The gaps were planted with longleaf pine
seedlings in January 2005 at a density of 400 per acre or
100 per gap.
The goal of designing the monitoring was to create a protocol that would capture a snapshot of long leaf survival,
hardwood control, and the plant community in a way that
could be performed in one day of sampling. Data was collected in a random subset of gaps in August 2005 and 2006.
This included longleaf survival (within two meters of the
center transect), hardwood growth (hardwood cover and
species were quantified using the line-intercept method),
and species richness (in two circular one meter plots, all
plants below one meter that were rooted within or hanging
in each plot were identified to species).
Longleaf survival looks good on the ground, with the trees
looking robust and some even beginning to bolt. Data,
however, show a survival rate of only 66%. Hardwood
control appears to be successful. Data show that hardwood
cover did not increase overall, and only infrequent seedlings
Foraging habitat has been improved for red-cockaded woodpeckers. Artificial cavities have been
installed in areas previously occupied by the birds.
Fire has been key in the restoration of natural habitats. A one to two-year return interval appears to be optimal.
Although it’s a hot job, tending to each turpentine tree during the first couple of burns is proving
to be successful in getting fire back into these stands.
Poster Presentations 69
of hardwood trees were found on the sampling transects.
Seedlings observed this year were winged elm and persimmon. The occasional survivors of the Chopper application
were water oak and darlington oak.
From 2005 to 2006, species richness tended to increase in
our plots, with an average percent increase of 55% (n=16).
Only one species is known to be exotic (Florida pusley),
and no invasive exotics were found. Most species are early
successional, such as dogfennel, fireweed, and 3-seeded
mercury. In 2006 there were more grasses, more Rubus
species, and less fireweed. The higher cover of Rubus could
interfere with prescribed fire, and emphasizes the importance of implementing regular prescribed fire in these gaps.
Some common legumes have become established, such as
creeping lespedeza and ticktrefoil.
Mayhaw Wildlife Management Area
This 4,700 acre Wildlife Management Area in Miller
County hosts a mosaic of natural pine, pine plantation, and
Species Found in Gaps*
Acalypha gracilens
Ambrosia artemesiifolia
Andropogon sp.
Aralia spinosa
Bignonia capreolata
Campsis radicans
Chamaecrista fasiculata
Cirsium sp.
Conyza canadensis
Croton glandulosus var. septentrionalis
Dichanthelium aciculare
Dichanthelium acuminatum
Dichanthelium laxiflorum
Dichanthelium oligosanthes
Dichanthelium scoparium
Erechtites hieraciifolia
Eupatorium capillifolium
Helianthemum rosmarinifolium
Hypericum gentianoides
Hypericum hypericoides
Juncus marginatus
Lechea mucronata
Lespedeza repens
Paspalum sp.
Phytolacca americana
Pinus palustris
Pinus taeda
Pseudognaphalium obtusifolium
Richardia sp.
Rubus sp.
Vitis rotundifolia
2
3
4
4
3
3
2
3
2
4
3
3
3
4
3
1
1
4
3
3
3
4
4
2
2
5
3
2
1
2
3
*Numbers indicate *quality* on a scale of 1 to 5, 5 is
best
flatwoods. LPER work was carried out on two tracts,
Oldhouse and Firetower. Timber thinning and prescribed burning have occurred over the past four years.
A monitoring scheme was set up in two stands with
longleaf gaps planted with different wiregrass treatments. The maps below show the distribution of the
gaps and the treatments ranging from no wiregrass,
seed, and densities of 1,000, 3,000, and 5,000 per acre.
This study will help managers determine
the most cost-effective density in which to plant wiregrass.
Data collected inside a three meter radius of the center
of the gap includes: number of wiregrass plugs, hardwood count, and number of live longleaf. Also recorded were height of five closest longleaf pine, percent
canopy cover, and density of understory (using a density board). All species were recorded in three randomly placed one meter circular plots in each of the
gaps. Two consecutive years of data have been collected, but analysis has not yet been conducted. The
vegetation community tends to be diverse and seems to
be correlated with hydrology. Gaps near roads seem to
have a higher number of early successional species.
Differences in wiregrass plugs from the two growers
were remarkable. Some plugs were large and robust
and brought a diversity of species to the gaps, particularly later successional composites, such as grass-leafed
golden aster and blazing star. Other plugs were 100%
wiregrass. Using a mixture of each would give you the
wiregrass fuels needed for fire along with the diversity.
It remains to be seen whether competition in the plugs
will be detrimental to the wiregrass.
Acknowledgements
Funding for the project was provided by the National
Fish and Wildlife Foundation and Southern Company’s
Longleaf Legacy. We also thank our Partners: Joseph
W. Jones Ecological Research Center and The Nature
Conservancy.
Americorps volunteers hand plant wiregrass to specified densities for the monitoring program. We found it much easier to
work with volunteers than with contracted crews when precise
and varied planting densities were required.
Years of prescribed burning have helped keep the groundcover in some of these stands healthy. Growing season fires
will be used to further enhance the diversity.
This stand has incredible groundcover. No wiregrass planting
was necessary here. We are experimenting with underplanting longleaf in areas where the canopy is more sparse.
Poster Presentations 70
Longleaf Pine Seedling Survivorship and Growth on Poorly Drained Soils
Susan Cohen1 and Joan Walker2
1
USDA Forest Service, Southern Research Station, Research Triangle Park, North Carolina, 27709, USA
2
Department of Forest Resources, Clemson University, South Carolina, 29634, USA
Abstract
Artificial regeneration is required to restore longleaf
pine where the natural seed source has been lost and site
preparation is often employed to establish seedlings.
This study evaluates the site preparation methods used
in restoring longleaf pine to poorly drained sites. The
study is a randomized block design with eight silvicultural treatments applied to ameliorate conditions commonly thought to impede longleaf survival and growth,
such as poor drainage and vegetative competition. The
treatments included an herbicide application or a singlepass chop prior to burning, followed by flat planting,
mounding and planting or bedding and planting. Site
preparation treatments did not have a significant effect on
seedling survivorship after two growing seasons. While
causes of mortality were not determined, field observations
suggest incorrect planting depth and planting into piled organic matter contributed heavily. Height and root collar
diameter, however, were significantly higher on the bedding, mounding and herbicide treatments. Bedding in combination with the chop plus herbicide application had the
greatest growth among all treatments. When combined
with herbicides, seedling growth in the bedded and
mounded treatments out performed seedlings the chop-bed
and the chop-mound treatments, respectively. Seedling
biomass, when compared between flat planting and bedding, was greater on bedded sites. In general, the more intensive treatments had greater seedling growth responses
and an earlier emergence from the grass stage.
Poster Presentations 71
Restoring and Managing Longleaf Pine Ecosystems in the Southern United States: Southern
Research Station Research Work Unit 4158 – Auburn, AL: Clemson, SC; Pineville, LA
K.F. Connor1, D.G. Brockway1, J.D. Haywood2, J.C.G. Goelz2, M.A. Sword-Sayer2, S-J.S. Sung2, and J.L. Walker3
1
U.S. Forest Service, Southern Research Station, Auburn, AL 36849
U.S. Forest Service, Southern Research Station, Pineville, LA 71360
3
U.S. Forest Service, Department of Forest Resources, Clemson, SC 29634
2
Background
Longleaf pine (Pinus palustris P. Mill.) forest ecosystems
once encompassed 37 million hectares in the southeastern
United States. These vast forests extended from Virginia
southward to central Florida and westward to eastern
Texas. Today, longleaf pine has all but disappeared as a
dominant species. Longleaf forests occupy fewer than 1
million fragmented hectares, less than 3% of its original
range, and are one of the most threatened ecosystems in
the U.S.
From 1870 to the early 1930s, the southern forest was cut.
Some inroads had been made prior to this but the advent
of railroads through the South opened the area to extensive logging (Jose et al., 2006). This was followed by
regeneration failure of the longleaf pine overstory. Seeds
of longleaf pine are irregularly produced and have many
predators. In addition, naturally occurring longleaf pine
seedlings may stay in the grass stage for several years.
When the longleaf forests were harvested, many of the
seedlings that did bolt from the grass stage were a favorite food source of the exploding feral hog population
(Jose et al., 2006). Wild hogs were at one time so prevalent they supported a meat-packing industry in the South.
When fire suppression became common practice, longleaf
pine stand regeneration was doomed.
By this time, longleaf pine was considered by many landowners to be inferior; it was perceived as difficult to regenerate, slow growing, unproductive, and chancy. When
an economic return on timber investment was the desired
end product, why plant longleaf pine when loblolly pine
(Pinus taeda L.), fast growing and easy to regenerate, was
there to fill the gap?
Current Status
The existing southern landscape is under ever-increasing
pressures, with social values, economic demands, and
natural events resulting in forest fragmentation, urban
sprawl, insect and disease outbreaks, and proliferation of
invasive species. In recent years, catastrophic hurricanes
have inflicted billions of dollars in damage on southern
states; and wildfire represents a constant threat to southern forest resources. Forest managers in the South often
face conflicting objectives. Once asked how to best
prepare a site, which genetically superior lines to plant,
when to thin, when to cut, and how to maximize forest
value for landowners, they are now asked how to best
defend against invasive species, what can be done to prevent major insect infestations, and when should forests be
cut and wood salvaged? How can timber resources be
protected from damage or loss when major disturbances,
such as wildfires and hurricanes, occur? In some cases,
the only answer might be that, if the forest is planted with
species vulnerable to insect attack or high winds, damage
will occur.
The Case for Longleaf Pine
When deciding what to replant on forested areas, landowners must now consider many risk factors that often
favor longleaf pine and its associated communities.
Longleaf pine evolved in the southern hurricane zone; its
seedlings are uniquely adapted to take advantage of gaps
created in the overstory. Longleaf pine trees stand up
well to hurricane-force winds, and longleaf pine ecosystems are fire-adapted. Longleaf pine is now recognized as
a species with great natural resilience.
Longleaf pine ecosystems are among the most diverse in
the continental U.S., often with 40 or more species of
higher plants per square meter (Walker and Peet 1984).
These ecosystems provide excellent habitat for many
game species and are home to numerous threatened and
endangered species of animals, including red cockaded
woodpeckers, pine snakes, and gopher tortoises. An important attribute of longleaf pine ecosystems is their
unique ability to resist and recover from what for other
southern pines would be catastrophic events. Adapted to
fire, responsive to gaps in the overstory, able to bolt from
the grass stage – their unique physiology prepares them
for rapid recovery after hurricanes and lightning-caused
fires. It was only exploitation and neglect that reduced
longleaf pine ecosystems from a dominant to marginal
existence.
Now, only our intervention can bring these valuable ecosystems back from obscurity. In the 1930s, it was thought
that excluding fire from longleaf pine stands would enable
them to recover. Now it is recognized that it is fire that
Poster Presentations 72
species and promoting longleaf pine growth and dominance
of the landscape. In the 1930s, longleaf pine was considered too slow to grow, too uneconomical to manage and not
productive. So loblolly and slash pine plantations proliferated throughout the South. Now it is known that these species, especially loblolly pine, are more susceptible to wind
damage and breakage in hurricane-force winds and are
highly vulnerable to insects, disease, and fire.
The resilience of longleaf pine ecosystems is what makes
them so attractive to landowners facing recovery from devastating timber losses after Hurricanes Ivan, Katrina, and
Rita. In addition to timber products, the landowner can
harvest intermediate products, such as pine straw, and draw
additional benefits from the wildlife that inhabit these
unique forest ecosystems. Further, not only does longleaf
pine have a high resistance to the southern pine bark beetle
but growth and yield models also show that productivity of
planted longleaf pine eventually catches and surpasses that
of loblolly pine. Lastly, through intensive research programs, we have overcome many of the difficulties surrounding regeneration of longleaf pine.
In response to the rapidly growing demand for information
about longleaf pine ecosystems, the U.S. Forest Service,
Southern Research Station, has established a new research
work unit, Restoring and Managing Longleaf Pine Ecosystems (SRS-4158). The unit has seven scientists with expertise in plant physiology, ecology, silviculture, and biometrics. Two experimental forests, the Palustris and the Escambia, provide a land base for practical experiments in and
demonstrations of longleaf pine establishment, development, and management. Headquartered in Auburn, AL,
scientists in the unit are also stationed at Clemson, SC and
Pineville, LA, providing broader customer access and research opportunities in a variety of longleaf pine ecosystems. For more information, contact Kristina Connor, Project Leader SRS-4158. Phone (334) 826-8700; fax (334)
821-0037; email [email protected]
References
Jose S., Jokela E.J., Miller D.L. (eds). The Longleaf Pine
Ecosystem. Ecology, Silviculture, and Restoration.
New York: Springer Science+Business Media, LLC.
438 p.
Walker J., Peet R.K. 1984. Composition and species diversity of pine-wiregrass savannas of the Green Swamp,
North Carolina. Vegetatio 55: 163-179.
Poster Presentations 73
South Carolina Lowcountry Forest Conservation Project
W. Conner1, T. Williams1, G. Kessler1, R. Franklin1, P. Layton1, G. Wang1, T. Straka1
B. Humphries , C. LeShack2, K. McIntyre3, R. Mitchell3, S. Jack3 W. Haynie4, A. Nygaard4, L. Hay4 D. Beach5, J.
Lareau 5, J. Johnson6, and M. Robertson7, M. Prevost7, M. Nespeca7
2
1
Clemson University, Clemson, South Carolina, 29634, USA
2
Ducks Unlimited, Memphis, Tennessee, 38120, USA
3
Joseph W. Jones Ecological Research Center, , Newton, Georgia, 39870, USA
4
Lowcountry Open Land Trust, Charleston, South Carolina 29403, USA
5
South Carolina Coastal Conservation League, Charleston, South Carolina, 29402, USA
6
The Conservation Fund, South Carolina Office, Columbia, South Carolina, 29201, USA
7
The Nature Conservancy, Columbia, South Carolina, 29250, USA
Introduction
The Lowcountry Forest Conservation Partnership has a
goal of protecting South Carolina Coastal Plain forest
ecosystems and their levels of biological diversity, rare
and special plants and animals. These forests represent
important ecosystems found in most Southern states. The
area’s forests are globally significant because they contain remnant longleaf pine forests that support one of the
richest plant communities known in the world. There are
also undisturbed wetland forests and isolated wetlands
such as Carolina Bays which are essential habitat for
many birds and water-loving animals. These systems exist because of a historic interaction of two ecological
processes, fire and water. The project area is at risk from
a combination of threats such as, urban development,
incompatible forestry practices, the loss of fire, alteration
of stream flow on major rivers and climate change. This
landscape can only be sustained through conservation and
restoration. The Lowcountry Forest Conservation Partnership was made possible as the result of a generous grant
from the Doris Duke Charitable Foundation.
•
•
•
•
Restoring and restructuring degraded bottomland
hardwood forests into productive forests; Using uneven-aged forest management in bottomland hardwoods stand; Using fire and other methods to preserve and develop herbaceous vegetation communities in longleaf pine stands; Use of fire as a tool in the
ecotone between pine uplands and bottomland forests; Change even-age loblolly and longleaf stands to
uneven-age stands; Transition loblolly stands to longleaf stands.
287 Natural Resource Professionals who manage
more than 4,238,800 acres have been trained in Conservation Forestry Practices in twelve programs.
363 landowners who own more than 226,970 acres
have attended thirteen workshops on Conservation
Forestry.
79 landowners have developed and are implementing
conservation forestry plans on 134,300 acres.
Conservation Forestry cost-share program in partnership with AFF & USF&WS.
Accomplishments
•
•
•
100,000 acres of land protected
from development through purchase and conservation easements
which allow the land to continue
to be managed for ecologicallysustainable forestry and wildlife
management.
The South Carolina Prescribed
Fire Council has been established
to promote and protect the use of
prescribed fire in the state. Three
annual meetings have been held.
Seven demonstration areas with a
total of 9,000 acres have been
established to show conservation
forestry concepts which include:
Poster Presentations 74
An Investigation of Old-field Longleaf Growth, Yield, Diameter Distributions, Product Class
Distributions, Pine Straw Production, and Economics of Management Intensities in Georgia
E. David Dickens1, Bryan C. McElvany1 and David J. Moorhead1
1
Daniel B. Warnell School of Forestry and Natural Resources, The University of Georgia Tifton, Georgia, 31793, USA
Introduction
Longleaf pine (Pinus palustris, Mill.) once occupied an
estimated 60 to 90 million acres in the SE US (Croker
1990, Engstrom et al. 2001). Today, longleaf pine covers
an estimated 3.5 million acres in the SE US (Kelley and
Bechtold 1990). Much of these acres are on military bases
and State or National Forests. Over 116,000 acres of
longleaf were planted on former row crop fields (oldfields), pastures, and hay fields in the late 1990's Conservation Reserve Program (CRP) in Georgia. Alabama,
Florida, North and South Carolina have also had large
acreage plantings of longleaf during this period.
Little is known of the growth rate and wood yields of
longleaf pine on these sites with typically good fertility,
no woody competition, and good soil tilth. Work done in
Georgia on CRP old-field planted slash and loblolly
(Dangerfield and Moorhead 1998) found that these two
species grew at a much faster rate and yielded more wood
than many researchers had anticipated using cut-over SI
curves. Goelz and Leduc (2001 a;b) studied longleaf in
Texas and Louisiana (Gulf Coastal Plain) on land that
was formerly in pasture for a relatively short period of
time. They found culmination of mean annual increment
(MAI) ranging from 0.70 (@ age 60-years SI<50' base
age 25-years), 1.6 (@ age 35-years, SI=50-60'), and 2.0
(@ age 30-years, SI>60') cords/ac/yr. Mean annual increment (MAI) through age 15-years from Goelz and Leduc
study ranges from 0.30 to 1.75 cords/ac/yr. Modeled oldfield longleaf MAI using WINYIELD v. 1.1 at age 15years produces less than one cord/ac/yr for SI levels of
63, 68, and 71 (the highest value WINYIELD will allow
for longleaf) feet.
Plots installed in an old-field longleaf stand in Screven
County, Georgia in 2001 (Dickens and Moorhead, unpublished data) had an MAI of 2.0 to 2.36 cords/ac/yr (1/10th
acre plots) where TPA was 370 to 560. The soils, Bonneau (loamy sand with argillic @ 34-39") and Blanton
(loamy sand with argillic @ 41-50") on this site are considered to have average to low native (no inputs) productivity. This old-field longleaf site in Georgia through age
15-years has from 20 to 160 TPA in the chip-n-saw size
class. The stand has been raked three times. In May 2002
the longleaf stand produced 290 to 360 bales/acre where
basal area ranged from 130 to 150 square feet/acre
(without fertilization). Pine straw collected in December
2004 ranged from 280 to
295 bales per acre. Using current per bale prices for raked
straw this would be $144 to $180/acre at $0.50/bale or
$290 to $360/acre at $1.00/bale for the May 2002 rake
and $140 to $147.50/acre at $0.50/bale or $280 to $295/
acre at $1.00/bale for the December 2004 rake.
There is potentially a great incentive (economic, aesthetic,
and environmental) for NIPF landowners to grow longleaf
on old-field, pasture, and hayfield sites. Early in the
stand’s life (age 10 to 25-years) income can be generated
annually or periodically from pine straw raking (Dickens
2000, Dickens 2001).
Wiregrass (Aristrida stricta,
Michaux) may be established in stand gaps during this
period. Once a stand is thinned, pine straw raking can
continue (Dickens 2001) with some clean-up or management can shift towards establishing wiregrass and growing quality timber. Economics of growing longleaf is
attractive using conservative WINYELD v. 1.1 growth
rates (Dangerfield et. al. 2002). Dangerfield et al. (2002)
found rates of return ranging from 12 to 15.4 percent
under various levels of management and rotation ages for
longleaf plantations.
We know little of the upper end of longleaf’s growth rate,
wood and pine straw yields, and profitability on these oldfield sites. We have installed two study areas in planted
(December 1986) old-field longleaf stands in Georgia
(Screven and Tift Counties). Soil series have been delineated by an NRCS soil mapper as Bonneau and Blanton at
the Screven County site and Albany and Leefield at the
Tift County site. Gross treated (1/4 acre) plots were installed with a 1/10th acre internal permanent measurement
plots (IPMP). There are 40 feet of untreated buffer between each plot. Each living tree in the IPMP was aluminum tree tagged, numbered, and measured for dbh, total
height, height to base of live crown, and fork or broken
top, and tree form, defects noted (Dec 2003 for the Screven County site and Feb 2004 for the Tift County site).
There are 3 (Tift county) or 4 (Screven County site) replications of each treatment per study area. Treatments include: control (no fertilization), full dose of NPK
(DAP+urea+muriate of potash; 150 N, 50 elemental-P and
50 elemental-K lbs/ac), and a half dose of NPK
(DAP+urea+muriate of potash) applied in mid-February
2004. The second 2 dose of the half dose treatment will
be applied in February 2007. Herbicides (glyphosate)
were used one-time (mid-summer 2004) in all study area
plots to keep the stand clean for straw production.
Poster Presentations 75
Baseline soil (0-6") and foliage samples have been collected
(December 2003 at the Screven county site and February
2004 at the Tift county site). Post fertilizer application foliage and soil samples have been collected every winter.
Leaf area has been estimated every summer along with digital photos of crowns from each plot for eventual longleaf
fertilization recommendations.
Planted longleaf volume equations from Baldwin and Saucier (1983) have been used to estimate volume/acre and
product class distribution on these old-field stands until we
get a volume equation that is more appropriate for old-field
longleaf stands. The economic profitability of growing
longleaf on old-field sites will be estimated from this study.
Baseline Growth, Yield, Pine Straw Production, Diameter Distributions and Volume
Diameter Classes are presented in tables 1, 2, and 3 and
figures 1 through 4. Mean trees per acre (TPA) were similar
at each site by the beginning of the 18th growing season
ranging from 303 to 347 (Tables 1 and 2). Diameters (@
4.5 ft above groundline) ranged from 7.96 to 8.26 inches.
Basal area ranged from 113 to 128 ft2/acre. Total heights
ranged from 49.3 to52.4 feet. Mean live crown ratios
ranged from 43 to 50 percent. Total volume ranged from
2548 to 2945 ft3/acre. Pulpwood volume (trees with a 5 to 9
inch dbh) ranged from 997 to1446 ft3/acre. Chip-n-saw
volume (trees with a dbh $9 inches) ranged from 979
to1341 ft3/acre (Tables 1 and 2). Mean pine straw production estimates prior to fertilizer treatments ranged from 143
to 175 bales/acre (Table 3). Diameter distributions and
volume by diameter class for the Screven and Tift County
sites are found in Figures 1 through 4. Two-year post application data will be presented as a poster at the Longleaf
Alliance Regional Annual meeting in Tifton, Georgia on
13-16 November 2006.
Figure 1. Old-field (planted December 1986) longleaf
diameter distributions Screven County site prior to at
the fertilizer treatments.
Figure 2. Old-field (planted December 1986) longleaf
volume by diameter class at the Screven County site
prior to fertilizer treatments (at age 18-years-old).
Figure 3. Old-field (planted December 1986) longleaf
diameter distributions at the Tift County site prior to
fertilizer treatments.
Table 1. Old-field planted (December 1986) longleaf pre-treatment growth and yield parameters
at the Screven County, Georgia site at the end of the 18th growing season.
Treatment
TP
A
dbh
(in)
BA/ac
(ft2)
total
height
(ft)
live
crown
ratio (%)
Total
volume
(ft3)
pulpwood
volume
(ft3)
chipnsaw
(ft3)
Control
325
8.37
121
50.4
43
2713
997
1341
2 NPK
303
8.47
114
50.1
45
2548
1035
1193
Full
NPK
328
8.32
120
49.3
44
2625
1341
986
Poster Presentations 76
Table 2. Old-field planted (December 1986) longleaf pre-treatment growth and yield parameters at
the Tift County, Georgia site at the end of the 18th growing season.
Treatment
TPA
dbh
(in)
BA/
ac
(ft2)
total
height
(ft)
live
crow
n
ratio
(%)
Total
volume
(ft3)
pulpwood
volume
(ft3)
chip-nsaw
(ft3)
Control
320
7.97
114
52.1
48
2641
1110
1193
2 NPK
347
7.96
113
51.6
50
2556
1280
979
Full NPK
317
8.26
128
52.4
49
2945
1466
1156
Literature Cited
Figure 4. Old-field (planted December 1986) longleaf
volume by diameter class at the Tift County site prior to
fertilizer treatments (at age 18-years-old).
Table 3.Pre-(January 2004) and post-treatment old-field
planted (December 1986) longleaf pine straw production estimates at the Screven and Tift County, Georgia
sites.
-------------Site (mo/yr) ---------------Treatment
Screven (1/04) , (12/04),
(1/06)
Tift (1/04),
(1/19/06)
Control
148, 282, 277
175, 328
2 NPK
160, 294, 355
166, 402
Full NPK
143, 289, 367
163, 367
Baldwin, V.C. and J.R. Saucier. 1983. Aboveground weight
and volume of unthinned, planted longleaf on W. Gulf
forest sites. USDA Forest Service Res. Paper SO-191.
New Orleans, LA Southern Exp. Stn. 25 p.
Croker, T.C. 1990. Longleaf pine myths and facts. In:
Farrar, R.M. ed. Proceedings of the symposium on the
management of longleaf pine; Long Beach MS. USDA
Forest Service GTR SO-75. New Orleans, LA Southern
Forest Exp. Stn: 2-10.
Dangerfield, C.W., Jr. and D.J. Moorhead. 1998. Intensive
forest management: Shifting row crop and pasture to
tree crops in Georgia - woodflow and financial returns
for old-field timber crops examined.
www.bugwood.org. 7 p.
Dangerfield, C.W., Jr., E.D. Dickens, and D.J. Moorhead.
2002. Changing cords to thousand board feet and
higher financial returns with management and time for
old-field longleaf pine timber crops. In: In: S. Graddo
ed. Proceedings of the 32nd Annual So. Forest Econ.
Workshop. Virginia Beach, VA. pp. 78-83.
Dickens, E.D. 2000. Effect of inorganic and organic fertilization in longleaf pine stands on deep sands on pine
straw and wood volume production. In: Proceedings of
the 10th Biennial So. Silvi. Res. Conf., Shreveport, LA.
Dickens, E.D, C.W. Dangerfield, Jr., and D.J. Moorhead.
2001. Short-rotation management options for slash and
loblolly pine in southeast Georgia. In: Zhang, D. and
S.R. Mehmood eds. Proceedings of the 31st Annual So.
Forest Econ. Workshop. Atlanta, GA. pp. 61-65.
Dickens, E.D. 2001. Fertilization options for longleaf pine
stands on marginal soils with economic implications.
In: Zhang, D. and S.R. Mehmood eds. Proceedings of
the 31st Annual So. Forest Econ. Workshop. Atlanta,
GA. pp. 67-72.
Poster Presentations 77
Engstrom, R.T; L.K. Kirkman, and R.J. Mitchell. 2001.
Natural history: Longleaf pine-wiregrass ecosystem.
Georgia Wildlife. Vol. 8, No. 2.Decatur, GA. pp. 5-18.
Goelz, J.C.G. and D.J. Leduc. 2000a. A model describing
growth and development of longleaf pine plantations:
consequences of observed stand structures on structure
of the model. In: Proceedings of the 11th Biennial So.
Silvi. Res. Conf. Shreveport, LA.
Goelz, J.C.G. and D.J. Leduc. 2000b. Long-tern studies on
development of longleaf pine plantations. In: Kush,
J.C. complier. Forest for your future- Restoration and
management of longleaf pine ecosystems. Proceedings
of the 3rd Longleaf Alliance Regional Conf. Alexandria, LA. Longleaf Alliance Report No. 5
Kelley, J.F; W.A. Bethtold. 1990. The longleaf pine resource. In: Farrar, R.M. ed. Proceedings of the symposium on the management of longleaf pine; Long Beach
MS. GTR SO-75. New Orleans, LA: USDA FS, Southern Forest Exp. Stn: 11-22.
Poster Presentations 78
Old Resinous Turpentine Stumps as an Indicator of the Range of Longleaf Pine
in Southeastern Virginia
Thomas L. Eberhardt1, Philip M. Sheridan2, Jolie M. Mahfouz1, and Chi-Leung So1,3
1
Southern Research Station, USDA Forest Service, Pineville, Louisiana;71360, USA
2
Meadowview Biological Research Station, Woodford, Virginia, 22580, USA;
3
School of Renewable Resources, LSU Ag Center, Baton Rouge, Louisiana, 70803, USA
Abstract
Wood anatomy cannot be used to differentiate between
the southern yellow pine species. Wood samples collected from old resinous turpentine stumps in coastal
Virginia were subjected to chemical and spectroscopic
analyses in an effort to determine if they could be identified as longleaf pine. The age and resinous nature of
the samples were manifested in high specific gravities,
the presence of oxidized monoterpenes, and the ability
to be grouped separately from wood from recently harvested trees by NIR spectroscopy. Since there are no
standards for old resinous pine stumps, studies are continuing to determine changes that occur in longleaf pine
stumps aged under field conditions.
Introduction
Longleaf pine (Pinus palustris Mill.) is the third most
abundant pine species in the southeastern United States
(Koch 1972). Straight growth, coupled with wood that
is strong and hard, made this pine species highly desirable for poles, construction lumber, and flooring. Longleaf pine also has a well established history in naval
stores production, from early turpentining operations to
the subsequent processing of residual stumps, especially
those from trees harvested in the late 19th and early 20th
centuries (Gardner 1989). The range for longleaf pine
spans from southeastern Virginia to eastern Texas
(Koch 1972). In Virginia, harvesting practices and
changes in land use since colonial settlement has dramatically reduced the presence of longleaf pine. Of the
original 1.5 million acres of longleaf forest estimated to
exist prior to colonial settlement, only 800 acres remain
(Sheridan et al. 1999). Longleaf pine restoration efforts
have initiated studies to verify its range by determining
the species of very old turpentine stumps. Our efforts
were directed towards determining if chemical and
physical characterizations of wood taken from selected
stumps could provide information indicating the likely
pine species.
Materials and Methods
Highly weathered wood specimens were collected from
stumps located in Caroline, Prince George, Southampton, and Sussex counties in Virginia. Wood shavings
were analyzed by near infrared (NIR) spectroscopy with
multivariate analysis, as described in Eberhardt and So
(2005). Samples of longleaf and loblolly pine, obtained
from recently harvested trees, were added to the NIR analysis to provide a reference point for these unknown stump
samples. For the GC-MS analyses, wood shavings (1 g)
from the specimens were steeped in methylene chloride (5
ml). GC-MS analyses of the resultant extracts were carried
out on a Hewlett Packard 6890 gas chromatograph equipped
with a Hewlett Packard 5973 mass selective detector and a
HP-INNOWax column (0.25 mm ID × 60 m length × 25 µm
film thickness). The temperature regimen of the column
was programmed to hold for 1 min at 40 ºC, increase to 80
ºC at a rate of 16 ºC min-1, and then to 240 ºC at a rate of 7
ºC min-1, with the final temperature being held for 10 minutes. The temperatures for the injector inlet and mass detector were maintained at 200 ºC and 225 ºC, respectively.
Peaks were identified by spectral match with NIST 98
(NIST, Gaithersburg, MD) and in-house chemical libraries.
Small wood blocks (ca. 1 cm3) were also cut from the samples using a band saw. Specific gravity measurements were
determined on weighed wood blocks by mercury displacement and also by careful measurement of block dimensions
with a caliper. Extractives contents were determined by
extracting wood blocks with methylene chloride for 3 days
in a Soxhlet apparatus.
Results and Discussion
Taking into consideration signs of stump scarification and/
or the occurrence of longleaf pine at the site (Southampton
specimen, only), along with the reported ranges for each of
the southern yellow pines, it appeared likely that the Southampton and Sussex county specimens were from longleaf
pine trees. On the other hand, the Caroline (Scholl) specimen had a greater probability of being loblolly pine (Pinus
taeda L.) because the collection site was outside the known
range of longleaf pine and in a mixed hardwood/loblolly
pine stand. Since wood structure cannot be used to differentiate between the southern yellow pines (Panshin and de
Zeeuw 1980), our objective was to assess whether reported
chemical and physical differences could be used for species
identification.
Principal component analysis (PCA) was applied to the NIR
spectra to observe any clustering and/or differences
Poster Presentations 79
Figure 1. Principal component analysis results from stump wood samples and longleaf and loblolly pine wood samples from recently harvested trees.
between the wood samples from the stumps and recently
harvested trees. PCA was hindered by a lack of control
samples, nevertheless, it was plausible that data gathered
might be either indicative of longleaf pine or allow the
elimination of other pine species. Several discrete groupings can be observed in the analysis of the PCA scores
(Figure 1). The highly weathered stump samples clearly
separate out from the recently harvested longleaf and loblolly pine samples. The samples from recently harvested
trees further separate into loblolly and longleaf pines. As
one would expect, the stump samples are closer to the longleaf heartwood sample than to the sapwood sample. Tentative groupings can be formed for the Sussex and Southampton samples. The Caroline (Scholl) sample appears closer
to the recently harvested trees than the stumps, and if all the
stump samples are assumed to be longleaf pine, the Caroline (Scholl) sample could possibly be another species such
as loblolly pine.
Given their fragrant nature, the stump wood samples were
also subjected to analysis by GC-MS to determine if significant amounts of monoterpenes remained despite many years
of weathering. The ability to obtain seemingly representative monoterpene compositions suggested an opportunity to
develop a chemotaxonomic approach to determine the
stump taxa. Reported analyses of fresh oleoresin from most
southern yellow pines (e.g., P. palustris, P. taeda, P. echinata, P. elliottii) have shown α-pinene to comprise 50-80%
of the monoterpenes detected (Hodges et al. 1979, Strom et
al. 2002). The second most abundant monoterpene, βpinene, ranges from 20-40%. Along with the pinenes, much
smaller amounts of camphene, myrcene, and limonene are
also typically reported. Pond pine (Pinus serotina Michx.)
is the exception among the southern yellow pines with
limonene comprising as much as 90% of the detected
monoterpenes (Mirov 1961). We hypothesized that comparisons of the monoterpene compositions with those
from other stumps, in conjunction with available data for
the oleoresin from recently harvested trees, might allow
the stump species identification.
We found α-pinene to be the most abundant monoterpene
in 4 of the 6 samples, comprising 40-50% of the volatiles
detected (Table 1). In contrast to that for fresh oleoresin,
the amounts of β-pinene in the stump wood samples were
greatly diminished. Since β-pinene has a higher boiling
point than α-pinene, the higher rate of loss of β-pinene
was attributed to its lower stability rather than higher
volatility. The second most abundant compound detected
for these samples was the oxidized monoterpene, αterpineol; significant amounts of other oxidized monoterpenes (e.g., camphor, fenchyl alcohol, borneol) were also
observed. This result was not surprising since wood naval
stores (i.e., that from old pine stumps) have been reported
to contain high amounts (50-60%) of α-terpineol
(Buchanan 1963). Given the similarity in the monoterpene compositions between samples taken from sites
within (Southampton and Sussex counties) and outside
(Caroline) the known range for longleaf pine, the similarity of the monoterpene compositions between the longleaf
pine and loblolly pine oleoresin from live trees, and
within-sample variability, it was not possible to identify
the stumps as longleaf pine apart from the other southern
pines. However, these data do suggest that none of the
original trees were pond pine for which limonene is the
predominant monoterpene. Limonene is a thermal isomerization product of α-pinene (Derfer and Traynor 1989,
Drew et al. 1971) and thus it is unlikely that the high relative amounts of α-pinene detected could be derived
Poster Presentations 80
Table 1. Percentage compositions of monoterpenes and methylchavicol detected in stump wood samples.
Stump Wood Samples
Prince
SouthGeorge
ampton
Caroline
(Scholl)
Caroline
(Pines)
47.37a
48.59
18.06
α-fenchene
0.80
0.42
camphene
3.59
0.24
β-pinene
1.55
myrcene
1.29
Monoterpene
α-pinene
Sussex
(John Hancock)
Sussex
(Joseph
Pines)
58.22
12.07
45.30
3.14
0.58
5.60
0.74
5.46
3.10
7.58
2.99
2.40
nd
1.25
nd
2.75
1.88
nd
0.03
nd
0.19
nd
α-phellendrene
nd
b
3.23
0.41
nd
nd
α -terpinene
nd
1.26
1.33
nd
nd
nd
10.96
8.80
1.63
9.29
0.43
4.61
limonene
ß-phellendrene
nd
6.58
nd
nd
nd
0.31
p-cymene
0.74
0.11
47.97
0.28
19.14
1.40
terpinolene
1.26
2.23
1.89
1.68
nd
1.11
fenchone
0.36
nd
2.88
0.26
13.89
2.32
camphor
1.10
nd
6.58
0.82
19.95
4.36
fenchyl alcohol
2.83
2.78
1.69
1.92
0.15
0.89
terpinen-4-ol
1.62
0.56
1.97
0.93
11.22
3.64
methylchavicol
0.20
0.63
nd
2.55
0.52
6.89
α-terpineol
23.55
17.03
4.72
16.18
7.27
21.58
borneol
2.78
3.26
2.27
2.91
2.18
0.92
a
percent peak area for identified compounds
nd (not detected)
b
from a monoterpene composition predominated by limonene. In addition to the monoterpenes, methylchavicol (pallylanisole) was detected in all but the Prince George sample. Its presence affords few clues to a specific pine species.
Analyses of the Prince George and Sussex (John Hancock)
samples were particularly interesting since they showed an
even greater degree of monoterpene oxidation. In these
samples, the amounts of α-pinene and α-terpineol were significantly lower while higher amounts of p-cymene, fenchone, camphor, and terpinen-4-ol were present. At this
juncture, it should be recognized again that very little data
is available relating monoterpene compositions to age and
species for very old southern yellow pine stumps. In one
case, it has been suggested that the inherent acidity of wood
promotes the conversion of α-pinene to cymene (Drew et al.
1971). Elevated temperatures have been shown to promote
monoterpene oxidation (McGraw et al. 1999). Accordingly,
it is speculated that these two trees (Prince George and Sussex (John Hancock)) were harvested much earlier than the
others and/or were exposed to high temperatures during
forest fires. In fact, burn scars on the Sussex (John Hancock) sample indicate the exposure to fires that one would
expect in a longleaf pine ecosystem. Reported specific gravity values for the wood from the southern yellow pines
show a lower value for loblolly pine as compared to
Table 2. Specific gravity and non-volatile extractives
contents of stump wood samples.
Stump
Wood
Sample
Southampton
Caroline
(Scholl)
Sussex
(John
Hancock)
Poster Presentations 81
Specific Gravity (gcm-3)
Before Extraction
After Extraction
0.94 ± 0.08
0.56 ± 0.03
NonVolatile
Extractives
(%)
42.98
0.70 ± 0.03
0.57 ± 0.02
10.44
0.76 ± 0.04
0.49 ± 0.03
35.29
longleaf pine (Wood Handbook 1974). Specific gravity
values determined for the stump wood samples by the two
different methods gave essentially the same result. All
specific gravity values were significantly higher than
those reported in the literature and reflect the very resinous nature of the samples (Table 2). These data illustrate
that measurement of specific gravity, which can easily be
carried out in the field, could be an alternative to extractions requiring solvents and laboratory facilities. Given
the small difference in specific gravity for longleaf and
loblolly pine woods, it is not surprising that the Southampton and Caroline (Scholl) samples have essentially
the same specific gravity values after extraction. Since
longleaf, and not loblolly pine, has an established history
of use in naval stores production, highly resinous samples
would seemingly have a greater likelihood of being longleaf pine. The high percentage of non-volatile extractives
in the Southampton and Sussex (John Hancock) samples
may reflect their use for naval stores production and provide a tantalizing clue that their identity may be longleaf
pine.
Conclusions
Similarities in the monoterpene compositions for the fresh
oleoresin of the southern yellow pines, and a lack of information about the volatilization and degradation of the
monoterpenes in the natural environment, greatly limit
our ability to assign the monoterpene compositions for
our stump wood samples to specific pine species. However, pond pine was excluded since it differs from most of
the other southern yellow pines with a monoterpene composition predominated by limonene. High extractive
yields from resinous stumps can be readily estimated by
specific gravity. A high extractive yield can be used to
infer those southern yellow pine species used for naval
stores production, specifically, longleaf and slash pines.
Gardner, F.H., Jr. 1989. Wood naval stores. In: Naval
Stores: Production, Chemistry, Utilization. Eds.
Zinkel, D.F., Russell, J. Pulp Chemicals Association,
New York. pp. 143-157
Hodges, J.D., Elam, W.W., Watson, W.F., Nebeker, T.E.
1979. Oleoresin characteristics and susceptibility of
four southern pines to southern pine beetle
(Coleoptera: Scolytidae) attacks. Can. Ent. 111:889896
Koch, P. 1972. Utilization of the Southern Pines, Agriculture Handbook No. 420, USDA Forest Service,
Washington, D.C.
McGraw, G.W., Hemingway, R.W., Ingram, L.L., Jr.,
Canady, C.S., McGraw, W.B. 1999. Thermal degradation of terpenes: ∆3-carene, limonene, and α–
terpinene. Environ. Sci. Technol. 33:4029-4033
Mirov, N.T. 1961. Composition of Gum Turpentines of
Pines, Technical Bulletin No. 1239, USDA Forest
Service, Washington, D.C.
Panshin, A.J., de Zeeuw, C. 1980. Textbook of Wood
Technology, Fourth Edition, McGraw-Hill, New
York
Sheridan, P., Scrivani, J., Penick, N., Simpson, A. 1999.
A census of longleaf pine in Virginia. In: Longleaf
Pine: A Forward Look. Proceedings of the Second
Longleaf Alliance Conference. Ed. Kush, J.S. Longleaf Alliance Report No. 4. Auburn, Alabama. pp.
154-162
Strom, B.L., Goyer, R.A., Ingram, L.L., Jr., Boyd, G.D.L.,
Lott, L.H. 2002. Oleoresin characteristics of progeny of loblolly pines that escaped attack by the southern pine beetle. For. Ecol. Manage. 158:169-178
Wood Handbook. 1974. Wood Handbook: Wood as an
Engineering Material. Agriculture Handbook No. 72,
Forest Products Laboratory, USDA Forest Service,
Washington, D.
Literature Cited
Buchanan, M.A. 1963. Extraneous components of wood.
In: The Chemistry of Wood. Ed. Browning, B.L.
Interscience Publishers, New York. pp. 313-367
Derfer, J.M., Traynor, S.G. 1989. Chemistry of turpentine. In: Naval Stores: Production, Chemistry, Utilization. Eds. Zinkel, D.F., Russell, J. Pulp Chemicals
Association, New York. pp. 225-260
Drew, J., Russell, J., Bajak, H.W. 1971. Sulfate Turpentine Recovery, Pulp Chemicals Association, New
York
Eberhardt, T.L., So, C. 2005. Variability in Southern
Yellow Pine Bark from Industrial Sources. Proceedings of the 59th Appita Conference Pre-Syposium,
Rotorua, New Zealand, May 12-13. pp. 109-112
Poster Presentations 82
Spatial Patterns of Fuels and Fire Intensity in Longleaf Pine Forests
B.L. Estes1, D.H. Gjerstad1, and D.G. Brockway2
1
School of Forestry and Wildlife Sciences, Auburn University, Alabama, 36849, USA
2
Southern Research Station, USDA Forest Service, Auburn, Alabama, 36849, USA
Introduction
Fire and hurricanes are two of the natural disturbances
that create variation at different temporal and spatial
scales (Turner et al. 1989, Whelan 1995). Frequent low
intensity burns, occurring every 1-3 years, are important
in maintaining variable horizontal structure in fire dependent communities such as longleaf pine forests
(Landers et al. 1995, Palik and Pederson 1996, Carter and
Foster 2004). Fire intensity is influenced by a number of
characteristics such as fuel load, type, arrangement, and
moisture, climate, and topography that vary on either
large or fine spatial scales (Hobbs and Atkins 1988, Whelan 1995, Archibold et al. 1998). Natural and imposed
disturbance regimes impact the accumulation and arrangement of fuel across landscapes producing variable fire
intensity.
Hurricanes cause both immediate and delayed mortality in
the overstory contributing to heavy fuel loads, affecting
fire intensity and residence time (Myers and van Lear
1998, Platt et al. 2002). Disturbance regimes play an important role in determining ecological patterns due to differential survival of plant species (Platt and Connell
2003). Fire and hurricane intensity can influence postdisturbance recovery of plant species patterns independently, but these disturbances can also have a distinct interactive effect (Passmore 2005). Characterizing the spatial
patterns of fire intensity has only been partially explored
but could be a useful tool in understanding patterns of
plant recovery. The objectives of this research were to 1)
Identify the presence of spatial patterns in maximum fire
temperature and 2) describe the functional relationship
between pre-fire fuel loads and subsequent maximum fire
temperatures.
Eight 200-m Transects (T), each surrounded by 9-ha compartments of high quality longleaf pine forest, were considered in the fire intensity analysis (Figure 1). The area surrounding each transect consisted of compartments utilized
by the CART study and represented five forest management
treatments (no harvest, single tree selection, group tree selection, uniform shelterwood, and irregular shelterwood).
One month following the completion of logging, Hurricane
Ivan (Category 3) struck the Gulf Coast on September 16,
2004. The EEF had substantial overstory mortality that
occurred in patches across the landscape. Forest management compartments harvested to a shelterwood were highly
vulnerable to windthrow and breakage, while the selectiontreated compartments sustained minor damage. Hurricane
impact resulted in reduced basal area across the forest management compartments compared to the projected treatment
residual BA including the “no harvest” plots. The remaining trees that were damaged during the storm were salvaged. The order of events (harvest, hurricane, and salvage) substantially increased fuel loads and created a more
discontinuous forest floor. Only eight transects were considered due to logistical constraints following Hurricane
Ivan and subsequent salvage operations that left compartments with high amounts of coarse woody debris and bare
soil, along with low residual basal area.
Figure 1. Plot layout and regeneration methods at the Escambia Experimental Forest.
Study Site and Methods
The Escambia Experimental Forest (EEF) (31° 01’ N, 87°
04’ W) is located ten kilometers south of Brewton, Alabama and consists of 1,214 hectares of second-growth
longleaf pine (Figure 1). The dominant tree species at the
EEF is second-growth longleaf pine (80%) that was naturally established from the 1958 seed crop (Boyer and
Miller 1994). In the longleaf pine stands, all stages of
growth are represented from seedlings to saplings to mature trees ranging from 9 to 88 years of age. The midstory
is occupied by a variety of scrub oaks and the understory
consists of grasses, forbs, shrubs, and vines.
T6
T7
T1 T2 T3
T8
T4
Poster Presentations 83
T5
Regeneration
Methods
No harvest
Single-tree
Group Selection
Shelterwood
Fuel components were sampled along the 200 m transects
and an overview of the data collected can be found in Table
1. Fire intensity was estimated by using pyrometers measuring the maximum temperature that occurred during the
prescribed fires (Table 1, Figure 2) (Wally et al. 2006). All
of the blocked variance analyses were performed using
Table 1. Data collected along 200-m transects
Variable
Data Collected
Sampling Interval
Fuels (1,10,100 hr)
Count of all fuels
Continuously along
200 m
Coarse Woody Debris
Distance along transect, length, species,
diameter, decay class
Continuously along
200 m
Litter Depth
Fuelbed Depth
Bare Soil (%)
Shrub Cover (%)
Maximum Fire Temperature (°C)
Depth in cm of identifiable needle and
leaf litter
Depth to tallest vegetation
1m
Results
1m
Measure of disturbed
soil
Continuously along
200 m
Measure of shrub
cover
Pyrometers with heat
sensitive paints (30
cm height)
Continuously along
200 m
PASSAGE: Pattern Analysis, Spatial Statistics And
Geographic Exegesis (Rosenberg 2001). In order to
identify the mean size of the patches and gaps, ThreeTerm Local Quadrat Variance (3TTLQV) was employed (Hill 1973, Dale 1999). The 3TTLQV anlayzes
the difference between overlapping blocks and identifies the pattern, avoiding the peak shift often seen in
Two-Term Local Quadrat Variance (TTLQV) (Guo and
Kelly 2004). The 3TTLQV analysis yields a distinct
variance peak that estimates the mean patch/gap size
exhibited in the data or the mean distance between the
center of a patch and gap (Guo and Kelly 2004). In
order to determine the relationship between the fuel
component predictors and the fire temperature, linear
stepwise regressions were performed for all seven transects and forest management treatments using SAS 9.1.
All models were selected based on an entry and exit
alpha levels of 0.05.
All of the management compartments were burned in
the winter of 2005 with the exception of the compartment with T6 that was burned in the spring after completion of salvage operations following Hurricane Ivan.
The winter burns had a mean air temperature range
from 2 – 12°C while the spring burn (T6) was 15°C and
mean relative humidity ranged from 63 – 82% during
both the winter
1m
Table 2. Mean and standard error of maximum temperature and fuel variables in plots at the Escambia
Experimental Forest.
Transect
T4
T7
T3
T6
T8
T5
T2
Median
and
Average
Range
of
Maximum T
(°C)
121
(121148)
204
(204231)
232
(232259)
260
(260287)
232
(232259)
121
(121148)
288
(288315)
Fine
Woody
Biomass
(g)
Coarse
Woody
Debris
(cm3)
Bare
Soil
(%)
Litter
(cm)
Fuelbed
Depth
(cm)
Shrub
Cover
(%)
Residual
Basal
Area
(m2/ha)
1-hr
10-hr
100-hr
9.3 ±
1.3
90.1 ±
6.6
64.0 ±
9.3
2340 ±
1050
14 ± 2
2.9 ±
0.3
22.7 ±
2.6
14 ± 1
1.4
7.1 ±
1.0
120.7 ±
8.1
124.2 ±
12.6
4453 ±
979
7±1
3.0 ±
0.1
26.0 ±
2.1
25 ± 2
2.4
9.7 ±
1.0
46.8 ±
4.1
23.4 ±
4.8
247 ±
138
11 ± 2
2.1 ±
0.1
14.8 ±
1.8
6±1
7.2
7.7 ±
1.0
67.3 ±
5.5
40.8 ±
7.0
3291 ±
1600
10 ± 2
3.2 ±
0.2
28.9 ±
2.4
23 ± 2
11.5
5.5 ±
1.1
53.9 ±
4.7
25.5 ±
5.6
1558 ±
1078
16 ± 3
2.4 ±
0.5
17.1 ±
1.6
11 ± 2
14.3
8.1 ±
1.0
66.5 ±
3.9
27.6 ±
4.6
2049 ±
1445
9±1
2.5 ±
0.1
11.7 ±
1.4
16 ± 2
19.3
12.8 ±
1.2
70.6 ±
4.5
16.8 ±
3.9
254 ±
170
1± 1
3.7 ±
0.1
19.6 ±
2.4
6± 1
20.9
Poster Presentations 84
Figure 2. Pyrometers used to estimate fire intensity
during prescribed burns at the EEF
and spring burns. Flame height and rate of spread ranged
from 0.5 – 1.0m/min and 0.5 – 1.3 m/min, respectively in
burn compartments with T5 and T4, while burn compartments with T6 and T3 had flame heights of 0.5 – 2 m and
flame spread of 1 -2 m/min. Compartments surrounding
T7 and T8 had maximum flame heights of 2m and a rate
of spread of up to 4 m/min.
The fuel components sampled along the transects varied
along a disturbance gradient with low disturbance in the
no harvest compartments and high disturbance in the compartments where forest management objectives required
overstory removal. The disturbance gradient was reflected in the residual basal area left within each compartment, following the multiple disturbances with T2 and T5
having 19-20 m2/ha, the selection compartments (T3, T6,
and T8) retained 7-14 m2/ha, and the shelterwood plots
had a low residual basal area ranging from 1.4-2.4 m2/ha
(Table 2). The 1-hr fuels were high along T5 (8.1 g) and
T2 (12.8 g) when compared to harvested plots (T4, T7,
T3, T6, and T8) (Table 2). The opposite was true of 100hr fuels as increased biomass was noted in the compartments that had some form of overstory removal (Table 2).
Litter depth was fairly consistent along the transects and
ranged from 2.1-3.7 cm (Table 2).
Table 3. Results from Three-Term Local Quadrat Variance (3TTLQV) with main peak and secondary peaks in
parentheses
Transects
Temp
1-hr
10-hr
100hr
Litter
Fuelbed
Shrub
T2
33
45
65
-
33
55
49
T3
49
(30)
25
-
62
(14)
46
(13)
10
45
(27)
T4
34
(18)
22
12
(23)
24
35
-
39
T5
34
(6,17)
46
(12,5
0)
9
-
31
26
17
-
37
23
15
44
(10)
49
(19)
58
(4)
62
(6,34)
29
(9)
29
(10)
-
37
38
36
28
42
(12)
17
(8)
57
(20)
T6
T7
T8
44
(14)
55
(11,33)
12
Maximum fire temperatures varied along the 200-m transects resulting in gaps of low temperatures and patches of
high temperatures. The highest fire temperature range occurred in T2 (288-315°C) while the lowest fire temperature
range occurred in T5 (121-148°C) (Figure 3). Burn compartments that were harvested to a shelterwood had two
varying ranges of maximum temperatures with T4 showing
a range of 149-176°C and T7 having a high temperature
range of 232-259°C (Figure 3). Burn and selection harvest
compartments had similar ranges of maximum temperatures
with T3 and T6 both harvested in groups of trees having
ranges of 204-231°C and 232-259°C and T8, in a single
tree harvest compartment, had a temperature range of 232259°C (Figure 3).
Peak variance according to Three Term Local Quadrat Variance (3TTLQV) occurred at block size of 33-34 in the no
harvest plots with T5 having several peaks indicating variable spatial influence (Table 3). Peak block size in the harvesting plots ranged from 34-55 indicating that fire temperature was operating at larger scales (Table 3).
Figure 3. Frequency of maximum fire temperature
throughout prescribed burns
The results of stepwise regression indicate different fuel
predictors explain the variability in fire intensity pattern that
was observed along the transects in the 3TTLQV analysis.
The model explaining fire intensity variability along the
transects sampled in the “no harvest” plots explained only a
small amount of variability (20%), with litter depth, bare
soil and shrub cover entering the model. When considering
T5 and T2 individually, neither model explained a substantial amount of variability, but litter depth and 1-hr fuel biomass were consistently important predictors. The models
explaining fire intensity along the transects in the selection
and shelterwood treatments explained 28-33% of the variation. The important factors in these models were 10-hr fuel
biomass, litter and fuelbed depth, as well as shrub cover
while the model explaining the variability in the fire temperature in the shelterwood treatment added 100-hr fuel
Poster Presentations 85
biomass. Bare soil (%) was a significant predictor in all
models, but explained more variability in the shelterwood
and selection plots. The fuel complex was extremely sensitive to harvesting and wind disturbance as was evident on
the Escambia following Hurricane Ivan and subsequent
salvaging operations.
Conclusions
Hurricane disturbance: Hurricanes had two major impacts
on the forested ecosystem: 1) reduction of overstory canopy
cover and 2) the addition of substantial fuel loads in the
form of pine straw and hardwood litter, downed woody
fuel, and coarse woody debris (Lugo 2000, Beckage et al.
2006).
Fire-temperature variability: There was considerable variation in temperatures recorded by the pyrometers at adjacent
locations. The selection and shelterwood plots had a larger
scale of pattern while the no harvest plots or low fire intensity plots varied on moderate scales (Franklin et al. 1997,
Rocca 2004, Kennard and Outcalt 2006).
Fuel variability: Bare soil and litter depth were important
predictors in all regression models. Bare soil was mainly
attributed to skid trails and logging decks as well as ground
disturbance from tip up mounds that occurred during Hurricane Ivan (Robichaud and Miller 1999). The plots that only
had minimal disturbance (BA) had flame temperature explained by litter depth, the selection plots or those with
moderate BA were influenced by 1, 10-hr fuels as well as
litter and shrub while those that were below 5 m2/ha basal
area were influenced by all fuel variables (Grace and Platt
1995).
Summary: Determining the spatial pattern and the predictors of fire intensity may be important in evaluating the
impacts of overstory removal, wind damage from hurricanes, and salvage operations on planning prescribed burns
and predicting post-fire effects.
Literature Cited
Archibold, O. W., L. J. Nelson, E. A. Ripley, and L. Delanoy.
1998. Fire temperatures in plant communities of the northern
mixed prairie. Can Field-Nat 112:234-240.
Beckage, B., L. J. Gross, and W. J. Platt. 2006. Modelling responses of pine savannas to climate change and large-scale
disturbance. Appl Veg Sci 9:75-82.
Boyer, W. D., and J. H. Miller. 1994. Effect of burning and brush
treatments on nutrient and soil physical properties in young
longleaf pine stands. For Ecol Manag 70:311-318.
Carter, M. C., and C. D. Foster. 2004. Prescribed burning and
productivity in southern pine forests: a review. For Ecol
Manag 191:93-109.
Dale, M. R. T. 1999. Spatial pattern plant analysis in plant ecology. Cambridge University Press, Cambridge.
Franklin, S. B., P. A. Robertson, and J. S. Fralish. 1997. Smallscale fire temperature patterns in upland Quercus communities. J Appl Ecol 34:613-630.
Grace, S. L., and W. J. Platt. 1995. Effects of adult tree density
and fire on the demography of pregrass stage juvenile longleaf pine (Pinus palustris Mill.). J Ecol 83:75-86.
Guo, Q., and M. Kelly. 2004. Interpretation of scale in paired
quadrat variance methods. J Veg Sci 15:763-770.
Hill, M. O. 1973. The intensity of spatial patterns in plant communities. J Ecol 61:225-235.
Hobbs, R. J., and L. Atkins. 1988. Spatial variability of experimental fires in south-west Western Australia. Aust J Ecol
13:295-299.
Kennard, D. K., and K. W. Outcalt. 2006. Modeling spatial patterns of fuels and fire behavior in a longleaf pine forest in
the southeastern USA. Fire Ecology 2:31-52.
Landers, J. L., D. H. Van Lear, and W. D. Boyer. 1995. The
longleaf pine forests of the Southeast: requiem or renaissance? J of For 93:39-44.
Lugo, A. E. 2000. Effects and outcomes of Caribbean hurricanes
in a climate change scenario. Sci Total Environ 262:243251.
Myers, R. K., and D. H. van Lear. 1998. Hurricane-fire interactions in coastal forests of the south: a review and hypothesis. For Ecol Manag 103:265.
Palik, B. J., and N. Pederson. 1996. Overstory mortality and
canopy disturbances in longleaf pine ecosystems. Can J For
Res 26:2035-2047.
Passmore, H. A. 2005. Effects of hurricanes and fires on southeastern savanna-forest landscapes. Louisiana State University, Baton Rouge.
Platt, W. J., B. Beckage, R. F. Doren, and H. H. Slater. 2002.
Interactions of large-scale disturbances: prior fire regimes
and hurricane mortality of savanna pines. Ecology 83:15661572.
Platt, W. J., and J. H. Connell. 2003. Natural Disturbances and
directional replacement of species. Ecol Monogr 73:507522.
Robichaud, P. R., and S. M. Miller. 1999. Spatial interpolation
and simulation of post-burn duff thickness after prescribed
fire. Int J Wildland Fire 9:137-143.
Rocca, M. E. 2004. Spatial considerations in fire management:
the importance of heterogeneity for maintaining diversity in
a mixed-conifer forest. Dissertation. Duke University.
Rosenberg, M. S. 2001. PASSAGE. Pattern analysis, spatial
statistics and geographic exegesis. in. Arizona State University, Tempe.
Turner, M. G., V. H. Dale, and R. H. Gardner. 1989. Predicting
across scales: Theory development and testing. Landsc Ecol
3:245-252.
Wally, A. L., E. S. Menges, and C. W. Weekley. 2006. Comparison of three devices for estimating fire temperatures in
ecological studies. Appl Veg Sci 9:97-108.
Whelan, R. J. 1995. The Ecology of Fire. Cambridge University
Press, New York.
Poster Presentations 86
Evaluating Forest Development and Longleaf Pine Regeneration at
Mountain Longleaf National Wildlife Refuge
Bill Garland1, John S. Kush2, and John C. Gilbert2,
1
2
U.S. Fish and Wildlife Service, Mountain Longleaf NWR, Fort McClellan, AL, USA
School of Forestry and Wildlife Sciences, Auburn University, Auburn, AL, 36849, USA
Abstract
The Mountain Longleaf National Wildlife Refuge
(MLNWR), formerly part of the Fort McClellan military
training installation, is located in northeast Alabama.
Mountain longleaf pine forests are a critically endangered
component of the once vast longleaf pine forests in the
Southeast. Unlike the Coastal Plain where small pockets
of the forest still remain, the MLNWR stands alone in the
mountain region. Previous research efforts have identified 9 old-growth tracts, lush herbaceous communities,
and several management regimes on the refuge. However, many of the stands of longleaf pine on MLNWR
have undergone various lengths of fires suppression and
degradation. This has led to hardwood encroachment and
longleaf pine regeneration failure. To evaluate the current condition of the forest, two old-growth stands were
resurveyed and assessed longleaf regeneration.
As part of an effort to determine the possible presence of
unexploded ordnance (UXO), the Army has randomly
sampled ¼ acre plots across the Refuge. Plots were
cleared of vegetation 4-inches DBH (diameter at breast
height; 4.5 feet) and smaller to aid equipment mobility.
A subset of these plots was examined to determine the
effects of these disturbances on forest composition, density, and regeneration.
study involves establishing a monitoring program in various stand situations, measuring existing impacts from
time of clearance (December 2002) and providing recommendations to biologists and managers on projected future
impacts in various forest types.
Using a study established in 1999 in two old-growth
stands and a subset of cleared UXO plots, forest development and longleaf pine regeneration at the MLNWR will
be addressed.
Study Area and Methodology
UXO Plots
The Army primary cleared all vegetation 4-inches DBH
and smaller from 1/4-acre randomly placed plots. A subset of plots on slopes with southerly aspects with an emphasis on areas predominated by longleaf pine were surveyed for their condition. The primary characteristics of
interest in this survey were composition, density, and
longleaf pine regeneration.
Species Composition
All stems on 1/10-acre plots located in the middle of the
1/4-acre cleared plots were stem mapped from plot center,
stems called and DBH measured.
Introduction
Since 1994, field reconnaissance on Fort McClellan
Army Base, now Mountain Longleaf National Wildlife
Refuge (MLNWR), by Auburn University’s School of
Forestry & Wildlife Sciences identified a number of oldgrowth longleaf pine stands. Many of these stands have
undergone various lengths of fire suppression and degradation. MLNWR’s longleaf pine forests provide the
“missing link” to scientists, land managers, and conservationists in the mountain region, providing the only information on 1) age and stand structure and dynamics of
frequently burned old-growth forests, 2) composition of
pristine plant communities, and 3) landscape extent of
mountain longleaf pine forests.
The U.S. Army is currently characterizing these lands to
determine the possible presence of unexploded ordnance
(UXO). This same sampling methodology represents one
option level for final remediation of the entire area. This
Longleaf Pine Regeneration
Within each 1/4-acre plot, four milacre quadrats were
placed in cardinal directions. Within each quadrat, all
longleaf pine seedlings were counted and mapped for future reference.
Old-growth stands
Two old-growth stands were re-surveyed. Originally,
these stands were selected because they had frequent
burning histories (1 to 3 year return interval) and good
accessibility. The two stands, hereafter referred to as
Caffey Hill and Red-tail Ridge, were located at ca. 1475
and 1150 feet above mean sea level. Caffey Hill is an
upper slope stand covering 3.7 acres and Red-tail Ridge
spans a mid- to upper slope position covering 4.5 acres.
Slopes of the two sites ranged from 40 to 60% at Caffey
Hill and 30 to 45% at Red-tail Ridge. Caffey Hill had a
Poster Presentations 87
SSE aspect, while Red-tail Ridge’s was WSW. Within each
stand, we re-measured all living longleaf pines > 1.0 inch
DBH (100% sample) for DBH (to nearest 0.1-inch). To
examine longleaf pine regeneration, transects which were
100 feet long by 3.0 feet wide were established at each
setup point in each stand. These transects were run in the
cardinal directions from each point. In addition, the 20 9foot X 9-foot quadrats were re-surveyed for longleaf pine
seedlings.
Results and Discussion
UXO Plots
Old-growth stands
Table 2 presents the stand data for Caffey Hill and
Red-tail Ridge with comparisons to the data collected in
1999. Stand density has decreased in both stands but
basal area increased at Red-tail Ridge while declining at
Caffey Hill. No longleaf pine regeneration was found
in the plots established in 1999 at either Caffey Hill or
Red-tail Ridge. No longleaf pine seedlings were observed in these transects on Caffey Hill, nor were any
observed near the vicinity of the stand. The conditions
at Red-tail Ridge were similar, though 4 seedlings were
found in the 16 transects, or the equivalent of 36 seedlings/acre.
Of the 26 UXO plots examined on southerly aspects, only 8
plots were dominated by longleaf pine based on basal area.
Table 2. Current stand characteristics for Caffey Hill
Table 1 presents the species present on each plot and its
and Red-tail Ridge, two old-growth longleaf pine
basal area and density. Across all plots, longleaf pine basal
stands, at Mountain Longleaf National Wildlife Refuge
area averaged 31.46 square feet/acre. This is the minimum
compared to the data collected in 1999 (Varner 2003a,
basal area required for adequate natural regeneration of
2003b).
longleaf pine. Only 12 of the 26 plots would have enough
Caffey Hill
longleaf pine basal area to obtain natural regeneration.
Character
Current
1999
Longleaf pine accounted for 35% of the basal area but only
15% of the total number of stems. Except for black cherry,
Stem density (trees/acre)
94.4
120.6
red maple, sweetgum, and yellow-poplar, the other species
would be expected in a montane longleaf pine-dominated
Basal area (square feet/acre)
32.63
34.84
ecosystem. Longleaf pine regeneration was absent from all
6.71
5.69
but 8 plots. Of these, only 3 plots had more than a few Arithmetic Mean DBH (inches)
seedlings and would be considered adequately stocked with
Maximum DBH (inches)
20.8
21.6
regeneration.
Table 1. Tree density (stems/acre) and basal area (square
feet/acre) for species located on 26 UXO plots on the
Mountain Longleaf National Wildlife Refuge.
Species
Black cherry
Blackgum
Density Basal area
7.7
2.79
14.6
2.35
Chestnut oak
25.4
10.95
Dogwood
1.9
0.31
Hickory spp.
23.1
6.65
Loblolly pine
6.2
3.65
Longleaf pine
33.1
31.46
Oak spp.
44.6
12.26
Post oak
3.1
2.08
Red maple
12.4
1.48
Shortleaf pine
12.4
6.61
Sourwood
13.1
3.21
Southern red oak
2.3
0.07
Sweetgum
1.9
0.35
White oak
1.5
0.41
Yellow-poplar
1.5
1.36
Red-tail Ridge
Character
Current
1999
Stem density (trees/acre)
105.2
114.5
Basal area (square feet/acre)
59.26
56.03
Arithmetic Mean DBH (inches)
8.87
8.03
Maximum DBH (inches)
27.4
27.9
Acknowledgements
The authors wish to thank the U.S. Geological Survey Alabama Cooperative Fish and Wildlife Research Unit and the
U.S. Fish and Wildlife Service.
Literature Cited
Varner, J.M., J.S. Kush, and R.S. Meldahl. 2003a. Vegetation of frequently burned old-growth longleaf pine
(Pinus palustris Mill.) savannas on Choccolocco
Mountain, Alabama, USA. Natural Areas Journal. 23
(1):43-52.
Varner, J.M., J.S. Kush, and R.S. Meldahl. 2003b. Structural characteristics of frequently-burned old-growth
longleaf pine stands in the mountains of Alabama. Castanea. 68(3):211-221.
Poster Presentations 88
Wiregrass – Overrated
John C. Gilbert1, John S. Kush1, and John McGuire1
1
School of Forestry and Wildlife Sciences, Auburn University, Auburn, Alabama, 36849, USA
Abstract
Ranges of Bluestem and Wiregrass
Aristida beyrichiana and Aristida stricta, commonly
referred to as wiregrass, are both conditional understory
species in the longleaf pine (Pinus palustris Mill.) ecosystem. A common but misleading association exists
between longleaf pine and wiregrass. The longleaf pine/
wiregrass ecosystem is often used synonymously with
the entire longleaf pine ecosystem. Another common
misconception is that the presence of wiregrass is necessary to maintain the longleaf pine ecosystem. In fact,
the understory of the longleaf pine ecosystem is composed of an extensive variety of species which is not
consistently dominated by a single species.
Pineland threeawn (Aristida beyrichiana Trin. & Rupr.) is
commonly referred to as wiregrass. This type of wiregrass
has a range from Mississippi to Florida and north into North
and South Carolina. Another similar species of wiregrass
that exists in North and South Carolina is Aristida stricta
Michx. Both species are part of the longleaf pine-wiregrass
range, but they have been classified as two different species
(Miller and Miller 1999; Peet 1993). However, Kesler et al.
2003 argued that on the species level the two were not distinctly different.
Wiregrass is one of the more famous of the understory
species, but an often forgotten but prominent understory
component of the longleaf pine ecosystem is bluestem
(Andropogon spp.). Bluestem is also referred to as
broomsedge, broomstraw, beardgrass, and a variety of
other common names. At least nine species of bluestem
have been identified in the longleaf pine ecosystem.
The range of bluestem stretches from Texas to Florida,
north to Maine, and west to North Dakota, which blankets the range of the longleaf pine ecosystem. The presence of bluestem provides many aesthetic, commercial,
and ecological values to the understory of longleaf pine
stands. With extensive restoration efforts of the longleaf pine ecosystem underway, bluestem is a valuable
understory component across the longleaf pine ecosystem.
Introduction
A common misconception about the longleaf pine ecosystem is that wiregrass is a necessary component of the
understory vegetation. Wiregrass is an integral understory component in the longleaf pine ecosystem, but it
is not present throughout the range of longleaf pine.
Even in the major range of wiregrass, it can be found as
a co-dominant with a variety of other understory species
(Outcalt 2000). In fact, the understory of the longleaf
pine ecosystem is not dominated by any single species
(Outcalt 2000). Another prominent understory component of the longleaf pine ecosystem is bluestem. The
importance of bluestem in the understory of longleaf
pine forests is often overlooked.
Bluestem is also referred to as broomstraw, beardgrass, and
a variety of other common names. The range of bluestem
stretches from Texas to Florida, north to Maine, and west to
North Dakota, which blankets the range of the longleaf pine
ecosystem (Miller and Miller 1999). Bluestem has several
species present in its range. Table 1 displays the nine separate species of bluestem that exist in the longleaf pinebluestem range (Grelen and Duvall 1966).
Both wiregrass and bluestem have extensive ranges in the
southeastern United States. Range maps vary from different
sources because natural and mechanical disturbances continually affect the composition of stands. There is also a
great deal of overlap between the extensive ranges of the
species of interest. Endozoochory with white-tailed deer as
the vector is a possible explanation for the colonization of
many forest species across landscapes (Myers et al. 2004).
Seed dispersal through endozoochory as well as ectozoochory with a variety of vectors has the potential to explain the spread of wiregrass and bluestem across this vast
ecosystem. This concept also gives rise to the threat of exotic encroachment on native understory species.
Table 1. Common Species of Bluestem in the Longleaf
Pine-Bluestem Range (adapted from Grelen and Duvall
1966).
big bluestem (Andropogon gerardii Vitm.)
pinehill bluestem (Andropogon divergens (Hack.))
little bluestem (Andropogon scoparius Michx.)
slender bluestem (Andropogon tener (Nees) Kunth)
broomsedge bluestem (Andropogon virginicus L.)
bushy bluestem (Andropogon glomeratus (Walt) BSP.)
Elliott bluestem (Andropogon elliotii Chapm.)
fineleaf bluestem (Andropogon subtenuis Nash)
paintbrush bluestem (Andropogon ternarius Michx.)
Poster Presentations 89
Range Types of the South
Figure 1 shows the range types of the south. Included in
these range types are the longleaf pine-bluestem and
longleaf-slash pine-wiregrass ranges. The longleaf pinebluestem range is largely located in southern Louisiana
and Texas, in southeastern Mississippi, and in southwestern and central Alabama. Small portions of this range are
also located in central Georgia, South Carolina, and North
Carolina (Williams et al. 1955). The longleaf-slash pinewiregrass range covers almost the entire state of Florida,
the southern portion of Alabama, and almost the southern
half of Georgia. Small portions of this range are also
located in the southwestern corner of South Carolina
(Williams and others 1955). This map also shows the
location of the shortleaf-loblolly pine-bluestem range.
This range stretches from eastern Texas to Virginia covering a vast amount of the southeastern United States
(Williams et al. 1955). This range covers the of the longleaf pine ecosystem, which makes bluestem a very integral understory component.
Figure 2. Map by Erin Hart.
Longleaf Pine - Bluestem Range
Figure 3 shows an estimation of the longleaf pinebluestem range in its virgin state. This range extends
from western Florida and southern Alabama west into
Texas. This range has been altered due to the clear cutting of old-growth pines and alterations of the natural
burn cycles. The natural stand composition that existed
in pre-settlement times exists in few places. There has
been a conversion of the overstory to loblolly pine
(Pinus taeda) and slash pine (Pinus elliottii) due to extensive planting and a failure to regenerate longleaf pine
forests. Even though there have been extensive disruptions to the stand composition, bluestem grasses have
continued to be persistent across this landscape. They
still represent one of the most important forage species
in this range (Grelen and Duvall 1966).
Figure 1. (Williams et al. 1955)
Bluestem and Wiregrass Ranges within the Longleaf
Pine Ecosystem
Figure 2 shows a current estimation of the range of bluestem and wiregrass species within the longleaf pine ecosystem. It includes more area dominated by blue stem in
central and north portions of Alabama, Georgia, South
Carolina, and Virginia than the other maps. This map
shows the prominence of bluestem as a dominant understory species in the longleaf pine ecosystem. The map
also provides a good representation of the overlap between the ranges of wiregrass and bluestem.
Poster Presentations 90
Figure 3. Grelen and Duvall 1966
WHY BLUESTEM SHOULD NOT BE OVERLOOKED
very important to landowners that are trying to restore
native understory species in longleaf pine forests.
1.) Loss of pristine old-growth forests
4.) Adaptations to prescribed fire
Bluestem is an integral understory component that should
not be overlooked. Ground cover degradation, increasing
urban interface, and the invasion of exotic species are all
threats to the existence of pristine old growth longleaf
pine stands (Varner and Kush 2004). States like Mississippi, Louisiana, and Texas contain little or no old growth
longleaf stands, which is the heart of the longleaf pinebluestem ecosystem (Varner and Kush 2004).
Wiregrass and bluestem are both well adapted to fire,
which is important to the survival of the longleaf pine
ecosystem. One advantage bluestem has is that it disperses its seeds by the wind unlike the fire dependent
wiregrass (Miller and Miller 1999). On the other hand,
wiregrass seed production requires fire to enter the stand
in the late spring and early summer (Miller and Miller
1999). Without frequent summer burns, the existence and
forage value of wiregrass can suffer while bluestem continues to flourish. Kush et al. (2000) found a higher species diversity for biennial winter burning than spring and
summer burns in bluestem dominated understories. An
advantage for bluestem is that it is not reliant on summer
burns for survival, which allows more flexibility in scheduling burns. External limitations on prescribed burning
continue to increase mainly as a result of conflicts with an
increasing urban interface and number of smoke sensitive
areas. Prescribed burn practioners do not need to place
further internal restrictions on themselves and narrow the
window of opportunity to burn a particular location only
in the summer. As it becomes more and more difficult to
burn, some fire is better than no fire.
2.) Species diversity
The (Peet and Allard 1993) characterized the longleaf
pine ecosystem into 5 xeric, 6 subxeric, 4 mesic, and 8
seasonly wet communities. The mesic longleaf pine
woodlands which have a mixture of bluestem and wiregrass in the understory contained between 100 and 140
vascular plants per 1000m2 which identifies it a one of the
most species rich communities in temperate North America (Peet and Allard 1993; Peet et al. 1990). In old
growth mountain longleaf pine stands located at Fort
McClellan in the Blue Ridge Physiographic Province of
Alabama, Varner et al. 2003 identified 72 native understory species. Paintbrush bluestem (Andropogon ternarius Michx.) was identified as one of the dominating
understory species. (Kush et al. 2000) found slender bluestem (Schizachyrium tenerum) to be the most frequent
occurring species across all treatments assoicated with the
study on the Escambia Expeimental Forest in Brewton,
AL. Brockway and Lewis (2003) documented 148 vascular plant species in the longleaf pine-bluestem ecosystem
in the Conecuh National Forest in Covington County, AL.
3.) Persistence
Even in the longleaf pine wiregrass range, bluestem can
still dominate the understory. Wiregrass does not reestablish itself once it has been removed (Grelen 1978). An
area where wiregrass has died out or been mechanically
removed can be occupied by bluestem. Bluestem can
also become a dominant component in the understory
depending on management strategies. An old growth
longleaf pine stand in Flomaton, AL had not been burned
for 45 years. After a prescribed burn regiment was implemented, Andropogon virginicus seeded in naturally
(Varner et al. 2000). To reestablish wiregrass after it has
died out or been removed, extensive planted operations
need to be executed. Bissett 1998 estimated it would cost
$3,365/ha to reestablish wiregrass with a direct seeding
method. Pittman and Karrfalt (2000) gave a production
cost of $170/1000 for wiregrass seedlings. These costs are
5.) Wildlife and production agriculture benefits
Wiregrass and bluestem are also sources of forage for
cattle and wildlife. After a burn, wiregrass and bluestem
are both preferred forage. However, wiregrass is not preferred 3-4 months after a burn, while bluestem is still preferred (White and Terry 1979). Along with the food and
fuel sources bluestem provides, it is also very useful as
cover to wildlife. Even through the lack of fire and extensive changes in overstory composition, bluestem has been
persistent in the longleaf pine-bluestem ecosystem
(Grelen and Duvall 1966).
Literature Cited
Bissett, N.J. 1998. Direct seeding wiregrass, Aristida
beyrichiana, and associated species. Pages 59-60 In:
Kush, J.S., compiler. Longleaf Alliance Report No.3.
Proceedings of the longleaf pine ecosystem restoration symposium presented at the society for ecological restoration 9th annual international conference;
1997, November 12-15; Fort Lauderdale, Florida,
USA.
Brockway, D.G. and C.E. Lewis. 2003. Influence of
deer, cattle grazing, and timber harvest on plant species diversity in a longleaf pine bluestem ecosystem.
Forest Ecology and Management 175:49-69.
Grelen, H.E. 1978. Forest grazing in the south. Journal
of Range Management. 31(4):244-250.
Poster Presentations 91
Grelen, H.E. and V.L. Duvall. 1966. Common plants of
longleaf pine- bluestem range. Southern Forest Exp.
Sta., New Orleans, Louisiana. 96 pp., illus. (U.S. Forest Serv. Res. Pap. SO-23).
Hart, E. from Managing the forest and the trees a private
landowner’s guide to conservation management of
longleaf pine. Theo Davis Sons, Inc., Zebulon, NC.
37p.
Kesler, T.R., L.C. Anderson, S.M. Hermann. 2003. A
taxonomic reevaluation of Aristida stricta (Poaceae)
using anatomy and morphology. Southeastern Naturalist. 2(1):1-10.
Kush, J.S. R.S. Meldahl, and W.D. Boyer. 2000. Understory plant community response to season of burn in
natural longleaf pine forests. Pages 32-39 in W. Keith
Moser and Cynthia F. Moser (eds.). Fire and forest
ecology: innovative silviculture and vegetation management. Tall Timbers Fire Ecology Conference Proceedings, No. 21. Tall Timbers Research Station, Tallassee, FL.
Miller, J.H. and K.V. Miller. 1999. Forest plants of the
southeast and their wildlife uses. Southern Weed Science Society. Craftmasters Printers, Incorporated Auburn, AL. 454p.
Myers, J.A., M. Vellend, S. Gardescu, and P.L. Marks.
2004. Seed dispersal by white-tailed deer: implications
for long-distance dispersal invasion, and migration of
plants in eastern North America. Oecologia 139: 35-44.
Outcalt, K.W. 2000. The longleaf pine ecosystem of the
south. Native Plants Journal 1:43- 44,47-53.
Peet, R.K. 1993. A taxonomic study of Aristida stricta and
A. beyrichiana. Rhodoara 95(881):25-37.
Peet, R.K. and D.J. Allard. 1993. Longleaf pine vegetation
of the southern atlantic and eastern gulf coast regions: a
preliminary classification. Pages 45-81 in Sharon M.
Hermann (ed). The longleaf pine ecosystem: ecology,
restoration, and management. Tall Timbers Fire Ecology Conference Proceedings, No. 18. Tall Timbers
Research Station, Tallassee, FL.
Peet, R.K., E. van der Maarel, E. Rosen, J. Willems, C.
Norquist, and J. Walker. 1990. Mechanisms of coexistence in species-rich grassland. Bulletin, Ecological
Society of America: 71:283.
Pittman, T. and R.P. Karrfalt. 2000. Wiregrass propagation
at the Andrews Nursery in Florida. Native Plants Journal 1:45-47.
Varner, J.M. III and J.S. Kush. 2004. Remnant old-growth
longleaf pine (Pinus palustris Mill.) savannas and forests of the southeastern USA: status and threats. Natural Areas Journal 24:141-149.
Varner, J.M. III, J.S. Kush, R.S. Meldahl. 2000. Ecological restoration of an old-growth longleaf pine stand
utilizing prescribed fire. Pages 216-219 in W. Keith
Moser and Cynthia F. Moser (eds.). Fire and forest
ecology: innovative silviculture and vegetation management. Tall Timbers Fire Ecology Conference
Proceedings, No. 21. Tall Timbers Research Station,
Tallassee, FL.
Varner, J.M. III, J.S. Kush, R.S. Meldahl. 2003. Vegetation of frequently burned old-growth longleaf pine
(Pinus palustris Mill.) savannas on Choccolocco
Mountain, Alabama, USA. Natural Areas Journal
23:43-52.
White, L.D. and W.S. Terry. 1979. Creeping bluestem
response to prescribed burning and grazing in south
Florida. Journal of Range Management. 32(5): 369371.
Williams, R.E., J.T. Cassady, L.K. Halls, and E.J.
Woolfolk. 1955. Range resources of the south.
Southern Section American Society of Range Management in cooperation with Georgia Agricultural
Experiment Stations, University of Georgia College
of Agriculture. Bulletin N.S. 9. 31p.
Poster Presentations 92
Longleaf Pine Re-Discovered at Horseshoe Bend National Military Park
John C. Gilbert1, Sharon M. Hermann2, John S. Kush1, Lisa McInnis3 and James Cahill3
1
Auburn University School of Forestry & Wildlife Sciences, Auburn, Alabama, 36849, USA
2
Auburn University Department of Biological Sciences, Auburn, Alabama, 36849, USA
3
US Department of Interior, National Park Service, Daviston, Alabama, 36256 USA
Abstract
Site
Horseshoe Bend National Military Park (HOBE) in
Tallapoosa County, near Daviston, AL, was the site of
the final battle of the Creek War in 1814. At that time
there had been little farming or logging on the 2,040
acre site. In 1814, a letter from General Coffee to General Jackson described the forest around the battlefield
as an “open hilly woodland”. During the next 140
years, HOBE experienced moderate grazing with
patches in agriculture. In 1905, timber survey records
from nearby property documented longleaf pine as the
dominant tree, but over the next quarter century, much
of the uplands were logged. In 1959, the National Park
Service (NPS) acquired HOBE. For the last 50 years,
there has been no wild or prescribed fire, and land management activity has been limited to mowing of the
battlefield and visitor areas. Past logging coupled with
prolonged fire suppression dramatically altered the site,
and it was thought that most of the longleaf pine had
been extirpated from HOBE. However, a recent search
of uplands uncovered three stands of suppressed longleaf pine plus an additional 280+ individuals of adult
trees scattered over the landscape. There is interest in
restoring the forest to its 1800’s condition, and the NPS
has begun re-introduction of fire in an effort to enhance
the remnants of the longleaf pine forest.
Horseshoe Bend National Military Park (HOBE) was the
site of the final battle of the Creek War in 1814. At that
time there was little farming or logging on the 825 ha site.
For next 140 years, HOBE experienced moderate grazing
with patches in agriculture. Over most of the region, much
of the uplands were logged in first quarter of 20th century.
In 1959, the NPS acquired HOBE. For the last 50 years,
there has been no wild or prescribed fire, and land management activity has been limited to mowing of battle field and
visitor areas.
Introduction
Throughout much of the late nineteenth and twentieth
centuries many sites in the northern (montane or Piedmont) range of longleaf pine (Pinus palustris) suffered
extensive cutting and prolonged periods of fire suppression that resulted in alteration of habitat structure and
species composition. One example of this alteration is
found in Tallapoosa County Alabama at Horseshoe
Bend National Military Park (HOBE), where it was
thought that longleaf pine had been extirpated from the
area in the early 20th century.
In 2004, the National Park Service (NPS) became interested in re-introducing fire to the fire suppressed stands
on HOBE in an effort to reduce hazardous fuels. As
part of the preliminary work for this effort, stands were
examined for their condition and the search uncovered
several stands supporting previously undocumented
fire-suppressed longleaf pine.
Longleaf Forest 100-200 Years Ago
Old-growth longleaf stands are rare and none remain in this
portion of range; however, early accounts of this region and
HOBE provide insight into forest structure and composition.
• In 1775, Bartram described a “vast open forest”
with longleaf, loblolly (Pinus taeda), and chestnut
on hills ~ 60 km south of HOBE.
• In 1814, General John Coffee wrote a letter to
General Andrew Jackson and noted that he had
established a battle line “in an open hilly woodland” associated with the site that is now HOBE.
• Detailed information was reported by Reed in
1905 for Coosa Co, Alabama, ~80 km northwest of
HOBE (see Hermann and Kush 2006); although
Reed describes forests almost a century after the
battle, large-scale logging and fire suppression had
yet to reach region.
• Early accounts plus data and photos from Reed
(1905) depict native upland open-canopied forests
dominated by longleaf. The detailed forest inventory data provided by Reed depicts most stands as
supporting multiple or all age-class trees.
Stand-Level Assessment: Methods and Results (20052006)
•
Following compass bearings, uplands of HOBE were
systematically searched for areas supporting dense
stands of residual longleaf pine.
• 3 stands (2-5 ha in area) were located; all longleaf
> 2.5 cm (1 in) diameter at breast height (DBH)
plus all other medium and large trees (> 15 cm
DBH) were stem-mapped.
Poster Presentations 93
•
•
Stems/ha for medium and large trees, > 15 cm (6
in) DBH, of all tree species on each stand was estimated and compared to data describing nearby
stands in the 1905 upland landscape (Reed 1905).
Size-classes of longleaf > 2.5 cm DBH in stands
were compared to data from 1905.
Landscape Level Assessment Compared to Stands:
Methods and Results (2006)
•
Large longleaf at landscape-level
•
•
•
•
Longleaf trees > 38 cm (15 in) DBH are large
enough to produce cones (Boyer 1990).
To estimate the number of residual large longleaf
over the entire site, the uplands of HOBE were
systematically searched, following compass bearings. GPS locations and DBH measurements were
recorded for of all large longleaf, > 38 cm DBH.
A few smaller longleaf trees were noted during the
landscape-level search but were not recorded. We
estimate that fewer than a dozen grass-stage longleaf persist at HOBE
Relict Oak Trees
•
Although historical descriptions and timber assessments indicate that uplands near HOBE were dominated by longleaf pine, Reed (1905) also indicated that
scattered hardwood trees were a natural part of this
forest.
• Some of the largest hardwoods in the three stemmapped stands were cored at breast height, and
growth rings were counted.
• Twelve individuals had rings > 75, suggesting that
these trees were present on-site when longleaf pine
was cut in the 1920’s.
•
9 post oaks (Quercus stellata) had ring
counts that ranged from 95-162.
•
2 southern red oaks (Q. falcata) had ring
counts of 86 and 133.
•
1 white oak (Q. alba) had a ring count of
103.
• Ages of these oaks suggest that these species were
likely part of the forest at the time of the Battle of
Horseshoe Bend and warrant inclusion in the upland forest restoration plan.
Conclusions
•
Longleaf persists at HOBE clustered in 3 stands
plus 280+ large trees scattered over the landscape.
• When compared to density of medium
and larger trees reported by Reed 100
years earlier, 3 existing stands support
•
•
similar numbers of longleaf, but loblolly and
and shortleaf pines are over-represented;
there has been invasion of numerous off-site
hardwoods.
• Stands lack smallest and largest longleaf
but support intermediate-sized (35-45
cm DBH) trees.
Residual mature longleaf have the potential to
serve as a seed source and promote forest restoration.
Relict oak trees should be included in the longleaf forest restoration plan for HOBE, however
the density of hardwoods must be greatly decreased to recreate an open-canopied forest.
Re-introduction of fire is necessary to promote
longleaf regeneration and eventually reduce
hardwood stems.
The first burns in over 50 years were applied in
spring 2006; more work is required to ensure fuel
reduction, adequate longleaf regeneration, and
decreases in hardwoods (Hermann et al. 2007).
Acknowledgements
National Park Service staffers are vital partners in the
project & we thank R. Howard, M. Lewis, & C. Noble.
Many people aided with field work. We appreciate efforts
of Auburn University School of Forestry & Wildlife Sciences senior projects (2005: J. Angel, J. McBrayer, R.
Musik, & P. Turner; 2006: C. Brannon, J. Davison, & S.
Partain). Additional assistance was provided by B. Estes,
V. Johnson, C. Newton, G. Sorrell, & J. Waites.
Literature Cited
Boyer, W.D. 1990. Longleaf pine. pp.405-412, In:
Burns, R.M.; Honkala, B.H. (tech. coord.), Silvics of
North America, Volume I Conifers. U.S. Department
of Agriculture, Forest Service Agriculture Handbook
654.
Hermann, S.M. & J.S. Kush, 2006. Assessment of restoration potential of residual stands of mountain
(piedmont) longleaf pine at Horseshoe Bend National
Military Park. Pgs 39-42 in (M.L. Cipollini, comp.)
Proceedings of Second Montane Longleaf Pine Conference Workshop; Longleaf Alliance Report No. 9
Hermann, S.M., J.C. Gilbert, J.S. Kush, C. Noble, and H.
Jerkins. 2007. Longleaf Pine Forest Restoration at
Horseshoe Bend National Military Park: Evaluation
of Residual Stands and Re-Introduction of Fire. In
Estes, B.L. and Kush, J.S Proceedings of the Sixth
Longleaf Alliance Regional Conference; November
13-16, 2006, Tifton, GA. Longleaf Alliance Report
No. 10.
Reed, F.W. 1905. A working plan for forest lands in
central Alabama. USDA Forest Service Bulletin No.
68.
Poster Presentations 94
A Container-Grown Seedling Quality DVD
Mark J. Hainds1, Elizabeth Bowersock2, and Dean Gjerstad2
1
2
The Longleaf Alliance, Solon Dixon Center, Andalusia, AL 36420, USA
School of Forestry and Wildlife Sciences, Auburn University, Auburn, AL, 36849, USA
Abstract
The Longleaf Alliance has produced a 44 minute DVD
on container-grown seedling quality. Numerous classes
or grades of seedlings were outplanted across the Southeastern US (VA, SC, AL, & GA) and tracked over time.
Information on grading and identification of various
seedling types (hybrids, doubles, willows, spiraled
roots, etc.) and effects on growth and survival is included in this DVD. This DVD is the first issued from
a four-part series of DVDs with subsequent issues to
cover site preparation, planting, and herbaceous release.
This DVD is the first of its kind. After viewing this product,
the audience should be able to easily identify various defects and characteristics of good and poor quality containergrown longleaf pine seedlings. Besides helping the audience
to identify seedling classes, grades, or defect, this DVD
provides:
•
Helpful remedies for doubles (Figure1, 3, 4) and wil
lows (Figure 2) Information on seedling care and stor
age.
A typical shipment of container-grown longleaf seedlings will contain many different grades and/or types of
seedlings (Table 1). Having the ability to identify the
various grades and defects of your container-grown
longleaf is a critical to assessing the over-all quality of
your purchase.
•
Seedling growth and survival over time across many
study sites in the Southeast (Table 1) (Figure 5)
•
The cost and benefits of planting or culling various
seedling grades, classes, or defects (Figure 6,7,8).
Figure 1. Two live seedlings in 1 plug (doubles)
Figure 2. When are willows
not problematical?
Figure 3. How should doubles be treated?
Poster Presentations 95
Figure 4. How does dividing
compare to clipping?
Table 1: Typical examples of seedling grades
from three different shipments.
100
80
60
Singles
40
Doubles
20
0
Samson
D. Ridge
Milledgev ille
Figure 5. % Survival Good vs. Doubles on 3 Sites
(age 3 Samson & D. Ridge, age 2 Milledgeville)
Nursery & Good/ Double Floppy Willow Hybrid Other
# Graded Target
(in
plug)
(X2 &
1 cut
off)
Nursery
3,282
45
64
101
13
19
A-2006
(93.1%) (1.3%) (1.7%) (3.1%) (0.4%) (0.6%)
(3,524)
Nursery B 3,209 257
184
20
0
-2005
(87.4%) (7.0%) (5.0%) (0.5%) (0%)
(3,670)
0
(0%)
91
149
4
6
7
Nursery B 3231
(92.6%) (2.6%) (4.3%) (0.1%) (0.2%) (0.2%)
-2002
(3,488)
Figure 6. Are there any hybrid seedlings
in this photo?
Subsequent DVDs to be issued by The Longleaf Alliance
will include:
• Planting
• Site Preparation
• Herbaceous Release.
Figure 7. What are the characteristics of a good plug and root system?
Figure 8. What are the two most dangerous defects in a container-grown
seedling?
Poster Presentations 96
Longleaf Pine Forest Restoration at Horseshoe Bend National Military Park: Evaluation of
Residual Stands and Re-Introduction of Fire
Sharon M. Hermann1, John C. Gilbert2, John S. Kush2, Caroline Noble3 and Herbert “Pete” Jerkins4
1
Auburn University Department of Biological Sciences, Auburn, Alabama, 36849, USA
Auburn University School of Forestry & Wildlife Sciences, Auburn, Alabama, 36849, USA
3
US Department of Interior, National Park Service, Tallahassee, Florida, 32312, USA
4
U.S. National Park Service, Cumberland Gap National Historic Park, Middlesboro, Kentucky, 40965, USA
2
Abstract
Forest restoration often encompasses two major activities: planting of dominant tree species and reintroduction of ecological processes. Appropriate areas
for planting as well as position of residual trees must be
factored into the plan. Re-introduction of fire following
fire suppression is also often a necessary part of restoration for longleaf pine forests. Application of this ecological process is relatively well-understood when forests have been frequently burned. However, reintroduction of burning in fire-suppressed stands requires special attention.
Horseshoe Bend National Military Park (HOBE) is in
initial stages of longleaf restoration. Silvicultural information on longleaf regeneration, density, and location
of existing trees will be used to determine the areas that
have a reasonable likelihood of recovery via natural
regeneration and to identify sites where hand planting
may be required. In 2006, the National Park Service
initiated prescribed burning as the first step in reduction
of fine fuel. A major challenge is how to remove excessive litter and duff and, at the same time, not destroy
mature longleaf with high value as future sources of
seed. Initial burns appear to be successful and removed
much of the litter. An immediate challenge will be to
check excessive sprouting observed on top-killed hardwood stems. Future fire management must target careful removal of duff. Current thought is that duff at the
base of many trees may remain for the near future and
that areas away from adult trees most suitable for natural regeneration may be the most important sites to target for duff removal.
general types ofactivities are being pursued: 1) evaluation
of residual longleaf to determine if supplemental planting is
necessary and 2) re-introduction of fire to remove litter/duff
and reduce stems of encroaching hardwood species. Important goals are reduction of the current off-site hardwood
species and enhancement longleaf pine. This open forest
was described in 1814, during the Battle of Horseshoe
Bend, and persisted near HOBE until the early 1900s (Reed
1905). Ultimately, the hope is that prescribed fire will be
the primary land management tool in the uplands.
Assessment of Residual Longleaf
(Hermann and Kush 2006, Gilbert et al. 2007)
•
•
•
Initial survey work documented ~ 800 longleaf,
including 280 large trees > 15 inches (38 cm) dbh.
Residual longleaf stands currently lack regeneration but the 280 larger trees are potential seed
sources.
To determine areas appropriate for future regeneration, soil type must be considered.
Soil Series Supporting Longleaf at HOBE
•
•
All large longleaf were mapped (sub-meter GPS)
and locations superimposed on HOBE soils map
(NRCS 2006).
Larger longleaf currently are found on 7 soils series; these soils provide an estimate of appropriate
areas to focus longleaf restoration efforts.
Seed Dispersal Distance
Introduction
Horseshoe Bend National Military Park (HOBE), located in Tallapoosa County Alabama, was once dominated by longleaf pine forest, but currently supports
only residual trees of this once common upland tree (see
paper by Gilbert et al. in this publication). Presence of
adult longleaf indicates that 75-100 years ago fire was
frequent. However the National Park Service (NPS) has
no records of wide-spread fire in more than 50 years.
The agency is in the preliminary stages of planning and
implementing ecosystem restoration at HOBE. Two
•
•
Dispersal distance determines the amount of appropriate area that has potential to be colonized by
natural regeneration.
Boyer (1963) documented that ~ 75% of all seed
falls within 20 m of second-growth trees and little
or no seed is found > 60 m. Grace et al. (2004)
suggested maximum dispersal distances in oldgrowth stands may exceed this but such distance is
unlikely for the shorter stature, second-growth
longleaf at HOBE.
Poster Presentations 97
•
We predict that central regions of gaps with radius
> 40 m will be devoid of longleaf for many decades in the future, even after appropriate exposed
soil conditions are created to promote natural regeneration.
•
Post-burn evaluation is on-going but to date few
hardwoods appear top-killed. However many
new basal sprouts were observed over the summer; 3-8 times the number present prior to fire.
Conclusions
Restoration Potential Based on Dispersal Distance of
Residual Trees
•
•
•
Relying on residual trees for regeneration is desirable. Planting in rocky soil will be difficult, even
if hand-dibbled, containerized seedlings are used.
In addition, seed from residual trees does not introduce off-site genetic material.
To determine if reliance on natural regeneration is
feasible over large areas at HOBE, dispersal distance of 95% of seed for each large tree was plotted on soil series currently supporting longleaf.
Although successful forest restoration does not
require fully stocked stands, a large area devoid of
longleaf is not desirable. In addition, fire management is likely to be difficult in large areas with no
longleaf.
•
•
•
•
Current Forest Composition and Fine Fuel
•
•
Following 50+ years of no burning, recent assessment of vegetation revealed many off-site hardwood stems (Hermann and Kush 2006, Gilbert et
al. 2007)are excessive. Fine fuel at the base of
residual large longleaf is in the range of 12 cm (4
in) deep and is of special concern.
•
•
Recent Fire at HOBE
•
•
1st prescribed fire occurred 3 April 2006. This is
later than recommended for burns in excessive
fuels; preferred dates = Dec 15 – Feb 15. However
careful application resulted in good initial results.
• Burn weather = Temp ~ 270C (800F), RH
~ 30 %, light wind
• NPS burn crew used hose lay to apply ~
200L (50 gal) water at base of many large
trees prior to fire.
• Fast-moving flanking strip fires minimized residence time near large trees.
Rapid and thorough mop-up targeted large longleaf
trees.In one burn unit, much of surface litter was
removed but little was burned near large trees.
• When large trees charred, effective mopup minimized duff consumption.
• No needle scorch was observed post-burn;
light wind during fire may have minimized this result.
•
•
Natural regeneration from existing trees has potential to enhance restoration efforts if the excessive litter and duff eliminated.
However, natural regeneration will not be sufficient for recovery of longleaf over the landscape
of HOBE; many areas on appropriate soil fall
outside the seed dispersal range of existing adult
trees.
Initial burns appear to produce desired results but
multiple fires are necessary to eliminate fuel.
Time-consuming application of water appears to
have minimized damage to residual trees but this
treatment was deemed too expensive for continued use. Alternative treatments and/or burn prescriptions that remove litter but not duff will be
explored.
Ultimate fire effects of the first burns currently
are unknown; Kush et al. (2004) reported mortality of larger trees occurring 2-4 years after reintroduction of fire into long-unburned longleaf
stands.
A pragmatic approach may be to permit some
duff to remain at base of large trees and focus
fire efforts on duff removal in areas away from
trees, in potential regeneration gaps.
Additional burns and related monitoring are
needed to determine if fire will decrease hardwoods and non-long pines (Hermann and Kush
2006, Gilbert et al. 2007). However, some scattered hardwoods, especially post oak (Quercus
stellata) and southern red oak (Q. falcata),
should be retained.
Short-term mechanical and/or chemical treatments may be required to restore forest composition.
At HOBE, decline of the longleaf forest occurred
over many decades. Residual, mature trees have
the potential to shorten the restoration period but
success will require time and persistence.
Poster Presentations 98
Acknowledgements
NPS staffers are vital partners in the project and we thank
J. Cahill, R. Howard, M. Lewis, and L. McInnis. Many
people contributed to the field work. We appreciate efforts of Auburn University School of Forestry and Wildlife Sciences senior projects (2005: J. Angel, J.
McBrayer, R. Musik, and P. Turner; 2006: C. Brannon, J.
Davison, and S. Partain). Additional field assistance was
provided by G. Sorrell, B. Estes, V. Johnson, C. Newton,
D. Tenaglia, and J. Waites.
Literature Cited
Boyer, W.D. 1963. Longleaf pine seed dispersal. US
Forest Service Research Note SO-3. USDA, New
Orleans, LA.
Gilbert, J.C., S.M. Hermann, J.S. Kush, L. McInnis, and
J. Cahill. 2007. Longleaf Pine Re-Discovered at
Horseshoe Bend National Military Park. In Estes,
B.L. and Kush, J.S Proceedings of the Sixth Longleaf
Alliance Regional Conference; November 13-16,
2006, Tifton, GA. Longleaf Alliance Report No. 10.
Grace, S.L., J.L. Hamrick, and W.J. Platt. 2004. Estimation of seed dispersal distance in an old-growth population of longleaf pine (Pinus palustris) using maternity exclusion analysis. Castanea 69(3):207-215.
Hermann, S.M. and J.S. Kush, 2006. Assessment of restoration potential of residual stands of mountain
(piedmont) longleaf pine at Horseshoe Bend National
Military Park. Pgs 39-42 in (M.L. Cipollini, comp.)
Proceedings of Second Montane Longleaf Pine Conference Workshop; Longleaf Alliance Report No. 9.
Kush, J.S., R.S. Meldahl, and C. Avery. 2004. A restoration success: longleaf pine seedlings established in a
fire-suppressed, old-growth stand. Ecological Restoration 22:6-10.
NRCS. 2006. Soil Survey Tallapoosa County, Alabama.
http://websoilsurvey.nrcs.usda.gov/app/.
Reed, F.W. 1905. A working plan for forest lands in
central Alabama. USDA Forest Service Bulletin No.
68.
Poster Presentations 99
What Happens to Top-Killed Seedlings?
Rhett Johnson1 and Mark J. Hainds1
1
The Longleaf Alliance, Solon Dixon Center, Andalusia, Alabama, 36240, USA
Abstract
In the initial study, two-year old longleaf seedlings were
prescribed burned in an operational fire in February of
2004. Top-killed seedlings which exhibited re-sprouting
from the root collar 4 months following the fire were
flagged and re-examined 18 months post fire. All of the
top-killed seedlings were alive18 months after the fire
with 80% in single stem height growth ranging from 0.4
to 2.7 feet tall with a mean of 1.06 feet. The remaining
20 % of the top-killed trees were alive, but had two live
stems and averaged only 0.3 feet in height. Additionally,
seedlings from three different studies: a planting depth
study, a herbaceous release study, and summer planting
study, were assessed for height, brown spot, and survival
immediately before a prescribed fire, approximately ½
year post-burn, and 1.5 years post burn.
Poster Presentations 100
Effects of Two Native Invasive Trees on the Breeding Bird Community of Upland Pine Forests
Nathan Klaus1 and Tim Keyes1
1
Georgia Non-game Endangered Wildlife Program, Georgia Department of Natural Resources,
Forsyth, Georgia, 31029, USA
Abstract
Land lottery surveys conducted prior to European settlement (circa 1820) reveal an upland pine forest unlike
contemporary forests in the Georgia Coastal Plain and
Piedmont. Specifically two species of tree, water oak
(Quercus nigra) and sweetgum (Liquidambar styraciflua) are entirely absent from all upland sites. We conducted point counts in paired sites with similar pine
stocking and size classes to investigate the effects of
this invasion on breeding bird communities. Bird species richness was 72% (p<0.001) higher and bird abundance was 68% higher (p<0.001) in pine stands that had
not been invaded by sweetgum or water oak. Eleven
pine specialists and early succession species were significantly more common in pine stands that had not
been invaded, and no species had a positive relationship
with sweetgum/water oak invasion, though many generalist
species used this habitat. Partners in Flight (PIF) scores
were summed across all points and an average
‘conservation value’ calculated by habitat type. Stands
without sweetgum or water oak scored a 76% higher conservation value than invaded stands. The sweetgum/water
oak/loblolly forest appears to be a recently emerged unnatural forest type and lacks a distinct bird community. Invasion of pine forests by these tree species substantially lowers the conservation value of a site.
Poster Presentations 101
The Regional Longleaf Pine Growth Study – 40 years old
John S. Kush1 and Don Tomczak2
1
Auburn University School of Forestry and Wildlife Sciences, Auburn, Alabama, 36849, USA
2
USDA Forest Service, Atlanta, Georgia, 30309, USA
Abstract
From 1964 to 1967 the USDA Forest Service established
the Regional Longleaf Pine Growth Study (RLGS) in the
Gulf States. The original objective of the study was to
obtain a database for the development of growth and yield
predictions for naturally regenerated, even-aged longleaf
pine stands. Plots were installed to cover a range of ages,
densities, and site qualities. The plots are inventoried on
a 5-year cycle and are thinned at each inventory, as
needed, to maintain the assigned density level. The study
accounts for growth change over time by adding a new
set of plots in the youngest age class every 10 years. The
project completed its seventh measurement period (35year measurement) in 2002 and started on the eighth
measurement (40-year measurement) in 2004. While the
resource base and public concerns have changed over the
last 40 years, so has the RLGS. In response to changing
questions, the objectives of the RLGS have been broadened. The major focus of the study is the longleaf pine
overstory, but details have been added to enhance our
understanding of stand dynamics. Among the significant
additions and related efforts to the RLGS have been: data
quantifying utility pole production, estimates of litter production, standing biomass, net-primary productivity and
leaf area index have been developed, models developed
from the RLGS have been implemented in the Forest Service’s Forest Vegetation Simulator program and the
RLGS plots are going to be used to determine the relationships between root biomass/carbon sequestration and
the density, site quality, and age of the longleaf pine overstory.
Introduction
project has begun its eighth measurement period (40-year
measurement).
Methods
The study currently consists of 292 1/5-acre and 13 1/10acre permanent measurement plots located in central and
southern Alabama, southern Mississippi, southwest Georgia, northern Florida, and the sandhills of North Carolina.
At the time of establishment, plots are assigned a target
basal area class of 30, 60, 90, 120, or 150 square feet/acre.
They are left un-thinned to grow into that class if they are
initially below the target basal area. Plot selection was
based upon a rectangular distribution of cells formed by: 6
age classes from 20 to 120 years, 5 site quality classes
ranging from 50 to 90 feet at 50 years, and 6 density
classes ranging from 30 to 150 square feet/acre and plots
left un-thinned to see how they grow.
In subsequent re-measurements, the plot is thinned back
to the previously assigned target if the plot basal area has
grown 7.5 square feet/acre beyond the target basal area.
The thinnings are generally of low intensity and are done
from below.
Net measurement plots are surrounded by a similar and
like-treated half-chain wide isolation strip. Plots are inventoried, and treated as needed, every 5 years. The
measurements are made during the dormant season and it
takes 3 year to complete a full measurement of all plots.
Results
In 1964, the USDA Forest Service established the Regional Longleaf Pine Growth Study (RLGS) in the Gulf
States. The original objective of the study was to obtain a
database for the development of growth and yield predictions for naturally regenerated, even-aged longleaf pine
stands. Plots were installed to cover a range of ages, densities, and site qualities. The study accounts for change in
growth by adding a new set of plots in the youngest age
class every 10 years.
In 1984, Auburn University, in a cooperative agreement
with the USDA Forest Service, re-measured the RLGS
plots for its fourth measurement period (20-year measurement). The cooperation continues today and the
Through the 30-year re-measurement, there are 32 publications and numerous presentations that are a direct result
of the RLGS. Another 18 related publications used information from the RLGS. Two computer software programs are available based on RLGS data.
The RLGS represents a stable, long-term database and an
active “field laboratory” for natural, even-aged longleaf
pine stands. The value of this project increases as more
and more ownership's in the South consider longleaf pine
management alternatives. The plots are available for cooperative studies that do not harm the plots or interfere
with future activities.
Poster Presentations 102
Utility Pole Production
Conclusion
Utility pole information is being used to develop relationships between stand characteristics, thinning activities,
and pole production. Across the study nearly 78% of
pole-size trees could be classed as poles.
The RLGS has adapted to changes in the resource base and
shifting public concerns over the last 40 years. The initial
installation in the mid-60’s resulted in 185 sample plots.
This number increased to 267 in 1987 and is now at 325.
As the number of plots have grown, and in response to
changing questions, the objectives of the RLGS have been
expanded. It is no longer meaningful to have growth projection models estimate only to stand-level merchantable
basal area and total volumes in pulp and sawtimber. Users
are demanding more information on multiple products, and
want trees/acre and merchantable volume by DBH classes
to answer their current questions. The RLGS is keeping
pace with ever-changing demands and is proving once again
that well designed, long-term studies are wise research investments.
Global Climate Change
Within this distribution are five time replications of the
youngest age class. All five replications are located on
the Escambia Experimental Forest in Brewton, AL. The
figure below indicates that there has been an increase in
growth among the time series plots located on the Escambia Experimental Forest.
Figure 1. Basal area increment/year (Square feet/acre)
10
Cooperators in the Regional Longleaf Pine Growth
Study
9
8
7
6
5
4
3
2
1
0
1
2
Tim e Se rie s
3
4
Carbon Sequestration
Longleaf pine is the most suited among the southern
pine species for carbon sequestration. It has a long life
span, up to 500 years. It grows on most sites in the
South and is relatively risk free. It has a high specific
gravity making it suited to long term storage in products. Research has shown it does well in an elevated
CO2 environment.
RLGS plots will be used to determine the relationships
between root biomass/carbon sequestration and the density, site quality, and age of the longleaf pine overstory.
Region 8 of the USDA Forest Service
Apalachicola National Forest - Wakulla District
Talladega National Forest - Talladega District
Talladega National Forest - Oakmulgee District
Homochitto National Forest - Homochitto District
DeSoto National Forest - Black Creek District
Conecuh National Forest - Conecuh District
Escambia Experimental Forest (Brewton, AL)
T.R. Miller Mill Company (Brewton, AL)
Florida Forest Service - Blackwater River State Forest
(Munson, FL)
Cyrene Turpentine Company (Bainbridge, GA)
Eglin Air Force Base (Niceville, FL)
Southlands Experimental Forest- International Paper
Company (Bainbridge, GA)
Gulf States Paper Corporation (Columbiana, AL)
Wefel Family Trust (Atmore, AL)
North Carolina Division of Forestry - Bladen Lake
State Forest (Elizabethtown, NC)
Kimberly-Clark Corporation (Weogufka, AL)
Management Perspective on the Regional Longleaf
Pine Growth Study
Managers have been hampered by a dearth of good scientific research to support growth estimates for longleaf
pine. In contrast to loblolly pine, on which ample research is in place to accurately predict growth, longleaf
pine growth has been speculative. The long-term Regional Longleaf Pine Growth Study (RLGS) results
have application in management decisions with both
timber and wildlife objectives. Because RLGS researchers have adapted to the changing needs of managers, study results now go well beyond plantation management and extend to longer rotations, where endangered species management may be a primary objective.
Poster Presentations 103
Chopper® Herbicide Site Prep Improves Quality of Weed Control
Dwight K. Lauer1 and Harold E. Quicke2
1
2
Silvics Analytic, Ridgeway, Virginia, 24148, USA
BASF Corporation, Research and Development Center Research Triangle Park, North Carolina, 27709, USA
Abstract
Two studies were installed to examine weed control and
growth of loblolly pine following different combinations of
site preparation and post-plant herbaceous weed control
(HWC) on cutover sites. Treatments included bedding
only, bedding + HWC, bedding + Chopper site preparation
and bedding + Chopper site preparation + HWC. These
experiments showed that site preparation with Chopper®
herbicide improved control of woody and herbaceous weeds
compared to bedding alone. Third year pine growth following bedding + Chopper herbicide was greater than for bedding + HWC. A synergistic response was observed with the
combination of Chopper® herbicide site prep and HWC in
that pine growth response to the combination was greater
than the sum of responses to individual treatments. This
improved quality of weed control has the potential to improve survival, shorten the grass stage period, and decrease
the woody fuel load at time of the first prescribed burn in
longleaf pine plantations
Introduction
The quality of vegetation control during the regeneration
phase of longleaf pine on cutover sites is important. Competing vegetation can lower survival and delay emergence
from the grass stage. Woody vegetation will decrease longterm growth and may make prescribed burning in established stands more hazardous due to increased fuel levels.
Vegetation control with chemical site prep is an attractive
option to control competing woody vegetation and improve
the quality of herbaceous weed control. Initial control of
herbaceous vegetation afforded by site prep is especially
important with longleaf pine because post-plant herbicide
options are more limited than with loblolly pine. Further, a
high quality site prep treatment may improve the quality of
post-plant herbaceous weed control.
A common management regime for loblolly pine is to use
chemical site prep to provide control of woody and herbaceous vegetation and then use post-plant herbaceous weed
control to extend the duration of herbaceous weed control.
A series of studies was initiated by BASF in 2001 and 2002
to examine the use of Arsenal® AC + Oust® post-plant
herbaceous weed control following Chopper® herbicide site
prep to establish loblolly pine stands. Two locations included a site prep treatment without chemical site prep.
These two locations provide a comparison of herbaceous
weed control quality with and without Chopper® herbicide
site prep. Although these experiments were with loblolly
pine, the findings concerning the quality of weed control
achieved by these different treatments can be adapted for
the establishment of planted longleaf pine as well.
Methods
Integrated systems of site prep and first-year herbaceous
weed control for loblolly pine plantations were examined
at two locations using a randomized complete block design with three replications. The Kings Ferry, FL location
was a lower coastal plain site with a poorly drained clay
soil. The Barnett Crossroads, AL location was an upper
coastal plain site with a moderately well drained clay soil.
The site prep herbicide treatment at both locations was a
tank mix of Chopper® herbicide with 1 or 2 pints Garlon® 4. The March applied first year herbaceous weed
control (HWC) treatment was 4 oz/ac Arsenal® AC + 2
oz/ac Oust®.
Integrated treatments differed by location. Treatments at
Kings Ferry were 1) double bed, 2) double bed + March
HWC, 3) single bed + Chopper® herbicide site prep, and
4) single bed + Chopper® herbicide site prep + March
HWC. Treatments at Barnett Crossroads were 1) single
pass rip and bed, 2) single pass rip and bed + March
HWC, 3) single pass rip and bed + Chopper® herbicide
site prep, and 4) single pass rip and bed + Chopper® herbicide site prep + March HWC.
Vegetation cover was assessed by vegetation type and
taxa in June, August, and October during the first year
after planting. Loblolly pine groundline diameter (gld)
and height were measured at the end of the third growing
season. Loblolly pine growth was compared using year 3
volume index computed as the volume of a cone using gld
and total height.
Vegetation Control
Chopper® herbicide site prep improved both woody and
herbaceous weed control in the first year and improved
the performance of the HWC treatment. At Barnett
Crossroads (Figure 1), total June cover was less than 25%
following Chopper® herbicide with or without HWC.
Total cover increased to only 14% by October for Chopper® herbicide + HWC. HWC without Chopper® herbicide had total cover of 63% by August and almost all of
Poster Presentations 104
this was woody vegetation such as gallberry. At Kings
Ferry (Figure 2), total cover in June for Chopper® herbicide, with or without HWC, and double bed + HWC did
not exceed 26%. By October, total cover was 72% for
double bed + HWC compared to 42% for Chopper® without HWC. The difference was due to woody cover. Total
cover was lowest for Chopper® herbicide + HWC
throughout the year with only 2% cover in woody vegetation in October.
Loblolly Pine Response
At Barnett Crossroads (Figure 3), Chopper® herbicide
performed better than HWC. HWC + Chopper® herbicide increased volume index by 250. This was greater
than the sum of the response from these treatments used
individually (53 + 113 = 166). At Kings Ferry (Figure 4),
Chopper® herbicide performed better than HWC. HWC
+ Chopper® herbicide increased volume index by 322.
This was greater than the sum of the response from these
treatments used individually (99 + 177 = 276). At both
locations, response to Chopper® herbicide without HWC
was greater than the response to HWC.
Third year loblolly pine volume index was greatest for the
Chopper® herbicide + HWC combination at both locations.
100
C o ver ( % )
80
None
60
HW C
40
Chopper
20
Chopper+HW C
0
June
August
October
Figure 1. Percent total vegetation cover during the first growing season at Barnett Crossroads, AL. All
site prep treatments included a single pass rip and bed. Treatments are designated as None for no further
treatment, Chopper for Chopper® herbicide site prep, HWC for first year March herbaceous weed control.
The woody cover component of total cover in October was 47%, 60%, 4%, and 6% for the treatments
None, HWC, Chopper, and Chopper + HWC, respectively.
100
C o v e r (% )
80
Double Bed
60
Double Bed+HWC
40
Chopper
Chopper+HWC
20
0
June
August
October
Figure 2. Percent total vegetation cover during the first growing season at Kings Ferry, FL. Chopper® herbicide site prep treatments were combined with single bedding. Treatments are designated as Double Bed for
double bedding, Chopper for Chopper® herbicide site prep, HWC for first year March herbaceous weed control. The woody cover component of total cover in October was 8%, 35%, 2%, and 2% for the treatments Double Bed, Double Bed + HWC, Chopper, and Chopper + HWC, respectively.
Poster Presentations 105
Stem Volume Index
350
300
+ 250
250
200
150
100
50
+ 113
+ 53
0
No TRT
MAR HWC Chopper Chopper +
MAR HWC
Stem Volume Index
Figure 3. Year 3 stem volume index for different combinations of site prep and herbaceous weed control at
Barnett Crossroads, AL.
600
500
400
300
200
100
0
+ 322
+ 99
Double
Bed
+ 172
Double
Single
Bed +
Bed +
MAR HWC Chopper
Single
Bed +
Chopper
+ MAR
HWC
Figure 4. Year 3 stem volume index for different combinations of site prep and herbaceous weed control at Kings
The quality of first year herbaceous weed control is compromised without proper site preparation. In these studies,
Chopper® herbicide site prep outperformed HWC in terms
of vegetation control and pine growth. HWC with only
mechanical site prep released competing woody vegetation.
The best performance was from the combination of Chopper® herbicide site prep + HWC. These results are consistent with longer-term studies (Miller et al. 2003) that demonstrate the importance of woody vegetation control at site
prep and the limited response to herbaceous weed control if
woody vegetation is not controlled.
Acknowledgements
Many thanks to Rayonier Forest Resources L.P. for providing study sites.
Herbicide Notes
Chopper® and Arsenal® are registered trademarks of
BASF Corporation.
Oust® is a registered trademark of Dupont.
Garlon® is a registered trademark of Dow Agrosciences, LLC.
Literature Cited
Miller, J.H., B.R. Zutter, S.M. Zedaker, M.B. Edwards,
R.A. Newbold 2003. Growth and yield relative to
competition for loblolly pine plantations to midrotation – A Southeastern United States regional
study. Southern Journal of Applied Forestry 27(4):
237-251.
Poster Presentations 106
Red-Cockaded Woodpecker Recovery and Longleaf Pine Ecosystem Conservation: Sharing
and Selling the Success through the Eyes of the Advocates
Jon Marshall1, Ralph Costa2, John Maxwell1 and Dave Case1
1
2
DJ Case & Associates, Mishawaka, Indiana, 46545, USA
U.S. Fish and Wildlife Service, Department of Forest Resources, Clemson University, Clemson, South Carolina,
29634, USA
Abstract
During the past 15 years, significant progress has been
made on recovery of the red-cockaded woodpecker
(RCW) and restoration and conservation of longleaf pine
forests. This progress can be measured in numbers of
new RCW groups, new acres planted in longleaf pine and
acres of existing longleaf pine improved, and importantly,
number of new partnerships focused on these efforts.
Although the success of RCW population and longleaf
pine restoration has been relatively recent, the foundation
for today’s success began many decades ago. Numerous
foresters, biologists, researchers, and administrators in the
private, state and federal sectors have made significant
contributions over those decades toward RCW and longleaf pine conservation. Their contributions have provided
ecological understanding, management direction, policy
solutions, economic incentives, cultural and social acceptance, and critically, a vision for the mission. We have
conducted 30-60 minute video interviews of approximately 40 of these key players. We have also filmed
about 30-40 hours of longleaf forest habitats (including
associated flora and fauna) and management practices
(e.g., logging, prescribed fire) throughout the southeast
and RCW conservation activities (e.g., capturing, banding, and translocating birds). Additionally, we took
over 500 production quality digital photographs of longleaf
habitats and management activities and RCW conservation
practices. The purpose of our project is to produce communications and outreach tools and materials using the interviews, longleaf/RCW video and photographs. The raw
video and photographs will be used in a variety of ways.
First, DJ Case & Associates, in cooperation with the U.S.
Fish and Wildlife, and attendees of the Longleaf Legacy
Forum, will produce several finished products, e.g., different length videos, to be used to inform and educate specific
target audiences about the value and importance of longleaf
pine ecosystems and RCWs. Second, still photographs and
video footage will be made available for use by conservation partners to produce customized videos, power point
presentations, and other educational materials, e.g., brochures, for selected constituencies. These constituents
could include, but will not be limited to, schools, trade organizations, congressional staff, media outlets, civic groups,
conservation organizations, and private landowners.
Through these focused efforts and venues it is our objective
to educate, enlighten, engage and empower more people to
become concerned about and interested in the restoration
and conservation of longleaf pine forests and their fauna
and flora.
Poster Presentations 107
Pathogenicity of Leptographium serpens to Longleaf Pine
George Matusick1, Lori Eckhardt1 and Scott Enebak1
1
School of Forestry and Wildlife Sciences, Auburn University, Alabama, 36849, USA
Abstract
Decline and mortality syndromes of southern pine species
have attracted an increasing amount of attention from
land managers over the past decade. The southern pine
decline disease syndrome is characterized by complex
interactions between several biotic and abiotic factors
including root inhabiting fungi from the genera Leptographium, Grosmannia and Ophiostoma. Leptographium
serpens is known to be associated with mortality of southern pines as part of the decline syndrome. The importance of longleaf pine to the southern pine ecosystem has
been realized by many as its native range has decreased
over the past
century. Longleaf pine with roots colonized by either L.
serpens or G. serpens have shown wilting symptoms preceding the death of the healthy foliage. Below-ground
symptoms are more apparent and often include resinous
lesions precluding dead roots, a distinct streaking through
the root flair, and a deterioration of fine root mass. A
more direct measure of pathogenicity of this pathogen on
longleaf pine will establish the exact relationship between
pathogen and host. The objectives of the project include
the measurement of pathogenicity of L. serpens in relation
to longleaf pine seedlings and mature roots.
Poster Presentations 108
An Economic Model for Multiple-Value Management of Longleaf Pine
B.B. McCall1, R. K. McIntyre2, S. B. Jack2, and R. J. Mitchell2
1
2
Larson & McGowin, Inc., Mobile, Alabama, 36603, USA
Joseph W. Jones Ecological Research Center, Newton, Georgia, 39870, USA
Abstract
There is growing interest in multiple-value forest management, particularly among private landowners. In addition to timber, many of these landowners may place
equal or greater value on wildlife habitat, recreational
use, aesthetics, and long-term asset appreciation. To balance multiple objectives, a low intensity silvicultural system featuring selection harvest and natural regeneration
may be most appropriate. However, little information
exists about the economic aspects of these integrated
management goals.
An interactive Excel spreadsheet model was developed
with initial data based on a standard 10% timber inventory of a moderately stocked 941 acre tract at the Joseph
W. Jones Ecological Research Center. Growth rates were
estimated from site-specific studies and applied to the
stand table to move trees through diameter classes over
time. Alternating halves of the tract were harvested using
single-tree selection every five years for 50 years. Three
harvest intensity scenarios were modeled to estimate timber revenues and a hunting lease value was factored into
the revenue stream. Expense assumptions were based on
common costs for management practices in the region.
The following values are presented from lower to higher
levels of harvest intensity, with figures in real values adjusted for inflation. Average present value ranged from
$1,428,903 to $1,721,283, with 20 year internal rates of
return from 3.58%to 3.86%. Accumulated cash flow
ranged from $2,309,050 to $5,499,282, with ending values from $4,907,850 to $3,118,557. Initial stocking was
4,672 mbf, with ending stocking ranging from 6,008 mbf
to 2,956 mbf.
This model is in the early stages of development and is
intended as a heuristic rather than a predictive tool. The
assumptions incorporated in the model represent our best
current information but can benefit from further refinement and validation. This modeling approach has also
pointed out information gaps and areas of needed research, such as growth and yield for multiage natural
longleaf pine stands, regeneration dynamics, and ingrowth
rates. Further refinement and development of models
such as this can help landowners understand trade-offs
and make informed decisions when balancing multiple
land management objectives.
Poster Presentations 109
Spatio-Temporal Patterns of Forest Structure and Understory Species
Composition in Longleaf Pine Flatwoods along Florida's Gulf Coast
George L. McCaskill1 and Shibu Jose1
1
School of Forest Resources, University of Florida, Gainesville, Florida, 32611, USA,
Abstract
Little information exists on spatio-temporal patterns of
overstory and understory characteristics in longleaf pine
flatwoods communities along the lower coastal plain of
the Gulf coast. Three representative sites along a spatial
gradient from Pensacola to Tampa Bay (720 km), each
containing stands of three distinct successional stages
(age groups) were used in this study. The different successional stages represented a chronosequence of 120
years applied across the spatial gradient. Stand structure
and plant species composition were measured using four
stands within each pre-defined age group at each site.
After establishment, understory species richness and diversity declined as stand age increased, but species redundancy and web complexity increased. While pioneer
(weedy) species dominated the early successional stages
of stand development, the understory species composition
resembled that of typical flatwoods communities within
thirty years of age. It appears that longleaf pine flatwoods
stands reach steady state equilibrium in terms of species
richness and diversity at a threshold age of ninety years
after establishment.
While overstory tree density decreased with stand age,
stand basal area increased as expected. The volume of
standing dead and downed coarse woody debris also increased with stand age, depending on fire history.
Poster Presentations 110
Tale of Two Forests: Light Environments in Slash and Longleaf
Pine Forests and Their Impact on Seedling Responses.
J.D. McGee1, R.J. Mitchell1, S.D. Pecot1, J.J. O'Brien2, L.K. Kirkman1, and M.J. Kaeser1
1
J.W. Jones Ecological Research Center, Newton, GA, 39870, USA
2
USDA Forest Service, Athens, GA, 30602, USA
Abstract
There is growing interest in multiple-value forest management, particularly among private landowners. In addition
to timber, many of these landowners may place equal or
greater value on wildlife habitat, recreational use, aesthetics, and long-term asset appreciation. To balance multiple
objectives, a low intensity silvicultural system featuring
selection harvest and natural regeneration may be most appropriate. However, little information exists about the economic aspects of these integrated management goals.
The following values are presented from lower to higher
levels of harvest intensity, with figures in real values adjusted for inflation. Average present value ranged from
$1,428,903 to $1,721,283, with 20 year internal rates of
return from 3.58%to 3.86%. Accumulated cash flow
ranged from $2,309,050 to $5,499,282, with ending values from $4,907,850 to $3,118,557. Initial stocking was
4,672 mbf, with ending stocking ranging from 6,008 mbf
to 2,956 mbf.
An interactive Excel spreadsheet model was developed with
initial data based on a standard 10% timber inventory of a
moderately stocked 941 acre tract at the Joseph W. Jones
Ecological Research Center. Growth rates were estimated
from site-specific studies and applied to the stand table to
move trees through diameter classes over time. Alternating
halves of the tract were harvested using single-tree selection
every five years for 50 years. Three harvest intensity scenarios were modeled to estimate timber revenues and a
hunting lease value was factored into the revenue stream.
Expense assumptions were based on common costs for
management practices in the region.
This model is in the early stages of development and is
intended as a heuristic rather than a predictive tool. The
assumptions incorporated in the model represent our best
current information but can benefit from further refinement and validation. This modeling approach has also
pointed out information gaps and areas of needed research, such as growth and yield for multiage natural
longleaf pine stands, regeneration dynamics, and ingrowth
rates. Further refinement and development of models
such as this can help landowners understand trade-offs
and make informed decisions when balancing multiple
land management objectives.
Poster Presentations 111
Linking State Prescribed Fire Councils as a Coalition: A Proposal to Promote Media and
Public Understanding of Rx Fire, and to Nationally Address Key Management, Policy, and
Regulatory Issues
Mark A. Melvin1, Johnny Stowe2, Frank Cole3, Lane Green4, Scott Wallinger5, and Lindsay Boring1
1
J.W. Jones Ecological Research Center, Newton, Georgia, 39870, USA
South Carolina Department of Natural Resources, Columbia, South Carolina, USA
3
For Land’s Sake, Thomasville, Georgia, 31799, USA
4
Tall Timbers Research Station and Land Conservancy, Tallahassee, Florida, 32312, USA
5
Forest Sustainability, Seabrook Island, South Carolina, 29445, USA
2
Abstract
•
The rural southern United States is experiencing rapid
changes in land use and demographics, with increased
challenges for landowners and managers of public and
private lands to conduct prescribed burning of pine woodlands and other pyric ecosystems. Across the country
there are common issues including public safety, ecological stewardship, liability, public education, and air quality
related regulations. Networking the organizations and
efforts together within the South, West, and other regions
that utilize prescribed fire will increase communication,
effectiveness of public education, and especially participation in fire policy decisions and regulatory outcomes.
While Florida pioneered the establishment of regional fire
councils, active or startup organizations are now emerging in most Southern states and several Midwestern and
Western states. A diverse group of private, public and
non-governmental leaders has reviewed the opportunities
for establishing a coalition of fire councils, as well as for
the need to initially examine the science and management
context for the new EPA particulate matter emission standards (PM2.5), which may place considerable new constraints upon land managers to achieve their prescribed
burning goals.
•
•
•
Developing Concerns on 2.5 Micron Particulate Matter (PM2.5) Regulations Intended to Improve Urban
Air Quality
•
•
•
•
Introduction
Prescribed fire managers across the nation continue to
face new and increasing challenges that limit or threaten
the use of prescribed fire. Creating a coalition of prescribed fire councils can prove instrumental in sharing
strategies, technology transfer, uniting on initiatives, and
public education.
Current trends that are cause for concern:
•
•
•
Loss of burning as a key element of rural culture
More roads and increased traffic in rural areas
Landowner constraints based on liability concerns
Lack of consulting practitioner’s capacity due in part
to liability concerns
Extended rural-urban interface zones as urban areas
push into the countryside
Increasing limitation of “burn days” due to environmental regulation originally intended to clean-up urban air quality
General lack of public understanding pertaining to the
role of fire in sustaining forest ecosystems –
“Smokey Bear” syndrome
•
•
•
High sulfate PM2.5 results from burning fossil fuels
(e.g. automobiles, diesel engines, and coal burning
power plants)
Smoke from prescribed burning isn’t a major cause of
emissions but its chemistry fits some of the PM2.5
pattern
New PM2.5 standards will put more urban areas into
“non-attainment” air quality status and apply pressure
to reduce emissions from fossil fuel burning
Most state air quality regulators will also focus upon
at least monitoring prescribed fire and some may attempt to limit prescribed fire activities
Dealing with PM2.5 must be state-by-state as it is
implemented and there is a need to ensure every state
has excellent information and plans
All councils should promote a “categorical exclusion” status from PM2.5 rules from the EPA
Negative news press based on poor information could
shape urban public opinion negatively about prescribed fire. We must promote the positive messages
about how it is used to protect human health and
safety by reducing harmful wildfires
Poster Presentations 112
What Can a Coalition Offer?
•
•
•
•
•
•
•
Bring active councils together to collaborate and begin
to work effectively
Seek to strengthen newly developing councils to make
them more effective players and allies
Promote councils in key states where they don’t exist
now
Engage other organizations that can play key roles,
(e.g. state forestry associations, forest landowner associations, NGO’s, state and federal agencies)
Create a “playbook” of exactly how to deal with state
environmental agencies related to PM2.5
Ensure national-level coordination with federal agencies
Promote public understanding of prescribed fire
Where Do We Go From Here?
1) Explore forming a core group of interested parties to
create a coalition of councils to effectively deal with
issues that impede the use of prescribed fire, which
promotes ecological function, public health and safety,
biological diversity, and the prevention of catastrophic
wildfires.
2) Place a high priority on encouraging and assisting newstates to create councils.
3) Secure funding and support to launch media campaign
to educate the public on the use, need, and forest health
values of prescribed fire and to differentiate between
prescribed fire vs. wildfire.
4) Schedule a conference in 2007 to broaden depth of
group to encompass the entire nation and move forward
with initiatives at a federal level.
Acknowledgements
The following organizations are playing a key role in the
advocacy of prescribed fire in forests: American Forest
Council, Environmental Defense, Georgia Forestry Commission, Joseph W. Jones Ecological Research Center,
Longleaf Alliance, The Nature Conservancy, Tall Timbers
Research Station, US Fish and Wildlife Service, The Wilderness Society, and North Florida, South Carolina and
Southwest Georgia Prescribed Fire Councils.
Poster Presentations 113
Longleaf Pine Genetics Research at the Harrison Experimental Forest
C.D. Nelson1, L.H. Lott1, J.H. Roberds1, T. L. Kubisiak1, and M. Stine2
1
Southern Institute of Forest Genetics, USDA Forest Service, Southern Research Station, Harrison Experimental Forest,
Saucier, Mississippi, 39574, USA
2
School of Renewable Natural Resources, Louisiana State University, Baton Rouge, Louisiana, 70803, USA
Introduction
Genetic and pathology studies of brown spot disease resistance and genetic studies of early height growth in
longleaf pine (Pinus palustris Mill.) have been ongoing at
the Harrison Experimental Forest (HEF, Saucier, MS) for
several decades. Fungicide/clay root-dip treatments were
formulated to provide brown spot control for several
years following outplanting. Genetic variation in brown
spot resistance was also quantified and utilized in advanced generation breeding. Recent analyses have shown
that resistance has been doubled in two-generations of
selection. A pioneering study into the genetic control of
several traits of longleaf pine was established in 1960 by
E.B. Snyder. This experiment, monitored for more than
40 years, is providing valuable information on genetic
variation in mature tree growth traits and resin properties.
The resin data are of particular interest to our collaborative research on mechanisms of resistance to southern
pine bark beetle. Finally, we have been hybridizing longleaf to slash pine in an effort to discover specific genes
that control the grass-stage trait and to develop a variety
of longleaf pine that initiates height growth in the first
year. Our extended abstract provides a brief summary of
each of these research areas.
Control of Brown Spot Needle Blight
Genetics of Brown Spot Resistance and Early Height
Growth
Improved silvicultural practices, including nursery practices, have provided a step-change in our ability to establish productive longleaf pine plantations. However, our
long-term interest is and has been in developing disease
resistant and fast growing longleaf pine populations
through genetic selection and breeding (Snyder & Derr
1972). We have found brown spot resistance and early
height growth to be under genetic control and have produced a second generation population of parents that
should produce seedlings with twice the resistance as the
first generation (Nelson et al. 2006).
Genetics of Life History Growth and Resin Yield
A pioneering study of genetic variation was established in
1960 by E.B. Snyder. Thirteen trees were randomly selected from natural forests on the HEF and mated to each
other in a full-diallel design (i.e., each parent mated in
both directions to all other parents plus 13 selfed matings,
resulting in 169 full-sib families). Progeny of 143 of
these matings were planted in replicated trials located 2
miles apart on the HEF (image below, Fig. 2). Early
measurements were used to estimate heritability (i.e., proportion
A.G. Kais (Kais 1981; Kais and Griggs 1984) developed
root-dip treatments (fungicide & clay slurry) that provide
control of brown spot needle blight (caused by (Scirrhia
acicola (Dearn.) Siggers)) for
several years following outplant- Figure 1. Susceptible and resistant longleaf pine families treated and untreated for
ing. The combination of high control of brown spot needle blight.
quality nursery stock and excellent weed control allows longleaf
pine to consistently emerge from
Resistant family
the grass stage in the second or
third year— greatly improving
the opportunity for plantation
establishment. The images below (Fig. 1) show treated and
untreated plots of the same longUntreated
leaf pine families in a field test Treated
on the HEF. Some families
Susceptible family
show genetic resistance and are
not appreciably benefited by the
Treated
Untreated
fungicide treatment.
Poster Presentations 114
of variation under genetic control) for several growth and
morphological traits (Snyder & Namkoong 1978). Recent
interest centers on changes in genetic variability for growth
as trees age as well as genetic variation in wood properties
and in oleoresin yield (i.e., resin flow), a trait believed to be
a major component of defense against bark beetle attack.
Since longleaf pine is highly resistant to colonization by
southern pine beetle and has high resin flow, it is of interest
to determine whether it also has low genetic variability for
this trait. Preliminary results (Table 1) for growth traits
show that heritability in these trials are relatively low, but
do vary over ages (Stine et al. 2001).
Table 1. Heritability (individual tree, narrow-sense) for
growth traits in longleaf pine.
Trait
Age 7
Age 17
Age 40
Height
0.13
0.06
0.15
DBH
0.07
0.14
0.11
Interspecies Hybrid Breeding and Mapping Genes for
Early Height Growth
Crosses between longleaf pine and slash pine (P. elliottii
var. elliottii) or loblolly pine (P. taeda) produce progeny
with intermediate early height growth. Backcrossing
these hybrids to both parental parents and analyzing the
means and variances of these populations can provide an
estimate of the number of genes controlling the trait
(Nelson et al. 2003). Our current estimate is relatively
small (on the order of less than 10), which suggests that a
backcross breeding program with the assistance of DNA
markers and accelerated breeding could produce a variety
of >98% longleaf that essentially lacks the grass-stage
(Fig. 3). This variety could prove very useful for plantation management of longleaf pine.
Longleaf diallel,
planted in 1960
Figure 2. Longleaf pine diallel planting, established on
the Harrison Experimental Forest in 1960.
Longleaf
X
100:0
Timeline w/
DNA markers &
topgrafting
Longleaf
100:0
Longleaf
100:0
2005-06
X
X
Slash
0:100
F1 hybrid
50:50
Backcross 1 progeny (BC1)
On average 75:25
Select trees w/ height growth and larger
proportion longleaf, based on DNA markers
For example 85% longleaf (85:15).
2010-11 Backcross 2 progeny (BC2)
On average w/out DNA selection 87.5:12.5
However, w/ DNA selection expect 92.5:7.5
Repeat to produce BC3 w/ >98% longleaf
Intercross BC3 to produce new variety
2015-16
Figure 3. Interspecies backcross breeding plan and
timeline for incorporating early height growth genes
into longleaf pine.
Literature Cited
Kais, A.G. 1981. Longleaf pine production-- a cooperative adventure. In: Proc. South. Nursery Conf.,
Sept. 2-4, 1980, USDA Forest Service Tech. Pub.
SA-TP-17: 73-85.
Kais, A.G. and M. Griggs. 1984. Control of brown spot
needle blight on longleaf pine through benomyl
treatment and breeding. In: Proc. IUFRO Conf.
Conifer Needle Diseases, Oct. 14-18, 1984, Gulfport, MS: 15-19.
Nelson, C.D., L.H. Lott, and D.P. Gwaze, 2006. Expected genetic gains and development plans for two
longleaf pine third-generation seedling seed orchards. In: Proc. 28th South. Forest Tree Improvement Conf., June 20-23, 2005, Raleigh, NC: 108114.
Nelson, C.D., C. Weng, T.L. Kubisiak, M. Stine, and
C.L. Brown. 2003. On the number of genes controlling the grass stage in longleaf pine. Journal of Heredity 94(5):392-398.
Snyder, E.B. and H.J. Derr. 1972. Breeding longleaf
pines for resistance to brown spot needle blight.
Phytopathology 62:325-329.
Snyder, E.B. and Namkoong, G. 1978. Inheritance in a
diallel crossing experiment with longleaf pine.
USDA Forest Service Research Paper SO-140, 31
p.
Stine, M., J.H. Roberds, C.D. Nelson, D.P. Gwaze, T.
Shupe, and L. Groom 2001. Quantitative trait inheritance in a forty-year old longleaf pine partial
diallel test. In: Proc. 26th South. Forest Tree Improvement Conf., June 26-29, 2001, Athens, GA:
101-103.
Poster Presentations 115
Long Term Research on the Effects of Fire Regime on Upland Longleaf Pine Forests
Thomas E. Ostertag1 and Kevin M. Robertson1
1
Tall Timbers Research Station and Land Conservancy, Tallahassee, Florida 32312, USA
The importance of fire in maintaining community structure and species composition in upland longleaf pine forests has long been understood. Less well known is what
fire regime is most appropriate for adequately controlling
hardwood competition, providing wildlife habitat structure and function, and maintaining the highest species
richness of indigenous plants. In particular it is important
to know at what point following fire hardwoods may no
longer be controlled using fire alone, at which point more
expensive management techniques (herbicide, mechanical
treatment) must be used to restore the habitat. Determining the effects of fire regime requires long-term research.
Such research in the past has not adequately addressed
native (never plowed) upland longleaf pine forests burned
with growing-season fires, alternating fire intervals, or
alternating season of fire.
With these points in mind, the Fire Ecology Laboratory at
Tall Timbers Research Station and Land Conservancy
(TTRS) has establish a long-term research project in upland longleaf pine forests, to quantify the effects of time
since fire (1-7 years) on variables that define healthy forests and suitable wildlife habitat, including hardwood
versus herb dominance, plant biodiversity, tree demographics, and fire behavior and effects. The study will
measure the effects of different soil types on these variables, as well as the effects of alternating intervals and
season of burn, and at what threshold of post-fire succession hardwoods can not be top-killed using fire alone.
These results will provide a "toolbox" for managers and
planners to achieve desired forestry and habitat management goals in the most cost-effective manner possible.
The study is located on Pebble Hill Plantation, a 1600 ha
property managed by TTRS near the Georgia-Florida
boundary. The study sites are within natural upland pine
forest, dominated by longleaf pine and wiregrass, and
never before plowed. The pines have been managed with
single-tree selection cuts to maintain an uneven size distribution that is relatively open (7-15 m2/ha basal area) to
mimic the natural forest structure.
Within these areas, 24 plots each measuring 35 x 35 m
(0.12 ha) have been permanently established. Plots were
all burned in spring 2005 and subsequently will be treated
with late spring or early summer (May-June) fires, unless
specified otherwise by the treatment. Each is to be
treated with one of the following fire intervals, randomly
assigned to plots within the three soil type blocks: 1 yr;
1.5 yr (alternating seasons); 2 yr; 3 yr; 1-3 yr random sequence; 4 yr; 5 year; and 7 year. For the 5 yr and 7 yr
treatments, if the first burn after interval does not return
the system to measured baseline conditions with regard to
hardwood dominance, the plot will be burned annually
until conditions are restored, after which the treatment
will be repeated.
The plots are clustered in three groups of the eight treatments and blocked according to soil type to incorporate
the range of sand depth from relatively deep (sandhills;
Arenic Paleudults) to shallow (clayhills; Plinthic Kandiudults). Plots not scheduled to burn are isolated by
mowed and raked firebreaks. Fuel load (4 x 0.25 m2
plots) and fuel moisture (from destructive samples) are
measured before burning and, flame length are recorded.
Post-burn residue is collected to estimate fuel consumption, particulate emissions, and fireline intensity. Rate of
spread and residence time will be measured using Thermocouples, placed at known distances apart and measuring temperatures every 1 second during the burn. Standard
weather variables (temperature, relative humidity, wind
speed and direction) are recorded using a portable weather
station.
Understory plant composition and hardwood sprouts are
monitored in each plot in the spring (prior to that year's
burns) and fall to cover any seasonal variation in the presence/absence of particular plant species. Within each plot,
two 100 m2 (10 x 10 m) vegetation subplots were permanently established using steel reinforcement bar, within
diagonal quarters of the plot. Each vegetation subplot has
two 1 m2 and 10 m2 nested subplots, to monitor vegetation
at different scales (modified Whittaker Plots). And to
sample vegetation near the plot corner as well as near the
plot center, to test for edge effects. Presence of each plant
species and approximate cover is recorded using a modified Daubenmire cover class method. From a marker at
the center of each vegetation subplot, all hardwood genetic individuals and sprouts are censused and measured
for diameter 3 cm above base within a 2 m radius of the
marker (28 m2). Hardwood sprouts are re-censused about
one month post-burn to measure percentage of hardwood
stem topkill. At the beginning of the study and at 3 year
intervals, all trees ≥4 cm at 1.5 m height within the plots
will be identified, mapped, and tagged to monitor longterm forest dynamics. Canopy cover is measured during
each census using a sighting tube, and tree mortality will
be noted following each burn.
Poster Presentations 116
Data will be analyzed using ANOVAs with fire regime as
the factor, plots as replicates, and soil type as a blocking
variable (or additional factor) to test the effects of fire regime on the various habitat characteristics measured. Also,
analyses of changes in species composition will be conducted using a DCCA analysis in CANOCO. However, the
most significant analysis will simply be a) whether or not
post-fire succession resulted in a change in habitat conditions and b) whether or not prescribed fire returns the habitat to suitable conditions under a given fire regime.
Poster Presentations 117
Spatial and Age Structure of Old-Growth Mountain Longleaf Pine, (Pinus palustris), Stands
in the Talladega National Forest of Northeastern Alabama
Brett Rushing1, Kevin Jenne2, and Robert Carter1
1
Department of Biology, Jacksonville State University, Jacksonville, Alabama, 36265, USA
2
Anniston Museum Of Natural History, Anniston, Alabama, 36201, USA
Introduction
The spatial distribution of trees and seedlings within forested stands can indicate past stand history and provide
guidance for making management decisions. Previous
research in Coastal Plain longleaf pine ecosystems has
found a negative relationship between mature tree and
seedling location (Palik et al. 1997, Brockway and Outcalt 1998, Grace and Platt 1995). Compared to Coastal
Plain longleaf, the spatial and temporal structure of montane longleaf pine ecosystems are poorly understood
(Varner et al. 2003). This research addresses this deficiency in montane longleaf pine stands.
Materials and Methods
The Talladega National Forest is located in the Ridge and
Valley and Blue Ridge Provinces of northeastern Alabama. This forest has several stands containing relic
(>100yr.) longleaf pine trees mixed with younger trees.
In the winter of 2004, data were collected for all living
members of the Pinus genus >11.4 cm DBH within four
stands containing relic longleaf pine trees. Within the
stands, a central tree was located as plot center from
which the distance and azimuth to trees within 40 m of
the plot center were measured. All trees were measured at
DBH (to nearest cm). All trees were cored at breast
height using an increment borer. Cores were stored,
dried, and the annual ring increments measured to the
nearest 0.01 mm using photo-lab equipment and ImagePro Plus 4.0 software. Mean growth for each specimen
was then calculated, and a master chronology was created
for each stand by averaging mean growth for each year.
During the winter of 2006, the stands were revisited and
the distance and azimuth to seedlings and saplings under
5 m in height were measured from plot center. The seedling/sapling plots had a radius of 20 m with the exception
of stand 62-29 which was limited to a 10 m radius due the
large number of seedlings.
The spatial relationship between mature trees was analyzed with Ripley’s univariate K(l) with 95% confidence
envelopes using 99 Monte Carlo Permutations with the
program Programita (Wiegand and Moloney 2004).
Cluster analysis of mature trees using x and y coordinates
and tree age was performed using Systat (2004). In addition, dispersion indices were calculated including Index of
Dispersion (ID), Index of Cluster Size (ICS), and Green’s
Index using the Program PASSAGE (Rosenberg 2001).
Ripley’s bivariate K(l) was calculated to determine the
relationship between mature trees and seedling locations
using the program Programita (Wiegand and Moloney
2004)
Results
The results varied with each stand reflecting the unique
stand histories and disturbance patterns. Dispersion indices indicated overall clumped distribution when all trees
were included (Table 1). Stands 64-1 and 31-9 have
lower clumping values due to the smaller number of trees
in the plot. When only trees greater than 90 years old
were analyzed, a similar pattern was found except all
clumping values were lower due to the lower number of
trees in the plot. The GI for trees greater than 90 years
old in Stand 31-9 was nearly uniform (Table 1).
Ripley’s univariate K(l) indicated clustering (above the
confidence envelope) of mature trees with the exception
of Stand 31-9. Stands 37-23 and 62-29 are consistently
clustered above a spatial scale of 9 to 10 m. Stand 64-1
was clustered at some points but not at others. It was the
only stand with mature hardwoods within the stand.
Stand 31-9 was consistently random in spatial distribution. This is also reflected in low dispersion indices
(Table 1).
Table 1. Dispersion indices for all trees and only trees
>90 years old.
Trees
<90
years
old
All
Trees
Stand
# of
Trees
64-1
45
12.36 11.36 0.47 4.36 3.36 0.14
37-23
97
24.25 23.25 0.97 7.21 6.21 0.26
31-9
37
62-29
90
Poster Presentations 118
ID
8.8
ICS
7.8
GI
ID
ICS
GI
0.33 2.67 1.67 0.07
23.24 22.24 0.93 5.72 4.72
0.2
Cluster analysis based on age and location indicated a
large number of clusters in Stands 37-23 and 62-29. The
clusters in 31-9 and 64-1 were less numerous and more
uniform in age.
Analysis of the relationship between mature trees and
seedlings using Ripley’s Bivariate K(l) also yielded variable results. In Stands 64-1 and 31-09, there was a negative relationship (below confidence envelope) between
mature tree and seedling location up to a spatial scale of
15-17 m. Thereafter the relationship is positive. Thus
seedlings and saplings are not found in close proximity to
trees. For Stands 62-29 and 37-23, there is no negative
relationship. However, Stand 31-9 never showed a large
negative relationship between trees and seedlings.
Discussion
Tree Spatial Pattern and Disturbance
Previous research has shown that longleaf pine trees and
seedlings have negative spatial relationships (Palik 1997,
Brockway and Outcalt 1998, Grace and Platt 1995).
Avery et al. (2004) found clustering of seedlings but no
spatial relationship with mature trees. This research indicates that the spatial distribution of trees and relationship
between the trees and seedlings varies depending on stand
history. Stands 31-9 and 64-1 had fewer trees and a
lower number of clusters. In 31-9, this can be attributed
to a lack of anthropogenic disturbances such as timber
harvesting. It is located in a remote area with no roads to
provide access. Stand 64-1 can be explained by hardwood trees within the stand that likely influenced spatial
geometry. This may explain the wider spatial distribution
of trees and fewer clusters. The stand is adjacent to a
road so past disturbances are expected.
Stands 37-23 and 62-29 both had a stronger degree of
clustering (Table 1) and number of clusters. These stands
showed more numerous periods of growth increase and
tree recruitment indicating a history of frequent disturbances (Jenne 2004). Disturbances have been found to
increase tree clustering during point pattern simulations in
Minnesota (Woodall and Graham 2004). The two periods
of drastic increase in mean growth seen in these stands
may be attributed to logging, fire suppression, and European settlement patterns. No large scale climatic disturbances have occurred in the study area (Climatology of
Alabama) to explain the drastic changes in growth and
recruitment. A drastic increase in mean annual growth
that occurred in 1833 coincides with Native American
removal and European settlement and resource exploitation (Yarnell, 1998). It can be speculated that this period
of land resource exploitation could have yielded vast
recruitment with less competition and increased growth
rates. It can also be speculated that the growth increase
starting in 1933 is due to disturbance by the utilization of
wood and other resources by the Civilian Conservation
Corps during the depression and also during World War
II. These growth responses were not found in Stand 31-9.
Its growth rate remained relatively uniform from 1825 to
the present. Its remote location likely reduced human
exploitation (Jenne 2004).
Seedling Spatial Pattern
The negative relationship between mature trees and seedlings in Stands 31-9 and 64-1 can be explained by the
lower number of trees and seedlings (66 seedlings for 319, 146 seedlings for 64-1). The fewer overstory trees and
their dispersion throughout the plot reduces that chances
of a clustering near trees. The plots had not experienced
recent prescribed burns needed for seedling establishment
(Wahlenberg 1946). Plot 64-1 also had some hardwood
trees within the stand that could reduce seedling survival.
Stands 37-23 and 62-29 both contained the largest number
of trees and seedlings (1987 for 62-29, 675 for 37-23).
The large number of seedlings can be attributed to prescribed fire within the previous 2 years and the large number of mature trees to provide seed.
Literature Cited
Avery, C.R., S. Cohen, K.C. Parker, and J.S. Kush. 2004.
Spatial patterns of longleaf pine (Pinus palustris)
seedling establishment on the Croatan National Forest, North Carolina. North Carolina Acad. Sci. 120:
131-142.
Brockway, D.G. and K.W. Outcalt. 1998. Gap-phase
regeneration in longleaf pine wiregrass ecosystems.
Forest Ecol. Mgt. 106: 125-139.
Climatological Summary , Alabama, Selected Cities, U.S.
Department of Commerce, Weather Bureau. Tuscaloosa Library Bindery, AL.
Palik, B.J., R.J. Mitchell, G. Houseal, and N.Pederson.
1997. Effects of canopy structure on resource availability and seedling responses in a longleaf pine ecosystem. Can. J. Forest Res. 27:1458-1464.
Grace, S.L. and W.J. Platt. 1995. Effects of adult tree
density and fire on demography of pregrass stage
juvenile longleaf pine (Pinus palustris Mill.). J. Ecol.
83: 75-86.
Rosenberg, M.S. 2001 PASSAGE. Pattern Analysis,
Spatial Statistics, and Geographic Exegesis, Ver.1.0.
Department of Biology, Arizona State University,
Tempe, AZ.
SYSTAT. 2004. SYSTAT 11.0 for Windows. SYSTAT
Software, Richmond, CA.
Poster Presentations 119
Varner, J.M., J.S. Kush, and R.S. Meldahl. 2003. Structural characteristics of frequently burned old-growth
longleaf pine stands in the mountains of Alabama.
Castanea 68: 211-221.
Wahlenberg, 1946. Longleaf pine: its ecology, regeneration, protection, growth, and management. Charles
Lathrop Pack Forestry Foundation, Washington, DC.
429 pp.
Wiegand, T. and K. A. Moloney. 2004. Rings, circles, and
null-models for point pattern analysis in ecology. Oikos
104: 209-229.
Woodall, C.W. and J.M. Graham. 2004. A technique for
conducting point pattern analysis of cluster stem-maps.
For. Ecol.. Mgt. 198: 31-37.
Yarnell S.L. 1998. The Southern Appalachians: A History
of the Landscape. USDA Forest Service Technical
Report. SRS-18. 43pp.
Poster Presentations 120
Spacing Recommendations for Longleaf Pine
David B. South1
1
School of Forestry & Wildlife Sciences, Auburn University, Auburn, Alabama, 36849, USA
My name is David Malone, and I am the Forest Silviculturist with the US Forest Service at Savannah River Site
south of Aiken, SC. At our location, we plant nearly 800
acres of bareroot longleaf seedling a year. The survival
rate 90% 1st yr, and 85% 3rd yr. Our spacing is 10x6
ft avg over 700 seedling/ac. We are facing a problem
with the high survival rates in our plantations. What to
do? Pre-commerical or let the stands go until 1st thin?
Do you know of any research papers on recommendations of actions for longleaf plantations?
Dear Mr. Malone:
If you plant too many trees per acre, then it will cost
you money to correct the mistake. A pre-commercial
thinning (to reduce stocking levels to 250 to 400 trees
per acre at age four) might cost $100 or more per acre.
Letting the stands go until age 20 or 25 yr (at the first
thin), will reduce average stand diameter (Lohery and
Bailey 1977) and will reduce the net present value of
the stand (Teeter and Somers 2005). If your objective
includes improving ecosystem restoration (Stainback
and Alavalapati 2004), improving wildlife habitat
(Thackston 2002), early production of sawlogs (South
2006), and avoiding an early thinning (Kush et al.
2006), be sure to not plant too many trees per acre.
When using good quality seedlings and with machine
planting, a 14-foot row spacing with 7 to 8 feet between
seedlings (450 to 390 trees per acre) should produce
about 300 to 380 live trees per acre at age 3 yr. Table 1
compares three spacing recommendations for either
container-grown longleaf pine, or machineLongleaf Pine
Seedlings planted per acre
Seedling cost/acre
Row spacing
Expected live seedlings - yr 1
Expected sawtimber trees 30 yr
Expected DBH at first thin
Expected DBH yr 30
Expected sawtimber yr 30
Rings per inch
Basal area yr 30
Pinestraw production yr 30
Poles at age 39
WHIP eligible? (subsidy)
Expected forage yr 30
N.P.V @5% (no straw or poles)
Favored by
Wide spacing
300 to 499
$39-$65
12-16 ft.
210 to 450/ac
90/ac
8 inches
9.1 inches
81 gtons/ac
6.6
116 sq.ft.
115 bales/yr
large diameter
YES
922 lbs/ac
$858/ac
Wildlife managers
planted large-diameter bareroot stock. As you know, it is
very important to define landowner objectives before recommending the desired tree spacing.
Literature Cited
Kush, J.S., J. C. G Goelz , R.A. Williams, D.R. Carter, and
P.E. Linehan. 2006 Longleaf Pine Growth and Yield.
In: The Longleaf Pine Ecosystem: Ecology, Silviculture and Restoration. S. Jose, E.J. Jokela, and D.L.
Miller, eds. Springer-Verlag, New York. Environmental Management Series. pp. 251-267.
Lohery, R.E. and R.L. Bailey. 1977. Yield tables and stand
structure for unthinned longleaf pine plantations in
Louisiana and Texas. USDA For. Ser. Res. Paper SO133. 53 pg. South. For. Exp. Sta., New Orleans, LA.
South, D.B. 2006. Planting longleaf pine at wide spacings.
Native Plants Journal 7(1): 79-88
Stainback, G.A. and J.R.R. Alavalapati. 2004. Restoring
longleaf pine through silvopasture practices: An economic analysis. Forest Policy and Economics 6 (3-4):
371-378.
Teeter LD, Somers G. 2005. Longleaf ecosystem restoration: long-rotation economics. In: Kush JS, compiler.
Symposium proceedings, fifth Longleaf Alliance regional conference; 2004 Oct 12–15; Hattiesburg, MS.
Auburn (AL): Longleaf Alliance. Report No. 8. p 129–
139.
Thackston, R. 2002. About the Bobwhite Quail Initiative.
Wildlife Resources Division.
Middle spacing
500 to 699
$65-91
10-12 ft.
350 to 630/ac
143/ac
6 inches
8.6 inches
55 gtons/ac
7.0
132 sq.ft.
130 bales
greatest # of poles
NO
820 lbs/ac
$802/ac
Values are site specific. Do not use above examples to predict future outcomes for all planted stands.
Poster Presentations 121
Close spacing
700 to 900
$91-117
8-10 ft.
490 to 810/ac
99/ac
5 inches
8.5 inches
40 gtons/ac
7.0
145 sq.ft.
145 bales
few poles
NO
650 lbs/ac
$777/ac
Nursery managers
Ichauway’s Prescribed Fire Management Program 1994-2006: A Balanced Approach
Jonathan M. Stober1 and Steven B. Jack1
1
Joseph W. Jones Ecological Research Center, Ichauway. Newton, Georgia, 39870, USA
Abstract
Ichauway, an 11,733 ha preserve in southwest Georgia,
contains significant remnants of the fire dependent longleaf pine-wiregrass (Pinus palustris-Aristida beyrichiana)
community. Prescribed fire is the principal forest management tool utilized at Ichauway for over 75 years. During the past 12 years the fire management program has
documented all prescriptions and evaluations for all fire
events. Fires are prescribed to meet specific objectives for
each burn unit with an overall objective to burn 4,5005,300 ha annually. Keystone objectives for every prescribed fire are safety, fire control, smoke management
and resource protection. After each fire event fire extent
and degree of crown scorch are mapped and placed in a
GIS. Weather conditions, fire objectives, fire origin, containment, and subjective evaluations of fuel consumption,
duff consumption, and woody plant top-kill are recorded
for each burn. Stated fire objectives are often focused on
fuel reduction and hardwood control, but can vary widely
from educational demonstrations to wiregrass seed production. Typically 99% of all fire events are prescribed
with containment median above 97% for over 1980 recorded fire events in the past 12 years. Two-thirds of the
acreage is burned in the dormant season (before April) in
a given year, and 87% of all prescribed fires occur with a
KBDI value below 400. Overstory crown scorch averages 5% of area burned. Analyses of 11 years’ data
found crown scorch to be dependent on understory type,
with 65% of all scorch occurring on wiregrass, 26% on
oldfield and 5% on shrub-scrub groundcovers. By focusing on burn unit objectives and frequency rather than season of burn the fire management program has, over the
past 12 years, consistently met the goal of burning 50% of
Ichauway’s landbase each year, including drought years.
The current management strategy provides a balanced
approach to meet objectives and sustain the ecosystem.
Introduction
The frequent application of prescribed fire to Ichauway
has created a property where high quality examples of
native communities endure today. Prescribed fire is the
one management tool that is uniformly applied across the
entire property. The property contains more than 9,700
ha (24,000ac) of upland pine grassland habitats, with the
remainder consisting of agricultural fields, wetlands, and
riparian hardwood hammocks. The pine forest at
Ichauway was intensively harvested early in the 20th century and currently has basal area ranging from 9-15 m2/ha
(40-60 ft2/ac) and higher with pines being widely spaced.
Upland pine habitats at Ichauway are dominated by mature 75-95 year old longleaf pine and either a wiregrass or
broom sedge (Andropogon virginicus) old field understory. Upland hardwood is generally localized to fire
shadows around roads, field edges, fire breaks, wildlife
food plots, old house sites and aquatic habitats.
The keystone objectives for every prescribed fire are
safety, fire control, smoke management and protection of
the resource. Each prescribed fire has specific objectives
that guide the application and purpose for the fire which
may include one or more of the following: fuel reduction
and hardwood control, perpetuating fire dependent species
and restoration, wildlife habitat management, research,
education and demonstration, seedbed or planting preparation, wiregrass seed production, wetland management,
boundary security, debris or slash burning, and hay production. The overall management strategy is to burn individual units on a 2-year rotation, but this can range anywhere from 8-months to more than 5-years depending on
the location, objectives and conditions of the burn unit. A
2-year rotation helps to maintain fuel loading within a
range that minimizes the risk and damaging effects of
wildfire.
Methods
At the beginning of the year a map (Figure 1) is given to
the Natural Resource Manager that identifies the annual
rough accumulation (fuel accumulation) for each burn
unit. The Natural Resource Manager identifies and coordinates all prescribed fires on the property. Prescribed
fires are executed with a team of 3-5 people, each having
a two-way radio, using All-Terrain Vehicles outfitted with
a drip torch and a water tank. Because current, accurate
fire weather is so critical to planning and executing a successful prescribed burn, weather data are collected to help
predict fire behavior. Each prescription records weather
forecast information on the burn plan with information
collected from on site weather stations and the Georgia
Forestry Commission Fire Weather web site. Before a
prescribed fire is ignited minimum weather conditions
must be met or exceeded: transport winds >14kph (>9
mph), mixing height >520m (>1700ft) and smoke dispersion index >40.
Poster Presentations 122
fires from 1994-1999 were primarily fuel reduction fires to
reduce duff accumulations so hotter maintenance fires could
be subsequently introduced. Crown scorch of forested areas
is divided into 5 categories: none, <1/3, 1/3-2/3, >2/3 but
not all, and complete. The later three categories are mapped
into a GIS. Crown scorch averages 5-6% of the total burned
acreage each year. On average 58% of the crown scorch is
in the 1/3-2/3 class, while the >2/3 and complete scorch
classes average 38% and 4% respectively. Trends indicate
that scorch has increased since 1994 and a greater proportion is scorched during the growing season. Crown scorch
is also dependent on understory type with wiregrass more
then twice as likely to result in scorched crowns.
Figure 1: years or rough map (fuel accumulation) for
2006 with 2 and 3 or more years of rough mapped by
burn blocks bounded by roads and firebreaks at Ichauway.
Approximately two weeks following the prescribed fire
the effects of the fire are evaluated by recording the
containment, origin, and assigning subjective classes for
duff consumption, woody understory kill and vegetative
fuel consumption. The extent of the fire and degree of
crown scorch are mapped and entered into a Geographic
Information System (GIS).
Much debate surrounds the need for exclusive use of growing season prescribed fire in the southeast. Ichauway has
been managed for over 75 years with March and April prescribed fires and has maintained and enhanced its fire maintained longleaf pine grasslands. It is the opinion of the
Jones Center staff that fire frequency is more important than
season of burn. Prescribed fires can occur any month at
Ichauway but generally occur during the first seven months
of the year, allowing vegetation time to recover before winter. Burn units are targeted by objective rather than season;
that is, maintaining manageable fuel loads and desired future conditions for the burn unit drive the decision to burn.
Increased frequency presents more opportunities to vary the
season, weather and type of fire needed to move the burn
unit toward maintenance condition.
Table 1: Total acreage burned during the dormant
(October-March) and growing season (April-September)
by year at Ichauway.
Results and Discussion
Prescribed fire objectives most often include reducing
fuel loads and controlling hardwoods. The principle
ignition source for >99% of all fire events are prescribed with the remainder either jumps from adjacent
properties or lightning strikes. Containment median is
above 97% (ranges 83-99% between years) with almost
all spot-overs contained during the fire. Since 2000
containment has improved with the removal of snags
along burn unit boundaries. Using a subjective evaluation of fuel removed by the fire, on average 30% of the
fires are categorized as “clean” meaning all fuel is removed with the remainder of the fires leaving “patchy”
fuel beds with a median of 20% of the fuel remaining.
Keeping the woody deciduous understory in shrub form
is a primary goal of the prescribed fire program. Over
the past six-years 80% of the burn units have achieved
over 95% control of the woody understory contrasted
with only 55% control from 1994-1999. The major
factor controlling this difference is that the prescribed
Year
2006
2005
2004
2003
2002
2001
2000
1999
1998
1997
1996
1995
1994
Averages
Poster Presentations 123
Dormant
Season
Growing
Season
Total
Acres
7765
11332
7054
9822
6830
8734
8796
6294
4943
3153
4994
8199
13473
5868
2726
4175
3886
3196
2623
2740
2936
5449
4813
5706
2377
1850
13632
14058
11229
13708
10026
11356
11535
9230
10393
7966
10700
10576
15323
Percent
Growing
Season
43
19
37
28
32
23
24
32
52
60
53
22
12
7799
3719
11518
34
Weather patterns in the southeast dictate that the most consistent weather to control and perform prescribed fires occur
from December through the end of March (Figure 2). Thus
the majority of prescribed fires take advantage of these conditions in order to meet yearly objectives for acreage
burned. After the end of March the drought index predictably increases and intensifies depending on spring and summer precipitation amounts. During the growing season as
drought index values increase prescribed fires become more
difficult to control and execute, especially if drought conditions occur and fuel loads are excessive. If management
executed prescribed fires exclusively during the growing
season weather conditions would dictate the acreage burned
each year. When drought conditions occurred this would
push the prescribed fire team further behind goals and result
in increased fuel loads during the next growing season.
Under the current management strategy at Ichauway an
average of 34% of all burning is conducted during the
growing season over the past 13 years (Table 1). Once fuel
loads were brought under control in 1994 and 1995 more
growing season prescribed fire has been utilized. From
1999-2001 growing season burning was reduced due to the
drought conditions in the region. The low acreage typically
burned during May and June are due to the spring droughts
that often suspend all prescribed burning in the region.
Summary and Conclusions
Prescribed fire is embedded in Ichauway's culture. For over
75 years prescribed fire has actively been used as a land
management tool. During each of the past 12 years our
objective of placing prescribed fire on approximately 50%
of the property has been achieved (Figure 3). The ability to
consistently meet objectives is due to manpower and equipment being readily available to exploit weather opportunities when they occur. Despite drought conditions from
1999-2001 the prescribed fire program continued to meet its
objectives by taking advantage of weather conditions as
they occurred regardless of season. By burning frequently
and not constraining management to growing season fires,
fuel loads are kept in check, fires
are kept under control and
smoke is managed when conditions are optimal. This keeps the
prescribed fire team setting prescribed fire rather than fighting
wildfire. The current management strategy offers a balanced
approach to meet objectives,
accomplish goals and sustain an
ecosystem. For more information visit www.jonesctr.org.
Figure 2: Summary of all KBDI values over the past 13 years for a calendar year are illustrated with a normalized
trend line and all prescribed fires (red dots) occurring between 1995 and 2006. White dots indicated fires set for
fuel reduction and wildlife management, particularly bobwhite quail management.
Figure 3: The percentage of prescribed fire acreage burned per year was based on a 5-year cumulative burn extent defined by burn mapping 1994-1998, 1999-2003 and the total cumulative extent thereafter divided by the
total acreage burned for each year.
Poster Presentations 124
Allatoona Lake Longleaf Pine Ecosystem Restoration Project
Terrell Stoves1
1
Allatoona Lake, Cartersville, Georgia, 30120, USA
Abstract
The US Army Corps of Engineers (USACE) manages
the 24,000 acres of land around Allatoona Lake, which
is 30 miles northwest of Atlanta, Georgia. In 2002,
USACE foresters began planning the restoration of a
sixty acre forested tract on a remote site in the heart of
the Allatoona Wildlife Management Area. USACE
foresters chose this site because a small number of longleaf pine trees were still present, showing that the site
one time contained a longleaf population. This site also
had a sandy clay, acidic soil with which longleaf pine is
associated.
A continuing challenge for both Allatoona Lake and the
Longleaf Pine Ecosystem Restoration Project is
neighboring development. The continued expansion
and development of the metro Atlanta area has put considerable pressure on USACE land and makes conducting prescribed fire operations increasingly difficult.
These prescribed fires play an essential role in maintaining the longleaf pine forest and its understory.
The US Army Corps of Engineers efforts at Allatoona Lake
have started the process of restoring a once dominant ecosystem to the area. This site is located at the northern tip of
the central portion of the longleaf pine’s original documented range and few forests in the region still contain
longleaf components. At Allatoona Lake, an attempt is
being made to turn a forested site that had only a remnant of
longleaf pine scattered throughout the landscape into a mature longleaf stand. With continued management, we hope
this site will serve as an opportunity to showcase a diverse
and important ecosystem in the heart of one of the fastest
growing parts of the country.
Poster Presentations 125
The Role of Ritual and Ceremony in Wildlands Conservation:
Reestablishing Primal Connections
Johnny Stowe1
1
South Carolina Department of Natural Resources, Columbia, South Carolina, 29224, USA
A society is ultimately measured not by what it develops
or consumes, but rather by what it has nurtured and preserved. - Jim Posewitz, Orion: The Hunter’s Institute
The worldview of a society is often written more truthfully
on the land than in its documents. - Kimmerer and Lake
(2001)
Natural resource managers tend to be trained in the scientific method and the natural sciences, with objectivity and
a “business-like” approach deemed paramount to their
professional duties. These qualities are no doubt imperative to conservation, and must always be part of land
manager’s training and operations. But despite the salient
conservation success stories in North America and other
areas, the outright loss of the rural and wild landscape and
the impaired state of those lands not lost indicate that
other tools are needed if we are to succeed in restoring
and maintaining the ecological integrity of the landscape.
I call for increased emphasis on the cultural attributes of
the landscape as an end in itself, as well as to serve as a
means to the end of conserving natural landscape attributes. I discuss the role of ritual and ceremony and their
attendant icons and symbols, in land protection and management, and suggest ways to incorporate them into the
conservationist’s “tool kit.”
Ritual and ceremony, and the icons (symbols) tied to
them, are integral but often taken-for-granted parts of
human ecology. All cultures consider them vital social
mechanisms for recognizing important events in time and
space, and in the past these events were often linked to
Nature. Today however -- especially in the Western
World where we are increasingly disconnected to the
natural world -- our rites and icons increasingly tend to be
artifacts.
Whereas throughout much of human existence phenomena such as births and deaths, adulthood, marriages, first
big game killed by a young hunter, first annual hunts for
certain species, first and last annual frosts, biannual solstices and equinoxes, moon phases and other phenological events were celebrated in a prescribed fashion aimed
at linking, emphasizing, celebrating and memorializing
key aspects of ontogeny and tribal events vis-à-vis the
observable natural environment, today we tend to ignore
or marginalize many of these events, and/or to recognize
and symbolize them with human-made products. Salient
symbols of contemporary Western Society include
cell phones, computers, email, iPods and other electronic
gadgets, SUVs, McDonald’s and such. The regular use
of, and association with these -- whether formalized and
prescribed, or through non-reflective habit -- could be
deemed key, albeit arguably unhealthy, rituals of our
times.
Loeffler (1989) described technofantasy as “a state of
mind where human attention is externalized and focused
on the process and product of human invention. In Western culture technofantasy has resulted in many ramifications, including rampant extraction of natural resources,
widespread pollution of natural environments, and a form
of cultural preoccupation with inorganic trivia.”
Even “outdoors folks” such as hunters seem these days to
be addicted to “technogadgets,” such as ATVs and other
“high-technology” in the forms of optics, firearms,
walkie-talkies, game cameras and game feeders. Hunters
also tend to be more interested in rituals such as planting
food plots (often with invasive exotic plant species) or
dumping out corn for bait than they are in restoring native
ecosystems. These practices can become entrenched inter-generationally in just a few years, and then they are
hard to stop. And they seem to often become ends in
themselves, rather than means (healthy and effective or
not) to an end.
My keen interest in the role of ceremony tied to wildlands
protection arose in the context of hunting, when in December 1993 I went deer (Cervidea) and boar (Sus scrofa)
hunting near the town of Pszczyna in southern Poland.
While there, I killed both a young red deer (Cervus
elaphus) and a mature fallow deer (Dama dama) stag, and
I experienced the ancient European ritual of honoring the
slain game and the hunt itself with the “broken branches”
ceremony (Stowe 1995). When I returned to the Southeastern U.S., I brought with me a deep appreciation for
the hunting ritual I learned in Poland, and have tailored it
to my taste and needs and continued it to this day. I conceived these words -- i.e. this prayer -- to memorialize the
game I kill and the landscape it lived in, and recite them
in the field before I gut the animal, and then again at the
table:
“In you (the game), I recognize and appreciate the land
that supports my existence. I am thankful for your beauty
and grace, and for the nourishment you will give me and
my family and friends. I hope your meat will give me
strength
Poster Presentations 126
and inspiration to restore and protect this landscape. And I hope
the paths of your kind and mine cross often, as they have in the
past, and that we may always be a blessing to one another."
Years after bringing the ceremony described above into
my life, I discovered fellow Southerner William Faulkner’s works on hunting, and was delighted to read of a
similar ritual practiced long ago in the Mississippi riverbottoms. In Go Down Moses, the Pulitzer Prize-winning
Faulkner (1990) described how Sam Fathers blooded The
Boy’s cheeks as he also honored his first white-tailed deer
(Odocoileus virginianus) buck: “I slew you; my bearing
must not shame your quitting life. My conduct forever
must become your death.”
Mautz (1995), in “So What’s Wrong with Hugging a
Tree?” -- one of the most eloquent and prescient essays I
have encountered -- points out how natural resource professionals have marginalized themselves by avoiding at
all costs - in curricula, avocationally, and on-the-job - any
subject matter approaching a spiritual connection to Nature. We seem to fear that expressing or even recognizing
such liberal-arts-type-sensitivity “suggests tree-hugging or
bambi-ism.”
I think Dr. Mautz is on to something. I believe that too
many scientists and other professionals working with
natural resources become almost robotic, and deprive
themselves of the ceremony and ritual that could enrich
their lives, as well as make their work more relevant and
effective on-the-ground. I am not a scientist, but the
poignant words of one of the best, and a hunter and philosopher as well -- Valerius Geist (1975) -- resonate
within me as I ponder how we can garner more support
for wildlands protection. Val wrote:
“In the past I revolted against ceremony, ridiculing it as
senseless, stupid, outmoded; but no more. Ceremonies
serve a good purpose. The forms are prescribed, and
through their performance give pleasure, security, identity, confidence in social relationships and, as such, an
‘inner strength.’ Many ceremonies permit persons to act
out their feelings in a meaningful way. Ceremonies bind
together those who participate; they delineate an in-group;
a gathering of like minds. Ceremonies often arouse emotions, and ensure that we shall remember the occasion as
well as the persons who participated. Trials by crisis in
fraternities did not arise by accident; they help strangers
become loyal friends. Ceremonies in war and peace, in
marriage, birth, and death, unite families, communities,
and nations; they help people to live in security and trust.
It is not surprising that even ancient philosophers of human nature expounded the need for ritual and order as a
prerequisite to civilized life.”
I had my way, no student would be allowed to study science
and be let loose on laboratories or field, without a degree in
liberal arts. And I do not say this in jest!” (Valerius Geist,
personal communication 2001). Anyone who wonders
about what type of scientist makes such a claim should do a
literature review of Val’s voluminous body of work on evolution and other such matters of “hard science.” I particularly recommend his magnum opus, Life Strategies, Human
Evolution, Environmental Design: Toward a Biological
Theory of Health.
Consider this. What single book has had an impact on conservation like Aldo Leopold’s (1949) A Sand County Almanac? This classic is used as a text in classes as sundry as -on the one hand, sociology, literature, visual arts, philosophy, psychology, and recreation -- and on the somewhat
diametric other, wildlife ecology and management, forest
policy, wilderness management, and restoration ecology.
And it is a book fraught with ritual and symbols linked to
the natural world and people’s role in that world.
Fran Hamerstrom (1989), the first female in the wildlife
management profession, described how her and her husband
sometimes purposely went afield hungry in order to more
fully appreciate the most atavistic role of hunting, that of
garnering meat for immediate consumption. She wrote, “A
good appetite – indeed sometimes real hunger – is part of a
real hunt for us. … those who have never known hunger,
and who do not understand the interrelationships between
hunting and hunger can barely understand and grasp hunting.”
Many of the Nature-based ceremonies that once were part
of our lives have been lost or usurped as we moved from
one continent to another, changed lifestyles and religions,
and otherwise changed our societies. Through interpretation of the archaeological record, and reviewing history, we
can revive the ancient rites that once sustained us. And
when the past doesn’t suffice to lead us into the future, we
can surely develop our own ways to honor and memorialize
our activities. Author, filmmaker and wildlands-loving
iconoclast Douglas Peacock (Loeffler 1989) described such
innovation, “I have my own little private ceremonies.
When my culture doesn’t provide ceremonies I just bloody
make up my own.”
The brilliant human ecologist Paul Shepard (1998), in Coming Home to the Pleistocene, suggests practical ways we
can incorporate key elements of our primal past into our
contemporary lives. Many of these elements are ritualistic
and symbolic by nature. Stowe et al. (2001) described how
the death and burial of loved ones could be dealt with in a
manner aimed at protecting the landscape as well as healing
grieved hearts and memorializing the deceased.
Poster Presentations 127
And So:
Stephen Pyne (2000), referring to the need to revive prescribed fire, wrote, “Prescribed fire doesn’t need a policy.
It needs a poet.” I agree. And we may soon have such a
stateswoman or man. Janisse Ray and others have captured
the hearts of Southerners with their place-based prose. By
linking place and culture and time through ritual and ceremony and icons we can create a setting in which local people can find inspiration, and out of that, we may one day
find our poet -- our contemporary Thoreau or Leopold, our
Faulkner or Rachel Carson.
I have sought to protect my native mountain longleaf pinelands of northwest Georgia and northeast Alabama by linking that unique landscape to the past, and to the cultures of
the various people who have lived there over the last dozen
or so millennia (Stowe 2004,2006). Rituals associated with
hunting -- as well as those I am developing for land management activities such as timber harvesting, firewood cutting, longleaf pine grassland restoration, prescribed burning,
and control of invasive exotic species -- are part of my land
management approach, and through them I hope to make
my hunting and restoration work, and the landscape they
center on more meaningful to myself, as well as to my family and friends and generations far into the future, and I
hope that my family and friends and others now unknown
will continue what I have started.
We Southerners tend to be mighty proud of place and identity, yet in land management and protecting our culture we
have made monumental mistakes – e.g., we latched onto the
once-novel idea of introducing invasive exotic species into
our landscape as putative panaceas and promoting them at
the expense of natives for “conservation;” we often become
so blinded by individual private property rights issues that
we cross the line beyond which the landscape as a whole,
including those private tracts, becomes threatened; we allowed the ancient multi-cultural tool of woods-burning to
be usurped by a cartoon bear and other carpetbaggers
(Stowe 2002); and we forgot that our native landscape, described by Leopold (1947) as the native soils, waters, flora
and fauna, as well as local people -- is our culture, heritage
and character. This upsets me. Part of that may be because
my people were farmers once allied with King Cotton, who
sent so much of our native topsoil down to the Atlantic
Ocean and Gulf of Mexico. But those earlier generations
did not have the scientific knowledge that we now have and
so are not culpable in the moral sense to anywhere near the
degree that we are, because we know quite clearly the effects of our actions.
I had the honor of corresponding with fellow Southerner
E.O. Wilson a few times (personal communication 2000,
2003), and when I bemoaned the lack of a conservation
ethic in so many of my fellow Southerners, he gave me
hope and inspiration when he responded: “Our fellow Southerners will come around to conservation, and big-time too – when
they realize it’s their heritage.”
What an amazing difference in my motivation such a
few words from a kindred spirit and fellow Son of the
Southland
had on me. So let’s “come around;” let’s restore our
heritage, those primal connections that once bound us to
the land in a way that made us who we are, those sacred
links that shaped our character and culture. Let’s do it
for the good of the land, as well as ourselves.
I am deeply obliged to Joyce Marie Brown of the University of
Central Florida for handling all aspects of the graphic design
and layout of this poster.
Literature Cited
Faulkner, W. 1990. Go Down Moses. Random House. 365
pp.
Geist, V. 1975. Mountain Sheep and Man in the Northern
Wilds. Cornell University Press. 248 pp.
Hamerstrom, F. 1989. Is She Coming Too? Iowa State University Press. 156 pp.
Kimmerer, R.W. and F.K. Lake. 2001. The Role of Indigenous Burning in Land Management. Journal of Forestry.
99:11. 36-41.
Leopold, A. 1947. The Ecological Conscience. Bulletin of
the Garden Club of America. September. 45-53.
Leopold, Aldo. 1968. A Sand County Almanac. Oxford
University Press. 226 pp.
Loeffler, J. 1989. Interviews with Iconoclasts. Harbinger
House. 194 pp.
Mautz, W. 1995. So What’s Wrong with Hugging a Tree?
Wildlife Society Bulletin. 23:1. 107-108.
Pyne, S.J. 2000. Green Skies of Montana. Forest History.
Spring. 37-38
Shepard, P. 1998. Coming Home to the Pleistocene. Island
Press. 195 pp.
Stowe, J. 1995. A Quality Hunt: European Style. 2:2. 11-12.
Stowe, J., E.V. Schmidt, D. Green. 2001. Toxic Burials: The
Final Insult. Conservation Biology. 15:6. 1817-1819.
Stowe, J. 2002. Woods Burning in South Carolina: The Nature and Culture of Wildland Fire and its Impact on Our
State’s Character. Unpublished speech presented at the
Annual Meeting of the SC Prescribed Fire Council. 10
November 2004 National Wild Turkey Fed. Headquarters: Edgefield, SC
Stowe, J. 2004. Wildlife and other Aspects of the Mountain
Longleaf Pine Forests and other Ecosystems of Northeast
Alabama and Northwest Georgia. Pages 1-27 in John S.
Kush. Compiler. Proceedings of the First Montane
Longleaf Pine Conference. 15-17 OCT 2003. Jacksonville State University, Jacksonville, Alabama. Longleaf
Alliance Rep No. 7.
Stowe, J. 2006. Connecting the Longleaf Pinelands of Northeast Alabama and Northwest Georgia: Our Last Chance
for Corridors to Protect the Special Things. Pages 24-28
in Martin L. Cipollini. Compiler. Proceedings of the
Second Montane Longleaf Pine Conference Workshop.
18-19 NOV 2005. Berry College. Mt. Berry, GA.
Longleaf Alliance Report No. 9.
Poster Presentations 128
Private Property Rights vis-à-vis Establishing and Maintaining Invasive Exotic Plant Species:
Legal and Ethical Ramifications of the “Right to Plant” versus Other’s “Right to Maintain
Landscape Integrity and Property Values”
Johnny Stowe1
1
South Carolina Department of Natural Resources, Columbia, South Carolina, 29224 USA
Conservation is paved with good intentions which prove
to be futile, or even dangerous, because they are devoid
of critical understanding either of the land, or of economic land-use. Aldo Leopold (1949)
Abstract
Here in the Southeastern United States the issue of private property rights (PPRs) is often an emotional and
divisive polemic. Many landowners resent perceived
intrusion of PPRs by government and other entities, and
the notion of “It’s my land and I’ll do what I please on
it,” resonates in many land use discussions. In this context, I call attention to the nascent, but rapidlyincreasing and ubiquitous, recognition of the pernicious
and insidious threat of invasive exotic species to the
natural and cultural heritage, as well as to the economy,
of the Southern landscape. Specifically, I review these
threats -- along with the four basic tenets comprising the
“sticks in the PPR bundle,” as they relate to introducing,
harboring, and promoting invasive exotic species -- and
call attention to the incongruities, inconsistencies and
contradictions inherent in certain PPR advocate’s positions.
Introduction
In the United States, few things are held as sacrosanct
as private property rights (PPR). The National Woodland Owner’s Association (2005) survey of its members
and affiliates -- The Top Ten Forestry Issues for 20052006 -- ranked “Private Property Rights” as second only
to “Fair Income, Inheritance and Property Taxes.”
While I have no data to support my contention, I doubt
that PPRs are more highly esteemed and defended anywhere in the nation than in the Southeastern U.S. (SE
US). Keville Larson, past president of the Forest Landowner’s Association, which represents landowners -small and large -- who own and manage more than 47
million acres in 17 SE states, told the group’s members
that “private property rights [are] the most fundamental element binding us together” (Larson 2003:
emphasis in the original).
Most landowners know in general what PPRs are, and
that level of understanding usually suffices for the situations most of us deal with. But as the landscape becomes
increasingly fragmented -- and as we get more-and-more
neighbors, and they are often “outside” (not local) folks that
we do not know -- the specifics and nuances of landowner
rights tend to take on greater importance. In this paper, I
will review the four “sticks” that represent the tenets comprising the “bundle” of PPRs, and use them as a lens
through which to examine the introduction and promotion
of, and failure-to-control, invasive exotic plant species in
the SE US.
Private Property Rights: The Precious Bundle
The principles of PPRs can be described as a bundle of
sticks that collectively represent the rights we hold so inviolate as landowners. They are exclusivity, specificity, enforceability, and transferability. Let’s examine them each.
Exclusivity refers to the nature of ownership, or more precisely, the level of exclusivity; e.g., single owner, and various partnerships such as joint tenancy and tenancy in common. Easements of ingress-and-egress and conservation
easements are examples of limited exclusivity.
Specificity refers to the particular rights assigned to the
owner. Most private landowners hold fee simple title to
their land, which represents absolute ownership, as opposed
to leasing, renting, life estates, and mineral, timber and
other rights of use. Easements are also defined in these
terms.
Enforceability refers to the means to enforce one’s landowner rights, i.e., if those rights are impugned, infringed
upon or usurped, then our legal system has, or should have,
mechanisms to protect the landowner.
Transferability refers to the ability to sell, give away, rent,
lease or otherwise divest a portion, or all, of the rights held.
These four principles overlap and underpin each other to
varying degrees depending on the specific issue.
Invasive Exotic Plant Species
Invasive, or alien, species, are: recognized by the scientific
community as harmful to “economic activity, ecosystems,
and human welfare” in the U.S. (Ecological Society of
America 2006); second only to habitat destruction and
Poster Presentations 129
degradation (and part of them) and more harmful than pollution, overexploitation and disease as a threat to imperiled
species in the U.S. (Wilcove et al. 1998); “cost[ing] the
U.S. $5.5-7.5 billion per year in economic losses” as of
1996 (Baskin 1996) a figure most likely much larger today;
a threat to homeland security by the U.S. Army War College (Pratt 2003); and “a significant component of humancaused global change” (Vitousek et al. 1997). These dire
appraisals from eminent scientists are only a sampling of
many, and include not only plants, but other invasive exotic
organisms, such as animals and fungi as well. However,
plants comprise a major part of the taxa that constitute these
threats. This paper deals only with invasive exotic plant
species.
U.S. Forest Service Research Ecologist Jim Miller’s (2003)
Nonnative Invasive Plants of Southern Forests: A Field
Guide
for
Identification
and
Control
(www.srs.fs.usda.gov/fia/manual/exotic_pest_plants.htm) is
one of the most significant conservation developments of
the last few decades. In it, Dr. Miller provides scientific, yet
practical information on invasive exotic plant species, including identification, ecology, nature of threat, principles
of control, and best-of-all, perhaps specific methods to combat them. He also covers rehabilitation of lands where infestations have been successfully diminished or eradicated.
Popular now among landowners, land managers, and a wide
array of conservationists -- from game and timber managers
and others with strong utilitarian interests, to native plant
enthusiasts, restoration ecologists, conservation biologists
and others whose chief interest is in ecosystem integrity -this book will in posterity be recognized as a paragon and
bellwether of conservation, and land restoration and management.
Professor of Silviculture Dave Moorhead and his colleagues
at the University of Georgia are also on the vanguard of the
nascent, but inevitably expanding movement to track, identify the threats and prevent new introductions of, and stymie
the spread of invasive exotic plants in the SE US. They too
are developing and using science as an underpinning from
which to launch pragmatic outreach projects (see, e.g., Evans and Moorhead 2006). Their publication Invasive Plant
Responses to Silvicultural Practices in the South (Evans et
al. 2006) is a synergistic complement to Miller’s field
guide, and ranks alongside it in importance.
This body of work is prescient and much-needed. Public
awareness of these threats are lacking (Colton and Alpert
1998), although a few land managers have for years tried to
publicize the issue in practical, landowner-oriented publications (see, e.g. Stowe 1998). If there has ever been a better
use of government tax dollars for conservation, the natural
and cultural heritage of our landscape, and our economy
than that of Miller, and Moorhead et al. described above, I
am not aware of it. If these publications are widely-used
with the attention and respect they deserve, then they will
result not in our society benefiting from the axiomatic
ounce of prevention instead of a belated pound of cure,
but rather a wise and timely ounce of prevention instead
of certain but untold megatons of cure!
A particularly insidious characteristic of invasive plant
species is that there may be a considerable time lag,
decades even, between introduction and invasion
(Baskin 1996, Randall and Marinelli 1996, David
Moorhead, personal communication, 2006). The Precautionary Principle, which suggests that we conservatively act in anticipation of harm in order to prevent it,
and that shifts the burden of proof to those who would
“develop” or alter otherwise a natural ecosystem (Noss
1999) applies especially in light of potential and known
time lags.
Your Right to Swing Your Fist Stops at My Nose:
The Law of Public Nuisance, If a Land Ethic Fails
Freyfogle (1998), in an essay titled “Land Ethics and
Private Property” published in the Society of American
Forester’s Forestry Forum on the Land Ethic: Meeting
Human Needs for the Land and Its Resources, discusses how the law of public nuisance decades ago
worked “to protect communities from bad land use,”
and maintains that the concept could today become a
“tool for discouraging environmentally unsound land
practices.” Once the legal system becomes involved, of
course, the matter lies at least partially outside the realm
of ethics, since the fear of legal penalties, rather than
any moral obligation, may be the primary impetus for
“right” behavior. Ideally, the less legal restraints we
have the better, but this works only as long as ethical
and other non-coercive societal mechanisms suffice.
The very existence of our legal system is evidence that
these mechanisms often do not -- whether we are dealing with land, or other issues. Granted, some laws are
superfluous, but most are not.
J. Owens Smith, Esquire (personal communication,
2004), who taught Natural Resources Law at the University of Georgia, introduced me to the term “Private
Property Perverts.” He defined them as landowners
who claimed total autonomy as to the use of their land.
The example he used was someone who insisted on the
unfettered right to, e.g., dump toxins in the creek because it ran through his land (i.e., he owned the land on
both banks) -- and maintained that the folks downstream must deal with it as best they can. I am a landowner myself -- owning and managing 104 acres in
Northwest Georgia -- and have strong convictions about
my private property rights, especially since, to pay the
property taxes of >$16 per acre last year, I had to borrow money from the bank! But I cannot fathom someone taking private property rights to the extreme
“perverted” end of the continuum. I have never actually
met anyone like this -- but I don’t doubt that they exist.
Poster Presentations 130
Most of the landowners in the SE US are reasonable folks
who, while standing firmly behind their PPRs, don’t irrationally insist that those rights extend to activities that
demonstrably and negatively impact their neighbor’s land
or public trust resources such as waterways. But the issue
is not as straightforward as it might seem: e.g., I maintain
I have a right to conduct prescribed fires on my land, but
my hypothetical “rurban” neighbor (Matthews 1992) may
state that s/he has a right to not be exposed to my smoke.
The intricacies of that issue are beyond the scope of this
paper, but you can see that this type polemic is often complex.
Let’s briefly and generally look at invasive plant species,
through the lens of the four tenets of PPRs:
Exclusivity: Landowners hold the PPRs, not others. I
have the right to manage my land for native species, and
to not have destructive invasive exotic species forced
upon my land.
Specificity: The specific usage rights of a tract, unless
legally designated to another party, are the landowner’s.
As in exclusivity above, I have the right to manage my
land for native species, and to not have destructive invasive exotic species forced upon my land.
Enforcability: At present there are few if any nuisance
laws to enforce, but if ethical and other societal constraints do not protect landowners from harm from other’s
actions, then this principle calls -- no, shouts -- for such
laws to be enacted to prevent harm and/or provide redress
for harm from invasive exotic species.
Transferability: The ability for landowners to transfer
their property, and at a fair price (i.e. at least market
value), can be infringed upon by other’s via invasive exotics being forced on them.
A Pernicious Example – Bicolor Lespedeza
Consider this: my neighbor plants an exotic plant species
known to be invasive, such as bicolor lespedeza
(Lespedeza bicolor) (Miller 2003, Evans et al. 2006), and
it invades my land.
My ability to burn (and thus my native vegetation) is affected, since fire -- the ecological imperative of many SE
US ecosystems (Wade et al. 2006) -- causes bicolor to
spread even worse (David Moorhead, personal communication, 2006). My soil is contaminated by alleopathic and
other chemicals produced by this pernicious invader. My
ability to practice forestry is impacted since the invader
stymies or prevents regeneration. The aesthetics of my
land are ruined since the native flora -- and thus the
beauty of seeing, hearing, smelling, touching, and tasting
them and their animal associates such as butterflies -- are
displaced.
My hunting and hiking are impacted because wildlife communities are affected, and also because I may be unable to
practically traverse my land because of the altered structure
of the vegetation. The market value and salability of my
land may also be decreased. And less tangible, but of paramount importance to me -- the land’s philosophical and
psychological values are damaged -- since to me, the goal
of restoring and maintaining a native ecosystem, with its
integrity (processes, species composition and structure)
intact -- is a primary value. Sadly, since wildlife managers
have long touted bicolor as a desirable species for wildlife,
many are loath to change their views on the matter, even in
the presence of clear and cogent evidence that it does not
bolster bobwhite quail (Colinus virginianus) populations
(which it had long-been touted to do).
From a moral standpoint, this is land-ethically wrong. John
Dewey described the concept of reflective morality, and I
(1997) reviewed the concept vis-à-vis hunting and land
management practices related to hunting. Dewey’s theory
held that we should be wary of becoming ossified in traditional morality (i.e., a practice is right simply because that’s
the way it has always been done), and that periodic reflection is vital and may from time-to-time show us that certain
values need to change. But alas, people are neo/xenophobic
by nature (Geist 1975). When the destructive impacts of
species like bicolor were unknown, then ignorance could be
claimed for introducing and promoting them, and thus
moral culpability lessened. But considering what we know
today, the blame is straightforward in my view. If reflection and accepting responsibility voluntarily continues to be
an anathema to some managers, then that is when the legal
system has its role.
What, Then, to Plant and Promote? A Simple Alternative – Go Native!
Dr. Chris Moorman (2003) of the NC State University Extension Service had the brilliant idea of appealing to folk’s
pride-of-place and heritage to encourage them to plant native species, rather than invasive exotics. His “Think
American: Manage Native Plants for Wildlife,” in Forest
Landowner magazine is a gem and I have latched onto the
concept. Moorman and his colleagues (Moorman et al.
2002) have produced a superb extension booklet, viz. Landscaping for Wildlife with Native Plants that provides detailed, yet user-friendly information on the topic.
Conclusion
As the landscape and demography of the SE US become
more fragmented and “rurban,” previously unconsidered
implications of PPRs must be openly discussed, so that the
assumed rights of one landowner do not impinge upon the
rights of others. The time to do this is now, before the
Poster Presentations 131
situation worsens. In light of the extremely destructive
nature of recently-introduced species like cogongrass
(Imperata cylindrica), this is a matter not only of PPRs,
but also one germane to the regional economic welfare of
the SE US and thus the nation. This issue leads into the
related one of urban sprawl, land use planning and the
sensitive topic of zoning. Although controversial, dealing
with these issues sooner rather than later will benefit us
all, and as change is inevitably thrust upon us, some folks
may find that they have views divergent, even diametric,
to the ones they thought they held.
Aldo Leopold (1949) pointed out that his land ethic operates like any other ethic – by “social approbation for right
actions: social condemnation for wrong actions.” Making
mistakes in land management is blameworthy but can be
understandable; denying and continuing these mistakes in
the face of the best-available-science compounds the culpability many-fold. We must either develop and implement a true land ethic, a holistic one, as Leopold implored
us to do, or we must take the less-effective, less-palatable
and more cumbersome path of legal coercion. The choice
is ours. Now or later.
This poster is dedicated to Dr. Larry Nelson (1950-2006)
of Clemson University. Larry was one of the first researchers and extension professionals to recognize and
address the threats of invasive exotic plant species to
native ecosystems, as well as forest and agricultural
lands, and his prescient and pioneering research and
outreach work was and is, not only monumentallybeneficial to the SE US landscape and its citizens per se,
but also responsible for catalyzing the efforts of others
who continue the work today. Larry, you are sorely
missed.
Literature Cited
Baskin, Y. 1996. Curbing Undesirable Invaders. BioScience. 46:10. 732-736.
Colton, T.F., and P. Alpert. 1998. Lack of Public
Awareness of Biological Invasions by Plants. Natural Areas Journal. 18:3. 262-266.
Ecological Society of America. 2006. The Ecological
Society of America Calls for Federal Leadership to
Control Invasive Species. ESA News: Media Advisory. 3 March 2006.
Evans, C.W., and D. J. Moorhead. 2006. Incorporating
Invasive Species Management into your Wildlife
Management Plan. Wildlife Trends. May 28-33.
Evans, C. W., D.J. Moorhead, C.T. Bargeron, and G.K.
Douce. 2006. Invasive Plant Responses to Silvicultural Practices in the South. The University of Georgia Bugwood Network. Tifton, GA. BW-2006-03.
52 pp.
Freyfogle, E. 1998. Land Ethics and Private Property.
Pages 49-70 in The Land Ethic: Meeting Human
Needs for the Land and Its Resources. The Society of
American Foresters Forestry Forum.
Geist, V. 1975. Mountain Sheep and Man in the Northern Wilds. Cornell University Press. 248 pp.
Larson, L.K. 2003. President’s Letter. Forest Landowner. 62:3. 4.
Leopold, Aldo. A Sand County Almanac. Oxford University Press.
Matthews, B.E. 1992. Rurbanites: The Problem of Keeping Two Roosters in the Same Henhouse. Proceedings of the Annual Conference of the Outdoor
Writer’s Association of America. Bismarck, ND. 30
June 1992.
Miller, James H. 2003. Nonnative Invasive Plants of
Southern Forests: A Field Guide for Identification
and Control. USDA Forest Service. GTR SRS-62.
99 pp.
Moorman, C., M. Johns, and L.T. Bowen. 2002. Landscaping for Wildlife with Native Plants. NCSU. 11
pp.
Moorman, C. 2003. Think American: Manage Native
Plants for Wildlife. Forest Landowner. 62:3. 5-9.
Pratt. 2003. Invasive Threats to American Homeland.
Parameters: U.S. Army War College Quarterly.
Spring. 44-61.
Noss, R.F. 1999. A Citizen’s Guide to Ecosystem Management. Biodiversity Legal Foundation and Wild
Earth Report, Special Paper 3. 33 pp.
National Woodland Owner’s Association. 2005. The Top
Ten Forestry Issues for 2005-2006. National Woodlands. 28:3. 8-11.
Randall, J. M. and J. Marinelli. 1996. Invasive Plants:
Weeds of the Global Garden. Brooklyn Botanic Garden. Handbook 149. 111 pp.
Stowe, J. 1997. Hunting in the Third Millenium And
Beyond? Quality Whitetails. 4:2. 10-14.
Stowe, J. 1998. Active Management Required on Heritage Preserves. Carolina For J. SEP. 18:9. 10-11.
Wade, D., S. Miller, J. Stowe, and J. Brenner. 2006. Rx
Fire Laws: Tools to Protect Fire: The “Ecological
Imperative.” Pages 233-262 In M.B. Dickinson, Editor. Fire in Eastern Oak Forests: Delivering Science
to Land Managers, Proceedings of a Conference;
2005 November 15-17; Columbus, OH. Gen. Tech.
Rep. NRS-P-1. Newtown Square, PA: U.S. Dept. of
Ag., Forest Service, Northern Research Station. 303
p.
Vitousek, P., C. D’Antonio, L. Loope, M. Redjanek, and
R. Westbrooks. 1997. Introduced Species: A Significant Component of Human-Caused Global Change.
New Zealand Journal of Ecology. 21:1. 1-16.
Wilcove, D., D. Rothstein, J. Dubow, A. Phillips, and E.
Losos. 1998. Quantifying Threats to Imperiled Species in the United States. Bioscience. 48:8. 607-615.
Poster Presentations 132
A Framework for Restoration: Increasing the Success of Longleaf Pine Restoration Projects
Rob Sutter1, Brett Williams2, Alison McGee3 and Michelle Creech4
1
The Nature Conservancy, Durham, North Carolina 27713, USA
2
The Nature Conservancy, Eglin AFB, 32542, Florida, USA
3
The Nature Conservancy, Darien, Georgia, 31305, USA
4
J.W. Jones Ecological Research Center, Newton, Georgia, 39870, USA
Abstract
Longleaf pine restoration projects in the southeastern US
have varied in their level of success. Projects can fall
short of success due to a lack of adequate planning, improper implementation (often related to the sequencing of
management actions) and a failure to detect and react to
changing conditions or unexpectedly high costs. A
poorly planned restoration project often results in unintended consequences, such as the elimination of desirable
species or the spread of invasive species. In response to
the growing interest in longleaf pine restoration, we developed a framework for ecological restoration that addresses many of these issues. The key components of a
framework for restoration include a thorough site assessment of current conditions, development of desired
ecological conditions and a plan to achieve them, implementation of the plan and then adaptation. The framework emphasizes the need to develop realistic desired
ecological conditions considering the initial starting point,
landscape context, desires of the landowner and what is
possible within the timeframe of the project. The framework also addresses the need to understand the relationship between the initial site condition and how to reach
the desired ecological condition, stressing the importance
of the proper sequencing of management actions. Restoration projects that follow the framework will be: 1) be
site-specific; 2) address landscape context; 3) consider
time as ecological factor; and 4) be based on realistic expectations. Four restoration case studies (Eglin AFB, FL;
Jones Ecological Research Center, GA; St. Marks NWF,
FL; and Cabin Bluff, GA) will be highlighted.
Poster Presentations 133
Repeated Fire Effects on Soil Physical Properties in Two Young Longleaf
Pine Stands on the West Gulf Coastal Plain
Mary Anne Sword-Sayer1
1
USDA Forest Service, Southern Research Station, Pineville, Louisiana, 71360, USA
Introduction
Repeated prescribed fire is a valuable tool for the management of longleaf and loblolly pine. When applied
every two to ten years, for example, prescribed fire perpetuates existing longleaf pine ecosystems (Outcalt
1997). Furthermore, the acceptance of fire as a management tool, together with recent improvements in longleaf
pine regeneration methods have aided efforts to restore
longleaf pine to its natural range (Outcalt 1997, Landers
et al. 1995). Low-intensity, prescribed fire every two to
five years is also commonly used to manage loblolly pine
on public and non-industrial, private land to reduce understory fuel and stimulate the development of wildlife
browse.
Malbis fine sandy loam complex. The Beauregard soil
forms the intermound and wetter portion of the site. The
Malbis soil forms slightly elevated pimple mounds. Site 2
has a slope of 1-5% and the soil is Ruston fine sandy loam
with some Malbis fine sandy loam and Gore very fine
sandy loam. A mixed pine-hardwood forest originally
occupied both sites. Site 1 was clearcut harvested,
sheared, and windrowed in 1991 and prescribe burned in
1993 and 1996. Understory vegetation at Site 1 is dominated by grasses. Site 2 was clearcut harvested in 1996
and roller-drum chopped and burned in August 1997.
Understory vegetation at Site 2 is dominated by woody
shrubs and herbaceous plants.
The response of understory vegetation to repeated prescribed fire over an extended period of time may affect a
suite of soil physical properties that influence both the
plant-available water holding capacity (PAWHC) and
bulk density (BD) of soil. These changes could negatively affect the sustainability of southern pine on the
west Gulf coastal plain for two reasons. First, the amount
of plant-available water in the soil during drought is already low and any further limitation would increase the
likelihood of reduced carbon fixation. Second, these soils
are often characterized by a subsurface BD that approaches the root growth-limiting value of 1.55 g/cm3
(Pritchett 1979). Fire-induced changes in soil porosity
that increase BD could restrict root system expansion and
therefore, access to water stored deep in the soil profile.
In this situation, access to deep water would depend on
interped spaces and old root channels in the subsoil (van
Lear et al. 2000). Because forest health and sustained
production are dependent on the expansion of tree root
systems and their acquisition of water and mineral nutrients, continued use of fire as a management tool requires
knowledge of its long-term effects on soil physical properties. It is hypothesized that long-term biennial prescribed fire decreases soil porosity which lowers PAWHC
and increases BD. The present objective is to summarize
the soil physical properties of two young stands of longleaf pine in response to two cycles of biennial prescribed
fire.
Treatment plots (22 x 22 m; 0.048 ha) were established
and blocks were delineated based on soil drainage and
topography. Three vegetation management treatments
were established: (1) Control (C)-- no management activities after planting, (2) Prescribed burning (B)-- plots were
burned using the strip headfire method in late spring every
two or three years, and (3) Herbicides (H)-- herbicides
were applied after planting for herbaceous and arborescent plant control. Specifically, the H plots at Site 1 were
planted in March 1997, and in May 1997 and April 1998,
sethoxydim (0.37 kg active ingredient (ai)/ha) and hexazinone (1.12 kg ai/ha) in aqueous solution were applied in
0.9-m bands over the rows of unshielded longleaf pine
seedlings. At Site 2, hexazinone (1.12 kg ai/ha) was
banded in April 1998 and 1999. At both sites in April
1998 and May 1999, triclopyr (0.0048 kg acid equivalent/
liter) was tank mixed with surfactant and water and applied as a directed foliar spray to competing arborescent
vegetation.
Materials and Methods
Two field sites are located on the Kisatchie National Forest in central LA. Three replications are located at Site 1,
and two replications are located at Site 2. Site 1 is gently
sloping (1-3%) and the soil is a Beauregard silt loam and
Recovering brush was cut by hand in February 2001. The
B plots were burned by the strip headfire method in May
1998 at Site 1, and in June 2000, May 2003, and May
2005 at both sites. Container-grown longleaf pine seedlings from a genetically improved, Mississippi seed
source (Site 1) and a Louisiana seed source (Site 2), were
planted at a spacing of 1.8 x 1.8 m in March 1997 and
November 1997, respectively. Treatment plots contained
12 rows of 12 seedlings each. The measurement plots
contained the innermost eight rows of eight seedlings in
each treatment plot.
In fall of 2004 and spring of 2006, one soil core (61 cm)
was extracted from a random location 1 m from the base
of three saplings per plot using a tractor-mounted
Poster Presentations 134
hydraulic probe equipped with an open-sided steel core
sampler (1.5 m), 4.1 cm in diameter (72 cores). One additional surface soil core (30.5 cm) was extracted per sapling (72 cores). Soil cores were stored in air-tight, plastic
liners and refrigerated until processing. From each 61 cm
soil core, three 10 cm depths increments were assessed for
physical properties. Depth increments represented the
surface soil (A horizon), the upper argillic horizon (Bt1
horizon), and the deeper argillic horizon (Bt2 horizon).
The A and Bt2 horizons were evaluated at 2-12 and 50-60
cm depths, respectively. The depth to the interface between the A, AB, E or EB horizon and the Bt1 horizon
was visually approximated. The 10 cm depth increment
beginning 2 cm beneath this interface was defined as the
Bt1 horizon. A second A horizon sample (2-12 cm) from
the 30.5 cm soil core was evaluated for soil physical properties.
The integrity of the 10 cm soil core increments was retained while two plastic rings, 1 cm in length and 4.1 cm
in diameter, were slid over the core increments. A band
saw was used to cut the ring-encased, 1 cm wide slices of
soil from the soil core increments. The two slices of soil
core from each soil core increment were placed on either a
–0.1 MPa or a -1.5 MPa equilibrated, ceramic pressure
plate. Total porosity fraction (TOP), microporosity fraction (MIP), macroporosity fraction (MAP), and PAWHC
were determined with data generated by the water retention method (Klute 1986) which requires determination of
soil water content at field capacity, –0.03 MPa (WATFC),
and permanent wilting point, –1.5 MPa (WATWP). Values of BD were determined by the core bulk density
method (Blake and Hartge 1986). The BD of the A, and
B horizons was calculated as the average of four and two
values, respectively. The TOP, MIP, MAP, and PAWHC
of the A horizon was calculated as the average of two
values.
Values of BD, WATFC, WATWP, TOP, MIP, MAP, and
PAWHC were transformed, as needed, to natural logarithms to establish normality, and evaluated by ANOVA
using a split plot in time, randomized complete block design with five blocks. Year was the whole plot effect and
vegetation management treatment was the subplot effect.
Effects were considered significant at P ≤ 0.05 unless otherwise noted. Means were compared by the Tukey test
and considered significantly different at P ≤ 0.05.
Results and Discussion
Year, block, and treatment significantly affected soil
physical properties in the A, Bt1, and Bt2 horizons. The
extent of these effects was greater in the A horizon than in
the Bt1 and Bt2 horizons. Values of BD in the A, Bt1,
and Bt2 horizons were 5, 7, and 8% less in 2006 compared to 2004 (A: 1.4 ± 0.03 g/cm3; Bt1: 1.6 ± 0.03 g/
cm3; Bt2: 1.7 ± 0.03 g/cm3). Similar trends were observed with WATFC and WATWP. It is speculated that
these effects were caused by soil water content at the time
of soil core collection. In 2004, soil cores were collected
when the soil was dry and in 2006, soil cores were collected
when the soil was wet. The Ultisol soils at the two study
sites are characterized by a suite of clay minerals dominated
by kaolinite and therefore, exhibit a low shrink-swell potential (Buol et al. 1980, Kerr et al. 1980). However, some soil
core expansion was expected after removal from the soil
profile due to the influence of organic matter and minor
clay minerals on expansion (Buol et al. 1980, Foth 1978).
Although small differences in WATFC and WATWP were
observed between years, PAWHC within a horizon was
similar between years with 19, 10 and 11% of the soil volume potentially accessible as plant-available water in the A,
Bt1, and Bt2 horizons, respectively.
Values of BD and WATFC in the A horizon were significantly affected by block. Subsequently, estimated values of
TOP, MAP, MIP, and PAWHC in the A horizon were significantly affected by block. These effects exhibited distinct site differences. Specifically, the two blocks at Site 2
were characterized by less WATFC in the A horizon compared to the three blocks at Site 1. This led to 23% less
PAWHC in the A horizon at Site 2 compared to Site 1.
Significant differences among blocks were also observed in
the Bt1 horizon. Both WATFC (P = 0.0613) and WATWP
were greater in the two blocks at Site 2 compared to the
three blocks at Site 1. This resulted in 55% less PAWHC in
the Bt1 horizon at Site 2 compared Site 1. It is proposed
that these effects were driven by soil texture and organic
matter differences between the two sites. Smaller WATFC
(24%) and MIP (24%) at Site 2 compared to Site 1 suggests
that fractions of silt and sand controlled soil physical properties in the A horizon. Larger WATWP (73%) at Site 2
compared to Site 1 suggests that the clay fraction controlled
soil physical properties in the Bt1 horizon. Site differences
in understory vegetation may have also affected soil physical properties. With more grass cover at Site 1 compared to
Site 2, for example, influences of fine root perturbation on
MIP in the A horizon may have been greater at Site 1 compared to Site 2 (Kramer 1983).
Vegetation management treatment significantly affected
WATFC and WATWP in the A horizon. Values of
WATFC were 16% less on the B and H plots compared to
the C plots. Values of WATWP on the H plots were 16%
less than that on the C plots, while WATWP was similar on
the B and C plots. As a result, estimated values of MAP,
MIP, and PAWHC in the A horizon were significantly affected by vegetation management treatment. Values of
MAP were 25% greater on the B plots compared to the C
plots, while MIP was 17% less on the B and H plots compared to the C plots (Figure 1). The effect of vegetation
management treatment on WATFC and MIP was apparent
in the PAWHC of the A horizon with 18% less PAWHC on
the B and H plots compared to the C plots (Figure 2).
Poster Presentations 135
Figure 1. Soil macroporosity (MAP) and microporosity
(MIP) of the A, Bt1, and Bt2 horizons in two stands of
young longleaf pine in response to three vegetation
management treatments. Variable means within a horizon associated with different letters are significantly
different at P = 0.05 by the Tukey test.
One significant effect of vegetation management treatment was observed in the Bt1 horizon, and two significant effects of vegetation management treatment were
observed in the Bt2 horizon. In the Bt1 horizon, the
WATWP of the H plots was greater (14%) than that of
the B plots. In the Bt2 horizon, WATFC and MIP on
the B plots were both 7% less compared to the C plots
(Figure 1). These effects on subsurface soil physical
properties, however, did not significantly influence
PAWHC in the Bt1 or Bt2 horizon (Figure 2). These
results suggest that frequent prescribed fire may affect
Plant-available water holding capacity
(% volume)
0.0
Burn
10.0
20.0
40.0
b
a
Control
Herbicide
30.0
A horizon
b
Burn
Control
Bt1 horizon
Herbicide
Burn
Control
Bt2 horizon
Herbicide
Figure 2. Plant-available soil water holding capacity
(PAWHC) of the A, Bt1, and Bt2 horizons in two
stands of young longleaf pine in response to three vegetation management treatments. Variable means within a
horizon associated with different letters are significantly
different at P = 0.05 by the Tukey test.
the physical properties that influence PAWHC in the surface soil on west Gulf coastal plain sites. After two cycles
of biennial prescribed fire, there was no evidence that these
effects had an impact on BD. The mechanism of B and H
reductions in WATFC, MIP, and PAWHC in the A horizon
may be linked to altered understory vegetation dynamics.
Significant block effects that separated soil physical properties by site support this proposition. It is hypothesized that
repeated burning in the B plots and chemical eradication of
understory vegetation in the H plots reduced fine root perturbation of the soil compared to the C plots. As the influence of fine root activity on soil porosity decreased, the
potential of the soil to store water that could be absorbed by
tree roots declined. Under normal environmental conditions, small decreases in PAWHC may not impact forest
production and health. However, when water availability is
limited by prolonged drought, small decreases in PAWHC
could create longer periods of water deficit that start earlier
in the growing season. We will continue to monitor the
long-term response of soil physical properties to B and H,
and the present observations will be combined with measurements of physiological function and biomass production
to assess the consequence of B and H on longleaf pine
physiological health.
Literature Cited
Blake, G.R.; K.H. Hartge. 1986. Bulk density. In: Klute A (ed)
Methods of Soil Analysis Part 1 physical and mineralogical
methods, 2nd ed. Madion, WI, Soil Science Society of America, Inc: 363-375.
Buol, S.W.; F.D. Hole; R.J. McCracken. 1980. Soil Genesis and
Classification, 2nd ed. Ames, IA, The Iowa State Univ. Press:
406 p.
Foth, H.D. 1978. Fundamentals of Soil Science, 6th ed. New
York, John Wiley & Sons: 436 p.
Kerr, A Jr; BJ. Griffis; J.W. Powell; J.P. Edwards; R.L. Venson;
J.K. Long; W.W. Kilpatrick. 1980. Soil survey of Rapides
Parish Louisiana. U.S. Dept. Agric. Soil Conservation Service and Forest Service in cooperation with Louisiana State
University. Louisiana Agricultural Experiment Station: 86 p.
Klute, A. 1986). Water retention: laboratory methods. In: Klute,
A (ed) Methods of Soil Analysis Part 1 physical and mineralogical methods, 2nd ed. Madison, WI, Soil Science Society of
America, Inc: 635-660.
Kramer, P.J. 1983. Water Relations of Plants. New York, Academic Press: 489 p.
Landers, J.L.; D.H. Van Lear, and W.D. Boyer WD. 1995. The
longleaf pine forests of the southeast: requiem or renaissance? J. For. 93:39-44.
Outcalt, K.W. 1997. Status of the longleaf pine forests of the
West Gulf Coastal Plain. Texas J. Sci. 49 (Supplement):5-12.
Pritchett, W.L. 1979. Properties and Management of Forest Soils.
New York, John Wiley & Sons: 500 p.
Van Lear D.H.; P.R. Kapeluck; W.D. Carroll. 2000. Productivity
of loblolly pine as affected by decomposing root systems.
For. Ecol. Manage. 138:435-443.
Poster Presentations 136
Preliminary Density Management Diagram for Naturally Regenerated Longleaf Pine
Curtis L. VanderSchaaf1, Ralph S. Meldahl2, and John S. Kush2
1
Department of Forestry, Virginia Polytechnic Institute and State University, Blacksburg, Virginia 24061, USA
2
School of Forestry and Wildlife Sciences, Auburn University, Alabama, 36849 USA
Abstract
Longleaf pine is one of the most commercially important tree species in the Gulf and Lower Atlantic Coastal
Plain regions. Density Management Diagrams are a
useful tool providing resource managers guidelines
when manipulating stand density to meet a particular
objective. This paper presents results of analyses conducted to determine the maximum size-density relationship (MSDR) and the threshold of self-thinning for
naturally regenerated longleaf pine stands in the Gulf
and Lower Atlantic Coastal Plain regions. Rather than
using regression analyses to estimate the slope of the
MSDR, an alternative procedure is presented. The alternative methodology MSDR slope estimate of -1.5942
is close to that originally proposed by Reineke while the
MSDR boundary line has an SDI of 425. The threshold
of self-thinning was determined to have an SDI of 195.
Thus, for naturally regenerated longleaf pine stands in
this region, density-dependent mortality (self-thinning)
is expected to begin at an SDI near 195.
Introduction
Stand density index (SDI) is a measure of density developed by Reineke (1933). Many Density Management
Diagrams (DMD) have been developed based on SDI
for southern pine species (Dean and Jokela 1992, Dean
and Baldwin 1993, Williams 1994, 1996, Doruska and
Nolen 1999, Dean and Chang 2002). However, none
have been created for longleaf pine (Pinus palustris
Mill.). This is a preliminary study to determine the
maximum size-density relationship (MSDR) boundary
level and slope coefficient (b), and when the Threshold
of self-thinning (Dean and Chang 2002) generally begins for naturally regenerated stands.
Reineke (1933) found when plotting the natural logarithm (Ln) of trees per acre (TPA) over the natural logarithm of quadratic mean diameter (QMD), stands selfthinned along a b of -1.6. This relationship can be algebraically manipulated to produce the equation [1]:
SDI = TPA*(QMD/10)b
[1]
The MSDR is assumed to be the maximum number of TPA
that can occur for a particular QMD throughout the Lower
Atlantic Coastal Plain and Gulf states regions in naturally
regenerated longleaf pine stands. Generally, there are three
other relative lines to the MSDR found on DMDs which are
referred to as management zones; canopy closure, full-site
occupancy, and the threshold of self-thinning. The threshold
of self-thinning management zone is important to avoid
competition-related mortality and to maintain tree vigor
(Dean and Jokela 1992). Stands managed beneath the
Threshold of self-thinning management zone would not be
expected to experience competition-induced mortality, or
density-dependent mortality (McCarter and Long 1986).
Due to time constraints, the Canopy closure and Full-site
occupancy lines are not included in this paper.
Methods
Data were obtained from naturally regenerated longleaf
pine stands in central and southern Alabama, southern Mississippi, southwest Georgia, and northern Florida. For a
more detailed description see Kush et al. (1987). Plots were
originally established from 1964 to 1967. Additional plots
have been established in stands of at least 20 years old. The
majority of the plots are 1/5th acre in size but some are
1/10th acre.
Only the ages that had no previous thinnings were included
in the b coefficient estimation, MSDR, and the Threshold of
self-thinning analyses. Average stand characteristics of the
plots and observations used in this current study are provided in Table 1. Site index was determined from an equation developed by Rayamajhi et al. (1999) using the measurement age for each individual stand closest to the base
age of 50.
In order to be consistent with the recommendation of
Weller (1987, 1991), visual inspection of all plots was conducted to ensure that only those measurement ages located
on the linear MSDR were included in determining the b
coefficient (Figure 1). Only those plots that had at least
four unthinned consecutive measurement ages were included for these analyses. Size-density trajectories (n =
Where:
b = MSDR beta coefficient.
Poster Presentations 137
Table 1. Average stand-level characteristics of the observations included in the 1.) Threshold of self-thinning analysis (n =
57 individual LnQMD-LnTPA observations from 57 plots) and 2.) b coefficient and the maximum size-density relationship
determination (n = 49 individual LnQMD-LnTPA observations from 18 plots). Where: Std dev – standard deviation, SDI –
stand density index calculated using b = 1.5942.
Variable
Age
Site index (ft)
QMD (in.)
TPA
BA (sq. ft./acre)
SDI
Age
Site index (ft)
QMD (in.)
TPA
BA (sq. ft./acre)
SDI
n
Mean
27
68
3.9
1462
102
280
57
Std dev
8
12
1.2
823
24
74
Beta coefficient and maximum size-density relationship
46
12
73
9
6.4
1.7
49
695
362
133
17
297
47
2
1.6
Determination of the threshold of self-thinning
Only those plots that had size-density trajectories with a
definite, constant curvature to the left (n = 57 plots) such
as seen in Figure 1 were included in the determination of
the Threshold of self-thinning. Table 1 gives a summary
of the stand-level characteristics for the initial age when
self-thinning began for the n = 57 plots used in determining the Threshold of self-thinning while Figure 2 shows
all measurement ages for the n = 57 plots equal to or
greater than the initial self-thinning age. The relative
Threshold of self-thinning SDI value was determined
graphically. A value was picked such that the majority of
measurement ages known to be self-thinning were above
this line (Figure 2).
Max
63
85
10.5
4385
156
485
31
49
3.8
165
99
230
88
85
12.7
2460
192
522
A
1.8
A
A
1.4
LnQMD
18 plots) were determined to have reached a MSDR if
they had two consecutive points that were self-thinning in
a straight line – a size-density trajectory moving in a
straight line (maximum size-density line) to the left and
not vertically. In the case with only two points comprising the maximum size-density line, only the latest measurement age was included; several stands had many
measurement ages that occurred along a MSDR. In these
cases, all but the first age were included in the analysis to
estimate the b coefficient.
Min
18
45
2.5
220
52
138
B
1.2
1
0.8
0.6
0.4
0.2
0
6.2
6.4
6.6
6.8
7
7.2
LnTPA
Figure 1. Demonstration of the LnQMD-LnTPA observations selected for the determination of the b coefficient (A) and the Threshold of self-thinning (B). The
line is the maximum size-density relationship for this
stand (MSDR dynamic thinning line). The last observation is not included in the determination of the b coefficient because it is falling away from the MSDR dynamic thinning line (Weller 1987).
Poster Presentations 138
7.4
4.5
4
4
3.5
3.5
3
LnQMD
LnQMD
3
2.5
2
1.5
2.5
2
1
1.5
0.5
0
4.5
5.5
6.5
7.5
8.5
9.5
LnTPA
1
4.5
5
5.5
6
6.5
7
7.5
8
LnTPA
Figure 2. Threshold of self-thinning (bottom line –
SDI value = 195) using a b coefficient of -1.5942 for
naturally regenerated longleaf pine. The upper line is
the maximum size-density line (SDI value = 425).
Only data from plots (n = 57 plots) used in determining the Threshold of self-thinning are included in the
figure.
Determination of the b coefficient and the maximum
size-density relationship
Rather than taking all LnQMD-LnTPA observations (n
= 49) occurring along an MSDR boundary and conducting regression analyses (whether linear or non-linear) of
LnTpa over LnQmd to determine an MSDR boundary
slope (Oliver and Powers 1978, Smith and Hann 1984,
Dean and Jokela 1992, Dean and Baldwin 1993,
Hynynen 1993, Williams 1994, 1996), we actually determined the slope of LnTPA (lnT) over LnQMD (lnD)
using equation [2]:
lnT 2 − lnT 1
ln D2 − ln D1
[2]
Using equation [2] provides a better quantification of
the relationship between tree density and mean diameter
across time because we do not assume that all specific
density-diameter relationships (or particular ages) occurring along a MSDR boundary are independent of one
another. Rather, the relationship at Time 2 is dependent
on the relationship at Time 1; this is the slope of the
change in TPA in relation to a change in QMD – the b
of Equation [1].
Reineke (1933) determined the b by visually placing the
MSDR boundary line above all data points and calculating the slope of this line. This method of estimating the
slope of the MSDR boundary may be biased because all
MSDR boundary observations are also treated as independent. The actual slopes of individual stand sizedensity trajectories are ignored. Similarly, when regression analyses are conducted to estimate the slope, the
Figure 3. Maximum size-density relationship (SDI =
425) using a b coefficient of -1.5942 for naturally regenerated longleaf pine. Observations are all unthinned
and thinned measurement ages from this study. n =
2107.
actual individual stand size-density trajectory slopes are
ignored. The average slope of values obtained using Equation [2] is –1.5942 (n = 49), which is close to the value proposed by Reineke (1933) and those values found by Zeide
(1985) using yield tables (USDA 1976) of naturallyregenerated stands, while the b obtained from Ordinary
least squares is -1.8714. Values from equation [2] ranged
from -1.0510 to -2.2997 which is a similar range found by
Zeide (1985) and Tang et al. (1994). Both of these b estimates (-1.8714 and -1.5942) correspond to the slope of the
MSDR species boundary line (Weller 1990). Although
statistically speaking the least-squares estimator (b = 1.8714) is unbiased when meeting standard regression assumptions, biologically it appears to be biased. Proc Model
of SAS (1988) was used to fit the equation of LnTPA over
LnQMD to determine the OLS b coefficient.
Results
Our average b coefficient value obtained from equation [2]
is very close to Reineke’s (Table 2). The MSDR species
boundary line was determined to have a SDI of 425 and the
Threshold of self-thinning was determined to have a SDI of
195. Figure 3 shows the MSDR in relation to all data (both
thinned and unthinned).
Table 2. Slope characteristics calculated using Equation [2]
for the LnQMD-LnTPA observations occurring along a
MSDR dynamic thinning line boundary. Where: Std dev –
standard deviation, value in parentheses is the standard error of the mean. n = 49.
Min
-1.0510
Poster Presentations 139
Mean
-1.5942
(0.0480)
Max
-2.2997
Std dev
0.34
Discussion
Reineke’s estimated value of b for naturally regenerated
longleaf pine was similar to ours (Table 2). Our MSDR
species boundary line was determined to have a SDI of 425
which is slightly greater than Reineke’s (1933) of 400. It
should be kept in mind that the MSDR species boundary
line is just that. It is assumed to be the maximum TPA that
can occur for a particular QMD across all naturally regenerated longleaf stands located in the Lower Atlantic Coastal
Plain and Gulf states regions. Due to differences in genetics and environmental conditions (Weller 1990, Hynynen
1993), and initial density (Weller 1990, VanderSchaaf
2003), not all longleaf stands within these regions will selfthin along the MSDR species boundary line. Weller (1990)
provides a good discussion on this topic. Most individual
stand size-density trajectories will self-thin along a line
lower than the species boundary line – what Weller called
the MSDR dynamic thinning line. It is assumed that the
Threshold of self-thinning holds constant across all individual stands.
Our relative value of the Threshold of self-thinning (195 of
425 = 46%) to the MSDR species boundary line is slightly
lower than reported for other southern yellow pine species
(Dean and Jokela 1992, Williams 1994, Dean and Chang
2002) and other North American conifers (Drew and Flewelling 1979, Long 1985, McCarter and Long 1986, Newton 1997). However, Dean and Baldwin (1993) and
Doruska and Nolen (1999) also found a relative value
(45%) for loblolly pine (Pinus taeda L.) very similar to
ours. Additional research is needed to determine the relative values of the canopy closure and the full-site occupancy (Dean and Chang 2002) management zones.
Literature Cited
Dean, T.J., and V.C. Baldwin. 1993. Using a stand densitymanagement diagram to develop thinning schedules for loblolly pine plantations. USDA For. Serv. Res. Pap. SO-275. 7
p.
Dean, T.J., and S.J. Chang. 2002. Using simple marginal analysis
and density management diagrams for prescribing density
management. South. J. Appl. For. 26: 85-92.
Dean, T.J., and E.J. Jokela. 1992. A density-management diagram for slash pine plantations in the lower Coastal Plain.
South. J. Appl. For. 16: 178-185.
Drew, T.J., and J.W. Flewelling. 1979. Stand density management: An alternative approach and its application to Douglas-fir plantations. For. Sci. 25: 518-532.
Doruska, P.F., and W.R. Nolen, Jr. 1999. Use of stand density
index to schedule thinnings in loblolly pine plantations: a
spreadsheet approach. South. J. Appl. For. 23: 21-29.
Hynynen, J. 1993. Self-thinning models for even-aged stands of
Pinus sylvestris, Picea abies, and Betula penula. Scand. J.
For. Res. 8: 326-336.
Kush, J.S., R.S. Meldahl, S.P. Dwyer, and R.M. Farrar, Jr.
1987. Naturally regenerated longleaf pine growth and yield
research. In: Phillips, Douglas R., comp. Proceedings of
the fourth biennial southern Silvicultural research conference; 1986 November 4-6; Atlanta. Gen. Tech. Rep. SE-42.
Asheville, NC: U.S. Department of Agriculture, Forest
Service, Southeastern Forest Experiment Station: 343-344.
Long, J.N. 1985. A practical approach to density management.
For. Chron. 61: 23-27.
McCarter, J.B., and J.N. Long. 1986. A lodgepole pine density
management diagram. West. J. Appl. For. 1: 6-11.
Newton, P.F. 1997. Algorithmic versions of black spruce stand
density management diagrams. The Forestry Chronicle 73:
257-265.
Oliver, W.W., and R.F. Powers. 1978. Growth models for ponderosa pine: I. Yield of unthinned plantations in northern
California. Res. Pap. PSW-133. Berkeley, CA: U.S. Department of Agriculture, Forest Service, Pacific Southwest
Forest and Range Experiment Station. 21 p.
Rayamajhi, J.N., J.S. Kush, and R.S. Meldahl. 1999. An updated site index equation for naturally regenerated longleaf
pine stands. In: Haywood, James D., comp. Proceedings
of the tenth biennial southern Silvicultural research conference; 1999 November 16-18; Shreveport. Gen. Tech. Rep.
SRS-30. Asheville, NC: U.S. Department of Agriculture,
Forest Service, Southern Research Station: 542-545.
Reineke, L.H. 1933. Perfecting a stand-density index for evenage forests. J. of Ag. Res. 46: 627-638.
SAS Institute, Inc. 1988. SAS/ETS user’s guide. Version 6. 1st
ed. Cary, N.C.: SAS Institute.
Smith, N.J., and D.W. Hann. 1984. A new analytical model
based on the -3/2 power rule of self-thinning. Can. J. For.
Res. 14: 605-609.
Tang, S., C.H. Meng, F. Meng, and Y.H. Wang. 1994. A
growth and self-thinning model for pure even-age stands:
theory and applications. For. Ecol. Manage. 70: 67-73.
USDA. 1976. Volume, yield, and stand tables for secondgrowth southern pines. USDA Misc. Publ. 50 (revision of
1929 edition).
VanderSchaaf, C.L. 2003. Can planting density have an effect
on the maximum-size density line of loblolly and slash
pine. Proceedings of the Northeastern Mensurationist Organization and Southern Mensurationists 2003 Joint Conference. pgs. 115-126.
Weller, D.E. 1987. A reevaluation of the -3/2 power rule of
plant self-thinning. Eco. Mon. 57: 23-43.
Weller, D.E. 1990. Will the real self-thinning rule please stand
up? – A reply to Osawa and Sugita. Ecology 71: 12041207.
Weller, D.E. 1991. The self-thinning rule: Dead or unsupported? – A reply to Lonsdale. Ecology 72: 747-750.
Williams, R.A. 1994. Stand density management diagram for
loblolly pine plantations in north Louisiana. South. J. Appl.
For. 18: 40-45.
Williams, R.A. 1996. Stand density index for loblolly pine
plantations in north Louisiana. South. J. Appl. For. 20:
110-113.
Zeide, B. 1985. Tolerance and self-tolerance of trees. For.
Ecol. Manage. 13: 149-166.
Poster Presentations 140
The East Gulf Coastal Plain Joint Venture: A Regional, Landscape-Scale
Approach to All-Bird Conservation
Allison Vogt1
1
East Gulf Coastal Plain Joint Venture, Auburn, Alabama, 36849, USA
Abstract
The East Gulf Coastal Plain Joint Venture (EGJV) is a
newly formed public/private partnership whose mission
is to enable integrated bird conservation at a regional
scale. The East Gulf Coastal Plain corresponds to that
portion of the Southeastern Coastal Plain Bird Conservation Region (BCR 27) west of the Georgia-Alabama
state line, and includes the western portion of the Florida panhandle, most of Alabama and Mississippi, portions of west Tennessee and Kentucky, and eastern portions of Louisiana.
The EGJV Management Board is composed of representatives from federal and state agencies, conservation
non-governmental organizations, and private industry.
Planning, implementation, monitoring, and evaluation
activities to conserve all priority birds and habitats
throughout the eco-region will begin during the summer
of 2006.
Priority habitats include coastal longleaf pine savannas,
grasslands, and maritime and bottomland forests. The most
heavily altered habitat type in this physiographic area is
pine forest, where conversion of longleaf to shorter rotation
pine species and fire suppression has changed species composition and vegetative structure. Priority species include
bobwhite quail, Swainson's warbler, Bachman's sparrow,
cerulean warbler, Bewick's wren, black rail, and reddish
egret.
The East Gulf Coastal Plain Joint Venture welcomes collaboration with all members of the bird conservation community.
Poster Presentations 141
A Continued Pinus palustris Burn Study Comparing Frequency and
Season of Fire to Basal Area Growth Loss
Ben Whitaker1, William D. Boyer2, and John S. Kush1
1
2
Auburn University, Auburn, Alabama, 36849, USA
USDA Forest Service, Southern Research Station, Auburn, Alabama, 36849, USA
Abstract
Prescribed fire has been greatly suppressed over the past
decades. Urbanization, short rotation fiber crops, false
propaganda, and mismanagement are factors in this fire
suppression. Pinus palustris, a fire dependent species,
has been taken out of its natural environment in many
places, often replaced with Pinus taeda and Pinus elliottii
in the sandy coastal plains and nutrient deficient ridge
tops and uplands farther inland. A continuing study in
Escambia County, Alabama focuses on Pinus palustris
with different burning regimens of 2, 3, and 5 year intervals conducted during both the winter and spring, as well
as a burn including the mechanical removal of hardwood
competition and a control plot where there was no intervention (Boyer 1995).
Field data is used to determine total basal area per acre
and site index in this continuing study. Three blocks containing 10 plots each with a 0.1 acre central plot (the
study area) within a 0.4 acre plot (the buffer zone) were
used to gather the data. The trees were established in
1973 from seed, liberated from the parent overstory in
1976, and the plots were established in 1984-1985 and
measured in 1987, 1990, 1994, and 2004 (Boyer 1995).
Basal area in 2004 for the spring burn was 113.32 ft2 BA/
AC, for the winter burn it was 106.39 ft2 BA/AC, for the
mechanical removal it was 116.74 ft2 BA/AC, and for the
control plot it was 108.64 ft2 BA/AC. Site index increased from year 1999 to 2004. The winter burn increased from 71 ft to 75 ft, the spring burn increased from
71 ft to 75 ft, the mechanical/burning operation increased
from 72 ft to 75 ft, and the control increased from 73 ft to
78 ft, on the average (Boyer 1995).
The results show that prescribed fire at various intervals
will increase vigor of a Pinus palustris stand by reducing
hardwood competition. The spring treatment yielded
more basal area per acre than the winter treatment. Overall, the mechanical/chemical treatment yielded the greatest amount of basal area.
When the European and British explorers arrived in
America, they often took note in their journals of the
seemingly continual smoky atmosphere (Bonnicksen
2000). This high frequency, low intensity burning regimen propagated a vast forest consisting of Pinus palustris
and a variety of understory herbaceous plants and wildlife. Today, prescribed fire is more scientific, but the
same principles are utilized. As a result of fire exclusion,
many forests are mismanaged with high fuel loads, low
reproduction rates and less diversity. This 1973 study by
Boyer focuses on different seasons of burns at varying
frequencies, plus mechanical/chemical treatments after
burning. The objective is to observe the amount of residual hardwood competition and the effects of fire on Pinus
palustis basal area growth loss.
Methods
The study site was in the Escambia Experimental Forest
in Escambia County, Alabama. The trees originated from
the 1973 seed crop and were regenerated using a shelterwood system, which was eventually removed in 1976 to
release the seedlings. Three blocks containing 10 plots
each were established with 4 square chain gross plots (0.4
acre). Within these plots, 1 square chain net plots (0.1
acre) were established and thinned to leave 400 trees/acre,
or 40 trees/net plot. The different treatments included
seasonal burns during the winter and the spring at 2, 3,
and 5 year frequencies plus an unburned control. Mechanical (axe) and chemical (glyphosate) treatments to
remove any hardwood over 4 feet in height were added to
the winter and spring burns at the 3 year interval, and
there was an unburned control, including the mechanical/
chemical treatment. This accounts for the 10 treatments
in every block (Boyer 1995). Fire type, flame length, air
temperature, relative humidity, fire-line intensity, and
crown scorch were measured and recorded to provide data
on the prescribed burns due to the variable nature of fire
depending on the environmental and weather conditions
of a particular day (Boyer 1995).
Results
Introduction
The 1999 treatment resulting in the greatest amount of
basal area/acre was the mechanical/chemical treatment
without any burning. The lowest amount of basal area
was the winter fire with a 3 year interval. The winter fires
resulted in 83.14 ft2 BA/AC for the 2 year interval, 80.90
ft2 BA/AC
Poster Presentations 142
Prescribed fire is as much a part of the history of the
South as Pinus palustris was centuries ago. Indians were
known to set fire to the woods to provide expanded and
better hunting habitat, increase their own food forage, and
to open the understory for ease of travel (Pyne 1982).
for the 3 year interval, and 81.53 ft2 BA/AC for the 5 year
interval. The spring fires resulted in 82.96 ft2 BA/AC for
the 2 year interval, 82.59 ft2 BA/AC for the 3 year interval, and 92.95 ft2 BA/AC for the 5 year interval. The
spring fire with a 3 year interval followed with the mechanical/chemical treatment had 94.05 ft2 BA/AC, and the
winter fire on the 3 year interval had 83.50 ft2 BA/AC.
The lone treatment of mechanical/chemical without fire
yielded 103.73 ft2 BA/AC, and the control plot had 86.97
Average site index (SI) increased in the 5 year span between year 1999 and year 2004. The average SI for the
1999 winter burns was 71 ft, and in 2004 it was 75 ft. In
the spring burns, the average SI was 87 ft in 1999, and 75 ft
in 2004. The burning plus mechanical/chemical treatment
had an average SI of 72 ft in 1999, and 75 ft in 2004. The
control plot also increased from an SI of 73 ft in 1999 to 78
ft in 2004 (Figure 3).
Figure 2. Avergae Basal Area for Average of Treatments
Figure 1. Average Basal Area for all treatments in 1999 and
2004
140
120
140
100
120
1999
80
100
80
60
1999
60
2004
40
2004
20
40
l
Co
nt
ro
ur
n
M
ec
ha
ni
ca
l/B
Sp
rin
g
W
in
te
rB
M
/
Co C
nt
ro
l
W
in
te
r
W 2
in
te
r
W 3
in
te
r
Sp 5
rin
g
Sp 2
rin
g
3
S
Sp prin
rin g
g 5
3
M
W /C
3
M
/C
0
Bu
r
ur
n
n
0
20
Tre atme nt
All data gathe re d fro m the Es c ambia Expe rime ntal
Fo re s t, Es c ambia Co unty, AL
Tre atme nt
All data gathe re d fro m the Es c ambia Expe rime ntal
Fo re s t, Es c ambia Co unty, AL
Figure 3. Average Site Index for Average of Treatments
The average of the 1999 treatments is 1.86 ft2 BA/AC
for the winter burn, 87.16 ft2 BA/AC for the spring
burn, 93.76 ft2 BA/AC for the spring and winter burn
including the mechanical/chemical treatments, and the
control plot was 86.97 ft2 BA/AC (Figure 2).
The average of the 2004 treatments is 106.39 ft2 BA/AC
for the winter burn, 113.32 ft2 BA/AC for the spring
burn, 116.74 ft2 BA/AC for the spring and winter burn
including the mechanical/chemical treatments, and
108.64 ft2 BA/AC for the control plot (Figure 2).
90
80
70
60
50
40
30
20
10
0
1999
Co
nt
ro
l
ur
n
M
ec
ha
ni
ca
l/B
Bu
rn
Sp
rin
g
B
ur
n
2004
W
in
te
r
In the 2004 continuation of the study, the mechanical/
chemical treatment had the greatest amount of basal
area/acre. The least amount was seen in the winter burn
with the 3 year interval. The winter fires resulted in
108.90 ft2 BA/AC for the 2 year interval, 102.23 ft2 BA/
AC for the 3 year interval, and 108.05 ft2 BA/AC for
the 5 year interval. The spring burns resulted in 106.88
ft2 BA/AC for the 2 year interval, 109.31 ft2 BA/AC for
the 3 year interval, and 123.77 ft2 BA/AC for the 5 year
interval. The treatment including fire on a 3 year interval in the spring followed by a mechanical/chemical
treatment yielded 109.54 ft2 BA/AC while the winter
burn of the same frequency and follow up treatment
yielded 112.55 ft2 BA/AC. The mechanical/chemical
treatment without fire yielded 128.14 ft2 BA/AC, and
the control plot had 108.54 ft2 BA/AC (Figure 1).
Tre atme nt
All data gathe re d fro m the Es c ambia
Expe rime ntal Fo re s t, Es c ambia Co unty, AL
Conclusion
The spring burns, or growing season burns, eliminate
more hardwood competition than the winter burns, or
dormant season burns. This is because the spring fires
burn at a higher temperature during warmer weather compared to less intense fires during the colder winter
weather. There is also a greater fuel load provided by the
herbaceous and woody plants that have broken bud. Because the buds have broken on the herbaceous and woody
plants, more of these are eliminated as well.
Poster Presentations 143
The mechanical/chemical treatments, both with and without
fire, resulted in the best increment growth because there
was much less stress to the residual stand of trees. When a
stand is burned, the competition is stressed to a point were it
cannot compete with the crop trees. However, even though
Pinus palustris tolerates fire, this does not mean its growth
is not affected. Fire affects all species, but some species are
able to resist it more efficiently. Pinus palustris is one of
these species. Because a mechanical treatment, chemical
treatment, or a combination of the two treatments is far
greater in cost to the landowner, fire is a better choice for
managing competition. This is true in the short run and in
the long run. Argument could be made that the mechanical/
chemical treatment is best early on because no stress is put
on the desired crop trees. However, the economics for this
are not sound and will not pay off because of the large cost
difference between this and fire, with fire being less expensive. For someone who is convinced that the results in the
long run will be best if fire is not used, the silvics of Pinus
palustris must be examined. The species must have bare
mineral soil to naturally regenerate. When fire is excluded,
a thick litter layer and duff layer accumulate on the forest
floor, making it nearly impossible for regeneration. There
is also a richness of understory species in ecosystems allowing fire. A number of grasses, plants and herbaceous materials will only survive in fire maintained landscapes (Kush
et al. 1999). Consequently, the plots without burning had
the fewest number of species with a higher percent of
woody species, and the winter burn had the highest number
of species (Kush et al. 2000).
Basal area was observed to be less in the 2 year interval
burns than in the 3 and 5 year intervals. This is due to the
greater frequency of stress placed on the trees, slowing
vigor, productivity, and growth. The results show that the
best outcome is seen when the stand was treated with prescribed fire at the 5 year interval. The 5 year interval reduced the stress and competition, and increased growth
when compared to the 2 and 3 year interval burning periods.
Negative effects of fire did not appear to affect the height
and diameter growth past the age of 24 years. However,
Boyer found negative results with basal area and volume
growth up to the age 30 years from prescribed fire (Kush et
al. 1999).
Different sites provide varying results in growth. Because
of this, individual sites must be observed and treated separately. A highly productive site, or one with a high site
index, will facilitate more growth to every herbaceous and
woody plant. This means that competition will grow more
vigorously and has a greater chance to compete with and
subdue the crop trees. Conversely, a low productivity site
cannot support as many fast growing competitors. Therefore, a high site index should be burned more frequently in
the early stages of the stand initiation phase of
development. This practice limits competition and provides the stand with the best chance of success. A low
site index should be burned less frequently, but regularly
enough to eliminate the woody competition.
As a stand matures and reaches the stem exclusion phase
of development, crown closure of the canopy will provide
the shade necessary to exclude many shade intolerant
competitors from the stand. As this occurs, fire can be
used as a treatment less frequently, which will limit the
stress to annual growth. However, it still must be used
with enough frequency to keep the litter and duff layers at
moderate levels and to protect the residual stand from
destructive wild fire. Since any choice can have both
beneficial and unwanted, unproductive effects, there are
advantages and disadvantages to prescribed fire. The site
characteristics, site index, landowner objectives, economics, and risks involved will all affect the usefulness of fire.
Acknowledgements
The authors wish to thank George Ward and Ronald
Tucker for their work on the Escambia Experimental Forest, and the cooperation of the T.R. Miller Mill Company.
Literature Cited
Boyer, WD. 1995. Progress Report: Timing of Prescribed Fire for Optimum Hardwood Control and
Minimum Impact on Pine Growth. Escambia
Experimental Forest, Principle Silviculturist.
U.S. Forest Service (FS-SO-4105-2.25, Problem
2).
Bonnicksen, T.M. 2000. America’s Ancient Forests:
From the Ice Age to the Age of Discovery. New
York, NY: John Wiley and Sons, Inc. 594 p.
Kush, J.S., Meldahl R.S., and Boyer W.D. 1999. Understory Plant Community Response after 23 years
of Hardwood Competition Control Treatments in
Natural Longleaf Pine (Pinus palustris) Forests.
Can. J. For. Res.29: 1047-1054.
Kush, J.S., Meldahl R.S., and Boyer W.D. 2000. Understory Plant Community Response to Season of
Burn in Natural Longleaf Pine Forests Pages 3339 in W. K. Moser and C. F. Moser, editors. Proceedings of the 21st Tall Timbers Fire Ecology
Conference, Fire and Forest Ecology: Innovative
Silviculture and Vegetation Management. Tall
Timbers Research Station, Tallahassee, FL.
Pyne, S.J. 1982. Fire in America. A Cultural History of
Wildland and Rural Fire. NJ: Princeton University Press, pp 143-160.
Poster Presentations 144
Landscape Scale Ecosystem Classification in Longleaf Pine Forests of the
Talladega Mountains, Alabama
Brent Womack1 and Robert Carter2
1
2
Georgia Department of Natural Resources, Dublin, Georgia, 31040, USA
Department of Biology, Jacksonville State University, Jacksonville, Alabama, 36265, USA
Introduction
Table 1. Mean of diagnostic environmental variables for
pine dominated sites in the Horseblock LTA.
Montane longleaf pine (Pinus palustris) ecosystems are
found in portions of northern Georgia and Alabama.
They are characterized by a mixture of Appalachian,
Piedmont, and Coastal Plain plant species. Vegetation
surveys have been conducted in areas such as Forest
McClellan, AL (Maceina et al. 2000) and Rome, GA
(Lipps and Deselm 1969), but there have been no attempts to study the interrelationship between forest
communities, soils, and landform variables. The objective of the study was to identify ecological land units
based on the discriminating vegetation, soils, and landform features in the Horseblock Mountain Landtype
Association (LTA) on the Talladega National Forest.
Methods
The study area was Horseblock Mountain Landtype
association located within the Shoal Creek Ranger District of the Talladega National Forest. The elevations in
the study site ranged from 600 feet above mean sea
level (msl) to over 1900 feet above msl. Topography
ranges from rolling hills reminiscent of the Piedmont
region to ridges with slopes approaching 70 percent. In
the summer of 2004, 44 plots were established in suitable forested sites. The sites were free of recent disturbance with the exception of fire. Tree, sapling, seedling,
and herbaceous strata were sampled following the
Carolina Vegetation Survey protocol (Peet et al. 1998).
Soils samples were collected by horizon from four locations within the plot to determine soil horizon depth and
chemical and textural properties. Landform variables
sampled included slope gradient, aspect, and landform
index (LFI).
Ecological land units were delineated through ordination and cluster analysis of presence/absence data. The
ordination programs employed were de-trended correspondence analysis and non-metric multidimensional
scaling (McCune and Grace 2002). Cluster analysis
was through TWINSPAN (Hill 1975, McCune and
Grace 2002). Environmental variables were related to
the ecological units through stepwise discriminant
analysis (p=0.10).
Longleaf Pine Longleaf pine - Longleaf Pine
Dwarf Huckleberry Shortleaf Pine - - Partridge
- Elliott’s Bluestem
Muscadine
Pea - Trefoil
Land19.36
form*
Index
27.1
Slope
A Horizon
Depth
4.42
(cm)*
B Horizon
Depth
13.64
(cm)
B Horizon
Ca (lbs/ac)
85.38
*
*Statistically Significant (p=0.10)
17.49
18.59
25.8
24.13
2.91
2.83
11.45
8.04
40
82.5
Table 2. Mean of diagnostic environmental variables of
hardwood dominated sites in the Horseblock Mountain
LTA.
American Hornbeam Chesnut Oak - Oakleaf
Sweetgum - Hayscented
Hydrangea - Wild Yam
Fern
Landform
40.75
34.26
Index Slope*
49.25
7.33
3.69
7.29
14.22
16.83
25
36.67
2.83
1
A Horizon
Depth (cm)
B Horizon
Depth (cm)
B Horizon
Mg (lbs/ac)*
B Horizon P
(lb/ac) *
*Statistically Significant (p=0.10)
Poster Presentations 145
Results
Literature Cited
Five landscape scale ecosystems were identified with a
unique species assemblages. Two were in hardwood dominated sites while three were in pine dominated sites. Each
community had diagnostic soil and landform variables as
well as diagnostic species (Tables 1, 2, 3, and 4). Species
abundant in all pine dominated sites included Diospyros
virginiana, Solidago odora, Quercus velutina, Andropogon
virginicus, Sorghastrum nutans, Pteridium aquilinum, and
Vaccinium pallidum. Species abundant in all hardwood
dominated sites included Quercus alba, Liriodendron
tulipifera, and Chasmanthium sessiliflorum.
Hill, M. O. 1979. TWINSPAN: A FORTRAN program
for arranging multivariate data in an ordered two-way
table by classification of the individuals and attributes. Dept. of Ecology and Systematics, Cornell University. Ithaca, NY. 90
Lipps, E.L. and H.R. Deselm. 1969. The vascular flora
of the Marshall Forest, Rome, Georgia. Castanea
34:414-432.
Maceina, E.C., J.S. Kush, and R.S. Meldahl. 2000.
Vegetational survey of a montane longleaf pine community at Fort McClellan, Alabama. Castanea 65:
147-154.
McCune, B. and J.B. Grace. 2002. Analysis of ecological
communities. MjM Software Design. Gleneden
Beach, OR. 300 pp.
Peet, R. K., T.R. Wentworth, and P.S. White. 1998. A
flexible multipurpose method for recording vegetation composition and structure. Castanea 63: 262274.
SPSS. 1996. SYSTAT 6.0 for Windows. Chicago IL.
Conclusions
The forest communities of the Talladega Mountain Range
represent a unique ecosystem blending of Coastal and Piedmont/Appalachian species. A complex interaction of fire
and landform has shaped the forest. The presence of longleaf pine on the upland sites indicates that fire played an
historical role in determining plant species distribution.
Table 3. Community type, habitat, and diagnostic species (percentage of plots present) for Hardwood Dominated Sites
in the Talladega Mountain Region of Northeast Alabama (1, 2, 3, 4, indicate tree, sapling, seedling, and herb, respectively).
Community:
Chesnut Oak - Oakleaf Hydrangea - Wild
Yam
American Hornbeam - Sweetgum - Hayscented Fern
Habitat:
Steep lower slopes near streams
Stream borders and small alluvial flats
43
40
14
20
Diagnostic Species:
Vitis rotundifolia 3
Chamaecrista nictitans spp. nictitans 4
Quercus prinus 2, 3
100
Antennaria plantaginifolia 4
57
Amphicarpa bracteata 4
71
20
Dioscorea villosa 4
71
20
Liquidambar styraciflua 2
14
100
Chasmanthium laxum 4
14
80
Dennstaedtia punctilobula 4
80
Fraxinus pennsylvanica 3
Acer barbatum 3
Carpinus carolina 3
14
43
Poster Presentations 146
100
100
80
Table 4. Community type, habitat, and diagnostic species (percentage of plots present) for Pine Dominated Sites in the
Talladega Mountain Region of Northeast Alabama (1, 2, 3, 4, indicate tree, sapling, seedling, and herb, respectively).
Longleaf Pine - Dwarf Huckleberry - Longleaf pine - Shortleaf Pine - Longleaf Pine - Partridge
Elliott’s Bluestem
Muscadine
Pea - Trefoil
Habitat:
Diagnostic Species:
Pinus palustris 2
Andropogon gyrans 4
Carex jamesii 4
Pinus echinata 2
Tephrosia virginiana 4
Vitis rotundifolia 3
Gaylussacia dumosa 3
Hypoxis hirsuta 4
Chamaecrista nictitans spp.
nictitans 4
Rocky upper slopes, deepest A horizon, low CA
Rolling hills, shallow soil,
higher CA
Mountain tops, shallow
soils
100
66
75
58
100
33
83
25
80
10
20
90
40
100
50
10
100
25
25
50
100
25
17
30
100
75
Desmodium marilandicum 4
75
Desmodium obtusum 4
75
Desmodium paniculatum 4
75
Quercus prinus 2, 3
33
60
Liquidambar styraciflua 2
8
10
Acknowledgements
This research was supported by Jacksonville State University, US Forest Service, and a grant from the National Fish
and Wildlife Foundation
Poster Presentations 147
50
Fuel Loads, Tree Community Structure, and Carbon Storage in Mountain
Longleaf Pine Stands Undergoing Restoration
Rebecca Worley1 and Martin Cipollini2
1
Warren-Wilson College, Asheville, North Carolina, 28815, USA
2
Berry College, Mount Berry, Georgia, 30149, USA
Abstract
Longleaf pine (Pinus palustris), once dominant in southern forests, has been reduced in range from 37 million ha
to less than 1.2 million ha due in large part to fire suppression. Healthy longleaf ecosystems are characterized
by an open under-story dominated by fire-resistant
grasses. Current longleaf stands are found mostly in the
costal plain but a few tracts remain on south-facing slopes
in mountainous regions of Alabama and Georgia. Little
research exists on montane longleaf and the majority of
what is known is based on studies of degraded stands.
The Berry College Longleaf Pine Project began in 1999
with the primary goal of restoring relict longleaf stands
on Lavender Mountain, Floyd County, Georgia. Since
then, restoration burns and hardwood control have been
initiated in mature stands, and planting has been initiated
in clear- and selective-cut areas. The goals of the current
research project were to assess changes in tree community
structure and fuel biomass in mature stands in the two
years following an April 2004 prescribed burn, and to
determine total biomass across all managed areas. A planar transect method was used to quantify downed woody
fuels, litter, herbs, duff, shrubs, and small trees. Large
tree biomass and community data were obtained using the
point-centered quarter method. Since the 2004 burn,
longleaf importance and the biomass of several woody
fuel categories, litter, and small trees increased, whereas
duff biomass decreased. Total biomass integrated across
the 309-acre managed area was over 15,000 tons (ca. 50
tons/acre).
Poster Presentations 148
Participant List
Robert N. Addington
P.O. Box 52452
Fort Benning, GA 31995
The Nature Conservancy
Phone: 706-544-7515
E-mail: [email protected]
Sara B. Aicher
US Fish And Wildlife Service
Route 2, Box 3330
Folkston, GA 31537
Phone: 912-496-7366
Fax: 912-496-3332
E-mail: [email protected]
Julius F. Ariail
Julius Ariail, Photographer
5802 Long Pond Road
Lake Park, GA 31636-2712
Phone: 229-563-0209
E-mail: [email protected]
Todd A. Aschenbach
University of Kansas
1200 Sunnyside Avenue
Lawrence, KS 66045
Phone: 785-864-5690
E-mail: [email protected]
Mark Atwater
Weed Control Unlimited, Inc.
Donalsonville, GA 39845
Phone: 229-524-6187
E-mail: [email protected]
Chadwick R. Avery
Eglin AFB
107 Highway 85 North
Niceville, FL 32578
Phone: 850-883-1141
E-mail:
[email protected]
Jason T. Ayers
176 Croghan Spur Rd., Suite 200
Charleston, SC 29407
Phone: 843-727-4707
Fax: 843-727-4218
E-mail: [email protected]
Charles Bailey
3005 Atlanta Highway
Gainesville, GA 30507
Phone: 770-531-6043
Kawika Bailey
865 Geddie Road
Tallahassee, FL 32304
Phone: 850-926-3073
Jill Barbour
USDA Forest Service
National Seed Laboratory
Dry Branch, GA 31020
Phone: 478-751-3553
E-mail: [email protected]
Chris Blackford
Roundstone Native Seed, LLC
9764 Raider Hollow Road
Upton, KY 42784
Phone: 270-531-2353
E-mail: [email protected]
Jason Blanton
1171 NE Daylily Ave.
Madison, FL 32340
Phone: 850-973-2967
Jim Bates
U.S. Fish and Wildlife
West Georgia Ecological Services
Ft. Benning, GA 31995
C.J. Blanton Jr.
117 NE Daylily Ave.
Madison, FL 32340
Phone: 850-973-2967
C. Victor Beadles
Beadles Lumber Company
P.O. Box 3457
Moultrie, GA 31776
Phone: 229-985-6996
E-mail: [email protected]
Dave G. Borden
P.O. Box 59
Pine Level, AL 36065
Phone: 334-224-1454
Fax: 334-420-2779
E-mail: [email protected]
Barbara Bell
USDA Forest Service
Calcasieu Ranger District
Boyce, LA 71409
Phone: 318-793-9427
Fax: 318-793-9430
E-mail: [email protected]
Lynda P. Borden
P.O. Box 59
Pine Level, AL 36065
Phone: 334-322-1486
E-mail: [email protected]
Wayne Bell
International Forest Co.
1265 GA Hwy 133 N.
Moultrie, GA 31768
Phone: 800-633-4506
E-mail: [email protected]
Amanda Bessler
Joseph W. Jones Ecological Research
Route 2
Newton, GA 39870
Phone: 229-734-4706
E-mail: [email protected]
Arvind A. Bhuta
Virginia Tech
442 E. Roanoke St. Apt. E
Blacksburg, VA 24060
Phone: 334-559-3265
E-mail: [email protected]
149
Larry Boulineau
P.O. Box 1033
Louisville, GA. 30434
Ft. Gordon, Georgia
Phone: 706-791-9927
E-mail: [email protected]
Elizabeth Bowersock
Auburn University
602 Duncan Drive
Auburn, AL 36849
E:mail: [email protected]
Allen Braswell
Installation Forester
Fort Gordon, GA 30905
Phone: 706-791-9932
E-mail: [email protected]
Hal Brockman
USDA Forest Service
1400 Independence Ave., SW.
Washington DC 20250-1123
Phone: 202-205-1694
Fax: 202-205-1271
E-mail: [email protected]
Dale Brockway
520 DeVall Dr.
Auburn, AL 36849-8700
Phone: 334-826-8700
E-mail: [email protected]
MaryJo Broussard
P.O. Box 8382
Mobile, AL 36689
Phone: 251-342-9233
Joyce Marie Brown
4000 Central Florida Blvd.
Orlando, FL 32816
Phone: 407-488-5590
Randy Browning
U.S. Fish and Wildlife Service
113 Fairfield Drive
Hattiesburg, MS 39402
Phone: 601-606-2622
E-mail: [email protected]
Frank Burly
1337 Long Horn Road
Middleburg, FL 32068
Phone: 904-291-5531
Shan Cammack
2065 US Hwy 278 SE
Social Circle, GA 30025
Phone: 770-918-6411
James Y. Campbell
1 Hawkinshurst Lane
Hopkins, SC 29061
Phone: 803-776-3671
E-mail: [email protected]
Steve Carpenter
3021 145th Road
Live Oak, FL 32060
Phone: 386-208-1460
Robert Carter
P.O. Box 122
Jacksonville, AL 36265
Phone: 256-782-5144
Paul Catlett
Camp Blanding
5629 SR 16 West Bldg 4540
Starke, FL 32091
Phone: 904-682-3453
E-mail: [email protected]
Jack Chappell
Meeks Farms & Nursery, Inc.
187 Flanders Rd.
Kite, GA 31049
Phone: 478-237-6863
Sabrina Clark
US Fish And Wildlife
6578 Dogwood View Pkwy
Jackson, MS 39213
Phone: 601-321-1135
E-mail: [email protected]
Allison Cochran
P.O. Box 278
Double Springs, AL 35553
Phone: 205-489-5111
E-mail: [email protected]
Joe Cockrell
176 Croghan Spur Rd., Suite 200
Charleston, SC 29407
Phone: 843-727-4707
Fax: 843-727-4218
E-mail: [email protected]
Frank T. Cole
1670 Meridian Rd.
Thomasville ,GA 31792
Phone: 229-377-8050
June Cole
1670 Meridian Rd.
Thomasville, GA 31792
Phone: 229-377-8050
Kristina F. Connor
USDA Forest Service
520 Devall Dr.
Auburn, AL 36849
Phone: 334-826-8700
Fax: 334-821-0037
E-mail: [email protected]
Matthew Corby
Camp Blanding
5629 SR 16 West Bldg 4540
Starke, FL 32091
Phone: 904-682-3243
Email:[email protected]
Ellen C. Corrie
1052 State St. NW
Atlanta, GA 30318-5345
Phone: 404-873-4957
E-mail: [email protected]
Tom Counts
P.O. Box 278
Double Springs, AL 35553
Phone: 205-489-5111
E-mail: [email protected]
F.G. Courtney
National Wildlife Federation
1330 W. Peachtree Street, Suite 475
Atlanta, GA 30309
Phone: 404-876-8733
E-mail: [email protected]
Jim Cox
Tall Timbers
13093 Henry Bendol Dr.
Tallahassee, FL 32301
Phone: 850-942-2487
E-mail: [email protected]
John M. Cox
Lolly Creek, LLC
1684 Wrights Chapel Road
Sumner, GA 31789
Phone: 229-776-2300
Fax: 229-776-8901
E-mail: [email protected]
Barbara Crane
1720 Peachtree Rd. NW Suite 816N
Atlanta, GA 30309
Phone: 404-347-4039
Jenny Crisp
Grey Moss Plantation
Lee County, AL 36849
Mike Connor
Scott Crosby
J.W. Jones Ecological Research Center 390 Holloway Road
Newton, GA 39870
Flornhome, FL 32140
Phone: 386-329-2552
150
Robert Cross
Rt. 1 Box 1097
Shellman, GA 39886
Phone: 800-554-6550
Stephen Crown
South Carolina Forestry Commission
353 Firetower Rd.
Orangeburg, SC 29118
Phone: 803-534-3543
E-mail: [email protected]
Lloyd Culp
U.S. Fish and Wildlife Service
7200 Crane Lane
Gautier, MS 39553
Phone: 228-497-6322
Fax: 228-497-5407
E-mail: [email protected]
Carol M. Daugherty
5297 Morgan
Milton, FL 32570
Phone: 850-777-9382
James Davis
Engineering & Environment, Inc.
20 Ces/cev
Shaw AFB, SC 29152
Phone: 803-895-9990
Fax: 803-895-5103
E-mail: [email protected]
Ted Devos
Bech of Devos Forestry & Wildlife
217 S. Count St
Montgomery, AL 36104
Phone: 334-269-2224
E-mail: [email protected]
David Dickens
UGA WSFNR
P.O. Box 8112 GSU
Statesboro, GA 30460
Phone: 912-681-5639
E-mail: [email protected]
Dan Dumont
Alabama Forest Resources Center
8 St. Joseph St., 2nd Floor
Mobile, AL 36602
Phone: 251-433-2372
E-mail: [email protected]
David Dyson
J.W. Jones Ecological Research Center
Rt. 2, Box 2324
Newton, GA 39870-9651
Phone: 229-734-4706
E-mail: [email protected]
Calvin Ernst
Ernst Southern Native Seeds
9006 Mercer Pike
Mcadville, PA 16335
Phone: 814-671-4840
E-mail: [email protected]
Thomas L. Eberhardt
2500 Shreveport HWY
Pineville, LA 71360
Phone: 318-473-7274
E-mail: [email protected]
Marcia L. Ernst
Ernst Southern Native Seeds
9006 Mercer Pike
Live Oak, FL
Phone: 814-720-2142
E-mail: [email protected]
Lori G Eckhardt
Auburn University
3301 School Of Forestry And Wildlife
Auburn, AL 36849
Phone: 334-844-2720
Fax: 334-844-1084
E-mail: [email protected]
Sharonte Edmond
Georgia Forestry Commission
3561 Georgia Highway
Camilla, GA 31730
Phone: 229-522-3580
E-mail: [email protected]
Neal Edmondson
Georgia Forestry Commission
P.O. Box 819
Macon, GA 31202
Phone: 478-751-3332
E-mail: [email protected]
James D. Elledge
212 Myrick Road
Lumberton, MS 39455
Phone: 601-796-5494
Fax: 601-796-5494
E-mail: [email protected]
Glenn E. Elms
394 FM 1375 West
New Waverly, TX 77358
Phone: 936-344-6205
E-mail: [email protected]
Alan Emmons
Southern Forestry Consultants, Inc.
305 W. Shotwell Street
Bainbridge, GA 39817
Phone: 229-220-1790
E-mail: [email protected]
151
Becky Estes
School of Forestry and Wildlife Sciences
Auburn University
Auburn, AL 36849
E-mail: [email protected]
Ray Evans
Northern Bobwhite Conservation Initiative
1995 Halifax
Holts Summit, MI 65043
Phone: 573-896-4836
E-mail: [email protected]
Bo Ewing
884 US HWY 280 E
Americus GA 31709
Phone: 229-942-2336
:
Greg Findley
3561 Hwy. 112
Camilla, GA 31730
Georgia Forestry Commission
Phone: 229-522-3580
Laura Fogo
P.O. Box 9
Bisco, NC 27209
Phone: 910-695-3323
Charles W. Fore, Jr.
State of Georgia
77 25th Ave.
Eastman, GA 31023
Phone: 478-374-6981
E-mail: [email protected]
Samuel Fowler
Auburn University
109 Duncan Hall
Auburn, AL 36849
Phone: 334-844-5546
William Frankenberger
Florida Dept. of Military Affairs
236 South Blvd.
Avon Pk. Air Force Rng., FL 33825
Phone: 863-452-4236
E-mail: william.frankenberger@avonpa
Robert M. Franklin
Clemson University Cooperative Extension
P.O. Drawer 1086
Walterboro, SC 29488
Phone: 843-549-2595
Fax: 843-549-2597
E-mail: [email protected]
Conrad J. Franz
73 Main Blvd.
Trenton, NJ 08618
Phone: 690-882-1519
Adam Gabryelski
Fort Gordon Forestry
143 Stone Mill Dr.
Augusta, GA 30907
Phone: 706-791-9930
E-mail:[email protected]
Stephen D. Gantt Jr.
USFS
9901 Hwy 5
Brent, AL 35034
Phone: 205-926-9765
E-mail: [email protected]
Jeff Gardner
45 HWY 281
Heflin, AL 36264
Phone: 256-463-2273
Bill Garland
P.O. Box 5087
Fort McClellan, AL 36205
Phone: 256-848-6833
E-mail: [email protected]
Robert P. Gehri
P.O. Box 2641
Birmingham, AL
Phone: 912-579-6518
Traci George
AL Wildlife & Freshwater Fisheries
64 North Union St.
Montgomery, AL 36130
Phone: 334-7-353-0503
E-mail: [email protected]
John Gilbert
School of Forestry and Wildlife Sciences
Auburn University
Auburn, AL 36849
Dean Gjerstad
School of Forestry and Wildlife Sciences
Auburn University
Auburn, AL 36849
Susan Glenn
51 Harbour Passage East
Hilton Head, SC 29926
Phone: 843-842-9696
Jeff Glitzenstein
Tall Timbers Research Station
Tallahassee, FL 32312
Jeffery C. Goelz
USDA FS
2500 Shreveport Highway
Pineville, LA 71360
Phone: 318-473-7227
Fax: 318-473-7273
E-mail: [email protected]
Doria Gordon
The Nature Conservancy
Gainseville, FL 32601
Chris P. Gowen
Toledo Manufacturing Co., Inc.
P. O. Box 488
Folkston, GA 31537
Phone: 912-496-7343
Fax: 912-496-4074
E-mail: [email protected]
Jennifer Greene
3125 Conner Blvd.
Tallahassee, FL 32301
Phone: 850-414-8602
Converse Griffith
USDA Forest Service
Calcasieu Ranger District
Boyce, LA 71409
Phone: 318-793-9427
Fax: 318-793-9430
E-mail: [email protected]
152
John C. Griffith
Georgia Forestry Commission
3561 Georgia Highway
Camilla, GA 31730
Phone: 229-522-3580
E-mail: [email protected]
Paige Grooms
305 Black Oak Rd.
Bonneau, SC 29431
Phone: 843-825-9987
Craig Guyer
Department of Zoology
Auburn University
Auburn, AL 36849
Phone: 844-9232
E-mail: [email protected]
Mark Hainds
Auburn University
Dixon Center
Andalusia, AL
Phone: 334-427-1029
E-mail: lla@ alaweb.com
Kent Hanby
431 Dogwood
Dadeville, AL 36853
Phone: 256-825-8593
Cathy Handrick
FL Fish & Wildlife Conservation
P.O. Box 177
Olustee, FL 32072
Phone: 386-758-5767
E-mail: [email protected]
Jon Handrick
Florida Division of Forestry
137 SE Forestry Circle
Lake City, FL 32025
Phone: 386-758-5713
E-mail: [email protected]
Larry Harris
UFL
Gainesville, FL
E-mail: [email protected]
Jennifer Hart
3742 Clint Dr.
Hilliard, FL 32046
Phone: 904-845-3597
Roger Hart
221 Airport Rd.
Fayetteville, NC 28306
Phone: 910-4372620
Lewis Hay
1706 Oak Grove Farm
Wadmalow Island, SC 29487
Phone: 843-559-0860
Lark Hayes
Southern Environmental Law Center
200 West Franklin Street
Chapel Hill, NC 27516
Phone: 919-967-1450
Fax: 919-929-9421
E-mail: [email protected]
Art W. Henderson
1001 North Street
Talladega, AL 35160
Phone: 256-362-2909
E-mail: [email protected]
Sandra Henning
1755 Cleveland Highway
Gainesville, GA 30501
U.S. Forest Service
Phone: 770-297-3064
E-mail: [email protected]
Nancy Herbert
200 Weaver Blvd.
Asheville, NC 28804
Phone: 828-257-4306
Sharon Hermann
Department of Biological Sciences
Auburn University, AL 36849
Phone: 334-844-3933
E-mail: [email protected]
George Hernandez
1720 Peachtree Rd. NW.
Atlanta, GA 30309
Phone: 404-347-3554
J. Kevin Hiers
JW Jones Ecological Research Center
Rt. 2 Box 2324
Newton, GA 39780
Phone: 229-734-4706
E-mail: [email protected]
John T. Hiers
Aucilla Pines, LLC
2409 Meadowbrook Drive
Valdosta, GA 31602
Phone: 229-244-5942
E-mail: [email protected]
Julie Hovis
US Air Force
345 Cullen Street
Shaw ABF, SC 29152
Phone: 803-895-9993
E-mail: [email protected]
Stanley Hinson
Southern Seed Company, Inc.
P.O. Box 340
Baldwin, GA 30511
Phone: 706-778-4542
E-mail: [email protected]
Stephen J. Hudson
Ecw Environmental Group, Llc.
534 Godfrey Lane
Auburn, AL 36830
Phone: 706-544-6263
E-mail:
[email protected]
Arthur Hitt
513 Madison Ave.
Montgomery, AL 36104
Phone: 334-240-9323
Malcolm Hodges
The Nature Conservancy
1330 West Peach Tree
Atlanta, GA 30309
Phone: 404-253-7211
E-mail: [email protected]
Joel Hodgson
Beaver Plastics LTD
12150 160 St
Edmonton, AB T5V 1H5
Phone: 604-552-1547
E-mail: [email protected]
David Hoge
USDA Forest Service
Atlanta, GA 30309
Phone: 404-347-1649
Mary Ann Hollenbeck
Joseph W. Jones Ecological Research
Route 2
Newton, GA 39870
Phone: 229-734-4706
E-mail: [email protected]
Gary Holmes
Osceola National Forest
24874 U.S. 90
Olustee, FL 32072
Phone: 386-752-2577
E-mail: [email protected]
153
Gwen Iacona
J.W. Jones Ecological Research Center
Rt. 2 Box 2325
Newton, GA 39870
Phone: 229-734-4706
E-mail: [email protected]
Lamar A. Isler
Georgia DNR
4073 East Gate Drive
Camilla, GA 31730
Phone: 229-220-2768
E-mail: [email protected]
Steven Jack
J.W. Jones Ecological Research Center
Rt. 2, Box 2324
Newton, GA 39870-9651
Phone: 229-734-4706
E-mail: [email protected]
Austin Jenkins
Clemson University
406 Watts Ave
Greenville, SC 29601
Phone: 864-313-4233
E-mail: [email protected]
Rhett Johnson
Solon Dixon Forestry Education Center
Andalusia, AL
Jamie Jones
South Carolina Forestry Commission
109 Hutson St.
Greenwood, SC 29649
Phone: 864-314-2028
E-mail: [email protected]
Bob Karrfalt
5675 Riggins Mill Rd.
Dry Branch, GA 31020
Phone: 478-751-3551
William Lamp
3561 HWY. 112
Camilla, GA 31730
Phone: 229-522-3580
Susan Kett
Osceola National Forest
24874 U.S. Hwy 90
Olustee, FL 32072
Phone: 386-752-2577
E-mail: [email protected]
Frank W. Lands
U.S. Army
1400 Cedar Hill Rd.
Douglasville, GA 30134
Phone: 404-464-1645
E-mail: [email protected]
Robert Kindrick
P.O. Box 2000
Pine Mountain, GA 31822
Phone: 706-663-3737
Paul J. Langford
PJ Langford Timber
9154 Woodrun Road
Pensacola, FL 32514
Phone: 850-477-6735
E-mail: [email protected]
L. Katherine Kirkman
Joseph W. Jones Ecological Research
Route 2
Albany, GA 39870
Phone: 229-734-4706
E-mail: [email protected]
Nathan Klaus
Georgia Department of Natural Resouces
Non-Game Endangered Wildlife Program
Forsyth, GA 31029
John Kush
School of Forestry and Wildlife
Auburn University
Auburn, AL 36849
Phone: 334-844-1065
E-mail: [email protected]
Cody Laird
Oakridge Farms/lolly Creek LLC
12 Rose Court
Atlanta, GA 30342
Phone: 404-316-3672
E-mail: [email protected]
Dobbs Laird
Oakridge Farms
1684 Wrights Chapel Rd.
Sumner, GA 31789
John Lambert
P.O. Box 3328
Sumrall, MS 39482
Phone: 601-758-4970
Keville Larson
2105 Venetia Rd.
Mobile, AL 36605
Phone: 257-476-4229
Weezie Larson
2105 Venetia Rd.
Mobile, AL 36605
Phone: 251-476-4229
Dwight K. Lauer
Silvics Analytic
Ridgeway, VA 24148
Gregory Lee
Moody Air Force Base
23 Ces/ceva
Moody AFB, GA 31699-1707
Phone: 229-257-5881
E-mail: [email protected]
Heather A. Lee
US Fish And Wildlife Service
Route 2 Box 3330
Folkston, GA 31537
Phone: 912-496-7366
Fax: 9124963332
E-mail: [email protected]
Tom Leslie
Georgia Engineering Alliance
542 St. Charles Avenue
Atlanta, GA 30308
Phone: 404-521-2324
154
Donzel Lewis
Lewis Forestry
182 Thomason Road
Cordele, GA 31015
Phone: 229-276-1541
E-mail: [email protected]
Lynn Lewis - Weis
P.O. Box 530
Edgefield, SC 29824
Phone: 803-637-3106
Michael Lick
P.O. Box 24B
Wiggins, MS 39577
Phone: 601-528-6173
Gerald W. Long
1830 Devils Backbone Road
Leesville, SC 29070
Phone: 803-532-4788
Joshua Love
Georgia Forestry Commission
P.O. Box 819
Macon, GA 31202
Phone: 478-751-3482
E-mail: [email protected]
Michael Low
107 HWY 85 North
Niceville, FL 32578
Phone: 850-883-1127
E-mail: [email protected]
Dwain Luce
Luce Packing Company
P.O. Box 8743
Moss Point, MS 39562
Phone: 251-343-3362
E-mail: [email protected]
Greg Luce
Luce Packing Company
P.O. Box 8743
Moss Point, MS 39562
Phone: 228-474-6383
E-mail: [email protected]
Margaret W. Luce
Luce Packing Company
P.O. Box 8743
Moss Point, MS 39562
Phone: 251-343-3362
E-mail: [email protected]
Susan Luce
Luce Packing Company
P.O. Box 8743
Moss Point, MS 39562
Phone: 228-474-6383
Bryan Maw
The University of Georgia
P.O. Box 748
Tifton, GA 31793
Phone: 229-386-3377
Steven Maharry
45 HWY 281
Heflin, AL 36264
Phone: 256-463-2273
Kirk McAlpin
Joseph W. Jones Ecological Research
Route 2
Newton, GA 39870
Phone: 229-734-4706
E-mail: [email protected]
Lawrence W. Mahler Jr.
P.O. Box 359
Summerdale, AL 36580
Phone: 281-423-3331
Raymond Majesty
1965 Tangewood Dr. Apt A
Glenview, IL 60025
Danny Marshburn
Camp Lejeune Marine Corps Base
Commanding Officer Attn
Camp Lejeune, NC 28542
Phone: 910-451-7223
Fax: 910-451-1787
E-mail: [email protected]
George L. McCaskill
University of Florida
281 Corry Village
Gainesville, FL 32603
Phone: 352-846-5950
E-mail: [email protected]
Joshua McCormick
Tall Timbers Research Station
13093 Henry Beadel Drive
Tallahassee, FL 32312
Phone: 850-893-4153
E-mail: [email protected]
James McHugh
Alabama Department Of Conservation
64 North Union Street
Montgomery, AL 36130
Phone: 334-242-3874
E-mail: [email protected]
Don McKenzie
2396 Cocklebur Rd.
Ward, AR 72176
Phone: 501-941-7994
Martha McKnight
113 W. Roanoke Dr.
Fitzgerald, GA 31750-8460
Phone: 229-423-2104
Thomas Meade
Avon Park Air Force Range
USAF
Avon Park, FL 33825
Phone: 863-385-7139
E-mail: [email protected]
Andy Meeks
Meeks Farms & Nursery, Inc.
187 Flanders Rd.
Kite, GA 31049
Phone: 912-536-3844
Jessica McCorvey
The Jones Center
Route 2, Box 2324
Newton, GA 39870
Steve Meeks
Phone: 229-734-4706
Meeks Farms & Nursery, Inc.
E-mail: [email protected] 187 Flanders Rd.
Kite, GA 31049
Howard E. McCullough
Phone: 877-809-1737
US Fish And Wildlife Service
E-mail: [email protected]
Katherine Martin
Route 2 Box 3330
J.W. Jones Ecological Research
Folkston, GA 31537
Mark A. Melvin
Rt. 2
Phone: 912-496-7366
J.W. Jones Ecological Research Center
Newton, GA 39870
Fax: 9124963332
Rt. 2, Box 2324
Phone: 229-734-4706
E-mail: [email protected] Newton, GA 39870-9651
E-mail: [email protected]
Phone: 229-734-4706
Jason McGee
E-mail: [email protected]
George Matusick
Joseph W. Jones Ecological Research
Auburn University
Route 2
Ben Miley
3301 School Of Forestry And Wildlife Newton, GA 39870
45 Highway 281
Auburn, AL 36849
Phone: 229-734-4706
Heflin, AL 36264
Phone: 334-844-1058
E-mail: [email protected]
Phone: 256-463-2273
Fax: 334-844-1084
E-mail: [email protected]
John McGuire
Kimberley D. Miller
School of Forestry and Wildlife SciUSFS- Calcasieu Ranger District
ences
9912 Hwy 28 West
Auburn University
Boyce, LA 71409
Auburn, AL 36849
Phone: 318-793-9427
Phone: 344-844– 1032
Fax: 318-793-9430
E-mail: [email protected]
E-mail: [email protected]
Joel Martin
Solon Dixon Forestry Education Center
12130 Dixon Center Road
Andalusia, AL 36420
Phone: 334-222-7779
E-mail: [email protected]
155
Manning Miller
St Joe Company
P.O. Box 400
Hosford, FL 32334
Phone: 850-379-8668
E-mail: [email protected]
Kim Mushrush
International Paper
P.O. Box 56
Bellville, GA 30414
Phone: 912-739-4721
Fax: 912-739-9409
E-mail: [email protected]
Weldon Miller
Ag-Renewal, Inc.
1519 E. Main Street
Weatherford, OK 73096
Phone: 580-772-7059
Fax: 580-772-6887
E-mail: [email protected]
Stefanie Nagid
FFWCC/ Restoration Ecologist
9225 CR 49
Live Oak, FL 32060
Phone: 386-362-1001
E-mail: [email protected]
Patrick J. Minogue
University of Florida
2319 B Via Sardina Street
Tallahassee, FL 32303
Phone: 530-604-8328
E-mail: [email protected]
Gil Nelson
Gil Nelson Associates
157 Leonard's Drive
Thomasville, GA 31792
Phone: 229-377-1857
E-mail: [email protected]
William Moody
125 Rose lake Rd.
Lexington, SC 29072
Phone: 803-808-2205
Lauren Newsome
Georgia Forestry Commission
1055 E. Whitehall Rd.
Athens, GA 30605
Phone: 706-542-9228
E-mail: [email protected]
Julie Moore
4401 N.Fairfax Drive, Rm 420
Arlington, VA 22203
Phone: 703-358-2096
Kathryn Mordecai
USDA Forest Service
Southern Research Station
Athens, GA 30602
Lee A. Mulkey
SEMP Director - UGA
P.O. Box 480
Demorest, GA 30535
Phone: 706-499-4493
E-mail: [email protected]
Keary Mull
SC Department Of Natural Resources
420 Dirleton Road
Georgetown, SC 29440
Phone: 843-546-3226
E-mail: [email protected]
Kenwood Nichols
2536 Centenary St.
Selma, AL 36701
Phone: 334-874-7167
E-mail: [email protected]
Joseph O’Brien
USDA Forest Service
Southern Research Station
Athens, GA 30602
Anna Osiecka
University of Florida
155 Research Rd.
Quincy, FL 32351
Phone: 850-875-7145
Fax: 850-875-7188
E-mail: [email protected]
Wimbric J Padgett Jr.
Padgett Tree Planting
Rt.1
Milan, GA. 31060
Phone:229-833-2088
Fax: 229-833-5573
E-mail: [email protected]
156
Eric Palola
58 State St.
Montpelier, VT 05602
Phone: 802-229-0650
Dale Pancake
12130 Dixon Center Rd.
Andalusia, AL 36420
Phone: 334-222-7779
James Parker
U.S. Army Infantry Center
Ft. Benning, GA 31905
Phone: 706-544-7081
Ronald A. Phernetton
US Fish And Wildlife Service
Route 2 Box 3330
Folkston, GA 31537
Phone: 912-496-7366
Fax: 9124963332
E-mail: [email protected]
Brad Phillips
P.O. Box 1463
Waynesboro, GA 30830
Phone: 706-554-2310
Bill Pickens
2411 Old US 70 West
Clayton, NC 27525
Phone: 919-553-6178
Caroline Polgar
Joseph W. Jones Ecological Research
Route 2
Newton, GA 39870
Phone: 229-734-4706
E-mail: [email protected]
Dotty S. Porter
Sessoms Timber Trust
3704 Dean Still Road
Blackshear, GA 31516
Phone: 912-449-8524
E-mail: [email protected]
Catherine Prior
The Nature Conservancy
113 8th St.
Columbus, GA 31901
Phone: 706-587-1395
E-mail: [email protected]
Harold E. Quicke
BASF Corporation
Raleigh, NC
Justin Qurey
221 Airport Rd.
Fayetteville, NC 28306
Phone: 910-437-2620
Wayne Rast
Carolina Heart Pine, Inc.
166 Winding Brook Rd.
Cameron, SC 29030
Phone: 803-534-0404
E-mail: [email protected]
Mark Register
International Paper
719 Southlands Rd
Bainbridge, GA 39819
Phone: 229-246-3642
E-mail: [email protected]
Roger Reid
The University of Alabama
P.O. Box 870 340
Tuscaloosa, AL 34487
Phone: 205-480-8443
roger.reid@discoveringalabama
Joe Reinman
St. Marks National Wildlife
St. Marks, FL 32355
Phone: 850-925-6121
Randy Roach
1208 B Main St.
Dophne, AL 36526
Phone: 251-441-5872
James Roberts
3125 Conner Blvd. C 25
Tallahassee, FL 32301
Phone: 850-414-9906
Darrell Russell
DOW Agrosciences
P.O. Box 1938
Roswell, GA 30077
Phone: 770-594-8949
E-mail: [email protected]
Buford Sanders
Georgia Forestry Commission
1055 E. Whitehall Rd.
Athens, GA 30605
Phone: 706-542-9939
E-mail: [email protected]
Jim Schlenker
221 Sirport Rd.
Fayetteville, NC 28306
Phone: 910-4372620
John Seymour
Roundstone Native Seed, LLC
9764 Raider Hollow Road
Upton, KY 42784
Phone: 270-531-2353
E-mail: [email protected]
Randy Seymour
9764 Raider Hollow Road
Upton, KY 42784
Phone: 270-531-2353
E-mail: [email protected]
Janet Sheldon
Georgia Conservancy
18 N. Main Street
Moultrie, GA 31768
Phone: 229-985-8117
E-mail: [email protected]
Richard Shelfer
325 John Knox Rd. Suite F 100
Tallahassee, FL 32303
Phone: 850-523-8553
E-mail: [email protected]
Geoff Rockwell
1609 Davis Ave.
Tifton, GA 31794
Phone: 229-388-9023
Steven J. Shephard
Camp Lejeune Marine Corps Base
Commanding Officer
Camp Lejeune, NC 28542
Phone: 910-451-7220
Lin Roth
Fax: 910-451-1787
Institute of Coastal Ecology and Forest E-mail: [email protected]
Research
Clemson University
Clemson, SC 29634
157
Margie Sheridan
8390 Fredericksbsrg Tppk.
Woodford, VA 22580
Phone: 804-633-4336
Phil Sheridan
8390 Fredericksberg Tppk.
Woodford, VA 22580
Phone: 804-633-4336
Gary Shurette
25 HWY 281
Heflin, AL 36264
Phone: 256-463-2273
Chuck Simon
County Extension Agent
Covington County
Alabama Cooperative Extension System
Andalusia, AL 36420
Terrell Simmons
Simmons Tree Farm
545 Snipesville Rd.
Denton, GA
Phone: 912-375-7520
Jo Ann Smith
US Forest Service
2500 Shreveport Hwy
Pineville, LA 71360
Phone: 318-473-7191
Fax: 318-473-7117
E-mail: [email protected]
Scott Smith
Joseph W. Jones Ecological Research
Route 2 Box 2324
Newton, GA 39870
Phone: 229-734-4706
Fax: 229-734-4707
E-mail: [email protected]
David South
Auburn University
School of Forestry and Wildlife
Auburn, AL 36830
Phone: 334-844-1022
E-mail: [email protected]
Eric Spadgenske
2100-1st Ave. North, Suite 500
Birmingham, AL 35203
Phone: 250-731-0874
Vaughan Spearman
P.O. Box Drawer 1086
Walterboro, SC 29488
Phone: 843-549-2595
Tommy Spencer
Osceola National Forest
24874 U.S. Hwy 90
Olustee, FL 32072
Phone: 386-752-2577
E-mail: [email protected]
Donna Streng
U.S. Fish and Wildlife
West Georgia Ecological Services
Fort Benning, GA 31995
Lee Stribling
School of Forestry and Wildlife
Auburn University
Auburn, AL 36849
Eric G. Strickland
Engineering And Environment Inc.
2320 Apartment C
Augusta, GA 30909
Phone: 706-791-9929
E-mail: [email protected]
Katharine L. Stuble
J.W. Jones Ecological Research Center
Rt. 2, Box 2325
Newton, GA 39870
Phone: 229-734-8026
E-mail: [email protected]
Eric Stuewe
Stuewe and Sons, Inc.
2290 SE Kiger Island Dr.
Corvallis, OR 97333
Phone: 541-757-7798
E-mail: [email protected]
John Sunday
1465 Tignall Rd.
Washington, GA 30673
Phone: 706-678-2015
Shi Jean S. Sung
USDA FS
2500 Shreveport Highway
Pineville, LA 71360
Phone: 318-473-7233
Fax: 318-473-7273
E-mail: [email protected]
Sammy Sweat
Simmons Tree Farm
545 Snipesville Rd.
Denton, GA
Phone: 912-422-3757
Mary Anne Sword-Sayer
USDA FS
2500 Shreveport Highway
Pineville, LA 71360
Phone: 318-473-7275
Fax: 318-473-7273
E-mail: [email protected]
Scotland Talley
FL. Fish And Wildlife Conservation
1600 NE 23rd Ave.
Gainesville, FL 32609
Phone: 352-955-2241
Fax: 352.955.2125
E-mail: [email protected]
Daniel Taylor
15019 Board St.
Broolsville, FL 34601
Phone: 352-754-6777
Deborah Taylor
Memorial Park Conservancy
5030 St. Kitts Calle
Dickinson, TX 77539
Phone: 713-863-8403
E-mail: [email protected]
Wayne Taylor
Avon Park Air Force Range
USAF
Avon Park, FL 33825
Phone: 863-452-4119
E-mail: [email protected]
Donald Temple
1800 Bunnlevel - Erwin Rd.
Bunnlevel, NC 28323
Phone: 910-591-9494
Gloria M. Thomasson
3719 Prentice Ave.
Columbia, SC 29205
Phone: 803-787-7046
William H. Thomasson
3719 Prentice Ave.
Columbia, SC 29205
Phone: 803-787-7046
Cindy Thompson
Osceola National Forest
24874 U.S. 90
Olustee, FL 32072
Phone: 386-752-2577
E-mail: [email protected]
Micah G. Thorning
USDA Forest Service
9901hwy 5
Brent, AL 35034
Phone: 205-926-9765
Fax: 205-926-9712
E-mail: [email protected]
Jeff Thurmond
NRCS
3381 Skyway Dr.
Auburn, AL 36830
Phone: 334-887-4560
Don Tomczak
1720 Peachtree Rd. NW,Room 816
Atlanta, GA 30309
Phone: 404-347-7475
Chris Trowell
Department of Social Science
South Georgia College
Douglas, Georgia 31533
Sandra Tucker
US Fish & Wildlife Service
105 Westpark Dr.
Athens, GA 30606
Phone: 706-613-9493
Fax: 706-613-6059
E-mail: [email protected]
Lex Tyson
J.W. Jones Ecological Research Center
Rt. 2, Box 2324
Newton, GA 39870-9651
Phone: 229-734-4706
E-mail: [email protected]
Joe Vanderwerff
5865 East HWY 98
Santa Rosa Beach, FL 32459
Phone: 850-231-5800
158
Merrill Varn
Varn Co.
P.O. Box 40965
Jacksonville, FL 32203
Phone: 904-356-4881
E-mail: [email protected]
Bennie F. Vinson
Alabama Power Company
600 N. 18th Street
Birmingham, AL 35291
Phone: 20-257-4622
Fax: 205.257.2764
E-mail: [email protected]
Allison Vogt
Auburn University
602 Duncan Drive, Room 3236
Auburn University, AL 36849-5418
Phone: 334-844-9219
Fax: 334-844-1084
E-mail: [email protected]
Joan Walker
Department of Forestry and Natural
Resources
Clemson University
Clemson, SC 29634
Don Wardlaw
BASF Corporation
3445 Meadowlane
Waycross, GA 31503
Phone: 912-663-8476
E-mail: [email protected]
Clay Ware
SFIM-AEC-TSR, Bldg.E4430
APG, MD 21010
Phone: 410-436-6463
Sarah L. Watkins
Luce Packing Company
P.O. Box 8743
Moss Point, MS 39562
Phone: 251-406-0074
E-mail: [email protected]
Holly Welch
South Carolina Forestry Commission
278 Spring Road
Laurens, SC 29360
Phone: 803-940-0980
E-mail: [email protected]
Micah White
Rt. 1 Box 67
Helene, GA 31037
Phone: 229-868-3385
John Whitesides
USAF
4 Winder Cres
Newport News, VA 23606
Phone: 757-764-2766
E-mail: [email protected]
F. Bennett Whitfield
Whitfield Farms & Nursery
2561 Lambs Bridge Road
Twin City, GA 30471
Marion S. Wiggers
J.W. Jones Ecological Research Center
Rural Route 2, Box 2324
Newton, GA 39870
Phone: 229-734-4706
E-mail: [email protected]
Harold P. Wilson
MFC/Waynesboro Nursery
1063 Buckatunna-Mt. Zion Road
Waynesboro, MS 39367
Phone: 601-735-9512
Fax: 601-735-3163
E-mail: [email protected]
Beth Wood
1550 Henly St.
Orangeburg, SC 29115
Phone: 803-534-6280
Keith Wooster
USDA-NRCS
355 East Hancock Avenue Ms. 207
Athens, GA 30601
Phone: 706-546-2114
E-mail: [email protected]
Rebecca M. Worley
Berry College
CPO 7155 P.O. Box 9000
Asheville, NC 28815-9000
Phone: 770-547-3897
E-mail: [email protected]
Beth Young
Cahaba River Publishing
2805 Shades Crest Road
Birmingham, AL 35216
Phone: 205-969-1800
159
James W. Zanzot
Auburn University
3301 School Of Forestry & Wildlife
Auburn University, AL 36849
Phone: 334-329-4121
E-mail: [email protected]
160