+ All Categories
Home > Documents > Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn...

Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn...

Date post: 20-Jul-2020
Category:
Upload: others
View: 0 times
Download: 0 times
Share this document with a friend
133
Wetland Vegetation Dynamics and Ecosystem Gas Exchange in Response to Organic Matter Loading Rates ___________ A Thesis Presented to The Faculty of the School of Marine Science College of William and Mary In Partial Fulfillment of the Requirements for the Degree of Masters of Science ___________ By David E. Bailey 2006
Transcript
Page 1: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

Wetland Vegetation Dynamics and Ecosystem Gas Exchange in Response to Organic Matter Loading Rates

___________

A Thesis Presented to

The Faculty of the School of Marine Science

College of William and Mary

In Partial Fulfillment of the Requirements for the Degree of

Masters of Science

___________

By David E. Bailey

2006

Page 2: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

1

APPROVAL SHEET

This thesis is submitted in partial fulfillment of

The requirements for the degree of

Master of Science

____________________________ David E. Bailey

Approved, August 15, 2006 ____________________________________ James E. Perry, Ph.D. Committee Chairman/Advisor ____________________________________ Patrick Megonigal, Ph.D.

Smithsonian Environmental Research Center Edgewater, MD

____________________________________ W. Lee Daniels, Ph.D. Virginia Tech Blacksburg, VA

____________________________________ Randolph M. Chambers, Ph.D.

____________________________________ David A. Evans, Ph.D.

Page 3: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

2

Table of Contents

Acknowledgements …………………………………………………………………... Page 5

Abstract ………………………………………………………………………………. Page 7

Chapter 1: A review of non-tidal wetland ecology and created wetlands ……. Page 8

Tables

Table 1: Redox Hierarchy …………………………………………………...... Page 13

Figures

Figure 1: Wetland Carbon Cycle ……………………………………………... Page 19

Chapter 2: Vegetation Dynamics …………………………………………………….. Page 42

Tables

Table 1: Soil Variables ……………………………………………………….. Page 75

Table 2: Plant Species List …………………………………………………… Page 76

Table 3: Dominant Species …………………………………………………… Page 79

Table 4: Vegetation Variables ………………………………………………... Page 80

Table 5: Similarity Table ……………………………………………………... Page 81

Figures

Figure 1 a & b: Map of Study Site ……………………………………………. Page 82

Figure 2: Sampling Design …………………………………………………… Page 83

Figure 3: Elevation vs. Loading Rate ………………………………………… Page 84

Figure 4: Soil Nutrients vs. Loading Rate ……………………………………. Page 85

Figure 5: Weighted Averages vs. Loading Rate ……………………………… Page 86

Figure 6: Plant Indicator Status ………………………………………………. Page 87

Figure 7a: Biomass vs. Elevation …………………………………………….. Page 88

Page 4: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

3

Figure 7b: Biomass vs. Elevation (outliers removed) ………………………... Page 89

Figure 8: PCA Loadings ……………………………………………………… Page 90

Figure 9: Tree Size vs. Loading Rate ………………………………………… Page 91

Figure 10: Tree Size vs. Elevation ……………………………………………. Page 92

Figure 11: Tree Size vs. Soil P ……………………………………………….. Page 93

Figure 12: Biomass vs. Soil P (0-12 mg kg-1) ………………………………... Page 94

Figure 13: Tree Size vs. Lower Soil P (0-15 mg kg-1) ……………………….. Page 95

Chapter 3: Ecosystem Gas Exchange ………………………………………………... Page 96

Tables

Table 1: CO2 Flux Variables …………………………………………………. Page 117

Figures

Figure 1: CO2 Sampling Equipment ………………………………………….. Page 118

Figure 2: Annual CO2 Fluxes ………………………………………………… Page 119

Figure 3: CO2 Flux vs. Loading Rate ………………………………………… Page 120

Figure 4: Monthly R ………………………………………………………….. Page 121

Figure 5: Monthly GPP ……………………………………………………...... Page 122

Figure 6: Monthly NEE …………………………………………………...….. Page 123

Figure 7: Respiration vs. Water Table ………………………………………... Page 124

Figure 8: GPP:R Ratio vs. Loading Rate ……………………………………... Page 125

Chapter 4: Summary and Conclusions ………………………………………………. Page 126

Appendices

Appendix 1: Organic Amendment Analysis ………………………………….. Page 128

Appendix 2: Soil Chemical Properties ……………………………………….. Page 129

Appendix 3: Monthly Total PAR …………………………………………….. Page 130

Appendix 4: Monthly Mean Temperature ……………………………………. Page 131

Page 5: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

4

Appendix 5: CH4 Flux Data …………………………………………………... Page 132

Page 6: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

5

ACKNOWLEDGEMENTS

I sincerely appreciate the fieldwork assistance of Peter Knowles, Azure Bevington, and J.

Paul Rinehimer, as well the laboratory support of Steve Nagle, Julia Burger, and Virginia Tech

Soil Testing Labs, the Biogeochemistry Lab of the Smithsonian Environmental Research Center,

Amber Hardison and the Iris Anderson Lab at VIMS, and Robert Atkinson of Christopher

Newport University. I am also indebted to Douglas DeBerry, both for fieldwork expertise and a

critical review of parts of this thesis, as well as to Cara Bergschneider, who built the

experimental plots and whose thesis was a strong foundation for this work. Lastly, I would like

to thank Committee members Jim Perry, Pat Megonigal, Lee Daniels, Randy Chambers, and

David Evans for their research suggestions and support, Patty Richards for vehicular support, as

well as Jennifer Bailey, Carl and Ruth Bailey, and Andrew and Kathy Bailey for moral support.

This research was supported in part by VDOT, whom I thank for the use of the site, as well as

the College of William and Mary, Virginia Institute of Marine Science.

Page 7: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

6

Wetland Vegetation Dynamics and Ecosystem Gas Exchange in

Response to Organic Matter Loading Rates

Page 8: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

7

ABSTRACT

Created wetlands are often limited in soil organic matter, usually a long-term product of ecosystem succession. Although many studies have tested the effect of adding organic material to these systems, few if any, have quantified the effect of various loadings of organic matter in created wetlands. The goal of this study was to determine how vegetation composition, standing crop biomass, woody vegetation development, and ecosystem gas exchange varied in a created freshwater wetland along a gradient of soil organic carbon (0 to 336 Mg ha-1 loading rates). Plot surface elevation varied positively with OM loadings, suggesting that inundation/aeration may modify OM effects. Soil nutrients (C, N, C:N, and P) also positively correlated with loading rate. Vegetation measurements suggested an overall similarity of plant assemblage composition and biomass regardless of loading rate, and a slight increase in tree size with loading rate. Gross primary production and net ecosystem exchange were weakly positively and negatively correlated with loading rate, respectively. Respiration was strongly positively correlated with loading rate, and was likely the controlling factor of CO2 gas flux among treatments. Soil nutrient values and vegetation composition, as well as ecosystem gas flux balance appear to be the best parameters upon which to base an organic matter loading rate decision. In this study, adding an organic matter amendment between LR 2 (56 Mg ha-1) and 3 (112 Mg ha-1) seems most appropriate, and may provide a “jumpstart” for the created non-tidal wetlands while also minimizing changes in surface elevation due to the added bulk material.

Page 9: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

8

Chapter 1

A review of non-tidal wetland ecology and created wetlands

Introduction

Freshwater wetlands are important and unique features of the global landscape. These

environments often provide favorable circumstances to facilitate physical and biogeochemical

functions that are critical to the maintenance of healthy environmental conditions in waterways

(Mitsch and Gosselink 2000). Some of these purported hydrologic and water quality functions

include short- and long-term surface water storage, maintenance of a high water table,

groundwater discharge or recharge points, transformation and cycling of elements, retention and

removal of dissolved elements, and accumulation of inorganic sediments (NRC 1995, Cole

2002). Other ecosystem functions performed by wetlands include fish and wildlife habitat,

primary production, waste assimilation, and global carbon and nutrient cycling (Odum 1978). It

is important to note that the ability of wetlands to perform certain ecosystem functions depends

on their extent and their location within the larger landscape. For example, the functionality of a

forested wetland will vary; if it lies along a river, it has a greater functional role in maintaining

water quality than if it were isolated from the stream (Mitsch and Gosselink 2000).

Recognition of the importance of wetland functions to our own well-being (i.e. ecosystem

services), and the acknowledged loss of 53% of the natural wetland area in the contiguous United

Stated (USEPA 1989, Dahl 1990), convinced the United States Congress to insure the

maintenance of adequate wetland area under the 1977 Clean Water Act (CWA). Section 404 of

the CWA is the primary guidance concerning wetland protection, administered under the

regulatory power of the US Army Corps of Engineers (USACE) overseen by the Environmental

Page 10: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

9

Protection Agency (EPA). In 1990, President George Bush, Sr. recommended a policy of ‘No

Net Loss’ of wetlands as a national goal (Mitsch and Gosselink 2000). As such, many permits

issued under section 404 of the CWA authorizing the filling or draining of wetlands are

accompanied by directives to create or restore a similarly sized or larger wetland area (i.e.

compensation). Importantly, however, compensatory wetland restoration and creation are “last

resort” regulatory options (Whigham 1999) following avoidance and minimization of wetland

disturbance. Wetland creation involves using engineering and construction practices to establish

conditions conducive to the development of wetland hydrology, soils, and vegetation in areas

where wetlands did not previously or historically exist. Creating a new wetland as a mitigation

practice is more difficult than wetland restoration due to the complexity of establishing the

necessary hydrologic regime (Stolt et al. 2000).

Succession in ecosystems, according to the framework of E.P. Odum (1969), involves the

transfer and flow of energy through time that shows perceivable directionality, is affected by

both the physical environment and community itself, and leads to a theoretical stabilized system

where energy efficiency for both biomass and symbiotic organism interaction are maximized. As

ecosystems mature, and energy usage shifts over time from high biomass production to

ecosystem maintenance, changes in the overall ecosystem energy balance should reflect a

progression from a net autotrophic regime (photosynthesis (P) / respiration (R) ratio >1) to a net

heterotrophic regime (P/R = 1). Such successional concepts, especially early successional

processes (primary succession) are rarely studied in wetlands (van der Valk 1981), and are only

now beginning to be applied to created wetlands (Noon 1996, DeBerry and Perry 2004).

Organic matter is important in the primary succession conceptual model because it

accumulates in developing ecosystems over time, mitigating perturbations caused by the physical

Page 11: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

10

environment (Odum 1969). Organic matter accumulation is an important and characteristic

feature of most wetland systems. However, created wetlands usually contain very low soil

organic carbon relative to natural wetlands; traditional creation practices such as topsoil scraping

remove not only the existent vegetation, but also any accumulated organic material in the surface

A horizon including seed banks. Many studies have advocated amending wetland creation sites

with organic material either in the form of salvaged natural wetland soils or mulches (Stauffer

and Brooks 1997, Whittecar and Daniels 1999, McKinstry and Anderson 2003, Anderson and

Cowell 2004, Bruland and Richardson 2004).

Monitoring organically amended created wetlands not only allows the comparison of

created wetland success between amended and unamended sites, it also facilitates the

characterization of primary succession processes at different organic mater loading rates by

simulating one aspect of the chronosequence of seral stages in created wetland development.

The purpose of this study was to evaluate the primary succession processes occurring in a

created forested wetland, in terms of both vegetative characteristics and ecosystem energy

exchange, along a gradient of soil organic carbon input that mimics the conceptual idea of

increasing organic matter with increasing ecosystem maturity. The parameters of interest in this

study were herbaceous vegetation community composition and primary productivity, woody

vegetation development, and the relative rates of photosynthesis and respiration (carbon flux) of

the vegetation-soil continuum. The goal of this study was threefold: 1) to determine how various

parameters (plant community composition, standing crop, woody vegetation growth, and Net

Ecosystem Exchange) in created freshwater wetlands vary with respect to different loadings of

organic matter amendments in wetland soils, 2) to monitor ecosystem function within a primary

successional system with respect to an added soil organic carbon gradient, and 3) to use the

Page 12: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

11

measured parameters to elucidate the effect of adding various loadings of an organic matter

amendment on created wetland success.

Literature Review

Successional processes in wetlands are the result of physical, chemical, and biological

factors that influence such ecosystem characteristics as plant community composition, primary

productivity, and energy exchange in the form of CO2 flux into and emitted from the wetland. In

created wetland systems, these factors act to control or influence pedogenesis and chemical and

nutrient cycling, which then shape the attendant vegetative community. The following sections

will explore the fundamental and relevant literature that describe the soil and chemical processes

affecting wetland primary succession, and will then focus on the historic and current research

concerning created wetland systems and primary succession within these systems.

Wetland Soils and Pedogenesis

Soil Development

Soil development tends to be more advanced on older sites – that is, soil horizonation and

strength of expression increases with time since the last major disturbance or since the start of

uninterrupted soil creation (Odum 1969, Marks and Bormann 1972, Odum 1985, Chadwick and

Graham 2000). In wetlands, this is due to the fact that soil-forming processes such as

weathering, incorporation of organics, and soil oxidation/reduction (redox) processes are time-

dependent (Jenny 1941, Stevens and Walker 1970, Mausbach and Richardson 1994) and are

inevitably linked to the colonizing vegetation (Craft 2001) and to the depletion of free oxygen in

the soil (Mitsch and Gosselink 2000, Megonigal et al. 2004). Thus, soil development in non-

Page 13: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

12

tidal, freshwater wetlands is defined by in situ physical, chemical, and biological attributes

present at a given point in time.

Wetlands are identified in part by the hydric characteristics of their underlying soil

substrate. Hydric soils are defined as soils that “formed under conditions of saturation, flooding,

or ponding long enough during the growing season to develop anaerobic conditions in the upper

part” (NTCHS 1985). Hence, anaerobiosis – the process of oxygen depletion in the soil via

biogeochemical pathways – is a critical defining factor in wetland soils.

Because molecular oxygen diffuses 10,000 times more slowly through water than air,

microbial metabolism in a saturated soil preferentially utilizes O2 faster than it can be replaced by

diffusion, and the system becomes anaerobic (Megonigal et al. 2004). Under inundated

conditions, oxygen may be present in shallow surface water due to the rapid transport of

atmospheric O2 across the air-water interface, low populations of oxygen-consuming organisms

in the water column, photosynthetic O2 production from algae, and surface water mixing

(Gambrell and Patrick 1978). As such, a continuous supply of dissolved oxygen is usually

present at the soil-water interface, resulting in a thin oxidized layer at the soil surface as O2

follows a concentration gradient downward (Ponnamperuma 1972).

As organic substrates are oxidized during microbial metabolism, free oxygen is

preferentially utilized as a terminal electron acceptor at a redox potential of 400 to 600 mV. As a

result, the O2 concentration below the thin oxidized layer drops abruptly in waterlogged soils,

and may be completely depleted within a matter of millimeters. At this point, a hierarchy of

potential terminal electron acceptors may be available to microorganisms that can respire under

increasingly anaerobic conditions, and these organisms preferentially reduce inorganic chemical

substrates based on the redox potential of the respective transformations (Table 1; Mitsch and

Page 14: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

13

Gosselink 2000, Wang and Patrick 2000). Chemical reactions that take place under anaerobic

conditions, or reduction-oxidation (redox) reactions, are a primary component in both the

creation of hydric soil redoximorphic features and the driving force behind nutrient cycles

responsible for selecting for and against certain plant species inherent in primary succession

(Mitsch and Gosselink 2000). These redox reactions are discussed below in the Chemical and

Nutrient Cycling section.

Table 1. The “redox hierarchy” following anaerobiosis (from Mitsch and Gosselink 2000) Element Oxidized Form Reduced Form Redox Potential (mV)

Nitrogen NO3- (nitrate) N2O, N2, NH4

+ (nitrous oxide, nitrogen gas, and ammonium)

250

Manganese Mn+4 (manganic) Mn+2 (manganous) 225

Iron Fe+3 (ferric) Fe+2 (ferrous) 120

Sulfur SO4-2 (sulfate) S-2 (sulfide) -75 to -150

Carbon CO2 (carbon dioxide) CH4 (methane) -250 to -350

Soil Organic Matter

The progressive increase in soil organic matter with site age is accompanied by increases in

denitrification capacity (Johns et al. 2004), C:N ratios (Nair et al. 2001), and plant species

diversity (Reinhartz and Warne 1993), and decreases in bulk density, pH, and soil chroma

(Bischel-Machung et al. 1996, Nair et al. 2001, Campbell et al. 2002). Soil organic matter

provides the materials and energy necessary to sustain metabolism within wetland ecosystems,

including providing plant nutrients through its decomposition and mineralization (Neue 1985),

Page 15: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

14

and provides the reducing power for microbially mediated chemical transformations under

anaerobic conditions (Vepraskas and Faulkner 2001).

Soil organic matter and organic carbon are coupled inexorably with nutrient cycling, and

regulate the conservation and recycling of nutrients between microbial, plant, and soil

communities (Collins and Kuehl 2001). As microorganisms oxidize organic compounds,

organically bound plant nutrients such as nitrogen and phosphorus are mineralized into their

inorganic forms. Concomitantly, microorganisms use carbon and other nutrients from plant

residues for their own growth, thereby temporarily immobilizing these compounds into microbial

biomass (Qualls and Richardson 2000). As microbes die and decompose, assimilated carbon and

nutrients are again mineralized and become available for further decomposition or successive

uptake. These processes occur in both aerobic and anaerobic environments; however, different

physiological groups of microorganisms are active in the two states, and the oxidized forms of

plant nutrients are reduced in anaerobic environments due to the loss of oxygen as an electron

acceptor (Collins and Kuehl 2001). Bulk organic matter decomposition by anaerobic microbes is

less efficient and much slower than in upland systems; therefore, wetland soils accumulate

organic matter at much higher rates than do uplands.

Chemical and Nutrient Cycling

Redox Reactions

Nitrogen

The first reaction following the onset of anaerobiosis is the reduction of nitrate (NO3-) to

nitrite (NO2-), and eventually to nitrogen gas (N2), which represents a potential loss of nitrogen

from the system (Phung and Fiskell 1973, Poe et al. 2003, Megonigal et al. 2004). This process

Page 16: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

15

is termed denitrification, and can reduce the bioavailability of nitrogen for plants in freshwater

wetland systems (Brinson et al. 1984, Ambus and Lowrance 1991, Groffman 1994). Several

other nitrogen transformations may take place, including nitrogen fixation, ammonification,

nitrification, and microbial immobilization (DeLaune et al. 1996, Cirmo 1998). The thin,

oxidized layer of soil at the soil-water interface is an important component of the nitrogen cycle

in freshwater wetlands, because the concentration gradients set up across this interface allow

diffusion of intermediary nitrogen forms upward or downward to the aerobic (e.g., nitrification)

and/or anaerobic (e.g., denitrification) sites of the major transformations (Gambrell and Patrick

1978, Mitsch and Gosselink 2000).

Manganese and Iron

After most of the nitrate-N in the system has been reduced, manganese and iron reduction

occur. Both elements behave in a similar manner in wetland substrates, with two valence states

resulting in either dissolved (reduced) or precipitated mineral (oxidized) forms (Gambrell et al.

1989, Nealson and Saffarini 1994, Vepraskas and Faulkner 2001). Importantly, however,

manganese reduces before iron and at higher redox potentials in most anaerobic environments.

Oxidized (manganic) manganese is black in color, and often forms hard nodules in the profile

where free oxygen is reintroduced into the soil matrix (e.g., fluctuating water tables, or

artificially drained hydric soils) (Ponnamperuma 1972). Oxidized (ferric) iron is red in color,

and zones of iron oxide precipitates may be found in fluctuating water table conditions similar to

manganese (Vepraskas 1994). These redoximorphic features, called “mottles” in earlier

literature and the 1987 COE Manual (Environmental Laboratory 1987), indicate reintroduction

of atmospheric oxygen into pore spaces within the reduced soil medium into which dissolved,

Page 17: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

16

reduced (ferrous) iron follows a concentration gradient to bind with the available oxygen. A

similar condition may occur in reduced soils when wetland plants oxidize the rhizosphere

surrounding the roots (Mendelssohn et al. 1995, Weiss et al. 2003,). Certain wetland plants have

the ability to transport atmospheric oxygen from the aerial organs (shoot) through airspace tissue

(aerenchyma) down to the roots, where the exuded oxygen forms iron oxide precipitates within

the adjacent soil around the root. These “oxidized rhizospheres” (also, oxidized root channels)

are of biological importance, because the presence of iron oxide can immobilize important

nutrients such as phosphorus, and can form plaques around the root which serve as barriers to

nutrients (Gambrell and Patrick 1978).

Sulfur

The reduction of sulfate typically follows iron and manganese in the redox hierarchy

(Table 1). Although sulfate reduction occurs in non-tidal, freshwater wetlands, it is more

characteristic of marine and estuarine wetland systems due to the prevalence of sulfate in

seawater (Megonigal et al. 2004). In freshwater systems, sulfate reduction is limited by the

comparatively low concentrations of sulfate (Craft 2001), and methanogenesis (see below) is the

predominant respiratory pathway at low redox states. However, in some wetland systems, sulfur

inputs from atmospheric deposition can elevate sulfate reduction rates in freshwater wetlands,

resulting in a concomitant reduction in methane production (Wieder et al. 1990, Gauci et al.

2004). Sulfate reduction is strongly linked to organic carbon oxidation, and usually requires a

refined, highly labile (easily broken down), short-chain carbon source (Megonigal et al. 2004).

Methanogenesis and Fermentation

Page 18: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

17

At the bottom of the redox hierarchy, carbon itself is reduced during microbial

respiration. Carbon reduction is a very low energy-yielding metabolic reaction (Table 1), and

typically occurs only when all other terminal electron acceptors have been exhausted within the

system. The two major anaerobic carbon transformations in wetland soils are methanogenesis

and fermentation. The former uses CO2 as a terminal electron acceptor, and the latter uses

simple, low molecular-weight organic compounds (Mitsch and Gosselink 2000).

Methanogenesis is the final step in organic carbon degradation within anaerobic environments,

resulting in the release of methane (CH4) gas as a waste product by methanogenic

microorganisms. Thus, for microbes to successfully utilize the methanogenic pathway, a source

of labile organic carbon must be readily available (Megonigal and Schlesinger 1997, Megonigal

et al. 2004). In freshwater wetlands, this source is typically the standing vegetation (although

phytoplankton also play a role in surface water systems). Plants supply organic compounds to

the substrate via decomposition of detritus and/or through the exudation of intracellular

constituents (Gambrell and Patrick 1978, Brinson et al. 1981, Moran et al. 1989). This link

between methane emissions and primary productivity has been demonstrated in natural and

anthropogenic freshwater wetland systems (Whiting and Chanton 1993, Megonigal et al. 2004).

Phosphorus

The phosphorus cycle is of interest in freshwater wetlands because it is often a limiting

nutrient in these systems (Aerts et al. 1992, Chapin et al. 2004), and because wetlands represent a

potential removal mechanism for elevated nonpoint source phosphorus levels in urbanized and

agricultural landscapes (Richardson 1985). The phosphorus cycle is unique among the other

major biogeochemical cycles in that it has no gaseous phase – phosphorus occurs as soluble or

Page 19: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

18

insoluble organic and inorganic complexes adsorbed to soil particles (Qualls and Richardson

1995, Mitsch and Gosselink 2000). Although redox potential does not directly affect phosphorus

transformations, an indirect effect may occur in the presence of other elemental redox forms such

as ferric iron, which immobilizes otherwise bioavailable phosphate by precipitation

(Ponnamperuma 1972, Mohanty and Dash 1982). Other removal mechanisms include

precipitation with calcium and aluminum, adsorption onto clay or organic particles and iron or

aluminum oxides (Gambrell and Patrick 1978), and incorporation in living biomass via direct

uptake primarily by microbes and to a lesser extent by plants. However, as anoxia proceeds in

saturated soils, iron-bound phosphorus may be released as bioavailable phosphate when ferric

iron is reduced to ferrous iron by anaerobic microbial respiration (Mitsch and Gosselink 2000).

Thus, freshwater wetlands can function as sources and/or sinks for phosphorus (Richardson and

Marshall 1986), depending on the biogeochemical setting and other factors such as pH

(Verhoeven et al. 1990).

The Carbon Cycle

In the biosphere’s carbon cycle, many wetlands, such as organic soil bogs and fens of

northern latitudes and pocosins of the southeastern US, function naturally as net carbon sinks by

accumulating carbon dioxide from the atmosphere (Armentano and Menges 1986). In fact,

northern bogs can have peat accumulation rates as high as 0.2 cm yr-1 (Collins and Kuehl 2001)

and carbon turnover times exceeding 2000 years (Oades 1988). However, wetlands are both

sources and sinks of organic material, and are affected by hydrologic characteristics such as

proximity to freshwater input, geomorphic orientation of marsh drainage system, and

meteorological phenomena (de la Cruz 1978). A diagram of the carbon cycle (DeBusk 1996 in

Page 20: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

19

Collins and Kuehl 2001) in wetlands is given in Figure 1, and is presented as a series of

compartments of carbon storage.

In the general cycle, inorganic carbon, in the form of CO2, is fixed through the

photosynthesis of plants and algae and stored as biomass. Litterfall and subsequent

decomposition of plant material through leaching, fermentation, and respiratory processes

transfers plant-bound carbon into storage as litter or detritus, microbial biomass, or dissolved

organic carbon (DOC). Incorporation of carbon substances into the soil matrix occurs through

the activity of invertebrates (i.e. bioturbation) or burial under deposited sediment or additional

litter. Soil organic matter can then be immobilized into microbial biomass or undergo further

decomposition, eventually combining with leached products from plant roots and exogenous

Figure 1. Wetland carbon cycle showing the transformations that take place in the production and accumulation of soil organic matter (from DeBusk 1996)

Page 21: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

20

sources to be finally converted and released as CO2 or CH4 by anaerobic respiration and

methanogenesis. The ecological value of carbon cycle processes in wetlands is based on

movement of detritus into neighboring bodies of water and into aquatic food webs, to be

consumed as secondary production by heterotrophic organisms, further broken down by

mechanical or biological processes, or mineralized by bacteria (de la Cruz 1978).

Plant Community Responses to Environmental Conditions

Soil properties can be affected by plant composition, species diversity, and successional

development of the standing vegetation (Marks and Bormann 1972, Parsons and Ware 1982,

Hooper and Vitousek 1997). As we have seen, accumulation of detritus in wetland systems is a

controlling factor in the development of wetland soils (Vepraskas 1994, Megonigal et al. 2004).

In this respect, vegetation provides a feedback mechanism for the development of substrates that

typically characterize natural wetland communities by providing organic matter in the form of

detritus to initiate microbially mediated reduction (Stauffer and Brooks 1997). Therefore, the

structure, age, composition, and density of the standing vegetation represent potential mediating

factors in the development of wetland soils.

Plant Succession

Few topics in ecology are as widely known and widely debated as plant succession. The

concept of succession – namely, the unidirectional, sequential replacement of species within a

community over time (Smith 1990) – has been criticized as a useful model for describing plant

communities (Bazzazz 1996). However, the fact that species shifts through time can be observed

and quantified in natural systems renders the concept useful in discussing ecosystem processes

Page 22: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

21

dependent on vegetative dynamics, particularly in wetlands (van der Valk 1981, Neiring 1987,

King and Allen 1996, Noon 1996, Mitsch and Gosselink 2000, Spencer et al. 2001). This is

perhaps because wetlands have been considered intermediary steps in a “hydrarch succession”

sequence (Wilson 1935, Mitsch and Gosselink 2000) that follows the development of vegetation

from an open water system (lake) to a terrestrial system.

The hydrarch sequence is a concept of autogenic succession, which presupposes that

changes in the community are brought about by the plants themselves (Smith 1990, Barbour et

al. 1999). In this context, plant-soil interactions are perceived as site-specific phenomena –

plants colonize a wetland substrate, contribute organic matter to the substrate, and the attendant

biogeochemical responses in the system influence successional changes to the community over

time. However, the recent conception of wetlands as pulsed systems (Niering 1987, Odum et al.

1995) limits the usefulness of traditional concepts of autogenic succession and the climax

community, in that the development of wetland vegetation is now viewed in response to

environmental conditions (i.e., allogenic succession) predicated by the hydrologic regime and

other geomorphic controls. Therefore, a discussion of vegetation dynamics and plant-soil

interactions in wetlands should consider the autogenic effects of initial floristic composition

(Niering 1987, Noon 1996) and site-specific organic matter inputs from the developing

vegetation community (Campbell et al. 2002), as well as allogenic processes related to other

environmental variables (Niering 1987, Craft 1997, Mitsch and Gosselink 2000).

Autogenic and Allogenic Relationships

Wetland plants retain a diversity of adaptations that allow establishment, growth, and

persistence in anaerobic soil conditions (Cronk and Fennessey 2001). Thus, the types of plants

Page 23: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

22

that may colonize a saturated or inundated soil must have adaptations that allow for rapid growth

and survival in a poorly oxygenated soil environment. The most extensive literature source on

early recruitment and colonization of recently disturbed non-tidal, freshwater wetland substrates

comes from studies in wetland creation and restoration sites, which implicate a diversity of

hydrophytes that can become established in such environments (Reinhartz and Warne 1993,

Noon 1996, Wilson and Mitsch 1996, Campbell et al. 2002, Whigham et al. 2002, DeBerry and

Perry 2004). Under these conditions, aboveground biomass equivalency with adjacent natural

wetlands can be achieved even in the early stages of plant development (Whigham et al. 2002,

DeBerry and Perry 2004), indicating that early colonizers allocate a significant proportion of

growth to aerial plant components. This is presumably facilitated by enhanced photosynthetic

capacity due to solar radiation exposure in the emergent macrophyte community (Brinson et al.

1981), and by the plants themselves – the colonizing species are generally annuals, or facultative

annuals, with the capacity to persist under potentially stressful, low-nutrient conditions (van der

Valk 1981, DeBerry and Perry 2004). As the vegetative community develops, biomass turnover

contributes organic matter to the soil, and the complex suite of biogeochemical transformations

described in previous sections are initiated.

Energy flow in non-tidal, freshwater wetland systems is detritus-based (Day 1984, Mitsch

and Gosselink 2000). As Odum (1969, 1985) points out, young sites typically exhibit open

mineral cycles with low structural complexity. As sites mature, biogeochemical energetics shift

toward closed mineral cycles with high structural complexity. This shift is facilitated by the

incorporation of biogenic organic products into the soil profile, which also functions to increase

water-holding capacity in the system, thereby influencing soil redox state, mineral cycling, and

microbial community development (D’Angelo et al. 2005). Therefore, as the vegetative

Page 24: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

23

community in non-tidal, freshwater wetland sites proceeds along a successional trajectory from

emergent to forested cover types, the potential sources of organic carbon will be augmented with

a parallel increase in structural complexity in the community. Forested systems support a

diversity of growth forms, including trees, shrubs, and understory herbaceous plants, and the

quality of the detritus improves accordingly (i.e., more protein- and nutrient-rich organic

products from leaves, fruits, flowers, tubers, etc.). In addition, plant community development

results in the production of a deep root system, which supports a diversity of soil microbiota and

further influences the redox state of soil via gas transport through the vascular tissues down to

the profile (Ehrenfeld and Toth 1997, Craft 2001,).

As the biogeochemical environment “improves” with respect to bioavailable nutrient and

organic carbon sources, a concomitant response in redox processes mediated by organic matter

inputs further influences the availability of growth-limiting nutrients such as N and P (Armstrong

and Boatman 1967, Gambrell and Patrick 1978, Koerselman et al. 1990, Aerts et al. 1992). The

structural complexity of the system increases, and nutrient availability gradients may become

established across the wetland substrate in response to hydrologic regime and other factors (e.g.,

pH, variable nutrient inputs, etc.) (Aerts et al. 1992, Bridgham et al. 1995, Bragazza and Gerdol

2002).

Such gradients are also influenced by allogenic processes such as nutrient inputs from

exogenous sources in open wetland systems (Craft and Richardson 1993a, Craft et al. 1995, Craft

and Richardson 1997, Cirmo 1998, Chiang et al. 2000). These inputs are regulated by physical

controls such as hydrologic regime and geomorphic setting (Megonigal and Day 1988,

Mausbach and Richardson 1994, Richardson et al. 2001), by the condition of the contributing

upgradient watershed (Brinson et al. 1984, Craft and Richardson 1993b, Newman et al. 2001,

Page 25: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

24

Qualls et al. 2001), and by the source-sink and redox functional status of the wetland for nutrient

subsidies (Bridgham and Richardson 1993,Cirmo 1998). Thus, the distribution and abundance

of plants may change in accordance with resource limitations established by such gradients

(Burke et al. 2003), and a feedback mechanism is established whereby organic carbon inputs

from the plant community moderate the soil biogeochemical setting, and the resultant

biogeochemical setting moderates the distribution and abundance of plant species over the

successional stages of vegetation community development.

Photosynthesis and Respiration

An understanding of rate determinants of photosynthesis and respiration is essential when

considering successional changes and ecosystem maturity based on energy flows in the form of

net carbon exchange in wetlands. Emergent wetlands are among the most productive plant

communities on Earth, owing mostly to near constant water supply and high nutrient

concentrations. Most emergent macrophytes exhibit sun plant characteristics and are capable of

utilizing full sun with minimal photoinhibition at high irradiance levels (Wetzel 2001).

Variations in the photosynthetic efficiency and overall photosynthetic rates of wetland plants are

associated with light (photosynthetically active radiation, PAR), temperature, and soil moisture,

though these variations are also affected by other allogenic and autogenic factors.

Gross photosynthetic production is controlled primarily by latitude and modified by

topography, both of which correspond to control solar radiation, mean annual temperature,

precipitation, and evaporation (Oades 1988). Mann and Wetzel (1999) showed that

photosynthetic efficiency in a common emergent hydrophyte, Juncus effuses, was highest (4.5%)

at the lowest irradiance levels and decreased to about 0.5% at full sunlight without any apparent

Page 26: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

25

photoinhibition. Temperature optima for maximum average net photosynthesis ranged from 35

to 40˚C for J. effuses, but photosynthetic efficiency decreased markedly at higher temperatures.

Similar light and temperature trends have been found for other emergent and floating plant

species, such as Typha latifolia (Wetzel 2001). Although water stress reduces photosynthesis,

growth, and total plant respiration (Ryan 2001), adequate soil moisture is necessary for large

photosynthetic uptake of CO2 (Schreader et al. 1998), as plants constantly lose water through

their stomata while they are photosynthesizing (Knorr 2000). In fact, the stomata of aquatic

macrophytes generally do not close to nearly the extent observed in terrestrial flora during

lighted hours, and thus higher transpiration rates occur in aquatic macrophytes than in land plants

(Wetzel 2001). Photosynthetic water loss rates are determined by vegetation temperature, as

evidenced by the strong temperature dependence of saturated vapor pressure observed in

stomatal cavities (Knorr 2000). Vegetation temperature differs from air temperature due to

sensible heat flux between the vegetation canopy and the free air, controlled largely by the net

radiation of the vegetation canopy and the opening and closing of the stomata (Knorr 2000).

There is also considerable evidence that photosynthetic capacity in both C3 and C4 plants is

closely correlated with leaf nitrogen content (Jones 1988). Indirect determinants of

photosynthesis and, consequently, primary productivity include water flow, and hence

geomorphic position (i.e. “open” or “closed” system), which act to replenish the nutrient and

oxygen supply to roots (Brinson et al. 1981). The more “open” a wetland, the greater the Net

Primary Productivity (NPP), as predictable periodic inundation brings in O2 and nutrient-rich

sediment, and flushes toxins (Craft 2001). “Closed” systems (i.e. bogs, pocosins) receive water

and nutrients only from precipitation, and thus have low NPP compared to “open” systems (Craft

2001).

Page 27: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

26

Ecosystem respiration is the net result of the microbially mediated respiration of soil

organic matter fractions and the metabolic respiration of plants (Trumbore 2000). Respiration

rate determinants in soils include organic matter quality (labile vs. refractory), hydrologic

conditions, and temperature (Chimner 2004), with temperature being the single most important

factor in soil organic matter loss (Brinson et al. 1981). Substrate quality influences carbon

mineralization rates, as greater amounts of labile and refractory material increase and decrease

mineralization rates, respectively. The labile fraction contains easily degradable polysaccharides

such as proteins, carbohydrates, and lipids, whereas refractory or recalcitrant compounds are

resistant to decomposition due to large fractions of high molecular-weight, high C:N ratio

structural components such as lignocellulose, hemicellulose, and complex polysaccharides like

chitin and waxes (Collins and Kuehl 2001, Megonigal et al. 2004). Many studies have found

conflicting results regarding the effect of moisture on litter decomposition (Brinson et al. 1981,

Bayley et al. 1985). Theoretically, the most rapid decomposition rates should occur with aerobic

conditions under a regime of wetting and drying, while continuously anaerobic conditions are

least favorable for decomposition (Brinson et al. 1981). Overall, lower water table depths have

been found to increase soil respiration rates in wetlands due to greater oxygen diffusion into

unsaturated soils leading to more efficient aerobic respiration and increased diffusion of CO2

(Bubier et al. 1998, Chimner 2004). Lastly, although soils formed in submerged conditions do

not follow climate patterns as closely as aerobic soils, as temperature is less influential over a

wider range (Neue 1985), warmer temperatures stimulate microbial activity, resulting in greater

CO2 production from microbially-mediated organic carbon mineralization (Chimner 2004).

Studies have shown that microbial activity and decomposition generally increase with increasing

Page 28: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

27

temperature up to ~35-40˚C (Craft 2001), while temperatures < 5 ۫ C inhibit microbially mediated

redox reactions (Megonigal et al. 1996).

Plants respire about 50% of the carbon derived through photosynthesis after

photorespiration, with the remaining 50% used for growth, propagation, nutrient acquisition, and

litter production (Ryan 1991). Plant dark respiration represents the process of glycolysis and the

oxidative pentose phosphate pathway, the Krebs cycle, and electron transport to oxidative

phosphorylation, with concomitant uptake of O2, and generation of CO2 (Ryan 1991).

Photorespiration is the oxidation of ribulose bisphosphate carboxylase in the presence of oxygen

(Ryan 1991). Rates of respiration, like all enzymatic chemical reactions, increase with

temperature, and show Q10 values (an empirical parameter that relates a 10˚C change in

temperature to a change in activity) of ~2 for herbaceous plants and ~2.3 for woody plants (Ryan

1991). Plant respiration, like photosynthesis is also strongly correlated to tissue nitrogen content,

as most of the organic nitrogen in plants is in protein and ~60% of maintenance respiration

supports protein repair and replacement (Ryan 1991, Reich et al. 1998). Further, atmospheric

CO2 concentrations can either increase (Amthor 1989) or decrease (Bunce 1990) plant

respiration rates, though there are confounding gaps in current knowledge that hamper the

development of adequate models of the response of plant respiration to atmospheric CO2 (Ryan

1991).

Created Forested Wetlands

Wetland creation and restoration in the United States, motivated in part by the

recognition of their many ecological and societal functions, but owing mostly to the policy of

‘No Net Loss’, has commonly occurred since the late 1980’s and early 1990’s. Though some

Page 29: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

28

created and restored wetlands have met mandated permit guidelines of functionality, failure of

these systems is common due mostly to a lack of proper hydrology (Mitsch and Gosselink 2000).

Other problems include improper geomorphic setting, chemical toxicity such as acid mine

drainage, and improper soil conditions including extreme compaction, high bulk density, and low

soil organic matter. The need to improve or boost created wetland functionality has led to the

contemporary research of created wetland systems and various suggested techniques to improve

the success of wetland creation projects. Herein, some aspects of created wetland research will

be explored including needs for further study.

The primary goal of compensatory wetland mitigation, whether creation or restoration, is

to achieve functional equivalency with natural wetlands. It follows, therefore, that the most

common approach to created wetland research is to compare these sites with similar and often

adjacent natural reference wetlands (Kentula et al. 1992, Bishel-Machung et al. 1996, Brinson

and Rheinhardt 1996, Cole and Brooks 2000, Stolt et al. 2000, DeBerry and Perry 2004, Brooks

et al. 2005). Advantages of using a created-reference wetland approach include: 1) specifying

mitigation goals through identifying reference standards from data that typify regional

conditions, 2) providing templates to which mitigation wetlands can be designed, and 3)

establishing a framework whereby a decline or recovery in functions following disturbance can

be estimated for single projects and over large areas through time (Brinson and Rheinhardt

1996). Various parameters have been sampled and compared between created and reference

wetlands, including soil parameters such as organic matter content, matrix chroma, temperature,

and redox potential (Bishel-Machung 1996, Stolt et al. 2000), hydrologic characteristics (Cole

and Brooks 2000), animal communities (Snodgrass et al. 2000, Snell-Rood and Cristol 2003,

Page 30: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

29

Stanczak and Keiper 2004), and a combination of vegetative and environmental characteristics

(Zampella and Laidig 2002, DeBerry and Perry 2004, Balcombe et al. 2005, Brooks et al. 2005).

However, few studies of created wetlands have attempted to apply a concept of primary

succession. Noon (1996) created a model of primary succession in for created wetlands under

the hypothesis that newly exposed hydric soils are colonized primarily by annual species in the

first growing season, perennial species in the second growing season, and vegetative perennials

in sites older than 5 growing seasons. Noon (1996) suggested that there are two phases in early

wetland primary succession: the Arrival and Establishment Phase in the first 3 years after

wetland creation, characterized by successional dependence on chance elements and physical

forces (allogenic), and the Autogenic Dominance Stage after the first 3 years, where succession

depends on plant community related processes. DeBerry and Perry (2004) observed a strong

presence of perennial emergents in a 2-year-old southeastern Virginia created wetland, and found

that the vegetative community did not fit the typical model of primary succession (van der Valk

1981), though it did seem to correspond to Noon’s (1996) model. Further studies are needed on

primary successional trends in created wetlands, as these systems are an example of one of the

only situations where early succession truly occurs, due to the removal of substrate including

seed banks.

Still fewer studies utilize the concept of net ecosystem CO2 exchange (Net Ecosystem

Exchange, NEE) as a tool to characterize the successional maturity of created wetland systems.

A variety of researchers have used NEE techniques to observe CO2 fluxes between wetlands and

the atmosphere (Bubier et al. 1998, Frolking et al. 1998, Schreader et al. 1998, Streever et al.

1998, Clark et al. 1999, Wickland et al. 2001). Most studies of this kind have been carried out in

boreal or subarctic peatland of North America (Bubier et al. 1998, Frolking et al. 1998,

Page 31: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

30

Schreader et al. 1998), and generally show that soil temperatures, incident solar radiation, trophic

status, and water table were closely coupled to NEE. Whereas intuitively carbon flux should

flow in the direction of wetland storage (Bubier et al. 1998), Shreader et al. (1998) showed that a

subarctic sedge fen was a net source of CO2 to the atmosphere during an unusually hot, dry

summer, suggesting that climate warming would need to be accompanied by a large rainfall

increase to maintain an average annual condition of net carbon gain. Similarly, Wickland et al.

(2001) measured CO2 and CH4 exchange between the atmosphere and a Rocky Mountain

subalpine wetland, and indicated that the wetland was a net source of carbon gas to the

atmosphere over the 3-year sampling period, despite the long-term carbon accumulation of ~0.7

mol m-2 yr-1 determined from 14C analysis. Clark et al. (1999) compared NEE between an

evergreen pine upland and deciduous cypress wetland ecosystem in Florida, and observed a

relatively low rate of annual carbon accumulation in the cypress wetland relative to the pine

ecosystem. Streever et al. (1998) described an in situ closed chamber technique to measure CO2

flux in marshes, for use in estimating marsh net productivity, comparing the productivity of two

of more marshes, or assessing the factors that influence productivity. Perhaps an underutilized

purpose for NEE measurements in wetlands is to determine the successional state of these

systems, where high biomass production in early successional stages should reflect a progression

from a net autotrophic regime (photosynthesis (P) / respiration (R) ratio >1) to a net

heterotrophic regime (P/R = 1) (Odum 1969). Such uses of NEE measurements have only

recently been applied to wetlands (Roggero 2003), but are absent from studies of created non-

tidal wetland systems.

Page 32: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

31

Objectives and Hypothesis

The objectives of this study were to measure and compare the effects of various loading

rates of an organic matter amendment in relation to 1) community composition and standing crop

of herbaceous vegetation, 2) development of planted woody vegetation, and 3) ecosystem carbon

flux measured as Net Ecosystem Exchange (NEE) of CO2, within a created freshwater wetland.

The preceding parameters were used to quantify the effects of incorporation of an organic

matter soil amendment on several metrics of vegetation growth and community structure as well

as carbon (CO2) flux in the created wetland. Comparing the vegetation and gas exchange

components among the various organic matter loading rates, relative soil nutrient concentrations,

and other parameters such as soil surface and water table elevation provided comparisons of

ecosystem function among the created wetland plots on a local scale, and provided insight into

carbon sequestration/release processes on a global scale. Lastly, this information was related to

ecological and management perspectives to discern how adding organic matter to created

wetlands influences created wetland success.

Null Hypothesis: Herbaceous plant community composition and biomass, planted woody

vegetation, and Ecosystem Gas Exchange will not vary with respect to organic matter

loading rate.

Alternate Hypothesis: Herbaceous plant community composition and biomass, planted

woody vegetation, and Ecosystem Gas Exchange will vary with respect to organic matter

loading rate.

Page 33: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

32

Vegetation community composition, biomass, and tree size are addressed in Chapter 2,

while ecosystem gas flux is addressed in Chapter 3. Chapter 4 presents a brief summary of the

combined work.

Page 34: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

33

Literature Cited

Aerts, R., B. Wallen and N. Malmer. 1992. Growth-limiting nutrients in sphagnum-dominated

bogs subject to low and high atmospheric nitrogen supply. Journal of Ecology 80:131-140.

Ambus, P. and R. Lowrance. 1991. Comparison of denitrification in two riparian soils. Soil Science Society of America Journal 55:994-997.

Amthor, J.S. 1989. Respiration and crop productivity. Springer-Verlag, New York. USA.

Anderson, C.J. and B.C. Cowell. 2004. Mulching effects on the seasonally flooded zone of west-central Florida, USA wetlands. Wetlands 24:811-819.

Armentano, T.V. and E.S. Menges. 1986. Patterns of Change in the Carbon Balance of Organic Soil-Wetlands of the Temperate Zone. The Journal of Ecology 74:755-774.

Armstrong, W. and D.J. Boatman. 1967. Some field observations relating the growth of bog plants to conditions of soil aeration. Journal of Ecology 55:101-110.

Balcombe, C.K., Anderson, J.T., Fortney, R.H., Rentch, J.S., Grafton, W.N., and Kordek, W.S. 2005. A comparison of plant communities in mitigation and reference wetlands in the mid-Appalachians. Wetlands 25:130-142.

Barbour, M.G., J.H. Burk, W.D. Pitts, F.S. Gilliam and M.W. Schwartz. 1999. Terrestrial Plant Ecology. Benjamin/Cummings, Menlo Park, CA.

Bayley, S.E., J. Zoltek, Jr., A.J. Hermann, T.J. Dolan and L. Tortora. 1985. Experimental manipulation of nutrients and water in a freshwater marsh: effects on biomass, decomposition, and nutrient accumulation. Limnology and Oceanography 30:500-512.

Bazzazz, F. A. 1996. Plants in Changing Environments: Linking Physiological, Population, and Community Ecology. Cambridge University Press, Cambridge, UK.

Bishel-machung, L., R.P. Brooks, S.S. Yates and K.L. Hoover. 1996. Soil properties of reference wetland and wetland creation projects in Pennsylvania. Wetlands 16:532-541.

Bragazza, L. and R. Gerdol. 2002. Are nutrient availability and acidity-alkalinity gradients related in sphagnum-dominated peatlands? Journal of Vegetation Science 13:473-482.

Bridgham, S.D. and C. J. Richardson. 1993. Hydrology and nutrient gradients in North Carolina peatlands. Wetlands 13:207-218.

Bridgham, S.D., J. Pastor, C.A. McClaugherty and C. J. Richardson 1995. Nutrient-use efficiency: A litterfall index, a model, and a test along a nutrient-availability gradient in North Carolina peatlands. American Naturalist 145:1-21.

Page 35: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

34

Brinson, M.M. and R. Rheinhardt. 1996. The role of reference wetlands in functional assessment and mitigation. Ecological Applications 6:69-76.

Brinson, M.M., Bradshaw, H.D., & Kane, E.S. 1984. Nutrient assimilative capacity of an alluvial floodplain swamp. Journal of Applied Ecology 21:1041-1057.

Brinson, M. M., A.E. Lugo and S. Brown. 1981. Primary productivity, decomposition and consumer activity in freshwater wetlands. Annual Review of Ecology and Systematics 12:123-161.

Brooks, R.P., D.H. Wardrop, C.H. Cole and D.A. Cambell. 2005. Are we purveyors of wetland homogeneity? A model of degradation and restoration to improve wetland mitigation performance. Ecological Engineering 24:331-340.

Bruland, G.L. and C.J. Richardson. 2004. Hydrologic gradients and topsoil additions affect soil properties of Virginia created wetlands. Soil Science Society of America Journal 68:2069-2077.

Bubier, J.L., P.M. Crill, T.R. Moore, K. Savage and R. K. Varner. 1998. Seasonal patterns and controls on net ecosystem CO2 exchange in a boreal peatland landscape. Global Biogeochemical Cycles 12:703-714.

Bunce, J.A. 1990. Short- and long-term inhibition of respiratory carbon dioxide efflux by elevated carbon dioxide. Annals of Botany 65:637-642.

Burke, M.K., S. LKing, D. Gartner and M.H. Eisenbies. 2003. Vegetation, soil, and flooding relationships in a blackwater floodplain forest. Wetlands 23:988-1002.

Campbell, D. A., C. A. Cole, and R. P. Brooks. 2002. A comparison of created and natural wetlands in Pennsylvania, USA. Wetlands Ecology and Management 10:41-49.

Chadwick, O.A. and R.C. Graham. 2000. Pedogenic processes. p. E41-E75. In: Sumner, M.E. Sumner (ed.). Handbook of Soil Science. CRC Press, Boca Raton, FL.

Chapin, C.T., S.D. Bridgham and J. Pastor. 2004. pH and nutrient effects on above-ground net primary production in a Minnesota, USA bog and fen. Wetlands 24:186-201.

Chiang, C., C.B. Craft, D.W. Rogers, and C.J. Richardson. 2000. Effects of 4 years of nitrogen and phosphorus additions on everglades plant communities. Aquatic Botany, 68:61-78.

Chimner, R.A. 2004. Soil respiration rates of tropical peatlands in Micronesia and Hawaii. Wetlands 24:51-56.

Cirmo, C.P. 1998. Linkages between hydrology and nitrogen biogeochemistry in riparian wetlands. In Majumdar, S.K., E.W. Miller, and F.J. Brenner (eds.). Ecology of Wetlands and Associated Systems. Pennsylvania Academy of Science.

Page 36: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

35

Clark, K.L., H.L. Gholz, J.B. Moncrieff, F. Cropley and H.W. Loescher. 1999. Environmental controls over net exchanges of carbon dioxide from contrasting Florida ecosystems. Ecological Applications 9:936-948.

Cole, C.A. and R.P. Brooks. 2000. A comparison of the hydrologic characteristics of natural and created mainstem floodplain wetlands in Pennsylvania. Ecological Engineering 14:221231.

Cole, C.A. 2002. The assessment of herbaceous plant cover in wetlands as an indicator of function. Ecological Indicators 2:287-293.

Collins, M.E. and R.J. Kuehl. 2001. Organic matter accumulation in organic soils. p. 137-162. In J.L. Richardson and M.J. Vepraskas (eds.) Wetland Soils: Genesis, Hydrology, Landscapes, and Classification. Lewis Publishers.

Craft, C.B. 1997. Dynamics of nitrogen and phosphorus retention during wetland ecosystem succession. Wetlands Ecology and Management 4:177-187.

Craft, C.B. 2001. Biology of wetland soils. p. 107-135. In J.L. Richardson and M.J. Vepraskas (eds.) Wetland Soils: Genesis, Hydrology, Landscapes, and Classification. Lewis Publishers.

Craft, C.B. and C. J. Richardson, 1993a. Peat accretion and N, P, and organic C accumulation in nutrient-enriched and unenriched Everglades peatlands. Ecological Applications 3:446-458.

Craft, C.B. and C. J. Richardson, 1993b. Peat accretion and phosphorus accumulation along a eutrophication gradient in the northern Everglades. Biogeochemistry, 22:133-156.

Craft, C.B., J. Vymazal and C. J. Richardson. 1995. Response of Everglades plant communities to nitrogen and phosphorus additions. Wetlands 15:258-271.

Cronk, J.K. and M.S. Fennessey. 2001. Wetland Plants: Biology and Ecology. Lewis Publishers, Boca Raton, Florida.

D’Angelo, E.M., A.D. Karathanasis, E.J. Sparks, S.A. Ritchey and S.A. Wehr-McChesney. 2005. Soil carbon and microbial communities at mitigated and late successional bottomland forest wetlands. Wetlands 25:162-175.

Dahl, T.E. 1990. Wetlands: losses in the United States 1780’s to 1980’s. U.S. Fish and Wildlife Service., Washington DDC, USA. NTIS Report PB91-169284.

Day, Jr., F.P. 1984. Biomass and litter accumulation in the Great Dismal Swamp. p. 386-392. In Ewel, K. C. and H. T. Odum (eds.). Cypress Swamps. Gainesville, FL: University Presses of FL.

de la Cruz, A.A. 1978. Production and transport of detritus in wetlands. p. 162-174. In Greeson, P.E., J.R. Clark and J.E. Clark (eds.) Wetland Functions and Values: The State of Our Understanding. American Water Resources Association. Minneapolis, MN.

Page 37: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

36

DeBerry, D.A. and J.E. Perry. 2004. Primary succession in a created freshwater wetland. Castanea 69:185-193.

DeBusk, W.F. 1996. Organic matter turnover along a nutrient gradient in the Everglades. Ph.D. dissertation. Univ. of Florida., Gainesville.

DeLaune, R.D., Boar, R R., Lindau, C.W. and B.A. Kleiss, 1996. Denitrification in bottomland hardwood wetland soils of the Cache River. Wetlands 16:309-320.

Environmental Laboratory. 1987. Corps of Engineers Wetland Delineation Manual. Technical Report Y-87-1, US Army Engineer Waterways Experiment Station, Vicksburg, Miss.

Ehrenfeld, J.G. and L.A. Toth. 1997. Restoration ecology and the ecosystem perspective. Restoration Ecology 5:307-317.

Frolking, S.E., J.L. Buiger, T.R. Moore, T. Ball, L.M. Bellisario, A. Bhardwj, P. Carrol, P.M. Crill, P.M. Lafleur, J.H. McCaughey, N.T. Roulet, A.E. Suyker, S.B. Verma, J.M. Waddington and G.J. Whiting. 1998. Relationship between ecosystem productivity and photosyntheitcally active radiation for northern peatlands. Global Biogeochemical Cycles 12:115-126.

Gambrell, R.P. and W.H. Patrick, Jr. 1978. Chemical and microbiological properties of anaerobic soils and sediments. In Hook, D. D. and R. M. M. Crawford (eds.). Plant Life in Anaerobic Environments. Ann Arbor Science, Ann Arbor, MI.

Gambrell, R.P., Patrick, W.H., Jr. and S.P. Faulkner. 1989. Extractable iron and manganese and redox changes in bottomland hardwood wetland-nonwetland transition zone soils. General Technical Report SE - U.S.Department of Agriculture, Forest Service, Southeastern Forest Experiment Station, (50), 117-122.

Groffman, P. M. 1994. Denitrification in freshwater wetlands. Current Topics in Wetland Biogeochemistry 1:15-35.

Hooper, D.U. and P.M. Vitousek. 1997. The effects of plant composition and diversity on ecosystem processes. Science 277:1302-1305.

Jenny, H. 1941. Factors of Soil Formation. McGraw-Hill, New York, NY.

Johns, D., H. Williams, K. Farrish, and S. Wagner. 2004. Denitrification and soil characteristics of wetlands created on two mine soils in east Texas, USA. Wetlands 24:57-67.

Jones, M.B. 1988. Photosynthetic responses of C3 and C4 wetland species in a tropical swamp. Journal of Ecology 76:253-262.

Kentula, M.E., R.P. Brooks, S.E. Gwin, C.C. Holland, A.D. Sherman amd J.C. Sinfneos. 1992. An approach to improved decision making in wetland restoration and creation. U.S. Environmental Protection Agency, Environmental Research Laboratory, Corvalis, OR, USA.

Page 38: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

37

King, S.L. and J.A. Allen. 1996. Plant succession and greentree reservoir management: implications for management and restoration of bottomland hardwood wetlands. Wetlands 16:503-511.

Knorr, W. 2000. Annual and interannual CO2 exchanges of the terrestrial biosphere: process-based simulations and uncertainties. Global Ecology and Biogeography 9:225-252.

Koerselman, W., S.A. Bakker and M. Blom. 1990. Nitrogen, phosphorus and potassium budgets for two small fens surrounded by heavily fertilized pastures. Journal of Ecology 78:428-442.

Mann, C. and R. Wetzel. 1999. Photosynthesis and stomatal conductance of Juncus effuses in a temperate wetland ecosystem. Aquatic Botany 63:127-144.

Marks, P. L. and F. H. Bormann. 1972. Revegetation following forest cutting: mechanisms for return to steady-state nutrient cycling. Science 176:914-915.

Mausbach, M.J., and J.L. Richardson. 1994. Biogeochemical processes in hydric soil formation. Current Topics in Wetland Biogeochemistry 1:68-127.

McKinstry, M.C. and S.H. Anderson. 2003. Improving aquatic plant growth using propagules and topsoil in created bentonite wetlands of Wyoming. Ecological Engineering 21:175-189.

Megonigal, J. P., and F.P. Day, Jr. 1988. Organic matter dynamics in four seasonally flooded forest communities of the dismal swamp. American Journal of Botany 75: 1334-1343.

Megonigal, J.P., M.E. Hines, and P.T. Visscher. 2004. Anaerobic metabolism: linkages to trace gases and aerobic processes. p. 317-424. In Schlesinger, W.H. (ed.). Biogeochemistry. Elsevier-Pergamon, Oxford, UK.

Megonigal, J. P. and W. H. Schlesinger. 1997. Enhanced CH4 emissions from a wetland soil exposed to elevated CO2. Biogeochemistry 37:77-88.

Megonigal, J. P., S. P. Faulkner and W. H. Patrick 1996. Microbial activity season in southeastern hydric soils. Soil Science Society of America Journal 60:1263-1266.

Mendelssohn, I.A., Kleiss, B.A., & Wakeley, J.S. 1995. Factors controlling the formation of oxidized root channels: A review. Wetlands 15:37-46.

Mitsch, W.J. and J.G. Gosselink. 2000. Wetlands. 3rd Edition. John Wiley and Sons, Inc.

Mohanty, S.K. and R.N. Dash. 1982. The chemistry of waterlogged soils. p. 389-396. In Gopal B., R.E. Turner, R.G. Wetzel and D.F. Whigham (eds.). Wetlands – Ecology and Management. Natural Institute of Ecology and International Scientific Publications, Jaipur, India.

Moran, M.A., R. Benner and R.E. Hodson, 1989. Kinetics of microbial degradation of vascular plant material in two wetland ecosystems. Oecologia, 79:158-167.

Page 39: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

38

Nair, V.D., D.A. Graetz, K R. Reddy, and O.G. Olila. 2001. Soil development in phosphate-mined created wetlands of Florida, USA. Wetlands 21:232-239.

National Research Council (NRC). 1995. Wetlands, Characteristics and Boundaries. National Academy Press, Washington, DC, 307 pp.

National Technical Committee for Hydric Soils (NTCHS). 1985. Hydric Soils of the United States. USDA Soil Conservation Service, Washington, D.C.

Nealson, K.H. and D. Saffarini. 1994. Iron and manganese in anaerobic respiration: environmental significance, physiology, and regulation. Annual Review of Microbiology 48:311-343.

Neue, H.U. 1985. Organic matter dynamics in wetland soils. p. 109-122. In Wetland Soils: Characterization, Classification, and Utilization. Proceedings of workshop, March 26-April 5, 1984, Manila, Philippines. International Rice Institute, Los Banos, Philippines.

Newman, S., S. Kumpf, J.A. Laing and W.C. Kennedy. 2001. Decomposition responses to phosphorus enrichment in an Everglades (USA) slough. Biogeochemistry 54:229-250.

Niering, W. A. 1987. Vegetation dynamics (succession and climax) in relation to plant community management. Conservation Biology 1:287-295.

Noon, K.F. 1996. A model of created wetland primary succession. Landscape and Urban Planning 34:97-123.

Oades, J.M. 1988. The retention of organic matter in soils. Biogeochemistry 5:35-70.

Odum, E.P. 1969. The strategy of ecosystem development. Science 164:262-270.

Odum, E.P. 1978. The value of wetlands: a hierarchical approach. In Gresson, P.E., J.R. Clark and J.E. Clark (eds.) Wetland Functions and Values: the State of Our Understanding, Proceedings from the National Symposium on Wetlands. American Water Resources Association, Lake Buena Vista, Florida.

Odum, E. P. 1985. Trends expected in stressed ecosystems. Bioscience 35:419-422.

Odum, W.E., E.P. Odum and H.T. Odum. 1995. Nature’s pulsing paradigm. Estuaries 18:547-555.

Phung, H.T. and J.G.A. Fiskell. 1973. A review of redox reactions in soils. Proceedings of the Soil and Crop Science Society of Florida 32:141-145.

Poe, A.C., M.F.Piehler, S.P. Thompson and H.W. Paerl. 2003. Denitrification in a constructed wetland receiving agricultural runoff. Wetlands 23: 817-826.

Ponnamperuma, F.N. 1972. The chemistry of submerged soils. p. 29-96. In N.C. Brady, (ed.). Advances in Agronomy. Academic Press, New York.

Page 40: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

39

Qualls, R.G. and C.J. Richardson. 1995. Forms of soil phosphorus along a nutrient enrichment gradient in the northern Everglades. Soil Science 160:183-198.

Qualls, R. G. and C. J Richardson. 2000. Phosphorus enrichment affects litter decomposition, immobilization, and soil microbial phosphorus in wetland mesocosms. Soil Science Society of America Journal 64:799-808.

Qualls, R.G., C. J. Richardson, and L.J. Sherwood. 2001. Soil reduction-oxidation potential along a nutrient-enrichment gradient in the everglades. Wetlands 21:403-411.

Reich, P.B., M.B. Walters, D.S. Ellsworth, J.M. Vose, J.C. Volin, C. Gresham and W.D. Bowman. 1998. Relationships of leaf dark respiration to leaf nitrogen, specific leaf area and leaf life-span: a test across biomes and functional groups. Oecologia 114:471-482.

Reinhartz, J.A. and E.L. Warne. 1993. Development of vegetation in small created wetlands in southeastern Wisconsin. Wetlands 13: 153-164.

Richardson, C. J. 1985. Mechanisms controlling phosphorus retention capacity in freshwater wetlands. Science 228:1424-1427.

Richardson, C.J., J.L. Arndt and J.A. Montgomery. 2001. Hydrology of wetland and related soils. p. 35-84. In Richardson, J.L. and M.J. Vepraskas (eds.). Wetland Soils: Genesis, Hydrology, Landscapes, and Classification. Lewis Publishers, Boca Raton, FL.

Roggero, M.M. 2003. Ecosystem gas exchange in natural and created tidal salt marshes of Tidewater, Virginia. Masters Thesis. The College of William and Mary.

Ryan, M.G. 1991. Effects of climate change on plant respiration. Ecological Applications 1:157-167.

Schreader, C.P., W.R. Rouse, T.J. Griffis, L.D. Boudreau and P.D. Blanken. 1998. Carbon dioxide fluxes in a northern fen during a hot, dry summer. Global Biogeochemical Cycles 12:729-740.

Smith, R. L. 1996. Ecology and Field Biology, 5th edition. HarperCollins College Publishers, New York, NY.

Snell-Rood, E.C. and D.A. Cristol. 2003. Avian communities of created and natural wetlands: bottomland forests in Virginia. The Condor 105:303-315.

Snodgrass, J.W., A.L. Bryan and J. Burger. 2000. Development of expectations of larval amphibian assemblage structure in southeastern depression wetlands. Ecological Applications 10:1219-1229.

Spencer, D.R., J.E. Perry, and G.M. Silberhorn. 2001. Early secondary succession in bottomland hardwood forests of Southeastern Virginia. Environmental Management 27:559-570.

Page 41: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

40

Stanczak, M. and J.B. Keiper. 2004: Benthic invertebrates in adjacent created and natural wetlands in northeastern Ohio, USA. Wetlands 24:212-218.

Stauffer, A.L. and R.P. Brooks. 1997. Plant and soil responses to salvaged marsh surface and organic matter amendments at a created wetland in central Pennsylvania. Wetlands 17:90-105.

Stevens, P.R. and T.W. Walker. 1970. The chronosequence concept and soil formation. The Quarterly Review of Biology 45:333-350.

Stolt, M.H., M.H. Genthner, W.L. Daniels, V.A. Groover, S. Nagel and K.C. Haering. 2000. Comparison of soil and other environmental conditions in constructed and adjacent palustrine reference wetlands. Wetlands 20:671-683.

Streever, W.J. A.J. Genders and M.A. Cole. 1998. A closed chamber CO2 flux method for estimating marsh productivity. Aquatic Botany 62:33-44.

Trumbore, S. 2000. Age of soil organic matter and soil respiration: radiocarbon constraints on belowground C dynamics. Ecological Applications 10:399-411.

United States Environmental Protection Agency (USEPA). 1989. Highlights of section 404: federal regulatory program to protect waters of the United States. Office of Wetlands Protection, Washington DC, USA.

van der Valk, A.G. 1981. Succession in wetlands: a Gleasonian approach. Ecology 62:688-696.

Vepraskas, M. J. 1994. Redoximorphic features for identifying aquic conditions. North Carolina Agricultural Research Service, Raleigh, NC. Technical Bulletin 301.

Vepraskas, M.J. and S.P. Faulkner. 2001. Redox chemistry of hydric soils. p. 85-105. In J.L. Richardson and M.J. Vepraskas (eds.) Wetland Soils: Genesis, Hydrology, Landscapes, and Classification. Lewis Publishers.

Wang, Z. and W.H. Patrick, Jr. 2000. Anaerobic microbially mediated processes. p. C120-C129. In M.E. Sumner (ed.). Handbook of Soil Science. CRC Press, Boca Raton, FL.

Weiss, J.V., D. Emerson, S.M. Backer, and J.P. Megonigal. 2003. Enumeration of Fe(II) and Fe(III)-reducting bacteria in the root zone of wetland plants: implications for the rhizosphere iron cycle. Biogeochemistry 64:77-96.

Wetzel, R.G. 2001. Limnology: Lake and River Ecosystems. 3rd Edition. Academic Press.

Wilson, L.R. 1935. Lake development and plant succession in Vilas County, Wisconsin. Part 1, the medium hard water lakes. Ecological Monographs 5:207-247.

Wilson, R. F. and W. J. Mitsch. 1996. Functional assessment of five wetlands constructed to mitigate wetland loss in Ohio, USA. Wetlands 16:436-451.

Page 42: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

41

Whigham, D.F. 1999. Ecological issues related to wetland preservation, restoration, creation and assessment. The Science of the Total Environment 240:31-40.

Whigham, D., M. Pittek, K. H. Hofmockel, T. Jordan and A.L. Pepin. 2002. Biomass and nutrient dynamics in restored wetlands on the outer coastal plain of Maryland, USA. Wetlands, 22:562-574.

Whiting, G.J. and J.P. Chanton. 1993. Primary production control of methane emission from wetlands. Nature 364:794-795.

Whittecar, G.R. and W.L. Daniels. 1999. Use of hydrogeomorphic concepts to design created wetlands in southeastern Virginia. Geomorphology 31:355-371.

Wickland, K.P., R.G. Striegl, M.A. Mast and D.W. Chow. 2001. Carbon gas exchange at a southern Rocky Mountain wetland, 1996-1998. Global Biogeochemical Cycles 15:321-335.

Wieder, R.K., J.B.Yavitt and , G.E. Lang. 1990. Methane production and sulfate reduction in two Appalachian peatlands. Biogeochemistry 10:81-104.

Zampella, R.A. and K.J. Laidig. 2003. Functional equivalency of natural and excavated coastal plain ponds. Wetlands 23:860-876.

Page 43: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

42

Chapter 2

Wetland vegetation dynamics in response to organic matter loading rates

INTRODUCTION

Permits to drain or fill wetlands granted under section 404 of the 1977 Clean Water Act

are often accompanied by compensatory mitigation requirements designed to replace the

functions [i.e. ecosystem services (sensu Odum 1978)] once performed by the disturbed area.

Thus, the primary goal of such final mitigation alternatives is to achieve functional equivalency

with natural wetlands. The failure of created wetlands to achieve this goal is due mostly to a

lack of proper hydrology (Mitsch and Gosselink 2000), but can also be attributed to improper

geomorphic setting, lack of microtopography, extreme soil compaction (i.e. high bulk density),

improper soil chemical conditions, and low soil organic matter (Whittecar and Daniels 1999).

Variable but low amounts of soil organic matter have been reported for created wetlands

in Pennsylvania [0.5-0.9% (Stauffer and Brooks 1997), 2.3-6.5% (Cole et al. 2001), 4% (Brooks

et al. 2005)], Florida [2.6 ± 0.3% (Anderson and Cowell 2004)], and Virginia [0.9-1.9%

(Whittecar and Daniels 1999), 3.5-7.2% (Bruland and Richardson 2004)]. These values can be

compared to organic matter amounts in natural reference wetlands of 21.7 ± 20.4% in

Pennsylvania (Bischel-Machung et al. 1996), 13.8% in Florida (Brown 1991), and 2.4-11% in

Virginia (Stolt et al. 2000).

Organic matter generally accumulates to a maximum in ecosystems as succession

proceeds (Odum 1969), leading many studies to advocate amending wetland creation sites with

organic material, either in the form of salvaged natural wetland soils or mulches, to help

Page 44: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

43

mitigation wetlands achieve functional equivalency sooner (Stauffer and Brooks 1997, Whittecar

and Daniels 1999, McKinstry and Anderson 2003, Anderson and Cowell 2004, Bruland and

Richardson 2004). Indeed, studies comparing mulched and non-mulched created wetlands are

common (Bischel-Machung et al. 1996, Brinson and Rheinhardt 1996, Cole and Brooks 2000,

Stolt et al. 2000, Brooks et al. 2005). Also, many parameters have been sampled and compared

between created and reference wetlands, including soil parameters such as organic matter

content, matrix chroma, temperature, and redox potential (Bischel-Machung et al. 1996, Stolt et

al. 2000), hydrologic characteristics (Cole and Brooks 2000), animal communities (Snodgrass et

al. 2000, Snell-Rood and Cristol 2003, Stanczak and Keiper 2004), and a combination of

vegetative and environmental characteristics (Zampella and Laidig 2003, DeBerry and Perry

2004, Balcombe et al. 2005, Brooks et al. 2005). However, few studies (Bergschneider 2005)

have explored the differences in created wetland ecosystem functions with several different

loadings of an organic soil amendment (e.g an organic carbon gradient). Early successional

processes (i.e. primary succession) are also rarely studied in wetlands (van der Valk 1981), and

are only now beginning to be investigated in created wetlands (Noon 1996, DeBerry and Perry

2004).

The purpose of this study was to examine the primary successional processes occurring in

a created forested wetland, in terms of vegetative characteristics, along a gradient of soil organic

carbon. Our goal was to determine how vegetation composition, standing crop biomass, and

woody vegetation development varied in a created freshwater wetland with respect to different

loadings of a soil organic matter amendment. Our hypothesis was that the vegetative parameters

would vary with respect to the soil organic carbon-loading gradient. Using the results of this

Page 45: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

44

study, we were able to develop recommendations for organic matter amendments in created

wetland systems.

SITE DESCRIPTION

The Charles City Wetland Mitigation Site (CCW) is a 21 ha constructed mitigation

wetland owned by the Virginia Department of Transportation (VDOT) in Charles City County,

VA (76˚55’33” W, 37˚20’37” N) (Figure 1a). The site can be classified as palustrine emergent

headwater wetlands (Cowardin et al. 1979, DeBerry and Perry 2004), with 18 ha designated as

forested wetlands. In 1996, the upper soil profile was excavated to an initial depth of -0.5 to -0.6

m (Bergschneider 2005). However, during summer and fall 2003 much of the site was regraded

to a lower elevation, and some areas were ‘ripped’ to an approximate depth of 0.25-0.5 m to

decrease the effects of compaction and create more favorable conditions for vegetation (Schmidt

2002). Although construction specifications called for the replacement of topsoil, there is no

evidence that such replacement occurred. CCW is characterized by surface exposure of a plastic

E or argillic horizon (Btg) that typically exceeds 1 m in depth (Bergschneider 2005).

Precipitation is the dominant hydrologic factor in CCW, and fall and winter months are generally

accompanied by up to 0.6 m of standing water (Schmidt 2002). In the summer and fall, surface

water appears to be perched over the restrictive Btg, particularly following heavy rain events.

Following initial grading, the site was stabilized with a seed mix of annual ryegrass (Lolium

perenne ssp. multiflorum) and switchgrass (Panicum virgatum) (DeBerry and Perry 2004).

The study site consisted of a 680 m2 area located along the northern edge of the CCW

(Figure 1b) that contained twenty 4.57 x 3.05 m (15 x 10 ft) plots separated by 3.05 m (10 ft)

Page 46: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

45

alleyways. In June 2002 each plot received one of five organic matter loading rates, ranging

from 0 to 336 Mg ha-1 (Table 1) intended to 1) bracket the currently utilized and recommended

rates employed by the wetland design and construction industry, and 2) coincide with rates

reported for other disturbed land re-vegetation studies in Virginia (Bergschneider 2005). The

plots were arranged in a randomized complete block design (Figure 2), creating four treatment

plot replicates per loading rate. Bergschneider (2005) tested for and found no block x treatment

effects within the experimental site. The organic amendment material consisted of dry, mixed

wood and yard waste compost processed by Grind-All LLC of Richmond, Virginia. This

product was chosen due to its relative stability, history of use, moderate degree of

decomposition, and relatively low total nitrogen content (Bergschneider 2005). Chemical

composition analysis of the amendment material is listed in Appendix 1. Amendments were

incorporated into the plots in June 2002 via disking and roto-tilling by tractor. Plots in loading

rates (LRs) 4 and 5 supported a layer of unincorporated organic amendment (compost) atop the

mineral soil surface, with a thicker layer in LR 5 than 4. Bergschneider’s (2005) study

confirmed an organic matter mounding effect in 2003 by showing that redox potential (EH)

readings in LRs 4 and 5 were consistently higher than those in LRs 2 and 3, presumably due to

gains in surface elevation and subsequent increases in oxygen diffusion rates.

Between December 1 and 15, 2002, five Betula nigra and Quercus palustris saplings

were planted on 1.2 m (4 ft) centers in each experimental plot. Each sapling was fertilized with

two (2) 16-8-12 controlled release fertilizer tablets buried within the planting pit near the tree

roots. (Note: In late September 2005, the authors noted that 3 of the assumed B. nigra saplings

[all in plot 20] were actually B. alleghaniensis). Further details regarding the preparation of the

experimental plots can be found in Bergschneider (2005).

Page 47: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

46

METHODS

Elevation

Relative elevations of each plot were measured on July 18, 2005 using a Topcon® AT-G7

Autolevel and a standard stadia rod. Elevation measurements were collected from the center

point of each plot and from the base of each B. nigra sapling (up to 6 total points per plot). The

relative measures were then compared to a known surveyed elevation benchmark and converted

to elevations above sea level. The average study site elevation was 10.4 m above sea level,

although significant variability existed among plots.

Soil Nutrients

One sample of the top 10 cm of the soil profile (from the soil surface) was collected in

each plot on August 22, 2005 using a posthole digger. Each sample was thoroughly mixed and

analyzed for total carbon (C) and nitrogen (N) content via a controlled combustion Elemental

Analyzer (Nelson and Sommers 1996). Due to the assumed lack of carbonates present in the soil

at CCW (Bergschneider 2005), total C was presumed to approximately equal organic carbon.

Dilute double-acid extractable phosphorus (P), along with other micronutrients (K, Ca, Mg, Zn,

Mn, Cu, Fe, and B) (Appendix 2), was determined by Inductively Coupled Plasma Spectroscopy

(Donohue and Heckendorn 1996). All soil nutrient analyses were performed by the Virginia

Tech Soil Testing Laboratories.

Plant Assemblage

Page 48: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

47

Herbaceous vegetation was sampled monthly (April-October 2005) in each of the 20

plots for community composition. Measurements of vegetative cover were collected from two

randomly placed 1 m x 1 m PVC quadrats in each plot. Percent cover per species was visually

estimated directly in the field as a value of 1 to 100% or trace (<1%) using a modified Braun-

Blanquet cover scale (Daubenmire 1959, DeBerry and Perry 2004) where: <1% = trace, 1 to 5%

= 3%, 5 to 25% = 15%, 25 to 50% = 37.5%, 50 to 75% = 62.5%, 75 to 95% = 85%, and 95 to

100% = 97.5%. Standing dead plant material and bare ground were treated as unique species.

Plant taxonomy and nomenclature followed Gleason and Cronquist (1991) or the Natural

Resources Conservation Service (2006) Plants Database.

Estimates of cover and frequency were converted to relative measures, and Importance

Values (IV) were calculated as the sum of relative cover and relative frequency for each species

(Atkinson et al. 1993). The IV of each treatment replicate were averaged over the growing

season to calculate IV at each loading rate. Dominant species for each treatment were selected

by ranking in order of decreasing IV, with dominants comprising the first 50% of the total and

any additional species greater than 20% (50:20 rule). Species Richness (SR) was determined

both as the total number of species for each loading rate during the 2005 growing season as a

whole and as a per quadrat (m-2) average. Evenness (J’) and the Shannon Diversity Index (H’)

values (Zar 1984) were calculated for each loading rate using IV and SR data. These values

were calculated with standing dead and bare ground removed from the calculations, but

including planted tree species Betula nigra, B. alleghaniensis, and Quercus palustris; removing

these species from SR, J’, and H’ calculations would inaccurately depict the current plant

community.

Page 49: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

48

The Ellenberg Community Coefficient Similarity Index (SIE) (Mueller-Dombois and

Ellenberg 1974) was used as a measure of plant community similarity among loading rates. This

calculation weights species presence and absence between two communities by IV, and is

summarized in the following equation:

where Mc is the sum of IV common to both loading rates, Ma is the sum of IV unique to loading

rate a, and Mb is the sum of IV unique to loading rate b.

Weighted Averages (WA) were calculated per species as the product of the sum of per

replicate IV and the indicator index of that species (Region 2, Reed 1988). Indicator values

ranged from 1 (OBL) to 5 (UPL), with intermediate indicators assigned in between (FACW+ =

1.67, FACW = 2, FACW- = 2.33, FAC+ = 2.67, FAC = 3, FAC- = 3.33, FACU+ = 3.67, FACU

= 4, FACU- = 4.33). These replicate values were then averaged over the growing season for

each loading rate. This calculation is summarized in the following equation:

where y1, y2, . . . , ym are the relative IV values for each species in a loading rate, and u1, u2, . . . , um

are the indicator values of each species. WA was calculated including the planted tree species.

Standing Crop Biomass

Aboveground standing crop measurements were obtained by clipping a sub-sample of all

living and standing dead plant material at the soil surface from two randomly placed 0.25 m2 (0.5

x 0.5 m) PVC quadrats in each plot. These samples were collected on August 22, 2005, as late

summer is thought to represent peak seasonal biomass in southeastern Virginia wetlands (Perry

SIE = (Mc/2)/(Ma + Mb + (Mc/2))

WA = (y1u1 + y2u2 + . . . + ymum) / Σ yiui m

i=1

Page 50: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

49

and Atkinson 1997, Perry and Hershner 1999, DeBerry and Perry 2004). Harvested material was

separated by species and dried at 40˚C until constant mass was achieved.

Woody Vegetation Development

Sampling of woody vegetation development characteristics was performed once on June

21, 2005 for all Betula nigra saplings within the 20 plots. Quercus palustris saplings were

omitted from these measurements due to high mortality. Morphometric characteristics measured

included total height, crown diameter, main stem diameter, and number of stems. Total height

was sampled using standard meter tape, while crown diameter and main stem diameter were

quantified using macro-calipers (Haglof, Inc. “Mantax Precision” Calipers) and micro-calipers

(SPI 6”/.1 mm Poly Dial Calipers), respectively. Crown diameter was measured at three

different angles at the visual diameter maximum and averaged to determine the final value.

Main stem diameter was measured once at the base of the main stem (trunk) or just above the

visual top of stem base swelling (hypertrophy) often accompanying trees growing in flooded

conditions (Cronk and Fennessey 2001).

Morphometrics were analyzed using Principal Components Analysis (PCA), which

enabled the data to be analyzed simultaneously, providing a visualization of data structure not

available using simple regression techniques. PCA was used to produce an index that

represented woody vegetation development, which could then be analyzed using statistical

techniques described in the following section. Development of Betula nigra saplings was

quantified using an index represented by the scores on the first Principal Component in the PCA.

Multivariate statistics were performed on MATLAB® software (The MathWorks 2005).

Page 51: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

50

Data Analysis

Data that did not meet conditions of normality were either log10 transformed or analyzed

using non-parametric statistics. Due to the gradient of organic matter loadings, simple regression

was the primary technique used to explore relationships between various parameters and loading

rate, plot elevation, and/or soil nutrients. In cases where data did not fit a linear/curvilinear

model, one-way Analysis of Variance (ANOVA) or the non-parametric Kruskal-Wallis test was

used to test for contrasts among treatments. Tukey’s family error rate was used for pair-wise

comparisons of elevation and tree development vs. loading rate, while the non-parametric

Wilcoxon Rank Sum Test was used for non-normal data including soil C and N. Tests for

normality, significance, and pair-wise comparisons were performed using Minitab®, release 14

(Minitab 2005), while various regressions were performed using SigmaPlot, v. 9.0 (Systat 2004).

Unless otherwise indicated, data are reported as means ± the Standard Error (SE).

RESULTS

Elevation

Linear regression revealed a positive relationship between loading rate and surface

elevation (p ≤ 0.001, R2 = 0.58) (Figure 3) with a maximum elevation increase of approximately

11 cm. Pair-wise comparison showed that plots in LRs 1-3 were not significantly different in

elevation, while LRs 4 and 5 were significantly higher than LRs 1-3, and LR 5 was higher than

4. Differences in plot elevation among loading rates were the result of incomplete incorporation

of the organic material as noted in Bergschneider (2005), particularly in the higher LRs (4 and 5)

Page 52: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

51

where conventional tillage (disking and roto-tilling) simply could not mix the large volumes

applied.

Soil Nutrients

Linear regression was positive between soil C (p ≤ 0.001, R2 = 0.60), N (p ≤ 0.001, R2 =

0.66), P (p ≤ 0.001, R2 = 0.66), and C:N (p ≤ 0.001, R2 = 0.49) and loading rate (Figure 4),

indicating that soil chemical parameters were affected by and increased with increasing organic

matter loadings. The C, N, P, and C:N relationship with loading rate appeared positive from 0 to

224 Mg ha-1. However, there was an obvious decrease or stasis in these properties from 224 to

336 Mg ha-1, likely related to the small sample size (n = 4) and disproportionate effect of one of

the LR 5 soil samples (plot 20). This data was not removed due to an inability to justify its

exclusion (i.e. no obvious reasons for differences between this replicate and others in LR 5).

Increasing C:N indicated that organic matter in the system may become more refractive with

increasing loading rate.

Plant Assemblage

Sixty-four (64) vascular plant species representing twenty-seven (27) families were

collected from the experimental block during the study (Table 2). Forty-five (45) were

perennial, fifteen (15) annual, six (6) perennial/annual, and one (1) biennial. Scirpus cyperinus

(FACW+) was the dominant or co-dominant species in each plot (IV range from 18.2 to 33.3),

with standing dead vegetation co-dominant (IV range from 13.8 to 19.5) (Table 3). Other co-

dominants included open unvegetated ground (bare ground) in LRs 1-4 (IV range from 7.8 to

12.8) and Eleocharis obtusa (OBL) in LRs 1 and 2 (IV range from 9.8 to 10.8). Among LRs 1-3,

Page 53: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

52

common subdominant species included Typha latifolia (IV range from 9.8 to 10.8), Juncus

acuminatus (IV range from 2.5 to 4.7), and Polygonum hydropiperoides (IV range from 1.9 to

4.2). Common subdominant species in LRs 4 and 5 included Juncus effusus (IV range from 4.6

to 7.6), Andropogon virginicus (IV range from 3.3 to 4.5), and Acalypha rhomboidea (IV range

from 2.0 to 3.8).

Diversity Measurements. Species richness (SR) values ranged from 7.8 ± 0.4 species m-2 in LR

2 to 5.3 ± 0.3 species m-2 in LR 5 (Table 4). A weak (R2 = 0.05) but significant (p ≤ 0.001)

decrease in SR with increasing loading rate was detected. Evenness (J’) values ranged from 0.92

± 0.01 in LR 2 to 0.86 ± 0.01 in LR 5 (Table 4), with linear regression suggesting a weak (R2 =

0.02) yet significant (p = 0.013) negative relationship with loading rate. Shannon index (H’)

values showed a high of 1.8 ± 0.1 in LR 2 and a low of 1.4 ± 0.1 in LR 5 (Table 4). Linear

regression suggested a weak (R2 = 0.05) yet significant (p ≤ 0.001) negative relationship between

H’ and loading rate.

Community Similarity. The Ellenberg Community Coefficient Similarity Indices (SIE) were high

(> 0.5) among all loading rate comparisons (Table 5). These indices, combined with the similar

ranges and weak correlations of SR, J’, and H’ vs. loading rate, suggested that species

composition of the plant communities of all loading rates were very similar to each other.

Weighted Averages (WA). The WA values ranged from 1.6 ± 0.03 in LR 2 to 2.1 ± 0.1 in LR 5

(Table 4) and showed a positive linear relationship with both loading rate (p ≤ 0.001, R2 = 0.27)

and surface elevation (p ≤ 0.001, R2 = 0.35) (Figure 5). The range in WA was narrow, and

generally contained a majority of wetland (OBL through FACW) plant species in all plots

(Figure 6). However, there was also an apparent increase in the abundance (i.e. IV) of upland

Page 54: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

53

species (FAC through UPL) versus wetland species (FAC+ through OBL), as evidenced by the

change in the ratio of wetland vs. upland species abundance from LR 1 (4.9) to LR 5 (1.2).

Standing Crop Biomass

Peak season biomass estimates ranged from 580 ± 97 in LR 2 to 790 ± 60 in LR 5 (Table

4), but showed neither a significant correlation with loading rate (p = 0.283) nor significant

differences among loading rates (Kruskal-Wallis Test, p = 0.205). However, a quadratic

regression model revealed a significant relationship (p = 0.007, R2 = 0.23) between biomass and

surface elevation (Figure 7a). These results indicate that peak biomass occurs at an optimum

plot elevation, with values decreasing at both higher and lower elevations. However, linear

regressions were not significant between biomass and either soil N (p = 0.547), P (p = 0.539), or

C:N (p = 0.845).

Closer inspection of biomass values showed that plot 14 (LR 1) biomass values (1360

and 1444 g m-2) were both much higher than the rest of the LR 1 replicates (120-560 g m-2).

Also, plot 17 (LR 4) replicate samples were very different from one another, with one sample

showing very high biomass (1912 g m-2) and the other showing very low values (608 g m-2)

Because the biomass values for these plots were the highest seen in all of the experimental block,

and are higher than what is usually reported in these systems, these data points were removed

and the analysis was rerun without these outliers. Under these conditions, biomass estimates

showed a significant linear relationship with both loading rate (p = 0.020, R2 = 0.15) and plot

elevation (p = 0.030, R2 = 0.13) (Figure 7b). Even with outliers removed, linear regressions

were not significant between biomass and either soil N (p = 0.203), P (p = 0.092), or C:N (p =

0.462). However, there were no known factors regarding the site preparation or ambient

Page 55: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

54

environmental characteristics of plots 14 and 17 that would explain such unusual biomass values.

Therefore, the discussion of standing crop biomass in the following sections will be based on

data including the outliers, with a supplemental section that discusses the modified data set.

Woody Vegetation Development

Principal Components Analysis revealed similar loadings on the first principal

component for total height, crown diameter, and main stem diameter (Figure 8). Due to the

similar magnitude of these loadings, as well as the common metric of size (length or width)

between these morphometrics, the scores on the first principal component were used as an index

of tree “size”. Pair-wise comparisons showed significant increases in tree size between LRs 1-3

and 4-5 (Figure 9). Linear regression showed a significant positive relationship of tree size

versus loading rate (p ≤ 0.001, R2 = 0.21) (Figure 9), surface elevation (p ≤ 0.001, R2 = 0.28)

(Figure 10), and soil P (p ≤ 0.001, R2 = 0.15) (Figure 11).

DISCUSSION

Loading Rate Effects on Elevation

Since elevation varied positively with loading rate (Figure 3), it became an important

environmental gradient to consider in our study. Differences in elevation among loading rates

reflected the incomplete incorporation of the organic amendment material, particularly in the

higher-level amendments (LRs 4 and 5), into each plot during site preparation (W.L. Daniels

pers. com.). This elevation effect was noted in an earlier study of the same experimental block

Page 56: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

55

by Bergschneider (2005), suggesting that organic matter amendments in LRs 4 and 5 have not

settled to a significant degree between studies.

Elevation can be used as a proxy for a soil moisture/water saturation gradient, driven by

inundation frequency and duration. This assertion is justified by 1) the relatively small aerial

extent (680 m2) of the experimental block, 2) the dominance of precipitation-driven hydrology,

and 3) the determination of CCW as a groundwater recharge site (Despres 2004). These three

factors lead us to assume that the CCW experimental block as a whole experienced a uniform

hydrologic input, and, as such, generally experienced hydrologic inundation more frequently and

for longer duration (i.e. hydroperiod) in lower elevation plots than in higher elevation plots.

Changes in physical and chemical conditions along hydrologic gradients are well known, and can

affect soil O2 (oxygen) availability (Gambrell and W. H. Patrick 1978, Nedwell 1984, Craft

2001), soil redox potential (EH) (Ponnamperuma 1972, Vepraskas and Faulkner 2001), and

nutrient availability (Bayley et al. 1985, Richardson 1985, Bedford et al. 1999). The potential

importance of elevation, and therefore hydrological differences among plots and loading rates, in

this study are discussed further within the following “Plant Assemblage”, “Standing Crop

Biomass”, and “Woody Vegetation Development” sections.

Soil Nutrients

Changes in soil chemical concentrations were expected to be a major driver of the

potential differences in the measured parameters among loading rates, especially in relation to

standing crop biomass and woody vegetation development. As expected, soil C content, as well

as N and P (Figure 4), generally increased with increasing loading rate. Increases in these

parameters logically reflected the addition of increasing volumes of coarse organic matter

Page 57: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

56

amendments, which also contained increasing volumes of N and P in their organic forms

(Bergschneider 2005).

The range of C (i.e. organic matter) values found in this study (1.7-17%) was similar to

those reported for surface horizons of southern forested wetlands on mineral soils (2.8-18%)

(Lockaby and Walbridge 1998). However, the average C value of LR 1 (1.7%) was below the

soil surface organic matter concentration of 2%, reported by Baker and Broadfoot (1979) as

indicative of nutritional limitation for many deciduous floodplain species. N levels ranged from

0.2-0.7% for LRs 2-5 in our study, values encompassing and extending higher than those

reported for reference wetlands in the Virginia Coastal Plain (0.2-0.4%) (Stolt et al. 2000).

However, N content of LR 1 (0.1%) was lower than these natural wetland values, and was

consistent with levels in Virginia unamended created wetlands (0.03-0.2%) (Stolt et al. 2000).

Available P levels measured in this study (2.8-18 mg kg-1) were within ranges reported from

numerous southeastern US floodplain forest soils (2-18 mg kg-1) (Lockaby and Walbridge 1998).

However, LRs 1 (2.8 mg kg-1) and 2 (6.3 mg kg-1) were below the “crude deficiency level” for

available soil P (7 mg kg-1) reported in Lockaby and Walbridge (1998) for hardwood floodplain

species.

The low levels of N and P in LR 1 confirm the general nutrient deficiency reported for

other created wetlands (Bischel-Machung et al. 1996, Whittecar and Daniels 1999, Stolt et al.

2000). Considering the increases in nutrient levels with increasing loading rates, our results

reinforce the importance of incorporating organic matter into created wetland soils in terms of

establishing a nutrient environment more similar to that of natural or reference wetland systems.

However, it is important to note that addressing potentially complicating factors such as

Page 58: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

57

hydrological changes (i.e. elevation) at higher loading rates should be of prominent concern

when designing such experiments. This is discussed further in the following section.

Plant Assemblage

Vegetation composition in the CCW experimental block (Table 2) was similar to that of

other created wetlands in Virginia (Atkinson et al. 1993, Stolt et al. 2000, DeBerry and Perry

2004, Atkinson et al. 2005). We had hypothesized that vegetation composition would vary

noticeably among different loadings of organic matter; however, we did not detect such

differences. The overall similarity in dominant vegetation among all loading rates was likely

attributable to the prevalent communities of Scirpus cyperinus bordering the experimental block

(DeBerry 2006). However, the increase in dominance of S. cyperinus and decrease of Eleocharis

obtusa with increasing loading rate suggested a potential successional trend with organic

amendment amount.

Scirpus cyperinus, an abundant native in many wetland types ranging from eastern

portions of Canada south to Texas and Louisiana (Godfrey and Wooten 1979), dominated the

CCW experimental block as a whole, and generally increased in dominance with increasing

loading rate (Table 3). The overall dominance of this species was consistent with several studies

of wetlands in Virginia (Jones et al. 1993, Yu et al. 1998, Atkinson and Cairns 2001, Atkinson et

al. 2005). This species is a biennial, and can produce hundreds of thousands of small (≤ 1 mm),

lightweight (1 x 10-5 g) seeds (Shipley and Parent 1991), which can be transported via wind,

water, or animals to suitable habitat patches (Larson 1999). In addition, it can tolerate a wide

range of environmental conditions (Kadlec 1958, 1961, Wilcox et al. 1985, Larson 1999). For

example, S. cyperinus is known to colonize dry sites in wetlands, such as during seasonal or

Page 59: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

58

managed draw downs, and also has the ability to tolerate prolonged inundation (Kadlec 1958,

1961). This pattern was consistent with the S. cyperinus dominance at CCW during the 2005

study; the experimental plots were inundated in all but the highest loading rates throughout the

winter and early spring 2005 (December-April), dried out completely at the surface in early

summer (June), and remained dry until late fall (November).

Eleocharis obtusa is a common, cespitose, obligate wetland annual that occurs

throughout the eastern U.S. in “muddy places” (Strausbaugh and Core 1977). This species co-

dominated in LRs 1 and 2, but decreased in abundance in LRs 3-5 (Table 2). Although little

ecological information is available for E. obtusa, it is possible that relatively large areas (high

IV) of bare ground (Table 3) could have encouraged its establishment as a dominant in LRs 1

and 2. The amount of bare ground decreased with increasing loading rate (IV range from 12.8 to

7.8). This result is consistent with Stauffer and Brooks (1997), who reported higher vegetative

coverage in organically amended plots than in unamended plots.

Larger areas of bare ground could potentially lead to higher and more diurnally

fluctuating soil temperatures in the upper profile (Stolt et al. 2000) due to the reduced amount of

plant shading in these areas (Aust and Lea 1991). Higher soil temperatures generally encourage

higher microbial activity (Zak et al. 1999), leading to faster organic matter decomposition rates

(Reth et al. 2005) and perhaps a more stressful plant environment in terms of rooting stress due

to higher respiration rates (Londo et al. 1999). Such temperature responses could potentially

initiate a positive feedback loop in low organic matter sites: open ground leads to higher soil

temperatures, increasing organic matter respiration rates, leading to decreased soil organic

matter, causing lower plant production, thereby leading to low plant coverage and more bare

ground. However, evaluation of several growing seasons would be required to test this

Page 60: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

59

hypothesis. As such, evidence from this study suggests that E. obtusa opportunistically

colonizes unutilized habitats, as represented in this study by bare ground.

DeBerry and Perry (2004), in a similar study of an unamended portion of the CCW site

and an adjacent reference wetland, found that S. cyperinus was a dominant plant in an adjacent

emergent reference wetland (IV = 29.6), while E. obtusa was dominant in the created wetland

(IV = 33.4). Therefore, the results of this study suggested that plots with higher loadings of

incorporated organic matter (i.e. LRs 3-5) more closely resembled adjacent reference wetlands,

as compared to plots with little or no organic matter amendments. In terms of similar IV, LR 3

most closely resembled DeBerry and Perry’s (2004) reference wetland, especially when

considering S. cyperinus [26.89 (this study), 26.7 (DeBerry and Perry 2004)] and Juncus effusus

[6.12 (this study), 5.1 (DeBerry and Perry 2004)].

Standing dead vegetation was prevalent in the experimental block, and visual observation

showed that much of it was senesced S. cyperinus material. Thus, the co-dominance of standing

dead vegetation in each loading rate (Table 3), and its positive relationship with loading rate,

most likely reflected the dominant production of S. cyperinus in previous growing seasons as

well as its relatively low decomposition rate (Kittle et al. 1995). The large volume of standing

dead vegetation in all loading rates suggested an abundance of refractive carbon within this

created wetland that could serve as a future autogenic soil carbon source. However, the portion

of this material that will be incorporated into the soil as organic matter is unknown. Further,

plant root material, not measured in this study, can be the major source of organic matter to the

soil in some ecosystems (Megonigal and Day 1988, Collins and Kuehl 2001).

We hypothesized that plant community diversity would vary with organic matter loading

rate, and to some extent this occurred; weak but significant negative relationships were seen

Page 61: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

60

among species richness (SR), evenness (J’), and diversity (Shannon Index, H’) (Table 4). These

trends agree with Anderson and Cowell (2004), who reported significantly lower SR and H’

values in mulched wetlands relative to non-mulched wetlands in Florida. Our results suggest a

potential relationship between nutrient availability to a few of cosmopolitan species and substrate

availability to several specialized niche species (sensu Moore et al. 1989). However, the narrow

ranges of diversity values may still indicate a similarity of diversity over all loading rates, as was

corroborated by direct comparisons of plant community similarity (SIE) (Table 5). This high

degree of similarity may have been an artifact of the overall dominance of S. cyperinus and co-

dominance of standing dead vegetation. Whether S. cyperinus is a species that facilitates plant

diversity (nurse species) or suppresses plant diversity (invasive species) over time warrants

further study.

Weighted Average (WA) values likely reflected changes in elevation among plots rather

than differences in loading rate. As would be expected, WA increased significantly with

increasing elevation (Figure 5), suggesting a negative relationship between the relative

abundance of hydrophytic species and elevation. This relationship was seen (Figure 6), as the

cumulative IV of upland species (FAC through UPL) increased with increasing LR (and

elevation), while the cumulative IV of wetland species (FAC+ through OBL) stayed relatively

constant. Indeed, many studies have reported vegetation composition changes along hydrologic

gradients (Beatty 1984, Messina and Conner 1998, Wall and Darwin 1999, Waddington et al.

2001, Burke et al. 2003, Nicol et al. 2003, Fraser and Karnezis 2005). More specifically, Vivian-

smith (1997) found that microtopographical differences, even as minor as 1-3 cm, produced

significant plant community structure differences in experimental wetland communities.

Therefore, the 11 cm difference in the average elevations among plots in loading rates 1 (10.36

Page 62: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

61

m) and 5 (10.47 m) could potentially have acted as a selective pressure on plant species

colonization and survival. Even so, our data indicated that all plots supported a majority of

hydrophytic plants (FAC+ through OBL) (55-83%). Thus, loadings of organic matter do not

solely determine or encourage the establishment of hydrophytes in created wetland systems, and

other factors such as elevation (i.e. hydroperiod/oxygen availability) and seed dispersal may be

more important determinants.

Standing Crop Biomass

Biomass ranges (including outliers) for this study (580-790 g m-2) fell within those for

natural inland freshwater marshes (500-5500 g m-2) (Mitsch and Gosselink 2000), were on the

low side of the range reported for several studies of created wetlands in central Pennsylvania

(520-1697 g m-2) (Cole et al. 2001), and were generally higher than those reported in an earlier

study from our created wetland site (146-896 g m-2) (DeBerry and Perry 2004) and mulched

wetland creation areas in west-central Florida (349 g m-2) (Anderson and Cowell 2004).

We hypothesized that aboveground standing crop biomass would vary positively with

organic matter loadings; however, this was not the case. Our results were consistent with others

(Cole et al. 2001, Anderson and Cowell 2004) who did not detect total biomass differences in

mulched and non mulched wetlands. In a 2003 study of the same experimental plots,

Bergschneider (2005) also failed to detect loading rate-based differences in biomass. These

results suggest that aboveground plant biomass may not be directly dependant on soil organic

matter, at least when considering early successional communities.

Elevation was a better predictor of biomass than loading rate (Figure 7), and the

relationship between the two suggested that at some elevation (~10.45 m) the predominant

Page 63: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

62

environmental conditions (i.e. nutrient availability, plant available water, redox potential, etc.)

were optimal for plant (aboveground) primary production. Plant biomass and growth differences

along hydrologic gradients have been well documented (Mitsch and Ewel 1979, Brinson et al.

1981, Craft 2001, Craft et al. 2002, Fraser and Karnezis 2005). Bayley et al. (1985) inferred that

sequences of drought and flooding provided a substantial nutrient source to vegetation that was

not available in continuously flooded marshes, leading to higher plant production. This finding

potentially explains results from the current study, in that plots with elevations corresponding

with the highest biomass values were such that occasional flooding during the growing season

(presumably by summer rains) may have made limiting nutrients (especially P) temporarily

available for vegetation, and that subsequent drying may have allowed for optimal aerobic, yet

moist, growing conditions.

Removing outlying points from the biomass dataset suggested linear relationships with

both loading rate and elevation. Because of the significance of each curve, and the similar

amount of data explained by each curve (i.e. R2 = 0.14 and 0.13), it is difficult to derive which

factor or suite of factors affected biomass production more. More data, perhaps over several

growing seasons, may be needed to explain why the outlying points were so much higher than

the rest of the dataset. It is possible that the linear relationships seen were artifacts of the

modified data, and, therefore, do not warrant further discussion.

We also anticipated that plant biomass would vary positively with soil nutrients,

particularly N and P, as macronutrient availability represents a potential constraint on vegetative

productivity (Lockaby and Walbridge 1998). However, we did not detect a correlation between

biomass and soil nutrients (N, P, C:N) in either the original or modified (outlier-removed)

dataset. It is possible that plant available nutrients were utilized by vegetation before the

Page 64: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

63

sampling period (August 2005). Also, levels of N and P in the soil (representing both plant

available and unavailable chemical species) may not have been indicative of the plant available

nutrient resource pool, perhaps due to immobilization within coarse-organic amendments (e.g.

ground-woody debris larger than a few centimeters in diameter). Further, perhaps the amount of

nutrients provided by the organic matter additions, especially the higher LRs (4 and 5), was in

excess of concentrations required by existing vegetation, and that some other physical or

chemical factor or suite of factors was limiting to growth. It is also possible that available soil

nutrients were allocated to belowground biomass by the plot vegetation (Day and Megonigal

1993), and therefore were not reflected in our biomass analysis.

Although the incorporation of organic matter into created wetland soils certainly has the

potential to improve the plant rooting environment, either through lowering soil bulk density,

muting soil temperature fluctuations, or providing a nutrient source, this early successional

created wetland appears to have responded more directly to soil surface elevation and

presumably saturation and inundation. Further investigation should consider sampling

belowground biomass production, analyzing soils for plant available macronutrient compounds

such as nitrate (NO3-), ammonium (NH4

+), and phosphate (PO43+), and perhaps analyzing plant

tissue nutrient ratios to determine limiting nutrient status in these systems.

Woody Vegetation Development

As with biomass, we expected tree sizes to vary positively with organic matter

amendment loadings. Although loading rate was a significant predictor (R2 = 0.21), our results

suggested that elevation differences among plots (R2 = 0.28) better explained the differences in

tree size. This finding was corroborated by the pair-wise analysis of tree size vs. loading rate,

Page 65: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

64

which showed a clear and significant separation between LRs 1-3 and 4-5 [an elevation

difference of 8 cm (10.37- 10.45 m)] (Table 1), but similar tree sizes within the two groups.

Bergschneider (2005) found a similar increase in tree size between LRs 2 and 3. Betula nigra is

a flood-tolerant species, and can withstand soil inundation for one to three months during the

growing season (Norby and Kozlowski 1983). However, Norby and Kozlowski (1983) showed

that dry weights of roots, stems, and leaves of flooded B. nigra individuals were reduced to 24,

76, and 73% of those of unflooded individuals after 5 weeks. Further, woody wetland species

are known to show slower growth rates in flooded bottomland hardwood forests vs. similar

wetlands with a lower hydroperiod (Malecki et al. 1983, Megonigal et al. 1997).

The relationship of tree size with elevation (Figure 10) was linear, and differed from that

of herbaceous plant biomass, which showed a quadratic relationship with elevation. The

difference in elevation relationship between the two structural forms may be explained by the

differences in rooting depth between the herbaceous and woody species. For example, in LR 5,

the set of plots with the highest average elevation (10.47 m), B. nigra individuals may be able to

send roots through the unincorporated organic amendment layer into the mineral portion of the

soil profile where available P presumably exists (Richardson 1985, Craft 2001), perhaps in

higher concentrations resulting from nutrient (P) leaching from decomposition of the highly

loaded organic material. Therefore, tree saplings in the higher loading rates may have been able

to exploit nutrient (P) sources in both the organic amendment layer and the mineral soil that may

have been unavailable to the more shallow-rooted herbaceous vegetation. This explanation is

plausible as tree size did show a significant positive relationship with P (Figure 11, p ≤ 0.001, R2

= 0.15). These findings suggest that early sapling development does not respond to soil organic

matter alone. Instead, it appears as though the combination of elevation-related hydrology and

Page 66: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

65

organic amendment-related nutrient gradients may better explain the linear tree size relationship

seen in this study. As succession proceeds within the experimental plots, and as the organic

matter amendments in the higher loading rates continue to settle due to physical breakdown and

microbially-mediated OM oxidation, hydrology could become less of a differentiating factor

among the loading rates. It is possible, then, that B. nigra sizes would more closely track organic

matter loading rate given more time for such processes to occur. Future studies would be

required to test this hypothesis.

Plant Production and Soil Phosphorus

The removal of the modified (i.e. outlier-removed) biomass data, and a closer inspection

of the tree size data, revealed a potential relationship between aboveground plant production of

both herbaceous and woody species and soil P, up to a saturation point. As shown in Figure 12,

herbaceous biomass appears to increase linearly with soil P up to a concentration of 10 mg kg-1

(p = 0.017, R2 = 0.25), after which biomass was variable but relatively constant. Further, tree

size linearly increased with soil P (p ≤ 0.001, R2 = 0.30), up to a concentration of 13-15 mg kg-1

(Figure 13), after which tree size was fairly constant. The Virginia Department of Conservation

and Recreation (2005) lists soil P concentrations of 11-15 mg kg-1 as indicative of “medium

fertility” for cultivated crops such as corn. However, it is thought that native plant species are

actually more efficient at P uptake and can tolerate much lower available P levels than crops

(W.L. Daniels, pers. com.). Thus, it is possible that the increases in aboveground production up

to an intermediate soil P concentration seen in the study reflects P limitation in the lower loading

rates, with increasing availability until plants could no longer take advantage of the superfluous

resource (i.e. saturation). This saturation effect could potentially explain the weak relationship

Page 67: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

66

between biomass and tree size with loading rate. That is, as increasing loadings of the

amendment material increased soil P, these factors did not respond linearly in the higher loading

rates. The potential relationship between plant production and soil P could use further study, in

part to determine a P fertility threshold for native wetland vegetation.

RECOMMENDATIONS

The lack of correlation between biomass and organic matter loading rate indicates that

aboveground plant biomass should not be used as a functional indicator of created wetland

success. This is corroborated by other studies that reported either no relationship between soil

organic matter and biomass (Cole et al. 2001, Anderson and Cowell 2004) or that created

wetland aboveground biomass equivalency with adjacent natural wetlands can be achieved even

in the early stages of plant development (Whigham et al. 2002, DeBerry and Perry 2004).

Further, adding high loadings of organic matter into created wetlands (i.e. LRs 4 and 5)

could inadvertently raise soil surface elevations and decrease the frequency of

inundation/anaerobic conditions. This is illustrated by the positive relationship of Weighted

Averages (WA) and loading rates. For example, Wentworth et al. (1988) suggested and

Atkinson et al. (1993) confirmed that WA ≤ 2.0 was an acceptable value on which to designate

wetlands on the basis of vegetation alone. Using this criterion, LRs 1-3 would be considered

“wetlands” and LR 4 and 5 would not (Table 4). As such, elevation changes due to organic

matter mounding in these higher loading rates could have an effect on regulatory wetland status

within the CCW experimental block. Therefore, we recommend that organic matter loadings

into created wetland soils be applied such that the material is fully incorporated into the soil

Page 68: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

67

profile, and that the loading rate used be moderate enough to not significantly alter soil surface

elevation.

Given that standing crop biomass did not vary with loading rate, and that the tree size-

loading rate relationship may be better explained by plot elevation, a recommendation for an

organic matter loading in similar created wetland systems should instead be based on soil

nutrient values, including the potential aboveground production-soil P relationship, and

vegetation composition. Therefore, an organic matter soil amendment similar to LR 3 (112 Mg

ha-1) is appropriate, as average nutrient levels (C = 6.7%, N = 0.3%, P = 8.3 mg kg-1) were within

range of natural systems and vegetation composition was most similar in this loading rate to that

of adjacent early-successional reference wetlands (DeBerry and Perry 2004). This organic

matter loading seemed to provide a “jumpstart” for the created plots, while also minimizing the

change in elevation (+2 cm from LR 1) due to the added bulk material.

Page 69: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

68

Literature Cited

Anderson, C. J. and B. C. Cowell. 2004. Mulching effects on the seasonally flooded zone of west-central Florida, USA wetlands. Wetlands 24:811-819.

Atkinson, R. B. and J. Cairns, Jr. 2001. Plant decomposition and litter accumulation in depressional wetlands: functional performance of two wetland age classes that were created via excavation. Wetlands 21:354-362.

Atkinson, R. B., J. E. Perry, and J. J. Cairns. 2005. Vegetation communities of 20-year-old created depressional wetlands. Wetlands Ecology and Management 13:469-478.

Atkinson, R. B., J. E. Perry, E. Smith, and J. C. Jr. 1993. Use of created wetland delineation and weighted averages as a component of assessment. Wetlands 13:185-193.

Aust, W. M. and R. Lea. 1991. Soil temperature and organic matter in a disturbed forested wetland. Soil Science Society of America Journal 55:1741-1746.

Baker, J. B. and W. M. Broadfoot. 1979. Site evaluation for commercially important southern hardwoods. USDA Forest Service, Southern Forest Experiment Station, General Technical Report SO-26, New Orleans, LA.

Balcombe, C. K., J. T. Anderson, R. H. Fortney, J. S. Rentch, W. N. Grafton, and W. S. Kordek. 2005. A comparison of plant communities in mitigation and reference wetlands in the mid-Appalachians. Wetlands 25:130-142.

Bayley, S. E., J. J. Zoltek, A. J. Hermann, T. J. Dolan, and L. Tortora. 1985. Experimental manipulation of nutrients and water in a freshwater marsh: effects on biomass, decomposition, and nutrient accumulation. Limnology and Oceanography 30:500-512.

Beatty, S. W. 1984. Influence of microtopography and canopy species on spatial patterns of forest understory plants. Ecology 65:1406-1419.

Bedford, B. L., M. R. Walbridge, and A. Aldous. 1999. Patterns in nutrient availability and plant diversity of temperate North America wetlands. Ecology 80:2151-2169.

Bergschneider, C. R. 2005. Determining an appropriate organic matter loading rate for a created coastal plain forested wetland. M.S. Thesis. Virginia Tech, Blacksburg, VA, USA.

Bischel-Machung, L., P. P. Brooks, S. S. Yates, and K. L. Hoover. 1996. Soil properties of reference wetlands and wetland creation projects in Pennsylvania. Wetlands 16:532-541.

Brinson, M. M., A. E. Lugo, and S. Brown. 1981. Primary productivity, decomposition and consumer activity in freshwater wetlands. Annual Review of Ecology and Systematics 12:123-161.

Page 70: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

69

Brinson, M. M. and R. Rheinhardt. 1996. The role of reference wetlands in functional assessment and mitigation. Ecological Applications 6:69-76.

Brooks, R. P., D. H. Wardrop, C. H. Cole, and D. A. Cambell. 2005. Are we purveyors of wetland homogeneity?: A model of degradation and restoration to improve wetland mitigation performance. Ecological Engineering 24:331-340.

Brown, M.T. 1991. Evaluating created wetlands through comparisons with natural wetlands. U.S. Environmental Protection Agency, Environmental Research Laboratory, Corvallis, OR, USA. EPA/600/3-91/058.

Bruland, G. L. and C. J. Richardson. 2004. Hydrologic gradients and topsoil additions affect soil properties of Virginia created wetlands. Soil Science Society of America Journal 68:2069-2077.

Burke, M. K., S. L. King, D. Gartner, and M. H. Eisenbies. 2003. Vegetation, soil, and flooding relationships in a blackwater floodplain forest. Wetlands 23:988-1002.

Cole, A. C., R. P. Brooks, and D. H. Wardrop. 2001. Assessing the relationship between biomass and soil organic matter in created wetlands of central Pennsylvania, USA. Ecological Engineering 17:423-428.

Cole, C. A. and R. P. Brooks. 2000. A comparison of the hydrologic characteristics of natural and created mainstem floodplain wetlands in Pennsylvania. Ecological Engineering 14:221-231.

Collins, M. E. and R. J. Kuehl. 2001. Organic matter accumulation in organic soils. p. 137-162 in J. L. Richardson and M. J. Vepraskas (ed.) Wetland Soils: Genesis, Hydrology, Landscapes, and Classification. Lewis Publishers.

Cowardin, L. M., V. Carter, F. C. Golet, and E. T. Laroe. 1979. Classification of wetlands and deepwater habitats. U.S. Department of the Interior, Fish and Wildlife Service, FWS/OBS-79/31, Washington, D.C.

Craft, C., S. Broome, and C. Campbell. 2002. Fifteen years of vegetation and soil development after brackish-water marsh creation. Restoration Ecology 10:248-258.

Craft, C. B. 2001. Biology of wetland soils. p. 107-135 in J. L. Richardson and M. J. Vepraskas (ed.) Wetland Soils: Genesis, Hydrology, Landscapes, and Classification. Lewis Publishers.

Cronk, J. K. and M. S. Fennessey. 2001. Wetland plants: biology and ecology. Lewis Publishers, Boca Raton, FL.

Daubenmire, R. F. 1959. Canopy coverage method of vegetation analysis. Northwest Science 33:43-64.

Page 71: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

70

Day, F. P., Jr and J. P. Megonigal. 1993. The relationship between variable hydroperiod, production allocation, and belowground organic turnover in forested wetlands. Wetlands 13:115-124.

DeBerry, D. A. 2006. Floristic Quality Index: ecological and management implications in created and natural wetlands. Ph.D. Dissertation. Virginia Institute of Marine Science, College of William and Mary, Gloucester Point, VA, USA.

DeBerry, D. A. and J. E. Perry. 2004. Primary succession in a created freshwater wetland. Castanea 69:185-193.

Department of Conservation and Recreation. 2005. Virginia nutrient management standards and criteria. Commonwealth of Virginia, Division of Soil and Water Conservation, Richmond, VA, USA.

Despres, A. D. 2004. Hydrologic variations within created and natural wetlands in southeastern Virginia. M.S. Thesis. Old Dominion University, Norfolk, VA, USA.

Donohue, S. J. and S. E. Heckendorn. 1996. Laboratory proceedures: Virginia Tech Soil Testing and Plant Analysis Laboratory. Virginia Coop. Ext. Bull.: 452-881.

Fraser, L. H. and J. P. Karnezis. 2005. A comparitive assessment of seedling survival and biomass accumulation for fourteen wetland plant species grown under minor water-depth differences. Wetlands 25:520-530.

Gambrell, R. P. and J. W. H. Patrick. 1978. Chemical and microbiological properties of anaerobic soils and sediments. p. 375-473. in D. D. Hook and R. M. M. Crawford (ed.) Plant Life in Anaerobic Environments. Ann Arbor Science, Ann Arbor, MI.

Gleason, H. A. and A. Cronquist. 1991. Manual of Vascular Plants of Northeastern United States and Adjacent Canada, 2nd edition. New York Botanical Garden, Bronx, New York.

Godfrey, R. and J. Wooten. 1979. Aquatic and Wetland Plants of Southeastern United States: Monocotyledons. University of Georgia Press, Athens, GA.

Jones, D. H., R. B. Atkinson, and J. Cairns, Jr 1993. The distribution of wetland macrophytes in accidental wetlands on surface mined land in southwest Virginia. Virginia Journal of Science 44:148.

Kadlec, J. A. 1958. An analysis of a woolgrass (Scirpus cyperinus) community in Wisconsin. Ecology 39:327-332.

Kadlec, J. A. 1961. A further comment on the ecology of woolgrass (Scirpus Cyperinus). Ecology 42:591-592.

Kittle, D. L., J. B. McGraw, and K. Garbutt. 1995. Plant litter decomposition in wetlands receiving acid mine drainage. Journal of Environmental Quality 24:301-306.

Page 72: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

71

Larson, J. L. 1999. Woolgrass: a plant profile. Ecological Restoration 17:210-216.

Lockaby, B. G. and M. R. Walbridge. 1998. Biogeochemistry. p. 149-172 in M. G. Messina and W. H. Conner (ed.) Southern Forested Wetlands: Ecology and Management. Lewis Publishers/CRC Press, Boca Raton, FL.

Londo, A. J., M. G. Messina, and S. H. Schoenholtz. 1999. Forest harvesting effects on soil temperature, moisture, and respiration in a bottomland hardwood forest. Soil Sci Soc Am J 63:637-644.

Malecki, R. A., J. R. Lassoie, E. Rieger, and T. Seamans. 1983. Effects of long-term artificial flooding on a northern bottomland hardwood forest community Forest Science 29:535-544.

McKinstry, M. C. and S. H. Anderson. 2003. Improving aquatic plant growth using propagules and topsoil in created bentonite wetlands of Wyoming. Ecological Engineering 21:175-189.

Megonigal, J. P., W. H. Connor, S. Kroeger, and R. R. Sharitz. 1997. Aboveground production in southeastern floodplain forests: a test of the subsidy-stress hypothesis. Ecology 78:370-384.

Megonigal, J. P. and F. P. Day, Jr. 1988. Organic matter dynamics in four seasonally flooded forest communities of the dismal swamp. American Journal of Botany 75:1334-1343.

Messina, M. G. and W. H. Conner. 1998. Southern Forested Wetlands: Ecology and Management. Lewis Publishers/CRC Press, Boca Raton, FL.

Minitab. 2005. Minitab® statistical software, release 14. State College, PA.

Mitsch, W. J. and K. C. Ewel. 1979. Comparative biomass and growth of cypress in Florida wetlands American Midland Naturalist 101:417-426.

Mitsch, W. J. and J. G. Gosselink. 2000. Wetlands, third edition. John Wiley and Sons, Inc., New York, NY, USA.

Moore, D. R. J., P. A. Keddy, C. L. Gaudet, and I. C. Wisheu. 1989. Conservation of wetlands: do infertile wetlands deserve a higher priority? Biological Conservation 47:203-217.

Mueller-Dombois, D. and H. Ellenberg. 1974. Aims and Methods of Vegetation Ecology. The Blackburn Press, Caldwell, NJ.

Natural Resources Conservation Service. 2006. Plants Database. United States Department of Agriculture, <http://plants.usda.gov/>.

Nedwell, D. B. 1984. The input and mineralization of organic carbon in anaerobic aquatic sediments. Advances in Microbial Ecology 7:93-131.

Page 73: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

72

Nelson, D. W. and L. E. Sommers. 1996. Total carbon, organic carbon, and organic matter. in D. L. Sparks (ed.). Methods of soil analysis, part 3. Chemical Methods. SSSA Book Series no. 5. SSSA-ASA, Madison, WI.

Nicol, J. M., G. G. Ganf, and G. A. Pelton. 2003. Seed banks of a southern Australian wetland: the influence of water regime on the final floristic composition. Plant Ecology 168:191-205.

Noon, K. F. 1996. A model of created wetland primary succession. Landscape and Urban Planning 34:97-123.

Norby, R. J. and T. J. Kozlowski. 1983. Flooding and SO2 stress interaction in Betula papyrifera and B. nigra seedlings. Forest Science 29:739-750.

Odum, E. P. 1969. The strategy of ecosystem development. Science 164:262-270.

Odum, E. P. 1978. The value of wetlands: a hierarchical approach. in P. E. J. R. C. a. J. E. C. Gresson (ed.) Wetland Functions and Values: the State of Our Understanding. American Water Resources Association.

Perry, J. E. and R. B. Atkinson. 1997. Plant diversity along a salinity gradient of four marshes on the York and Pamunkey Rivers in Virginia. Castanea 62:112-118.

Perry, J. E. and C. H. Hershner. 1999. Temporal changes in the vegetation pattern in a tidal freshwater marsh. Wetlands 19:90-99.

Ponnamperuma, F. N. 1972. The chemistry of submerged soils. p. 29-96 in N. C. Brady, (ed.) Advances in Agronomy. Academic Press, New York.

Reed, P.B., Jr. 1988. National List of Plant species that Occur in Wetlands: 1988, Virginia NERC-88/18.46. U.S. Fish and Wildlife Service, St. Petersburg, Florida. 115 p.

Reth, S., M. Reichstein, and E. Falge. 2005. The effect of soil water content, soil temperature, soil pH-value and the root mass on soil CO2 efflux - A modified model. Plant and Soil 268:21-33.

Richardson, C. J. 1985. Mechanisms controlling phosphorus retention capacity in freshwater wetlands. Science 228:1424-1427.

Schmidt, J. M. 2002. Litter Decomposition in Created and Adjacent Forested Wetlands of the Coastal Plain of Virginia. M.S. Thesis. Virginia Tech, Blacksburg, VA, USA.

Shipley, B. and M. Parent. 1991. Germination responses of 64 wetland species in relation to seed size, minimum time to reproduction and seedling relative growth rate. Functional Ecology 5:111-118.

Snell-Rood, E. C. and D. A. Cristol. 2003. Avian communities of natural and created wetlands: Bottomland forests in Virginia. Condor 105:303-315.

Page 74: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

73

Snodgrass, J. W., M. J. Komoroski, A. L. Bryan, Jr., and J. Burger. 2000. Relationships among isolated wetland size, hydroperiod, and amphibian species richness: Implications for wetland regulations Conservation Biology 14:414-419.

Stanczak, M. and J. B. Keiper. 2004. Benthic invertebrates in adjacent created and natural wetlands in northeastern Ohio, USA. Wetlands 24:212-218.

Stauffer, A. L. and R. P. Brooks. 1997. Plant and soil responses to salvaged marsh surface and organic matter amendments at a created wetland in central Pennsylvania. Wetlands 17:90-105.

Stolt, M. H., M. H. Genthner, W. L. Daniels, V. A. Groover, S. Nagle, and K. C. Haering. 2000. Comparison of soil and other environmental conditions in constructed and adjacent palustrine reference wetlands. Wetlands 20:671-683.

Strausbaugh, P. D. and E. L. Core. 1977. Flora of West Virginia, 2nd edition. Senecca Books, Inc., Morgantown, WV.

Systat. 2004. SigmaPlot ® Software for Windows version 9.01. Systat Software, Inc, Point Richmond, CA.

The MathWorks. 2005. MATLAB®, v. 7.0.4. The MathWorks, Inc. Natick, Massachusetts.

van der Valk, A. G. 1981. Succession in wetlands: a Gleasonian approach. Ecology 62:688-696.

Vepraskas, M. J. and S. P. Faulkner. 2001. Redox chemistry of hydric soils. p. 85-105 in J. L. Richardson and M. J. Vepraskas (ed.) Wetland Soils: Genesis, Hydrology, Landscapes, and Classification. Lewis Publishers.

Vivian-smith, G. 1997. Microtopographic heterogeneity and floristic diversity in experimental wetland communities. The Journal of Ecology 85:71-82.

Waddington, J. M., P. A. Rotenberg, and F. J. Warren. 2001. Peat CO2 production in a natural and cutover peatland: implications for restoration. Biogeochemistry 54:115-130.

Wall, D. P., and S. P. Darwin. 1999. Vegetation and elevational gradients within a bottomland hardwood forest of southeastern Louisiana. The American Midland Naturalist 142:17-30.

Wentworth, T. R., G. P. Johnson, and R. L. Kologiski. 1988. Designation of wetlands by weighted averages of vegetation data: a preliminary evaluation. Water Resources Bulletin 24:389-396.

Whigham, D. F., M. Pittek, K. H. Hofmockel, T. E. Jordan, and A. L. Pepin. 2002. Biomass and nutrient dynamics in restored wetlands on the outer coastal plain of Maryland. Wetlands 22:562-574.

Whittecar, G. R. and W. L. Daniels. 1999. Use of hydrogeomorphic concepts to design created wetlands in southeastern Virginia. Geomorphology 31:355-371.

Page 75: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

74

Wilcox, D. A., N. B. Pavlovic, and M. L. Mueggler. 1985. Selected ecological characteristics of Scirpus cyperinus and its role as an invader of disturbed wetlands. Wetlands 5:87-97.

Yu, S. L., G. M. Fitch, and T. A. Earles. 1998. Constructed wetlands for nonpoint source pollution control. Virginia Transportation Research Council. Charlottesville, VA, USA. VTRC 99-R14.

Zak, D. R., W. E. Holmes, N. W. MacDonald, and K. S. Pregitzer. 1999. Soil temperature, matric potential, and the kinetics of microbial respiration and nitrogen mineralization. Soil Science Society of America Journal 63:575-584.

Zampella, R. A. and K. J. Laidig. 2003. Functional equivalency of natural and excavated coastal plain ponds. Wetlands 23:860-876.

Zar, J. H. 1984. Biostatistical Analysis, second edition. Prentice-Hall, Inc., Englewood Cliffs, NJ, USA.

Page 76: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

75

Table 1. Summary of mean (±1 SE) variation over loading rates for surface elevation, soil carbon (C), nitrogen (N), phosphorus (P), and C:N collected on July 18, 2005 (elevation) and August 22, 2005 (soil nutrients) in CCW, Charles City County, VA. LR 1 = 0 Mg ha-1, LR 2 = 56 Mg ha-1, LR 3 = 112 Mg ha-1, LR 4 = 224 Mg ha-1, LR 5 = 336 Mg ha-1. --------------------------Loading Rates (LR) -------------------------- Soil Variable LR1 LR2 LR3 LR4 LR5 n Loading Rate (Mg ha-1) 0 56 112 224 336 Surface Elevation (m) 10.36 ± 0.004 10.36 ± 0.01 10.38 ± 0.01 10.43 ± 0.01 10.47 ± 0.01 24 C content (%) 1.66 ± 0.28 3.37 ± 0.14 6.66 ± 1.08 16.53 ± 3.06 15.08 ± 3.51 4 N content (%) 0.10 ± 0.01 0.18 ± 0.004 0.29 ± 0.04 0.68 ± 0.09 0.64 ± 0.13 4 P content (mg kg-1) 2.75 ± 0.23 6.25 ± 1.33 8.25 ± 2.08 18.25 ± 2.37 17.75 ± 3.01 4 C:N 17.06 ± 0.98 18.29 ± 0.57 20.94 ± 0.76 23.89 ± 1.18 22.58 ± 1.35 4

Page 77: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

76

Table 2. Plant species occurring in collecting plots and relative IV during the 2005 growing season (April-October) in CCW, Charles City County, VA. Species names are accompanied by wetland indicator status (Indicator), plant life strategy (Duration), and are separated by each loading rate (LR) they occurred in. Nomenclature follows NRCS (2006). Planted species are denoted by *. LR 1 = 0 Mg ha-1, LR 2 = 56 Mg ha-1, LR 3 = 112 Mg ha-1, LR 4 = 224 Mg ha-1, LR 5 = 336 Mg ha-1.

Species Indicator Duration LR1 LR2 LR3 LR4 LR5

Aceraceae Acer rubrum L. FACW+ Perennial 0.30 0.94

Alismataceae Alisma plantago-aquatica L. OBL Perennial 0.17

Apocynaceae Apocynum cannabinum L. FACU Perennial 0.39

Asteraceae Baccharis halimifolia L. FACW Perennial 0.10 Bidens aristosa (Michx.) Britt. FACW- Annual/Biennial 1.18 1.81 2.29 0.52 0.26 Bidens frondosa L. FACW Annual 0.83 1.12 0.75 0.06 Eclipta prostrata (L.) L. FAC Annual/Perennial 0.28 0.36 0.63 0.69 0.94 Eupatorium perfoliatum L. FACW+ Perennial 0.18 Euthamia graminifolia (L.) Nutt. FAC Perennial 0.45 0.07 0.53 0.46 0.41 Solidago canadensis L. FACU Perennial 0.09 0.45 0.80 1.51 0.65 Symphyotrichum lateriflorum (L.) A.& D. Löve

FAC Perennial 1.20 0.25

Betulaceae * Betula alleghaniensis Britt. FAC Perennial 0.13 * Betula nigra L. FACW Perennial 3.10 2.86 3.06 4.28 4.77

Bignoniaceae Campsis radicans (L.) Seem. ex Bureau

FAC Perennial 0.07 0.09 0.09 0.32

Clusiaceae Hypericum crux-andreae (L.) Crantz FACU Perennial 0.21 Hypericum hypericoides (L.) Crantz FACU Perennial 0.12 Hypericum mutilum L. FACW Annual/Perennial 0.63 0.38 0.42 0.24

Commelinaceae Murdannia keisak (Hassk.) Hand.-Maz.

OBL Perennial 0.99 0.40 0.25

Cyperaceae Carex lurida Wahlenb. OBL Perennial 0.10 0.33 0.25 Carex vulpinoidea Michx. OBL Perennial 0.00 0.30 0.47 0.20 Cyperus pseudovegetus Steud. FACW Perennial 1.20 0.47 0.59 0.32 Cyperus strigosus L. FACW Perennial 0.19 Eleocharis obtusa (Willd.) J.A. Schultes

OBL Annual 9.89 10.79 4.24 1.24 0.65

Rhynchospora capitellata (Michx.) Vahl

OBL Perennial 0.09

Rhynchospora corniculata (Lam.) Gray

OBL Perennial 0.22

Scirpus cyperinus (L.) Kunth FACW+ Perennial 20.73 18.17 26.89 27.31 33.30

continued next page…

Page 78: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

77

Species Indicator Duration LR1 LR2 LR3 LR4 LR5

Euphorbiaceae Acalypha rhomboidea Raf. FACU- Annual 0.61 0.10 1.33 2.04 3.81 Chamaesyce maculata (L.) Small FACU- Annual 0.14 0.11 0.32 0.34 Phyllanthus caroliniensis Walt. FAC+ Annual 0.10

Fabaceae Lespedeza cuneata (Dum.-Cours.) G. Don

NI Perennial 0.30 0.12 0.55

Fagaceae * Quercus palustris Muenchh. FACW Perennial 1.27 1.46 1.41 1.70 2.26

Hamamelidaceae Liquidambar styraciflua L. FAC Perennial 0.24

Juncaceae Juncus acuminatus Michx. OBL Perennial 3.78 4.65 2.45 0.79 0.48 Juncus effusus L. OBL Perennial 3.54 6.25 6.12 7.57 4.60 Juncus tenuus Willd. FAC- Perennial 0.87 0.94 0.76 1.44 0.52

Lamiaceae Prunella vulgaris L. FACU+ Perennial 0.39 0.31 0.46 0.09

Lythraceae Rotala ramosior (L.) Koehne OBL Annual 0.20

Onagraceae Epilobium coloratum Biehler OBL Perennial 1.67 1.35 1.46 1.37 1.48 Ludwigia alternifolia L. FACW+ Perennial 0.53 0.59 1.28 0.80 0.66 Ludwigia glandulosa Walt. OBL Perennial 0.63 0.29 0.26 0.19 Ludwigia palustris (L.) Ell. OBL Perennial 2.79 3.90 2.43 1.29 0.57

Oxalidaceae Oxalis stricta L. UPL Perennial 0.11 0.22

Platanaceae Platanus occidentalis L. FACW- Perennial 0.30

Poaceae Andropogon virginicus L. FACU Perennial 0.75 0.87 0.38 3.27 4.45 Arthraxon hispidus (Thunb.) Makino None Annual 0.14 Dichanthelium scoparium (Lam.) Gould

FACW Perennial 0.20 0.08 0.46 0.00

Digitaria ischaemum (Schreb.) Schreb. ex Muhl.

UPL Annual 0.67 0.30

Echinochloa muricata (Beauv.) Fern. FACW+ Annual 3.23 2.61 0.99 0.65 0.17 Panicum dichotomiflorum Michx. FACW- Annual 0.46 0.66 1.04 0.07 0.22 Saccharum giganteum (Walt.) Pers. FACW+ Perennial 0.09 0.07 0.21 0.69 0.39 Setaria parviflora (Poir.) Kerguelen FAC Perennial 1.62 1.79 1.62 2.73 2.29

Polygonaceae Polygonum hydropiperoides Michx. OBL Perennial 3.76 4.21 1.86 2.03 1.61 Polygonum lapathifolium L. FACU+ Annual 0.35 Polygonum pensylvanicum L. FACW Annual 0.09 Polygonum persicaria L. FACW Annual/Perennial 1.19 0.92 0.43 0.26 0.69 Polygonum punctatum Ell. OBL Annual/Perennial 0.15 Rumex crispus L. FACU Perennial 0.41 0.44 0.50 0.40 0.99

Rosaceae Rubus argutus Link FACU Perennial 1.17 0.12

continued next page…

Page 79: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

78

Species Indicator Duration LR1 LR2 LR3 LR4 LR5

Rubiaceae Diodia virginiana L. FACW Annual/Perennial 0.25

Salicaceae Salix nigra Marsh. FACW+ Perennial 0.32 0.03 0.13 0.52 0.15

Scrophulariaceae Agalinis purpurea (L.) Pennell FACW- Annual 0.17

Typhaceae Typha latifolia L. OBL Perennial 4.24 5.99 3.14 2.48 2.01

Vitaceae Parthenocissus quinquefolia (L.) Planch.

FACU Perennial 0.10

Vitis rotundifolia Michx. FAC- Perennial 0.12

Non-Categorical Standing Dead 14.53 13.75 17.85 19.53 19.25 Bare Ground 12.75 11.53 10.38 7.80 8.58

Page 80: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

79

Table 3. Summary of dominant species in each loading rate based on IV using the 50:20 rule for the 2005 growing season (April-October) in CCW, Charles City County, VA. LR 1 = 0 Mg ha-1, LR 2 = 56 Mg ha-1, LR 3 = 112 Mg ha-1, LR 4 = 224 Mg ha-1, LR 5 = 336 Mg ha-1.

-----------------------------------Loading Rates (LR)------------------------------------ Species LR1 LR2 LR3 LR4 LR5

Scirpus cyperinus 20.73 18.17 26.89 27.31 33.30 Standing Dead 14.53 13.75 17.85 19.53 19.25 Bare Ground 12.75 11.53 10.38 7.80 Eleocharis obtusa 9.89 10.79

Page 81: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

80

Table 4. Summary of mean (±1 SE) variation over loading rates for total species richness (TS), per quadrat species richness (SR), evenness (J’), Shannon Diversity Index (H’), weighted average (WA), and standing crop biomass over the 2005 growing season (April-October) in CCW, Charles City County, VA. Values do not including Bare Ground or Standing Dead categories. LR 1 = 0 Mg ha-1, LR 2 = 56 Mg ha-1, LR 3 = 112 Mg ha-1, LR 4 = 224 Mg ha-1, LR 5 = 336 Mg ha-1. ------------------------------Loading Rates (LR) ------------------------------ Vegetation Variable LR1 LR2 LR3 LR4 LR5 n Total Species (TS) 39 35 41 48 39 Species Richness (SR) 7.39 ± 0.45 7.75 ± 0.37 6.71 ± 0.49 7.30 ± 0.40 5.29 ± 0.26 56 Evenness (J’) 0.89 ± 0.01 0.92 ± 0.01 0.87 ± 0.01 0.89 ± 0.01 0.86 ± 0.01 56 Shannon Index (H’) 1.71 ± 0.07 1.82 ± 0.05 1.54 ± 0.08 1.70 ± 0.05 1.38 ± 0.05 56 Weighted Averages (WA) 1.63 ± 0.03 1.57 ± 0.03 1.76 ± 0.04 2.02 ± 0.05 2.12 ± 0.06 56 Standing Crop Biomass 603.5 ± 179.8 580.0 ± 96.9 601.5 ± 111.0 623.0 ± 193.5 789.5 ± 60.2 8

Page 82: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

81

Table 5. Matrix of Ellenberg Community Coefficient Similarity Index (SIE) from relative IV of species, calculated for each month of the 2005 Growing Season (April-October) and averaged for each treatment comparison ) in CCW, Charles City County, VA. LR 1 = 0 Mg ha-1, LR 2 = 56 Mg ha-1, LR 3 = 112 Mg ha-1, LR 4 = 224 Mg ha-1, LR 5 = 336 Mg ha-1.

LR 5 LR 4 LR 3 LR 2 LR 1 0.76 0.82 0.88 0.92 LR 2 0.76 0.85 0.90 LR 3 0.80 0.84 LR 4 0.86

Page 83: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

82

Figure 1. Charles City Mitigation Wetland (CCW) site location, Charles City County, VA (Bergschneider 2005). Part a represents the location of Charles City County in Virginia. Part b shows the location of the experimental block (denoted by the “W”) within CCW.

a.

b.

Page 84: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

83

Figure 2. Schematic design of CCW experimental plots in Charles City County, VA.

Page 85: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

84

Loading Rate (Mg ha-1)

0 100 200 300

Ele

vatio

n (m

)

10.25

10.30

10.35

10.40

10.45

10.50

Figure 3. Mean (±1 SE) elevation vs. Loading Rate in CCW, Charles City County, VA, from April to October 2005. Different letters above the treatment means denote significant differences based on Tukey’s Family Error Rate.

a a a

b

c

p ≤ 0.001 R2 = 0.58 n = 24

Page 86: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

85

Loading Rate (Mg ha-1)

0 100 200 300 400

Soil

C:N

14

16

18

20

22

24

26

28

a ac

ad

bd cd

p ≤ 0.001 R2 = 0.49 n = 4

Figure 4. Linear regression of mean (±1 SE) soil Carbon (C), Nitrogen (N), Phosphorus (P) content and C:N vs. Loading Rate in CCW, Charles City County, VA, on August 22, 2005. P-values reported indicate significance of the linear regression at the α = 0.05 level. Different letters above the treatment means denote significant differences at the α = 0.05 level based on Wilcoxon Rank Sum Test for C and N and Tukey’s family Error Rate for P and C:N.

Loading Rate (Mg ha-1)

0 100 200 300 400

Soil

P c

onte

nt (m

g kg

-1)

0

5

10

15

20

25

30

a a

ac

b bc

p ≤ 0.001 R2 = 0.66 n = 4

Loading Rate (Mg ha-1)

0 100 200 300 400

Soil

N c

onte

nt (%

)

0.0

0.2

0.4

0.6

0.8

1.0

a b

c

d cd p ≤ 0.001 R2 = 0.66 n = 4

b

Loading Rate (Mg ha-1)

0 100 200 300

Soi

l C c

onte

nt (%

)

0

5

10

15

20

25

30

a

c

d cd

p ≤ 0.001 R2 = 0.60 n = 4

Page 87: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

86

Elevation (m)

10.35 10.40 10.45 10.50 10.55

WA

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

Figure 5. Linear regression of mean (±1 SE) per plot weighted average (WA) values vs. elevation in CCW, Charles City County, VA, from April to October 2005. The p-value reported indicates significance of the linear regression at the α = 0.05 level.

p ≤ 0.001 R2 = 0.35

Page 88: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

87

Loading Rate (Mg ha-1)

0 56 112 224 336

Impo

rtanc

e Va

lue

(IV)

-80

-60

-40

-20

0

20

40

60

80

100

FAC+ FACW OBL

FAC FACU UPL

Figure 6. Total IV of general wetland indicator categories (Reed 1988) for different loading rates in CCW, Charles City County, VA, April to October 2005. FACW and FACU categories include +/- modifiers, while the FAC category includes the + modifier.

Page 89: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

88

Elevation (m)

10.35 10.40 10.45 10.50

Biom

ass

(g m

-2)

0

200

400

600

800

1000

1200

1400

1600

1800

2000

2200

Figure 7a. Biomass values vs. Elevation in CCW, Charles City County, VA, on August 22, 2005. The p-value reported indicates significance of the quadratic regression at the α = 0.05 level.

p = 0.007 R2 = 0.23

Page 90: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

89

Elevation (m)

10.35 10.40 10.45 10.50

Bio

mas

s (g

m-2

)

0

400

800

1200

1600

2000

Figure 7b. Linear regression of biomass values (outliers removed) vs. elevation in CCW, Charles City County, VA, on August 22, 2005. The p-value reported indicates significance of the linear regression at the α = 0.05 level.

p = 0.030 R2 = 0.13

Page 91: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

90

Figure 8. Sorted loadings on first principal component from a Principal Components Analysis calculated using log transformed morphometric data. Total height, crown diameter, and main stem diameter show similar loadings on the first principal component (PC1), and have the common metric of size (length or width), therefore the scores on PC1 were used as an index of tree development.

PC1 Loadings

0.0 0.1 0.2 0.3 0.4 0.5 0.6

Crown Diameter

Stem Diameter

Number of Stems

Total Height

Page 92: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

91

Loading Rate (Mg ha-1)

0 100 200 300

Tree

"Siz

e" (S

core

s on

PC

1)

-4

-2

0

2

4

Figure 9. Linear regression of mean (±1 SE) tree “size” vs. Loading Rate in CCW in Charles City County, VA on June 21, 2005. Different letters above the treatment means denote significant differences based on Tukey’s family error rate. The p-value reported indicates significance of the linear regression at the α = 0.05 level.

a a a

b b

p ≤ 0.001 R2 = 0.21 n = 20

Page 93: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

92

Elevation (m)

10.30 10.35 10.40 10.45 10.50 10.55

Tree

Siz

e (S

core

s on

PC

1)

-6

-4

-2

0

2

4

6

Figure 10. Linear regression of PC1 Scores (tree size) vs. elevation in CCW, Charles City County, VA, on June 21, 2005. The p-value reported indicates significance of the linear regression at the α = 0.05 level.

p ≤ 0.001 R2 = 0.28

Page 94: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

93

Soil P (mg kg-1)

0 5 10 15 20 25 30

Tree

Siz

e (S

core

s on

PC

1)

-6

-4

-2

0

2

4

6

Figure 11. Linear regression of mean (±1 SE) tree “size” (PC1) values vs. soil phosphorus (P) content in CCW, Charles City County, VA, on August 22, 2005. The p-value reported indicates significance of the linear regression at the α = 0.05 level.

p ≤ 0.001 R2 = 0.15

Page 95: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

94

Soil P (mg kg-1)

0 2 4 6 8 10 12

Biom

ass

(g m

-2)

0

200

400

600

800

1000

1200

1400

Figure 12. Linear regression of mean (±1 SE) biomass values vs. soil phosphorus (P) content (0 to 12 mg kg-1) in CCW, Charles City County, VA, on August 22, 2005. Biomass values are modified such that 4 outlying data points have been removed. The p-value reported indicates significance of the linear regression at the α = 0.05 level.

p = 0.017 R2 = 0.25

Page 96: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

95

Soil P (mg kg-1)

0 2 4 6 8 10 12 14 16

Tree

Siz

e (s

core

s on

PC

1)

-4

-2

0

2

4

6

Figure 13. Linear regression of mean (±1 SE) tree “size” (PC1) values vs. soil phosphorus (P) content (0 to 15 mg kg-1) in CCW, Charles City County, VA, on August 22, 2005. The p-value reported indicates significance of the linear regression at the α = 0.05 level.

p ≤ 0.001 R2 = 0.30

Page 97: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

96

Chapter 3

Wetland ecosystem gas exchange in response to organic matter loading rates

INTRODUCTION

Mitigation requirements based on section 404 of the 1977 Clean Water Act often

require the replacement of specific acreages of wetland habitat to compensate for the loss

of natural wetlands and their associated functions. Wetland creation is typically more

difficult than restoration due to the complexity of establishing the necessary hydrology,

soils, and vegetation in areas where wetlands did not previously or historically exist (Stolt

et al. 2000). Thus, failure (i.e. not achieving functional status of natural wetlands) of

these systems is common (Zedler 1997, Whigham 1999).

One reason for created wetland failure has been improper soil conditions,

including low soil organic matter (Stauffer and Brooks 1997). Although organic matter

accumulation is a characteristic feature of most wetland systems, traditional creation

practices such as top soil scraping typically remove up to 1 meter of surface soil (W.L.

Daniels, pers. com.), including organic-rich material in the A horizon (Whittecar and

Daniels 1999, Bergschneider 2005). Indeed, lower organic matter than that of natural

wetlands has been reported for created wetlands in Pennsylvania (Stauffer and Brooks

1997, Cole et al. 2001, Brooks et al. 2005), Florida (Anderson and Cowell 2004), and

Virginia (Whittecar and Daniels 1999, Bruland and Richardson 2004). As such, many

studies have advocated amending wetland creation sites with organic material either in

Page 98: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

97

the form of salvaged natural wetland soils or mulches (Stauffer and Brooks 1997,

Whittecar and Daniels 1999, McKinstry and Anderson 2003, Anderson and Cowell 2004,

Bruland and Richardson 2004). Because organic matter accumulation occurs over

centuries, the hope is that adding organic amendments to created wetland soils will help

to “jump start” the maturation process, and achieve functional equivalency (to natural

wetlands) sooner.

Since construction practices usually remove vegetation, soils, and existing seed

banks, most created wetlands occur as primary successional ecosystems. Organic matter

(OM) is important in these ecosystems since its accumulation over time acts to mitigate

perturbations caused by the physical environment; OM therefore acts as a stabilizing

force, theoretically leading to a “climax” system where energy efficiency for both

biomass and symbiotic organism interaction are maximized (Odum 1969). Based on the

work of E.P. Odum (1969), as ecosystems mature, and energy use shifts over time from

high biomass production to ecosystem maintenance, changes in the overall ecosystem

energy balance should reflect a progression from a net autotrophic regime

(photosynthesis (P) / respiration (R) ratio >1) to a balanced state (P/R = 1). As such, one

could theoretically use the P/R ratio as a functional index of the relative maturity of

ecosystems.

Few studies have utilized Odum’s concepts as a tool to characterize ecosystem

successional maturity. However, several studies have utilized techniques of measuring

carbon gas (CO2, CH4) fluxes as C uptake (Community Production, Gross Primary

Production, or Photosynthesis), C emission (Respiration), and their net difference (Net

Ecosystem Exchange, NEE) to observe energy fluxes between wetlands and the

Page 99: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

98

atmosphere (Bubier et al. 1998, Frolking et al. 1998, Schreader et al. 1998, Streever et al.

1998, Clark et al. 1999, Neubauer et al. 2000, Wickland et al. 2001). These techniques

provide an opportunity to use Odum’s (1969) bioenergetics theory to observe the relative

successional maturity of wetland ecosystems. Such uses of CO2 flux measurements have

only recently been applied to wetlands (Roggero 2003, Cornell et al. 2006), but are

absent from studies of created non-tidal wetland systems.

The purpose of this study was to examine the primary succession processes

occurring in a created forested wetland, in terms of ecosystem gas exchange, along a

gradient of soil organic carbon. Our goal was to determine how Gross Primary

Production, Respiration, and Net Ecosystem Exchange varied in a created freshwater

wetland with respect to different loadings of organic matter in the soil. We hypothesized

that Ecosystem Gas Exchange parameters would vary with respect to organic carbon

loading. Using the results of this study, we were able to compare how various organic

matter amendments affect energy flow in a primary successional system, as well as to

develop management recommendations for organic matter loading rates in created

wetland systems.

SITE DESCRIPTION

The Charles City Wetland Mitigation Site (CCW) is a 21 ha constructed

mitigation wetland owned by the Virginia Department of Transportation (VDOT) in

Charles City County, VA, USA (76˚55’33” W, 37˚20’37” N) (Chapter 2, Figure 1a). The

site can be classified as palustrine emergent headwater wetlands (Cowardin et al. 1979,

Page 100: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

99

DeBerry and Perry 2004), with 18 ha designated as forested wetlands. CCW was initially

constructed in 1996, although sections have been regraded as recently as 2003 (see

Bergschneider 2005). This site is characterized by silty clay loam soils typically

exceeding 1 m in depth. Precipitation is the dominant hydrologic factor in CCW, and fall

and winter months are generally accompanied by up to 0.6 m of standing water (Schmidt

2002).

The study site consisted of a 680 m2 area located along the northern edge of the

CCW (Chapter 2, Figure 1b) containing twenty 4.57 x 3.05 m (15 x 10 ft) plots separated

by 3.05 m (10 ft) alleyways. In June 2002 each plot received one of five organic matter

(wood/yard waste compost, Appendix 1) loading rates, ranging from 0 to 336 Mg ha-1

arranged in a randomized complete block design (Chapter 2, Figure 2), creating four plot

replicates per treatment. Amendments were incorporated into the plots via disking and

roto-tilling by tractor. Plots showed a positive correlation between loading rate and

surface elevation due in part to incomplete incorporation of the amendment material,

especially in Loading Rates 4 and 5 (Chapter 2, Table 1, Figure 3). Further details

regarding the preparation of the experimental plots as well as previous studies at CCW

can be found in Chapter 2 as well as in Bergschneider (2005).

METHODS

Water Table Depth

Elevations of each plot were determined as described in Chapter 2 (Elevation).

The depth of the water table was measured periodically (2-4 times per month) in one of

Page 101: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

100

~30 wells maintained by the Virginia Department of Transportation (VDOT) at the

CCW. Well # 2 data (VDOT, unpublished data) were chosen as the approximate level of

the water table in the experimental plots due to the close proximity of the well to the plots

(~10 m north of the experimental block). The water table elevation was estimated using

relative water table height data from Well # 2 and surface water depth data taken from an

adjacent plot (Plot 12) during CO2 flux sampling. Since the plot surface elevation was

known, elevation of the water table during flooded periods could be determined by a

regression of water table surface (during flooded periods) of Plot 12 (i.e. plot surface

elevation + depth of surface water) vs. relative water table depth in Well # 2 (p = 0.004,

R2 = 0.96). Plot elevations were then subtracted from the water table elevation to

determine water table depths relative to the surface of each plot.

CO2 Flux

Ecosystem exchange of CO2 was measured using one of two aluminum-framed

chambers (Figure 1) (e.g. Streever et al. 1998, Roggero 2003, Cornell et al. 2006). Each

chamber was covered in airtight, transparent, Tefzel® film (DuPont, Inc., Circleville,

OH). The chambers were clamped onto 0.25 m2 aluminum bases (0.5 x 0.5 m), one

installed in the center of each plot two to four weeks prior to sampling. The bases were

inserted to a point at which approximately 3 cm of the base was above the soil surface.

The height of vegetation within the chamber base determined which chamber was used

for each plot (1.0 or 1.5 m tall).

The chambers were connected via bev-a-line tubing to a LI-6200 Portable

Infrared Gas Analyzer (IRGA) (LI-COR, Inc., Lincoln, NE), which cycled air between

Page 102: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

101

the chamber and the IRGA and periodically measured and recorded the concentration of

CO2. Prior to measurement, the IRGA was zeroed with ambient air passing sequentially

through a CO2 scrub (soda lime) and a desiccant (magnesium perchlorate, Mg(ClO4)2).

The IRGA was then calibrated with a gas standard containing 500 ppmv CO2 in N2 (Scott

Specialty Gases, Inc. Plumsteadville, PA).

In the field, CO2 levels were monitored with the chambers in place, but open to

the atmosphere, until a steady-state concentration had been reached. After this

equilibration period (usually 5 -10 minutes), the chamber lids were closed, the change in

[CO2] was allowed to stabilize (usually ~1 minute), and the sampling period was started.

[CO2] was measured and logged at ~30 second intervals during a five minute sampling

period, and thus 11 measurements were recorded by the IRGA for each light level.

Incident irradiance (Photosynthetically Active Radiation, PAR) was measured

concurrently with CO2 flux using a LI-190 Quantum Sensor (LI-COR, Inc., Lincoln, NE)

placed at the top of the inside of each chamber. These measurements were recorded at 1

minute intervals using a LI-1000 data logger (LI-COR, Inc., Lincoln, NE). Air

temperature within the chamber was also concurrently recorded at 5 second intervals

using a HOBO® U10 data logger (Onset, Inc., Bourne, MA).

CO2 fluxes were sampled monthly in each plot from June 2005 through May

2006. The majority of sampling was performed in full light, between 0900 and 1600

(Eastern Standard Time), to maximize light intensities. Four sampling treatments were

performed in each plot each month: one in full (ambient) light, two at different

intermediate lights levels (using two separate jackets of window screening), and one in

the dark (using an opaque blanket). Due to low photosynthetic activity, only the dark

Page 103: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

102

measurements were performed in December 2005 and January 2006. Measurements at

all light levels for each plot were measured on the same day, and monthly sampling for

all 20 plots was completed within two to four days.

Calculations and Modeling

Short-term field measurements of CO2 fluxes were scaled to monthly and annual

rates using a CO2 flux model. In the model, CO2 flux rates at various time scales are

controlled by changes in PAR and air temperature, measured at the Virginia Institute of

Marine Science (VIMS, ~40 km east of CCW) and at a nearby NOAA weather station,

Williamsburg, VA (~20 km east of CCW, 76˚71’70” W, 37˚23’30” N) (NOAA 2006),

respectively. CO2 fluxes were estimated for a one year period from June 2005 through

May 2006.

Monthly measurements of gross primary production (GPP) were calculated from

field measurements at full light and two shaded treatments, plus the respiration

measurements (dark fluxes) taken immediately following light measurements. Changes

in GPP over various time scales were determined by fitting GPP vs. PAR curves, one per

plot, per month. Curves were fit by hyperbolic curves using SigmaPlot®, v. 9.0 (Systat

2004). GPP was modeled as:

where I = average hourly irradiance (PAR), and a and b are empirically-derived constants

with the units of μmol C m-2 h-1 and μE m-2 h-1, respectively (Neubauer et al. 2000).

Respiration (R) rates were calculated using hourly temperature data as:

GPP = [(a x I) / (b + I)]

R = y0 + ax

Page 104: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

103

where y0 is the slope of an Arrhenius plot of the natural log of CO2 flux versus the

inverse of temperature in degrees K, a is an empirically-derived constant, and x is

average hourly air temperature in degrees K.

Hourly GPP and R rates were summed to obtain monthly and annual rates, and

NEE was calculated as:

Gas exchange was expressed relative to the ecosystem with GPP as positive values and

respiration as negative values.

Additional Carbon Sources

CH4 (methane) fluxes were sampled monthly in each plot from July 2005 to June

2006. Instantaneous CH4 flux rates were determined to be insignificant (monthly average

range from -0.05 to 0.09 mg C m-2 min-1). Therefore, CH4 data were not included in the

C flux analysis for CCW. However, monthly instantaneous CH4 fluxes and the methods

used are reported in Appendix 5.

Data Analysis

Simple regression was used to explore relationships among various parameters

(GPP, R, NEE, and GPP:R) and loading rate. Tests for normality were performed using

Minitab®, release 14 (Minitab 2005), while various regressions were performed using

SigmaPlot, v. 9.0 (Systat 2004). Unless otherwise indicated, data are reported as means ±

the Standard Error.

NEE = GPP + R

Page 105: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

104

RESULTS

Annual CO2 Fluxes

Respiration (R) values ranged from 796 ± 171 to 1912 ± 182 g C m-2 yr-1 in LRs 1

and 5 (Table 1, Figure 2), respectively, and correlated positively with loading rate (p =

<0.001, R2 = 0.62) (Figure 3). Gross primary production (GPP) ranged from 938 ± 108 g

C m-2 yr-1 (LR 1) to 1452 ± 96 g C m-2 yr-1 (LR 5), with values positively correlating with

loading rate (p = 0.002, R2 = 0.41). Annual NEE ranged from -459 ± 240 to 141 ± 67 g C

m-2 yr-1 in LRs 5 and 1, respectively, displaying a negative correlation with loading rate

(p = 0.006, R2 = 0.34). The weak regression of NEE and loading rate was possibly due to

high variability in GPP and R among treatment replicates, which were used to calculate

NEE by difference.

Monthly CO2 Fluxes

Loading rates 1 and 5 showed the lowest and highest R rates, respectively,

regardless of month, with the other loading rates sequentially falling in between (Figure

4). The highest R rates occurred in July with slightly lower rates in August. These high

summer rates coincided with high seasonal temperatures, and may also have been

influenced by hydrologic variation (i.e. soil moisture). That is, although the highest

annual temperatures typically occur in August, it is possible that the relatively lower

water table in July (see Figure 4) enhanced aerobic, and therefore more efficient,

respiration, creating the highest R rates observed in this study. This hydrologic effect on

respiration is consistent with a regression of monthly respiration rates versus water table

Page 106: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

105

depth (Figure 7), which shows a significant negative correlation between respiration rate

and water table height (p < 0.001, R2 = 0.60). The lowest R rates in the study occurred in

the winter months (December-February).

Monthly GPP values (Figure 5) showed more variability among replicate plots

than R. Generally, LRs 1 and 5 showed the lowest and highest GPP, respectively, in any

given month, with LRs 2 - 4 following more-or-less sequentially. However, there were

exceptions, such as LR 4 with the highest GPP in June, and LR 1 with the second highest

GPP in April. The highest overall GPP rates occurred from July to August, while the

lowest measured GPP occurred in February. Because measurements on a subset of plots

showed no significant photosynthesis in any loading rate during December and January,

we did not measure GPP during these months. However, we assume that the actual

lowest GPP (near zero) occurred in December and January.

All loading rates showed generally high NEE (Figure 6) in late spring (May and

June) and early fall (September), with lower NEE from late fall to early spring (October

to April) and the lowest NEE in mid summer (July). Low NEE (net heterotrophic

activity) from late fall to early spring was caused by low GPP due to seasonal senescence

of vegetation (data not shown). Further decreases in NEE in early spring (March and

April) were likely due to the effect of increases in temperature on respiration rate, while

GPP was still relatively low. Increasing NEE in the late spring (May) coincided with full

leaf-out of vegetation and the resulting increased GPP. Peak NEE for all loading rates

except LR 4 occurred in May; this was likely due to low R controlled by a high water

table and moderately low temperatures, as well as increasing GPP due to longer periods

of high solar irradiance.

Page 107: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

106

All loading rates had their lowest NEE values during July (range from -38 ± 40

[LR 1] to -102 ± 40 g C m-2 yr-1 [LR 3]). These low values reflected high respiration

brought on by high summer temperatures as well as dry conditions (as shown by water

table depth), which would have allowed aerobic respiration of organic matter. Indeed, all

loading rates showed higher R in July than any other month (see Figure 4). The positive

spike in NEE in September can be attributed to a decrease in R due to an increase in soil

moisture, while conditions were still optimal for high GPP (i.e. warm temperatures, long

periods of high irradiance).

Most loading rates showed net heterotrophic activity (i.e. NEE < 0) the majority

of the year, with LRs 2, 3, 4 and 5 showing net autotrophic activity (i.e. NEE > 0) for

only four, three, four, and two, months, respectively. LR 1, on the other hand, was net

autotrophic for one half (6 months) of the year. All loading rates were net autotrophic

during September, while all except for LR 4 were net autotrophic in May. All plots were

net heterotrophic in July and from December through March, while most plots showed

NEEs near or below 0 in October, November, and April.

Ratio of Photosynthesis to Respiration (GPP:R)

GPP:R ranged from 0.8 ± 0.1 in LR 5 to 1.3 ± 0.1 in LR 1, and appeared to trend

negatively with loading rate (Figure 8). LRs 1 and 2 were above Odum’s (1969)

theoretical line of 1:1, while LRs 3 -5 fell below the 1:1 line, indicating that they were

annually net autotrophic and net heterotrophic, respectively.

Page 108: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

107

DISCUSSION

CO2 Fluxes

The range of annual GPP and R rates measured in this study (938-1452 and 796-

1912 g C m-2 yr-1, respectively) were on the high side of annual rates measured in other

studies. For example, Hirota et al. (2006) reported GPP at 51 and R at 37 g C m-2 yr-1 in

a marsh on the Qinghai-Tibetan Plateau, Clark et al. (1999) found GPP to be 647 and R

to be 562 g C m-2 yr-1 in a Florida Cypress ecosystem, and Neubauer et al. (2000)

measured gross community GPP and R rates of 1062 and 1269 g C m-2 yr-1, respectively,

in a freshwater tidal marsh in Virginia. Mean NEE rates in this study were also highly

variable among loading rates (range of 141 to -459 g C m-2 yr-1), and encompass the

range of those of the previous studies listed (84 to -207 g C m-2 yr-1). One major

difference between the current study site and the other studies is age. Hirota et al. (2006),

Clark et al. (1999), and Neubauer et al. (2000) all studied relatively mature, natural

ecosystems compared to our ecosystem (CCW), which partially explains the near-zero

NEEs in those studies. The slightly more negative NEE in Neubauer et al. (2000) was

explained in part by tidally-controlled lateral transport of organic debris into the marsh,

thereby increasing the respiration component. CCW, on the other hand, represents an

early successional system, and would be expected to have NEE values far from

equilibrium. The different loading rates appear to modify whether gas flux will trend

positively or negatively with maturity; however, all trends would be expected to approach

equilibrium.

Page 109: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

108

High variability in annual NEE within treatment replicates may be due to slight

differences in plant community composition and/or plant sizes within chamber bases,

which could then affect photosynthetic and plant respiratory rates. Species richness data

(Chapter 2, Table 4) corroborates this hypothesis, as standard deviations ranged from 36

to 55 % of the mean (n = 56) of growing season plant community diversity measurements

among loading rates sampled in the same plots (though not within the chamber bases) as

the current study. Further, standing crop biomass standard deviations (Chapter 2, Table

4) ranged from 22 to 88% of the mean (n = 8) among loading rates. Thus, it is feasible

that within-treatment plant communities could produce highly variable GPP and plant R

due to differences in species composition and biomass. Further, typically diverse non-

tidal freshwater wetlands (e.g. sedge meadows, this study) could reasonably be expected

to have higher spatial CO2 flux variability than would wetlands with lower plant diversity

(e.g. salt marshes).

The significant positive correlation between annual GPP and loading rate

indicates that increasing amounts of organic material encouraged higher plant production,

perhaps due to increasing soil nutrients N and P (see Chapter 2, Figure 4) or inundation

frequency differences driven by plot elevations (see Chapter 2, Figure 3). However,

mean annual GPP only increased by a factor of 1.5 from LR 1 (938 g C m-2 yr-1) to LR 5

(1452 g C m-2 yr-1). The narrow ranges in GPP among loading rates agree with the results

of studies comparing standing crop biomass, another measure of annual plant production,

between natural and created wetlands (DeBerry and Perry 2004), amended and

unamended wetlands (Anderson and Cowell 2004), and along gradients of soil organic

matter (Cole et al. 2001, Bergschneider 2005). In fact, standing crop data from this study

Page 110: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

109

showed no significant correlation or significant differences among loading rates (Chapter

2, Table 4). In addition, using similar field methods as the current study, both Cornell et

al. (2006) and Roggero (2003) found similar rates of annual GPP between created salt

marshes (low soil organic matter) and natural reference marshes (high soil organic

matter).

Annual respiration values showed a positive relationship with loading rate, most

likely accounted for by microbial respiration of the increasing organic amendments.

Given adequate soil moisture and temperature, acceptable organic matter quality, and an

available heterotrophic microbial community (Craft 2001), rates of organic matter

decomposition would logically increase with increasing organic material availability. In

addition, increased soil organic matter usually results in lower soil bulk densities (Collins

and Kuehl 2001). In fact, a recent study of the same research plots Bergschneider (2005)

measured bulk densities that were significantly different among, and negatively

correlated with, loading rate (range from LR 1 = 1.3 g cm-3 to LR 5 = 0.5 g cm-3). Such

decreases in bulk density likely have the effect of increasing the depth of aeration in the

soil profile, thereby exposing more of the available organic matter to oxidation

(particularly in dry conditions).

Further, as plot elevation and loading rate were positively correlated, it is also

possible that some of the increase in respiration with loading rate was related to a

decrease in soil inundation. For example, studies have shown that the mean water table

position is a good predictor of mean respiration rates (Bubier et al. 1998, Wickland et al.

2001). Other studies suggest that wetland ecosystems can change from net CO2 sinks to

Page 111: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

110

net sources primarily based on changes in water table (Shurpali et al. 1995, Lafleur et al.

1997, Suyker et al. 1997).

Importantly, however, ecosystem respiration is the result of the metabolic

respiration of plants, in addition to the microbially mediated respiration of soil organic

matter fractions (Trumbore 2000). Although plant respiration was not directly measured

in this study and could not be estimated as a percentage of the total, it is generally

accepted that plants respire about 50% of the carbon derived through photosynthesis after

photorespiration, with the remaining 50% used for growth, propagation, nutrient

acquisition, and litter production (Ryan 1991). Because of this proportionality, the

narrow annual GPP range among loading rates, and the fact that mean annual R increased

by a factor of 2.5 from LR 1 (796 g C m-2 yr-1) to LR 5 (1912 g C m-2 yr-1), it can be

inferred that plant-related respiration could not totally explain the differences in total

respiration among treatments.

The negative trend in NEE with loading rate seems related mostly to differences

in R rather than GPP among loading rates. Thus, respiration was likely the dominant

CO2 flux determinant (i.e. controls NEE) in this study.

Successional Maturity

The negative correlation between the photosynthesis to respiration ratio (GPP:R)

and loading rate generally followed the predictions of Odum (1969). That is, if one

considered high soil organic matter (OM) as a proxy for successional maturity, one would

expect the ratio of photosynthesis to respiration to decrease (approach 1) from early to

later successional stages. Of course, OM accumulation is only one of many attributes

Page 112: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

111

(i.e. plant diversity and structure, biogeochemical cycling, etc.) expected to change as an

ecosystem matures. As such, OM loading rate is not a perfect proxy for successional

state, and care must be taken to cautiously interpret ecosystem function when using this

parameter.

LRs 1 and 2 showed average GPP:R above 1, and thus displayed characteristics of

an early successional system. LR 1 had no OM amendments and very low soil organic

carbon content (1.7 ± 0.3%, Chapter 2, Table 1, Figure 4), and thus would be expected to

have higher production than respiration. Judging from the low organic content of created

wetland soils (Whittecar and Daniels 1999, Stolt et al. 2000) as well as the quickly

accumulating plant biomass of southeastern Virginia created wetlands (DeBerry and

Perry 2004), the LR 1 ratio is likely representative of many non-tidal created wetlands in

this area. The slight decrease in the LR 2 ratio was probably related to the increased soil

organic content from LR 1 (see “CO2 Fluxes”), but organic matter is apparently still

accumulating in these plots (i.e. ratio > 1). Of course, LR 2 is not in a late successional

stage, and obviously lacks the plant structure of mature forested wetlands. However, the

energy flow is balanced, and appears to mimic that of a climax community.

LRs 3-5 all show GPP:R ratios below 1. Ratios below 1 are typically produced in

cases of organic pollution (Odum 1969). It is difficult to imagine a “naturally”

functioning ecosystem that could maintain such a ratio, as the respiration component

would be directly proportional to the amount of organic matter produced in the system.

Other anthropogenic circumstances can lead to unusually low GPP:R ratios. For

example, forest clearing can produce P/R ratios below 1, due to decomposition of dead

organic matter immediately following clear cutting (Smith 1996). Just as cleared forests

Page 113: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

112

will eventually reach a balance between production and respiration (if regeneration is

allowed), so too should LRs 3-5. It appears that the high loadings of organic matter in

these treatments added more organic carbon to the system than can be maintained under

the ambient environmental conditions. The response, then, was high respiration rates

(relative to production), acting to return the system to a carbon (i.e. energy) equilibrium.

The decreasing ratios with increasing loading rate can be explained simply by the

successively higher concentrations of organic carbon.

SUMMARY AND RECOMMENDATIONS

Annual fluxes of CO2 at the CCW indicated generally decreasing NEE of carbon

with increasing organic amendment loadings. These net fluxes were controlled primarily

by changes in respiration. The photosynthesis to respiration ratios (GPP:R) appeared to

mimic the theoretical decrease with successional maturity suggested by Odum (1969).

However, the higher loading rate ratios fell below the 1:1 relationship indicative of

climax systems, and instead were more suggestive of organically polluted environments.

The lower loading rates were suggestive of early successional systems, however, and

appear to be accumulating carbon as is typical of primary successional systems.

Comparisons of carbon flux components (GPP, R, NEE) as practiced in this study

showed sensitivity to many environmental gradients, and provide an opportunity to study

successional processes and differences in ecosystem parameters on a quantitative basis.

Using these techniques, wetland function can potentially be monitored on theoretical

grounds, instead of using only superficial indicators such as standing crop biomass.

Page 114: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

113

Our results indicate that organic amendments do affect CO2 flux processes in

created, early successional wetland systems. However, organic amendment loading rate

recommendations must be based on the specific goals of individual mitigation projects.

In a general sense, the higher rates (LRs 3-5) would most likely prove unnecessary since

the ecosystem energy balance would act to oxidize excess organic material until

equilibrium was reached (i.e. P/R = 1). If the primary mitigation goal is to maximize

carbon sequestration, not incorporating an amendment (i.e. LR 1) may produce best

results. However, opting for this option would ignore many of the positive effects that

studies such as Bergschneider (2005) have shown to accompany organic amendments.

Thus, a low to moderate amendment such as LR 2 could potentially balance various

mitigation goals such as plant diversity and soil fertility and bulk density, along with

providing net carbon sequestration or balance.

Page 115: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

114

Literature Review Anderson, C. J., and B. C. Cowell. 2004. Mulching effects on the seasonally flooded zone

of west-central Florida, USA wetlands. Wetlands 24:811-819.

Bergschneider, C. R. 2005. Determining an appropriate organic matter loading rate for a created coastal plain forested wetland. Thesis. Virginia Tech, Blacksburg, VA.

Brooks, R. P., D. H. Wardrop, C. H. Cole, and D. A. Cambell. 2005. Are we purveyors of wetland homogeneity?: A model of degradation and restoration to improve wetland mitigation performance. Ecological Engineering 24:331-340.

Bruland, G. L., and C. J. Richardson. 2004. Hydrologic gradients and topsoil additions affect soil properties of Virginia created wetlands. Soil Science Society of America Journal 68:2069-2077.

Bubier, J. L., P. M. Crill, T. R. Moore, K. Savage, and R. K. Varner. 1998. Seasonal patterns and controls on net ecosystem CO2 exchange in a boreal peatland landscape. Global Biogeochemical Cycles 12:703-714.

Clark, K. L., H. L. Gholz, J. B. Moncrieff, F. Cropley, and H. W. Loescher. 1999. Environmental controls over net exchanges of carbon dioxide from contrasting Florida ecosystems. Ecological Applications 9:936-948.

Cole, A. C., R. P. Brooks, and D. H. Wardrop. 2001. Assessing the relationship between biomass and soil organic matter in created wetlands of central Pennsylvania, USA. Ecological Engineering 17:423-428.

Collins, M. E., and R. J. Kuehl. 2001. Organic matter accumulation in organic soils. Pages 137-162 in J. L. Richardson and M. J. Vepraskas, editors. Wetland Soils: Genesis, Hydrology, Landscapes, and Classification. Lewis Publishers.

Cornell, J. A., C. Craft, and P. Megonigal. 2006. Carbon cycle succession in created salt marshes. Wetlands (in press).

Cowardin, L. M., V. Carter, F. C. Golet, and E. T. Laroe. 1979. Classification of wetlands and deepwater habitats. U.S. Department of the Interior, Fish and Wildlife Service, FWS/OBS-79/31, Washington, D.C.

Craft, C. B. 2001. Biology of wetland soils. Pages 107-135 in J. L. Richardson and M. J. Vepraskas, editors. Wetland Soils: Genesis, Hydrology, Landscapes, and Classification. Lewis Publishers.

DeBerry, D. A., and J. E. Perry. 2004. Primary succession in a created freshwater wetland. Castanea 69:185-193.

Page 116: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

115

Frolking, S. E., J. L. Buiger, T. R. Moore, T. Ball, L. M. Bellisario, A. Bhardwj, P. Carrol, P. M. Crill, P. M. Lafleur, J. H. McCaughey, N. T. Roulet, A. E. Suyker, S. B. Verma, J. M. Waddington, and G. J. Whiting. 1998. Relationship between ecosystem productivity and photosyntheitcally active radiation for northern peatlands. Global Biogeochemical Cycles 12:115-126.

Hirota, M., Y. Tang, Q. Hu, S. Hirata, T. Kato, W. Mo, G. Cao, and S. Mariko. 2006. Carbon dioxide dynamics and controls in a deep-water wetland on the Qinhgai-Tibetan planteau. Ecosystems 9:673-688.

Lafleur, P. M., J. H. McCaughey, D. W. Joiner, P. A. Bartlett, and D. E. Jelinski. 1997. Seasonal trends in energy, water, and carbon dioxide fluxes at a northern boreal wetland. Journal of Geophysical Research 102:29009-29020.

McKinstry, M. C., and S. H. Anderson. 2003. Improving aquatic plant growth using propagules and topsoil in created bentonite wetlands of Wyoming. Ecological Engineering 21:175-189.

Minitab. 2005. Minitab® statistical software, release 14. State College, PA.

Neubauer, S. C., W. D. Miller, and I. C. Anderson. 2000. Carbon cycling in a tidal freshwater marsh ecosystem: a carbon gas flux study. Marine Ecology Progress Series 199:13-30.

NOAA. 2006. DS3505 - Surface Data, Hourly Global 72029799999. National Climatic Data Center, NNDC Climate Data Online.

Odum, E. P. 1969. The strategy of ecosystem development. Science 164:262-270.

Roggero, M. M. 2003. Ecosystem gas exchange in natural and created tidal salt marshes of Tidewater, Virginia. Dissertation/Thesis. The College of William and Mary.

Ryan, M. G. 1991. Effects of climate change on plant respiration. Ecological Applications 1:157-167.

Schmidt, J. M. 2002. Litter Decomposition in Created and Adjacent Forested Wetlands of the Coastal Plain of Virginia. Virginia Tech, Blacksburg, VA.

Schreader, C. P., W. R. Rouse, T. J. Griffis, L. D. Boudreau, and P. D. Blanken. 1998. Carbon dioxide fluxes in a northern fen during a hot, dry summer. Global Biogeochemical Cycles 12:729-740.

Shurpali, N. J., S. B. Verma, and J. Kim. 1995. Carbon dioxide exchange in a peatland ecosystem. Journal of Geophysical Research 100:14319-14326.

Smith, R. L. 1996. Ecology and Field Biology, 5th edition. HarperCollins College Publishers, New York, NY.

Page 117: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

116

Stauffer, A. L., and R. P. Brooks. 1997. Plant and soil responses to salvaged marsh surface and organic matter amendments at a created wetland in central Pennsylvania. Wetlands 17:90-105.

Stolt, M. H., M. H. Genthner, W. L. Daniels, V. A. Groover, S. Nagle, and K. C. Haering. 2000. Comparison of soil and other environmental conditions in constructed and adjacent palustrine reference wetlands. Wetlands 20:671-683.

Streever, W. J., A. J. Genders, and M. A. Cole. 1998. A closed chamber CO2 flux method for estimating marsh productivity. Aquatic Botany 62:33-44.

Suyker, A. E., S. B. Verma, and T. J. Arkebauer. 1997. Season-long measurement of carbon dioxide exchange in a boreal fen. Journal of Geophysical Research 102:29021-29028.

Systat. 2004. SigmaPlot ® Software for Windows version 9.01. Systat Software, Inc, Point Richmond, CA.

Trumbore, S. 2000. Age of soil organic matter and soil respiration: radiocarbon constraints on belowground C dynamics. Ecological Applications 10:399-411.

Whigham, D. F. 1999. Ecological issues related to wetland preservation, restoration, creation and assessment. The Science of the Total Environment 240:31-40.

Whittecar, G. R., and W. L. Daniels. 1999. Use of hydrogeomorphic concepts to design created wetlands in southeastern Virginia. Geomorphology 31:355-371.

Wickland, K. P., R. G. Striegl, M. A. Mast, and D. W. Chow. 2001. Carbon gas exchange at a southern Rocky Mountain wetland, 1996-1998. Global Biogeochemical Cycles 15:321-335.

Zedler, J. 1997. Restoring tidal wetlands: a scientific view. National Wetlands Newsletter 19:8-11.

Page 118: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

117

Table 1. Summary of mean (±1 SE) annual variation over loading rates for gross primary production (GPP), respiration (R), net ecosystem exchange (NEE = GPP - R), and the photosynthesis to respiration ratio (GPP:R) from June 2005 to May 2006 in CCW, Charles City County, VA. --------------------------Loading Rates (LR) -------------------------- CO2 Flux Variable LR1 LR2 LR3 LR4 LR5 n

GPP (g C m-2 yr-1) 937.5 ± 108.2 1014.5 ± 133.4 1150.0± 49.3 1192.2 ± 162.8 1452.2 ± 95.8 4 R (g C m-2 yr-1) 796.1 ± 170.6 984.7 ± 159.0 1333.9 ± 195.7 1493.2 ± 134.7 1911.6 ± 182.1 4 NEE (g C m-2 yr-1) 141.4 ± 66.5 29.9 ± 66.1 -184.0 ± 178.0 -301.0 ± 200.0 -459.4 ± 240.2 4 GPP:R 1.25 ± 0.1 1.06 ± 0.1 0.92 ± 0.1 0.81 ± 0.1 0.78 ± 0.1 4

Page 119: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

118

Figure 1. Schematic of carbon flux sampling equipment. Air is pulled from the chamber into the IRGA, where CO2 content is measured. The air is then returned to the chamber. A photosynthetically active radiation (PAR) meter and air temperature sensor also gather information.

Temperature Sensor

Fans

PAR Meter

Base

IRGA

Air to IRGA

Air from IRGA

Li-Cor 6200

Page 120: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

119

Amendment

LR 1 LR 2 LR 3 LR 4 LR 5

CO

2 Flu

x (g

C m

-2 y

r-1)

-2000

-1000

0

1000

2000GPP Respiration NEE

Figure 2. Annual gross primary production (GPP), respiration (R) and net ecosystem exchange (NEE) for each organic matter amendment treatment at the Charles City Mitigation site (CCW), Charles City Co. Virginia. Error bars represent ± one Standard Error (n = 4). LR 1 = 0 Mg ha-1, LR 2 = 56 Mg ha-1, LR 3 = 112 Mg ha-1, LR 4 = 224 Mg ha-1, LR 5 = 336 Mg ha-1.

Page 121: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

120

Loading Rate (Mg ha-1)

0 100 200 300

CO

2 Flu

x (g

C m

-2 y

r-1)

-2000

-1000

0

1000

2000

GPPRespiration NEE

Figure 3. Linear regressions of mean (±1 SE) annual gross primary production (GPP), respiration (R) and net ecosystem exchange (NEE) vs. organic matter loading rate at the Charles City Mitigation site (CCW), Charles City Co. Virginia. R is presented as negative C flux, relative to gains and losses from the ecosystem. Nee represents the difference between GPP and R (NEE = GPP – R). All p-values reported represent significance at the α = 0.05 level.

p = 0.002, R2 = 0.41, n = 4

p = <0.001, R2 = 0.62, n = 4

p = 0.006, R2 =0.35, n = 4

Page 122: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

121

Jun 0

5

Jul 0

5

Aug 05

Sep 05

Oct 05

Nov 05

Dec 05

Jan 0

6

Feb 06

Mar 06

Apr 06

May 06

Wat

er T

able

(msl

)

10.27

10.29

10.31

10.33

10.35

10.37

10.39

10.41

10.43

10.45R

(g C

m-2

mon

th-1

)

0

100

200

300

400

LR1 LR2 LR3 LR4 LR5

Figure 4. Monthly respiration (R) for each Loading Rate from June 2005 to May 2006 at the Charles City Mitigation site (CCW), Charles City Co. Virginia. The heavy blue line represents water table elevation above sea level in meters (msl).

Water Table

Page 123: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

122

Jun 0

5

Jul 0

5

Aug 05

Sep 05

Oct 05

Nov 05

Dec 05

Jan 0

6

Feb 06

Mar 06

Apr 06

May 06

Wat

er T

able

(msl

)

10.27

10.29

10.31

10.33

10.35

10.37

10.39

10.41

10.43

10.45G

PP (g

C m

-2 m

onth

-1)

0

100

200

300

400LR1 LR2 LR3 LR4 LR5

Figure 5. Monthly gross primary production (GPP) for each Loading Rate from June 2005 to May 2006 at the Charles City Mitigation site (CCW), Charles City Co. Virginia. The heavy blue line represents water table elevation above sea level in meters (msl). GPP data were not collected from December 2005 to January 2006 due to very low photosynthetic rates.

Water Table

Page 124: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

123

Jun 0

5

Jul 0

5

Aug 05

Sep 05

Oct 05

Nov 05

Dec 05

Jan 0

6

Feb 06

Mar 06

Apr 06

May 06

Wat

er T

able

(msl

)

10.27

10.29

10.31

10.33

10.35

10.37

10.39

10.41

10.43

10.45N

EE (g

C m

-2 m

onth

-1)

-100

-50

0

50

100

150LR1 LR2 LR3 LR4 LR5

Figure 6. Monthly net ecosystem exchange (NEE) for each Loading Rate from June 2005 to May 2006 at the Charles City Mitigation site (CCW), Charles City Co. Virginia. The heavy horizontal line represents a CO2 flux of 0 g C m-2 month-1. The heavy blue line represents water table elevation above sea level in meters (msl). Positive NEE (> 0) indicate a net autotrophic system (i.e. carbon uptake) while negative NEE (< 0) indicates a net heterotrophic system (i.e. carbon release).

Water Table

Page 125: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

124

Water Table Depth (cm)

-20 -10 0 10

R (g

C m

-2 m

onth

-1)

0

100

200

300

400

500

Figure 7. Linear regression of respiration (R) CO2 flux vs. water table height relative to plot soil surface in CCW, Charles City County, VA, for June-September 2005. Positive water table depths are above the soil surface (0 cm). The p-value reported indicates significance of the linear regression at the α = 0.05 level.

p < 0.001 R2 = 0.60

Page 126: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

125

Loading Rate (Mg ha-1)

0 100 200 300

GPP

:R

0.4

0.6

0.8

1.0

1.2

1.4

1.6

Figure 8. Annual photosynthesis (GPP) to respiration (R) ratio vs. Loading Rate in CCW, Charles City County, VA, from June 2005 to May 2006. The dashed line represents a GPP:R of 1:1. Error bars represent ± one Standard Error (n = 4).

Page 127: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

126

Chapter 4

Summary and Conclusions

Chapters 2 and 3, though focusing on different suites of parameters, combine to

provide a reasonably complete picture of several created wetland ecosystem functions at

the Charles City Wetland Mitigation Site, and how those functions were effected by and

varied among organic matter amendment loadings. Plant communities, regardless of

loading rate were similar; there were few dominant and co-dominant species relative to

the overall site species richness. Standing crop biomass was similar among loading rates,

while both Gross Primary Production and tree size both varied positively (although

weakly) with loading rate. Respiration data showed a positive correlation with loading

rate, as modeled by air temperature. However, water table depth also appeared to play a

critical role moderating summer respiration rates. Growing season water table depths

(using plot elevation as a proxy) also seemed to explain trends in biomass and tree size,

suggesting that hydrology may be an important determinant for these parameters.

The overall similarity of plant community composition and standing crop among

loading rates, as well as the narrow range in gross primary production all seem to suggest

that soil amendments may not provide the predominant environmental forcing or may

exert conflicting forcings on these factors. However, significant trends in respiration and

NEE with loading rate support the assertion that organic matter does control ecosystem

energy exchange to some extent. That is, adding organic matter to an early successional

wetland system can change the carbon balance from one focused on production and

carbon sequestering to one dominated by respiration and carbon export. In either case,

Page 128: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

127

ecosystems not in production-respiration balance will approach that balance as succession

proceeds.

Both the vegetation and the CO2 flux analyses show that adding higher end (i.e.

LRs 4 and 5) loadings of organic matter amendments to created wetland soils is

unsubstantiated due to the lack of differences in the vegetation and the negative NEE

results. In fact, the negative NEEs indicate that excess organic matter is being oxidized

in these treatments because environmental conditions cannot support it. Thus, from a

management perspective, adding such high volumes of material may not be an efficient

use of resources, and could in fact add needless cost to wetland construction. The

combined results from Chapters 2 and 3 thus suggest an amendment loading rate between

LR 2 (56 Mg ha-1) and 3 (112 Mg ha-1), based on the recommendations in Chapter 2

tempered by the apparent carbon balance noted in Chapter 3.

Page 129: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

128

Appendices

------------Yard-waste Compost Samples------------ Analysis Units 1 2 3 4 5 Solids % 56 60 50 53 53 TKN % 0.85 0.84 0.89 0.80 0.92 P % 0.09 0.07 0.08 0.08 0.08 K % 0.40 0.36 0.36 0.41 0.38 S % 0.12 0.11 0.12 0.11 0.11 Ca % 1.36 1.18 1.27 1.21 1.22 Mg % 0.20 0.18 0.18 0.20 0.20 Fe mg kg-1 6250 6490 5950 5880 5820 Al mg kg-1 7230 6880 7560 7350 6770 Mn mg kg-1 407 360 366 395 398 Cu mg kg-1 38 33 51 29 28 Zn mg kg-1 111 96 100 92 90 Organic N mg kg-1 8400 7200 8700 7800 9100 Organic C mg kg-1 369850 306800 381104 310459 426627 C/N 44 43 44 40 47 EC mS cm-1 0.56 0.66 0.64 1.20 0.74

Appendix 1. Results of analysis of organic matter amendments performed by A and L Laboratories, Inc. Samples collected on July 9, 2002 (from Bergschneider 2005)

Page 130: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

129

Appendix 2. Table of chemical parameters for surface soil (top 10 cm) samples collected on August 22, 2005 in CCW, Charles City County, VA. Data are separated by plot (i.e. treatment replicate), with four replicates per organic matter amendment loading rate (LR). Soils were analyzed and data were provided by the Virginia Cooperative Extension Soil Testing Laboratory, Virginia Tech University, Blacksburg, VA.

(Mg ha-1) ----------------- ppm in soil (mg kg-1) ------------------- Mg 100g-1 -------------------------% ------------------------- Plot # LR pH BpH P K Ca Mg Zn Mn Cu Fe B Est. CEC Acidity Base Sat Ca Sat Mg Sat K Sat

1 112 6.46 6.18 6 74 1874 450 6.1 41.0 5.5 73.9 0.6 14.5 9.0 91.0 64.3 25.5 1.3 2 336 6.19 6.18 17 105 2886 458 12.9 44.0 1.0 29.2 1.0 19.7 6.6 93.4 72.9 19.1 1.4 3 0 5.98 6.13 3 36 774 240 3.6 38.6 12.7 116.9 0.3 7.5 21.3 78.7 51.3 26.2 1.2 4 56 6.40 6.28 4 52 1374 394 4.5 52.7 8.3 93.6 0.5 10.9 6.5 93.5 62.6 29.6 1.2 5 224 6.18 5.95 21 72 3169 346 19.5 36.3 0.7 20.3 0.8 21.5 12.4 87.6 73.5 13.2 0.9 6 224 6.45 6.22 15 111 2546 424 12.0 42.4 1.9 33.4 0.8 17.5 6.1 93.9 72.4 19.9 1.6 7 56 6.31 6.20 4 58 1227 251 4.3 54.9 7.3 68.7 0.5 9.5 12.5 87.5 64.3 21.7 1.5 8 112 6.29 6.11 6 69 1884 347 7.7 50.6 5.0 73.4 0.5 14.2 12.2 87.8 66.4 20.2 1.2 9 0 5.85 6.07 2 28 634 299 2.1 34.2 6.1 41.2 0.2 7.6 25.6 74.4 41.3 32.1 0.9 10 336 5.98 5.95 21 48 3132 336 19.7 40.9 0.4 20.2 0.7 21.2 12.6 87.4 73.8 13.1 0.6 11 336 6.26 6.07 24 45 3226 414 17.4 30.5 0.3 15.1 1.2 21.6 9.1 90.9 74.6 15.8 0.5 12 56 6.34 6.26 10 66 1160 334 6.9 65.6 20.6 116.5 0.4 9.5 8.7 91.3 60.7 28.8 1.8 13 224 6.64 6.19 24 120 3145 504 16.1 46.7 0.8 18.5 1.0 21.4 5.8 94.2 73.4 19.4 1.4 14 0 6.83 6.45 3 31 882 272 3.2 51.0 8.9 51.2 0.3 6.8 0.7 99.3 65.0 33.1 1.2 15 112 6.59 6.28 6 36 1478 336 5.5 59.9 5.2 49.3 0.5 10.9 6.5 93.5 67.4 25.3 0.8 16 0 7.01 N/A 3 16 900 295 4.2 57.4 13.8 42.2 0.4 7.0 N/A 100.0 64.5 34.9 0.6 17 224 6.44 6.17 13 64 2616 407 11.8 57.6 1.2 27.9 0.7 17.9 7.6 92.4 72.8 18.7 0.9 18 112 5.96 6.01 15 81 2011 259 10.9 47.0 2.1 54.4 0.5 14.7 15.8 84.2 68.3 14.5 1.4 19 56 6.31 6.25 7 38 1250 243 5.7 52.8 6.2 65.6 0.5 9.2 9.7 90.3 67.6 21.7 1.1 20 336 6.64 6.26 9 130 1949 435 6.5 45.9 4.2 39.1 0.7 14.5 5.7 94.3 67.2 24.8 2.3

Page 131: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

130

Jun 0

5

Jul 0

5

Aug 05

Sep 05

Oct 05

Nov 05

Dec 05

Jan 0

6

Feb 06

Mar 06

Apr 06

May 06

PAR

( m

ol m

-2 m

onth

-1)

1.0e+6

1.5e+6

2.0e+6

2.5e+6

Appendix 3. Total monthly PAR from June 2005 to May 2006 measured at the Virginia Institute of Marine Science, Gloucester Point, Virginia.

µ

Page 132: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

131

Jun 0

5

Jul 0

5

Aug 05

Sep 05

Oct 05

Nov 05

Dec 05

Jan 0

6

Feb 06

Mar 06

Apr 06

May 06

Tem

pera

ture

(°C

)

0

5

10

15

20

25

30

Appendix 4. Mean monthly air temperature from June 2005 to May 2006 measured at the NOAA National Climatic Data Center station, Williamsburg, Virginia (76˚71’70” W, 37˚23’30” N).

Page 133: Wetland Vegetation Dynamics and Ecosystem Gas …web.vims.edu/library/Theses/Bailey06.pdfIn wetlands, this is due to the fact that soil-forming processes such as weathering, incorporation

132

* CH4 data were collected in the same static chambers used for CO2 flux measurements. The chamber (either 1 or 1.5 m tall, depending on height of vegetation

was secured to the chamber base using clamps and covered with an opaque blanket to shade out light and reduce temperature fluctuations during measurements. For each plot, 10 ml of gas were withdrawn from a septum in the wall of the chamber at 0, 5, 10, 20, and 40 minutes. Samples were collected during the same weeks that CO2 flux measurements were performed (usually the first week of the month), from July 2005 to June 2006. Rubber bands were placed around the plunger and stopcock of each syringe to insure that leakage would not result in a change in CH4 concentration before analysis. Gas samples were analyzed within five days of collection using a Hewlett-Packard® model 5890 gas chromatograph, equipped with a 2.6 mL sampling loop and Flame Ignition Detector (FID) with Molecular Sieve 13x. The instrument was calibrated routinely before and during analysis using 1.00, 9.07, and 100.00 ppmv CH4 in N2 standards (Scott Specialty Gasses, Inc.). CH4 concentrations were regressed over the 40 minute sampling period to calculate a slope, i.e. the change in CH4 concentration per minute.

Appendix 5. Table of seasonal instantaneous (i.e. per minute) methane (CH4) fluxes collected from July 2005 to June 2006 in CCW, Charles City County, VA. Data are separated by plot (i.e. treatment replicate), with four replicates per organic matter amendment loading rate (LR). CH4 collection and analysis methods are described below the table*.

Mg ha-1 ----------------------------------------------------------------------------------------CH4 Flux (mg C m-2 min-1) -------------------------------------------------------------------------------------- Plot # LR July-05 Aug-05 Sept-05 Oct-05 Nov-05 Dec-05 Jan-06 Feb-06 Mar-06 Apr-06 May-06 June-06

1 112 -0.006 0.009 0.067 -0.005 -0.007 -0.002 N/A 0.001 0.003 -0.005 0.012 0.000 2 336 0.010 0.000 0.074 0.003 -0.005 0.000 N/A -0.001 0.002 0.000 0.004 0.000 3 0 -0.021 0.006 0.093 -0.001 0.000 0.003 N/A 0.005 0.004 0.001 0.000 0.000 4 56 0.034 0.009 0.056 -0.012 0.002 0.000 N/A 0.003 0.004 -0.001 0.001 -0.001 5 224 0.005 0.002 0.171 -0.003 0.003 -0.204 N/A -0.005 0.001 0.001 0.000 -0.019 6 224 -0.005 -0.003 0.004 0.003 -0.004 0.005 N/A 0.000 -0.001 0.000 0.004 -0.008 7 56 -0.054 -0.006 0.071 0.002 0.000 -0.001 N/A 0.001 0.001 0.000 0.000 0.000 8 112 0.002 0.000 0.104 -0.002 -0.001 -0.001 N/A 0.000 0.000 0.001 0.006 0.003 9 0 0.006 0.003 0.034 -0.002 0.003 -0.002 N/A 0.002 0.001 -0.002 0.009 0.000 10 336 -0.051 -0.013 0.082 -0.008 -0.001 0.002 N/A 0.001 0.005 0.001 -0.001 -0.017 11 336 N/A 0.048 0.162 0.000 0.001 -0.048 N/A -0.003 0.003 0.001 0.002 0.006 12 56 N/A 0.003 0.097 -0.001 0.000 0.003 N/A 0.001 0.002 -0.001 0.017 -0.001 13 224 N/A -0.002 0.009 -0.005 0.007 0.001 N/A 0.001 0.002 0.000 -0.001 0.000 14 0 N/A -0.001 0.031 0.000 0.000 0.002 N/A 0.002 0.001 0.001 0.001 0.000 15 112 N/A 0.008 -0.007 0.000 -0.001 -0.008 N/A 0.000 0.001 0.002 -0.006 0.002 16 0 N/A -0.002 -0.010 0.000 0.002 0.001 N/A 0.000 0.001 0.005 0.008 0.000 17 224 N/A 0.003 0.049 -0.002 -0.004 0.002 N/A 0.002 0.002 -0.072 -0.001 0.006 18 112 N/A 0.005 0.013 0.005 0.003 -0.018 N/A 0.002 0.002 -0.001 0.007 0.002 19 56 N/A 0.010 0.011 -0.002 0.010 0.005 N/A 0.003 0.001 0.001 0.010 0.002 20 336 N/A 0.000 -0.001 -0.005 0.000 -0.001 N/A -0.001 0.000 0.001 -0.001 0.001


Recommended