&EPA
  United States
  Environmental Protection
  Agency
  Development Document for Proposed
  Effluent Guidelines and Standards
  for the Construction and Development
  Category

  June 2002

-------


-------
Development Document for Proposed Effluent Guidelines
  and Standards for the Construction and Development
                         Category

                         June 2002
              United States Environmental Protection Agency
                      Office of Water (4303T)
                   1200 Pennsylvania Avenue, NW
                      Washington, DC 20460
                   www.epa.gov/waterscience/guide/

                       [EPA-821-R-02-007]

-------
Acknowledgments and Disclaimer

The Construction and Development Effluent Guidelines proposed rule and support documents
were prepared by the C&D Project Team:  Eric Strassler, Project Manager; Jesse .Pritts, P.E.,
Engineer; George Denning, Economist; Karen Maher, Environmental Assessor; and Michael G.
Lee, Attorney. Technical support for this Development Document was provided by Tetra Tech,
Inc.

Neither the United States government nor any of its employees, contractors, subcontractors or
other employees makes any warranty, expressed or implied, or assumes any legal liability or
responsibility for any third party's use of, or the results of such use of, any information,
apparatus, product or process discussed in this report, or represents that its use by such a third
party would not infringe on privately owned rights. Mention of trade names or commercial
products does not constitute-endorsement by EPA or recommendation for use.

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

 Contents

 SECTION 1: SUMMARY AND SCOPE

 1.1     Introduction	1-1
 1.2     Summary and Scope of Proposal	1-1


 SECTION 2: BACKGROUND

 2.1     Legal Authority	_	2-1
 2.2     Clean Water Act ,	'..'..'	2-1
        2.2.1   Best Practicable Control Technology Currently Available	2-2
        2.2.2   Best Conventional Pollutant Control Technology		2-3
        2.2.3   Best Available Technology Economically Achievable	2-3
        2.2.4   New Source Performance Standards	2-3
        2.2.5   Pretreatment Standards for Existing Sources and Pretreatment Standards for New
               Sources	2-4
        2.2.6   Effluent Guidelines Schedule	2-4
        2.2.7 .  NPDES Phase I and n Storm Water Rules	2-5
 2.3     Pollution  Prevention Act of 1990  		 2-5
 2.4     State Regulations  ...:	2-6
 2.5     References .	2-6

 SECTION 3: DATA COLLECTION

 3.1     Introduction .	;	 3-1
 3.2     Literature Search	 3-1
 3.3     Compilation of State and Municipal Existing Control Strategies, Criteria,
        and Standards	   3-1
 3.4     Other Data Sources	3-5
        3.4.1    Phase II Storm Water Rule Economic Analysis	".. 3-5
        3.4.2   USDA National Resource Inventory ..	3-5
        3.4.3   National Storm Water BMP Database . .'.	3-5
        3.4.4   BMP Design Manuals and Guidance Documents Developed by Governmental and
               Other Organizations	3-6
 3.5     References	,..-.. 3-6

 SECTION 4: INDUSTRY PROFILE

4.1     Introduction	4-1
4.2     Industry Description	4,4
        4.2.1   Industry Definition and Classification of Subsectors by NAIC and SIC Codes	4-1
        4.2.2   Residential Building Construction Group	4-6
        4.2.3   Nonresidential Building Construction Group  	.'.. ,	4-13
June 2002

-------
Development Document for Construction and Development Proposed Effluent Guidelines	

       4.2.4   Heavy Construction Subsector	4-21
4.3    Industry Practices and Trends	4-23
       4.3.1    Overview of Construction Land-disturbing Activities	4-23
       4.3.2   Construction Site Size Categories and Estimates of Amount of Disturbed Land  .... 4-26
               4.3.2.1 National Estimates of Disturbed Acreage	4-26
               4.3.2.2 Distribution of Acreage by Project Type	• • 4-28
               4.3.2.3 Distribution of Developed Acreage by Project Size and Geography	4-34
4.4    References	4-39

SECTION 5: TECHNOLOGY ASSESSMENT

5.1    Construction Erosion and Sediment Controls	5-1
       5.1.1    Introduction	.5-1
       5.1.2   Procedure for Technology Assessment	•	5-2
               5.1.2.1 Identification of Performance Goals	5-2
               5.1-.2.2 Goals, Environmental Impact Areas, and Assessment Scales  	5-3
               5.1.2.3 Qualitative Versus Quantitative Assessment	5-5
       5.1.3   Review of Historical Approaches to Erosion and Sediment Control 	5-5
       5.1.4   Goals, Control Strategies, Criteria, and Standards	5-8
               5.1.4.1 Goals, Control Strategies, Criteria, And Standards: How They Relate	5-8
               5.1.4.2 Levels of Performance or "How Well Do The Strategies Work?"		5-10
               5.1.4.3 Strategies, Criteria, Standards, And Enforcement	,	5-10
5.1.5  Control Techniques, BMP Systems 	5-14
               5.1.5.1 Erosion Control and Prevention	5-14
                      5.1.5.1.1        Planning, Staging, Scheduling	•	5-14
                      5.1.5.1.2        Vegetative Stabilization	-.		5-17
                                      5.1.5.1.2.1     Grass-lined Channels  	5-21
                                      5.1.5.1.2.2     Seeding	.'	...5-24
                                      5.1.5.1.2.3     Sodding	5-27
                                      5.1.5.1.2.4     Mulching	5-29
                                      5.1.5.1.2.5     Geotextiles  	5-33
                                      5.1.5.1.2.6     Vegetated Buffer Strips ..	5-35
                                      5.1.5.1.2.7     Erosion Control Matting		5-37
                                      5.1.5.1.2.8     Topsoiling       	5-40
               5.1.5.2 Water Handling Practices ..:		5-42
                      5.1.5.2.1        Earth Dike	5-42
                      5.1.5.2.2        Temporary Swale ..	5-44
                      5.1.5.2.3        Temporary Storm Drain Diversion  	5-51
                      5.1.5.2.4        Pipe Slope Drain	5-53
                      5.1.5.2.5        Stone CheckDam	5-57
                      5.1.5.2.6        Lined Waterways	 5-59
               5.1.5.3 Sediment Trapping Devices	 5-63
                      5.1.5.3.1        SiltFence	:. .	5-64
                      5.1.5.3.2        Super Silt Fence	5-71
                      5.1.5.3.3        Straw Bale Dike  	5-73
                      5.1.5.3.4       Sediment Trap	5-76
June 2002

-------
Development Document for Construction and Development Proposed Effluent Guidelines	

                     5.1.5.3.5       Sediment Basin	,	5-81
              5.1.5.4 Other Control Practices	5-88
                     5.1.5.4.1       Stone Outlet Structure	5-88
                     5.1.5.4.2       Rock Outlet Protection	5-90
                     5.1.5.4.3       Sump Pit	-.	'. :	5-93
                     5.1.5.4.4       Sediment Tank	 5-95
                     5.1.5.4.5       Stabilized Construction Entrance	 5-96
                     5.1.5.4.6       Land Grading  	-....'	5-98
                     5.1.5.4.7       Temporary Access Waterway Crossing	5-100
                     .5.1.5.4.8       Dust Control	5-103
                     5.1.5.4.9       Storm Drain Inlet Protection  .	5-106
                     5.1.5.4.10      Polyacrylamide (PAM)	5-108
       5.1.6   Summary		• • ... 5-112
5.2    References	....,.:	 5-124

SECTION 6: REGULATORY DEVELOPMENT AND RATIONALE

6.1    Identification of Industry Impacts	,	6-1
       6.1.1   Pollutant Indicators	6-1
       6.1.2   Physical/habitat Indicators	6-4
6.2    Development of Regulatdry Options	 . .. .	6-5
6.3    Regulatory Options Developed for the Proposed Rule	6-7
       6.3.1   Option 1  Inspection and Certification^	,	6-7
       6.3.2   Option 2 Codify EPA CGP Requirements with Site Inspection and Certification
              Provisions	6-9
       6.3.3   Option 3 No Regulation		.'	6-15
6.4    References	•	6-15

SECTION 7: APPROACH TO ESTIMATING COSTS

7.1    Overview	7-1
7.2    Methods for Estimating Erosion and Sediment Control Costs	 ..	7-1
       7:2.1   Overview ..	7-1
       7.2.2   Erosion and Sediment Control Costs	7-4
7.3    Methods for Estimating Administrative Costs	7-13
       7.3.1   Overview	••	7-13
       7.3.2   Administrative Costs to Permittees	-	 7-13
       7.3.3   Administrative Costs for General Permit Revisions	7-14
7.4    References	>	• • •	• •	7-15


Appendices

A.     State Regulations on the Control of Construction Storm Water  	  A-l
B.     Supporting Cost Data	.,.	  B-l
June 2002
                                                                                           in

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	_^	


 Tables

 Table 3-1.      State or Regional Planning Authority Requirements for Water Quality Protection  ... 3-3
 Table 3-2.      Municipal or Regional Planning Authority Requirements  	: . 3-4
 Table 4-1.      1997 NAICS Subsectors, Industry Groups, and Industries Performing Construction
                Activities That Might Disturb Land	4-3
 Table 4-2.      ,1987 SIC Industry Groups Performing Construction Activities
                That May Disturb Land			4-5
 Table 4-3.      Annual Housing Construction Starts by Type and Region	4-7
 Table 4-4.      Residential Construction Industry Profile for 1997	4-9
 Table 4-5.      Busiest Markets  for Single-Family Housing Permits for 1999	.. 4-10
 Table 4-6.      Busiest Markets  for Multifamily Housing Permits for 1999	4-10
 Table 4-7.      Changes in Housing Starts by Region (1989 and 1999)  	4-12
 Table 4-8.      Value of Construction Work for Manufacturing and Industrial Building Construction
                Establishments With Payroll by Type of Construction	'. ... 4-15
 Table 4-9.      Value of Manufacturing  and Industrial Building Construction Work for
                Establishments With Payroll by Location of Construction Work, 1997	4-16
 Table 4-10.     Value of Construction Work for Commercial and .Institutional Building Construction
                Establishments With Payroll by Type of Construction, 1997		4-18
 Table 4-1,1.     Value of Commercial and Institutional Building Construction Work for
               Establishments With Payroll by Location of Construction Work, 1997  	4-20
 Table 4-12.    Overview of Heavy Construction Industry,  1992 and 1997	4-22
 Table 4-13.    Acres Converted from Undeveloped to Developed State, 1992-1997		4-27
 Table 4-14.    New Single-Family and Multifamily Housing Units Authorized, 1995-1997	4-28
 Table 4-15.    Average and Median Lot Size for New Single-Family
               Housing Units Sold, 1995-1997	4-29
 Table 4-16.    Average Building Square Footage	4-31
 Table 4-17.    Typical Building Sizes and Size Ranges by Type of Building	,	4-32
 Table 4-18.    National Estimates of Land Area Developed Per Year, Based on Building
               Permit Data	,	4.33
 Table 4-19.     National Estimates of Land Area Disturbed Based on National Resources Inventory
               Totals	.,.	 4-34
 Table 4-20.     Distribution of 14 Community Survey Permits by Site Size	4-36
 Table 4-21.     Distribution of National Construction by Site Size and Development Type	4-38
 Table 5-1.      Description of Levels of Performance of Three Control  Strategies	 5-10
 Table 5-2.      Descriptions of Levels of Difficulty in Enforcement  	5-11
 Table 5-3.      Scheduling Considerations for Construction Activities Enforcement	5-16
 Table 5-4.      Conditions Where Vegetative Streambank Stabilization Is Acceptable	5-20
 Table 5-5.      Maximum Permissible Velocities for Individual Site Conditions
               for Grass Swales	5-22
 Table 5-6.      Typical Mulching Materials and Application Rates 	5-31
 Table 5-7.      Measured Reductions in Soil Loss for Different Mulch Treatments 	5-32
 Table 5-8.      Cubic Yards of Topsoil Required for Application to Various Depths		5-41
Table 5-9.      Grassed Swale Pollutant Removal Efficiency Data	5-49
Table 5-10.     Average Annual Operation and Maintenance Costs for a Grass Swale  	5-5:1
June 2002
                                                                                            IV

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

 Table 5-11.    Recommended Pipe/Tubing Sizes for Slope Drains 	5-54
 Table 5-12.    Slope Drain Characteristics	5-55
 Table 5-13.    Maximum Slope Lengths for Silt Fences 	5-65
 Table 5-14.    Typical Requirements for Silt Fence Fabric  	5-68
• Table 5-15.    Slope Lengths for Super Silt Fences	5-71
 Table 5-16.    Minimum Requirements for Super Silt Fence Geotextile Class F Fabric	5-72
 Table 5-17,    Maximum Land Slope and Distances Above a Straw Bale Dam	5-74
 Table 5-18.    Weir Length for Sediment Traps	'	5-78
 Table 5-19.    Range of Measured.Long-Term Pollutant Removal for Sediment Detention Basins  . 5-79
 Table 5-20.    Common Concerns Associated with Sediment Traps	,..	5-80
 Table 5-21.    Common Concerns Associated with Sediment Basins	 5-86
 Table 5-22.    Riprap.Sizes and Thicknesses (SHA Specifications)  	...... 5-91
 Table 5-23.    Application Rates for Spray-On Adhesives	 5^105
 Table 5-24.    Turbidity Reduction Values from PAM	.'.. ..,. 5-110
 Table 5-25.    Summary of Information on Erosion Control and Prevention
               BMPs (Sub-section 5.1.5.1)	5-112
 Table 5-26.    Summary of Information on Erosion Control and Prevention
               BMPs (Sub-section 5.1.5.2)	•	5-116
 Table 5-27.    Summary of Information on Erosion Control and Prevention
               BMPs (Sub-section 5.1.5.3)	5-119
 Table 5-28.    Summary of Information on Erosion Control and Prevention
               BMPs (Sub-section 5.1.5.4) ..,	5-122
 Table 7-1.      Total Costs of Proposed Rule Options  	7-1
 Table 7-2.      Regional  compliance Cost Adjustment Factors	.7-4
 Table 7-3.      Construction Site ESC BMP Descriptions and Site Thresholds		7-5
 Table 7-4,      Components of Existing State Erosion and Sediment Control Requirements	7-7
 Table 7-5.      State Acreage Equivalent to Proposed Option 2  .	 7-9
 Table 7-6.      Construction ESC BMP Design and Operation and Maintenance Costs as a
               Percentage of Capital Costs	—	7-10
 Table 7-7.    .  Evaluated construction Site BMPs that Augment the Suite of
               Baseline BMPs  	7-11
 Table 7-8.      BMP Quantity. Adjustment Factors for Baseline and Proposed Options . .'.	.7-12
 Table 7-9.      National Cost Estimates for Proposed Rule Options	7-13
 Table 7-10.    One-Time Hours and Costs to Incorporate Erosion and Sediment Control
               Effluent Guidelines Requirements into General Permits	 7-14


 Figures

 Figure 4-1.     Annual Housing Starts	4-8
 Figure 4-2.     Bureau of Census Housing Regions	:..,....	..4-11
 Figure 4-3.     Annual Housing Starts by Region	4-13
 Figure 4-4.     Value of Heavy construction work by region, 1997 	:.......	. 4-23
 Figure 5-1.     Flow Diagram Showing Relationship Among Goals, Strategies, Criteria,
               and Standards	5-9
 Figure 7-1.     EPA Ecoregions	,	 7-3
June 2002

-------

-------
Development Document for Construction and Development Proposed Effluent Guidelines	

SECTION 1: OVERVIEW

1.1  INTRODUCTION

This document presents technical information to support the Agency's analyses and complements
"Economic Analysis of Proposed Effluent Guidelines and Standards for the Construction and
Development Category," EPA-821-R-02-008, and "Environmental Assessment for Proposed
Effluent Guidelines and Standards for the Construction and Development Category," EPA 821-
R-02-009.

A summary of the information contained in the chapters of this document is as follows:

     •  Chapter 2 presents background information on the legal authority for effluent limitation
       guidelines and the existing EPA storm water program.

     •  Chapter 3 presents a summary of the data collection activities conducted to support the
       proposal.

     •  Chapter 4 summarizes the characteristics of the construction and development industry,
      , including major indicators of industry size and annual construction activity.
                           >
     •  Chapter 5 presents information and data on erosion and sediment control (ESC) best
       management practices (BMPs) used by this industry, including applicability, costs, and
       efficiencies.    '

     •  Chapter 6 presents a description of the regulatory options considered by EPA for
       developing the proposal, as well as a walk-through of the provisions of each proposed
       option.

     •  Chapter 7 presents the methodology used by the Agency to estimate the costs of the
       proposed options.

1.2  SUMMARY AND SCOPE OF PROPOSAL

The proposed rule contains three options for controlling storm water discharges from
construction sites.

     •  Option  1 would establish inspection and certification provisions to ensure proper
       implementation of controls.  This option would apply to all construction sites disturbing
       one or more acres of land required to obtain a permit under the existing National
       Pollutant Discharge Elimination System (NPDES) storm water regulations. This option
June 2002
1-1

-------
 Development Document for Construction and Development Proposed Effluent Guidelines           	

        would amend the NPDES regulations at 40 CFR Part 122, but would not create effluent
        limitation guidelines.

      •  Option 2 would add minimum requirements for preparation of a Storm Water Pollution
        Prevention Plan (SWPPP) as well as minimum requirements for sizing sediment basins,
        installing erosion and sediment controls, providing temporary stabilization to exposed
        soils, and conducting regular inspections.  Option 2 would apply to all sites that disturb
        five or more acres of land, consistent with the permitting requirements of the Phase I
        NPDES storm water regulations. This option would create a new effluent guidelines
        category at 40 CFR Part 450 and would also modify 40 CFR Part 122.

      •  Option 3 would not establish any new requirements.

 EPA estimated that Option 1 would cost approximately $130 million annually, while preventing
 the annual discharge of approximately 5.25 million tons of Total Suspended Solids (TSS) and
 associated turbidity to surface waters.  The estimated annual monetized benefits of this option are
 $10.4 million. Option 2 is estimated to cost approximately $505 million annually, while
 preventing the discharge of approximately 11.1 million tons of TSS and associated turbidity to
 surface waters annually. The estimated annual monetized benefits of Option 2 are $22.0 million.
 Option 3 is not expected to have any costs or benefits.
June 2002
                                                                                     1-2

-------
Development Document for Construction and Development Proposed Effluent Guidelines	

SECTION 2: BACKGROUND

2.1    LEGAL AUTHORITY

The Environmental Protection Agency (EPA) is proposing Effluent Limitation Guidelines for
discharges associated with construction and development activities under the authority of
Sections 301, 304, 306, 308, 402, and 501 of the Clean Water Act (CWA) (the Federal Water
Pollution Control Act), 33 United States Code (U.S.C.) 1311, 1314,  1316, 1318, 1342, and 1361.
This section describes EPA's legal authority for issuing the regulation, existing state regulations,
and other federal regulations associated with construction and development activities.

2.2    CLEAN WATER ACT

Congress adopted the-Clean Water Act (CWA) to "restore and maintain the chemical, physical,
and biological integrity of the nation's waters" (Section 101 (a), 33 U.S.C. 1251(a)). To achieve
this goal, the CWA prohibits the discharge of pollutants into navigable waters except in
compliance with the statute. CWA sec. 402 requires "point source" discharges to obtain a permit
.under the National Pollutant Discharge Elimination System (NPDES). These permits are issued
by EPA regional offices or authorized State agencies.

Following enactment of the Federal Water Pollution Control Amendments of 1972 (Pub.L. 92-
500, October 18, 1972), EPA and the States issued NPDES permits to thousands of dischargers,
both industrial (e.g. manufacturing, energy and mining facilities) and municipal (sewage.
treatment plants).  As required under Title HI of the Act, EPA promulgated effluent limitation
guidelines and standards for many industrial categories, and these requirements are incorporated
into the permits.

The Water Quality Act of 1987 (Pub.L. 100-4, February 4, 1987) amended the CWA. The
NPDES program was expanded by defining municipal and industrial storm water discharges as
point sources.  Industrial storm water dischargers, municipal separate storm sewer systems and
other storm water dischargers designated by EPA must obtain NPDES permits pursuant to
Section 402(p) (33 U.S.C. 1342(p)).
June 2002
2-1

-------
Development Document for Construction and Development Proposed Effluent Guidelines
2.2.1
BEST PRACTICABLE CONTROL TECHNOLOGY CURRENTLY
AVAILABLE
In guidelines for a point source category, EPA may define BPT effluent limits for conventional,
toxic,1 and non-conventional pollutants.  In specifying BPT, EPA looks at a number of factors.
EPA first considers the cost of achieving effluent reductions in relation to the effluent reduction  .
benefits. The Agency also considers the age of the equipment and facilities, the processes
employed and any required process changes, engineering aspects of the control technologies,
non-water quality environmental impacts (including energy requirements), and such other factors
as the Agency deems appropriate (CWA sec. 304(b)(l)(B)).  Traditionally, EPA establishes BPT
effluent limitations based on the average of the best performance of facilities within the category
of various ages, sizes, processes or other common characteristics. Where existing performance is
uniformly inadequate, EPA may require higher levels of control than currently in place in a
category if the Agency determines that the technology can be practically applied. (US Senate,
1973, p.  1468).

In addition, the Act requires a cost-reasonableness assessment for BPT limitations,  hi
determining the BPT limits, EPA considers the total cost of treatment technologies  in relation to
the effluent reduction benefits achieved.  This inquiry does not limit EPA's broad discretion to
adopt BPT limitations that are achievable with available technology unless the required
additional reductions are "wholly out of proportion to the costs of achieving such marginal level
of reduction." (US Senate, 1973,  p. 170)  Moreover, the inquiry does not require the Agency to
quantify benefits in monetary terms. See, for example, American Iron and Steel Institute v. EPA,
526 F. 2d 1027 (3rd Cir., 1975).

In balancing costs against the benefits of effluent reduction, EPA considers the volume and
nature of expected discharges after application of BPT, the general environmental effects of
pollutants, and the cost and economic impacts of the required level of pollution control, hi past
effluent limitation guidelines and standards, BPT cost-reasonableness removal figures have
ranged from $0.21 to $33.71 per pound removed in year 2000 dollars,  hi developing guidelines,
the Act does not require consideration of water quality problems attributable to particular point
sources, or water quality improvements in particular bodies of water. Accordingly, EPA has not
considered these factors in developing the limitations being proposed today. See Weyerhaeuser
Company v. Costle, 590 F. 2d 1011 (B.C. Cir. 1978).
   1 In the initial stages of EPA CWA regulation, EPA efforts emphasized the achievement of BPT limitations for
control of the "classical" pollutants (e.g., TSS, pH, BOD5). However, nothing on the face of the statute explicitly
restricted BPT limitation to such pollutants. Following passage of the Clean Water Act of 1977 (Pub.L. 95-217,
December 27,1977) with its requirement for point sources to achieve best available technology limitations to control
discharges of toxic pollutants, EPA shifted its focus to developing BAT limitations for the listed priority toxic
pollutants.
June 2002
                                                                                       2-2

-------
Development Document for Construction and Development Proposed Effluent Guidelines
2.2.2
BEST CONVENTIONAL POLLUTANT CONTROL TECHNOLOGY
The 1977 amendments to the CWA required EPA to identify effluent reduction levels for
conventional pollutants associated with BCT technology for discharges from existing point
sources.  BCT is not an additional limitation, but replaces Best Available Technology (BAT) for
control of conventional pollutants. In addition to other factors specified in sec. 304(b)(4)(B), the
CWA requires that EPA establish BCT limitations after consideration of a two- part "cost-
reasonableness" test.  EPA explained its methodology for the development of BCT limitations in
July 1986 (51FR 24974).

Section 304(a)(4) designates the following as conventional pollutants: biochemical oxygen
demand (BOD5), total suspended solids (TSS), fecal coliform, pH, and any additional pollutants
defined by the Administrator as conventional. The Administrator designated oil and grease as  an
additional conventional pollutant on July 30, 1979 (44 FR 44501). A primary pollutant of
concern at construction sites, sediment, is measured as TSS.
2.2.3
BEST AVAILABLE TECHNOLOGY ECONOMICALLY ACHIEVABLE
In general, BAT effluent guidelines (CWA sec. 304(b)(2)) represent the best existing
economically achievable performance of direct discharging plants in the subcategory or category.
The factors considered in assessing BAT include the cost of achieving BAT effluent reductions,
the age of equipment and facilities involved, the processes employed,  engineering aspects of the
control technology, potential process changes, non-water quality environmental impacts
(including energy requirements), and such factors asjhe Administrator deems appropriate. The
Agency retains considerable discretion in assigning the weight to be accorded to these factors.
An additional statutory factor considered in setting BAT is "economic achievability." Generally,
EPA determines the economic achievability on the basis of the total cost to the subcategory and
the overall effect of the rule on the industry's financial health. The Agency may base BAT
limitations upon effluent reductions attainable through changes in a facility's processes and
operations. As with BPT, where existing performance is uniformly inadequate, EPA may base
BAT upon technology transferred from a different subcategory or from another category.  In
addition, the Agency may base BAT upon manufacturing process changes or internal controls,
even when these technologies are not common industry practice.
2.2.4
NEW SOURCE PERFORMANCE STANDARDS
New Source Performance Standards (NSPS) reflect effluent reductions that are achievable based
on the best available demonstrated control technology.  New facilities have the opportunity to
install the best and most efficient production processes  and wastewater treatment technologies.
As a result, NSPS should represent the greatest degree of effluent reduction attainable through
the application of the best available demonstrated control technology for all pollutants (i.e.,
conventional, non-conventional, and priority pollutants). In establishing NSPS, CWA sec. 306
June 2002
                                                                       2-3

-------
Development Document for Construction and Development Proposed Effluent Guidelines	

directs EPA to take into consideration the cost of achieving the effluent reduction and any non-
water quality environmental impacts and energy requirements.
2.2.5
PRETREATMENT STANDARDS FOR EXISTING SOURCES AND
PRETREATMENT STANDARDS FOR NEW SOURCES
The CWA also defines standards for indirect discharges, i.e. discharges into publicly owned
treatment works (POTWs). These are Pretreatment Standards for Existing Sources (PSES) and
Pretreatment Standards for New Sources (PSNS) under sec. 307(b). Because EPA has identified
no deliberate discharges directly to POTWs, EPA is not proposing PSES or PSNS for the
Construction and Development Category. The information reviewed by the Agency indicates
that the Vast majority of construction sites discharge either directly to waters of the U.S. or
through MS4s. In some urban areas, construction sites discharge to combined sewer systems
(i.e., sewers carrying both storm water and domestic sewage through a single pipe) which lead to
POTWs.  Sediment is susceptible to treatment in POTWs, using technologies commonly
employed such as primary clarification, and EPA has no evidence of interference, pollutant pass-
through or sludge contamination.
2.2.6
EFFLUENT GUIDELINES SCHEDULE
Clean Water Act section 304(m) requires EPA to publish a plan every two years that consists of
three elements. First, under sec. 304(m)(l)(A), EPA is required to establish a schedule for the
annual review and revision of existing effluent guidelines in accordance with sec. 304(b).
Section 304(b) applies to ELGs for direct dischargers and requires EPA to revise such regulations
as appropriate. Second, under sec. 304(m)(l)(B), EPA must identify categories of sources
discharging toxic or nonconventional pollutants for which EPA has not published BAT ELGs
under sec. 304(b)(2) or new source performance standards under sec. 306. Finally, under sec.
304(m)(l)(C), EPA must establish a schedule for the promulgation of BAT and NSPS for the
categories identified under subparagraph (B) not later than three years after being identified in the
304(m) plan. Section 304(m) does not apply to pretreatment standards for indirect dischargers,
which EPA promulgates pursuant to sec. 307(b) and 307(c) of the Act.

On October 30, 1989, Natural Resources Defense Council, Inc.  (NRDC), and Public Citizen,
Inc., filed an action against EPA in which they alleged, among other things, that EPA had failed
to comply with sec. 304(m).  Plaintiffs and EPA agreed to a settlement of that action in a consent
decree entered on January 3,1, 1992.  (Natural Resources Defense Council et al v. Whitman,
D.D.C. Civil Action No. 89-2980). The consent decree, which has been modified several times,
established a schedule by which EPA is to propose and take final action for eleven point source
categories identified by name in the decree and for eight other point source categories identified
only as new or revised rules, numbered 5 through 12. EPA selected the Construction and
Development category as the subject for New or Revised Rule #10. The decree, as modified,
calls for the Administrator to sign a proposed ELG for the C&D category  no later than May 15,
June 2002
                                                                     2-4

-------
Development Document for Construction and Development Proposed Effluent Guidelines	

2002, and to take final action on that proposal no later than March 31, 2004. A settlement
agreement between the parties, signed on June 28, 2000, requires that EPA develop regulatory
options applicable to discharges from construction, development and redevelopment, covering
site sizes included in the Phase I and Phase IINPDES storm water rules (i.e. one acre or greater).
EPA is required to develop options including numeric effluent limitations for sedimentation and
turbidity; control of construction site pollutants other than sedimentation and turbidity (e.g.
discarded building materials, concrete truck washout, trash); BMPs for controlling post-
construction runoff; BMPs for construction sites; and requirements to design storm water
controls to maintain pre-development runoff conditions where practicable. The settlement also
requires EPA to issue guidance to MS4s and other permittees on maintenance of post-
construction BMPs identified in the proposed ELGs. Further discussion of approaches not
pursued by EPA at this time may be found in the docket for today's proposal.
2.2.7
NPDES PHASE I AND II STORM WATER RULES
The National Pollutant Discharge Elimination System (NPDES) is a permit system established
under the GWA to enforce effluent limitation. Operators of construction activities, including
clearing, grading and excavation are required to apply for permit coverage under the NPDES
Phase I and II storm water rules. Under the Phase I rule (promulgated in 1990), construction sites
of 5 or more acres must be covered by either a general or an individual permit. General permits
covering the Phase I sites have been issued by EPA regional offices and state water quality
agencies. Permittees are required to develop storm water pollution prevention plans that include
descriptions of BMPs employed, although actual BMP selection and design are at the discretion
of permittees (in conformance with applicable state or local requirements).

Construction sites between 1 and 5 acres in size are subject to the NPDES Phase n storm water
rule (promulgated in 1999). The construction activities covered under Phase n are termed small
construction activities and exclude routine maintenance that is performed to maintain the original
line and grade, hydraulic capacity, or original purpose of the facility. Under the Phase II program,
NPDES permit requirements for construction activities are similar to the Phase I requirements
because they will be covered under similar general permits.

2.3   POLLUTION PREVENTION ACT OF 1990

The Pollution Prevention Act of 1990. (PPA) (42 U.S.C. 13101 et seq., Pub. L. 101-508,
November 5, 1990) makes pollution prevention the national policy of the United States. The
PPA identifies an environmental management hierarchy in which pollution "should be prevented
or reduced whenever feasible; pollution that cannot be prevented should be recycled in an
environmentally safe manner, whenever feasible; pollution that cannot be prevented or recycled
should be treated in an environmentally safe manner whenever feasible; and disposal or release
into the  environment should be employed only as a last resort..." (42 U.S.C. 13103).  In short,
preventing pollution before it is created is preferable to trying to manage, treat or dispose of it
June 2002
                                                                                     2-5

-------
 Development Document for Construction and Development Proposed Effluent Guidelines            	

 after it is created.  According to the PPA, source reduction reduces the generation and release of
 hazardous substances, pollutants, wastes, contaminants or residuals at the source, usually within
 a process. The term source reduction "...includes equipment or technology modifications,
 process or procedure modifications, reformulation or redesign of products, substitution of raw
 materials, and improvements in housekeeping, maintenance, training, or inventory control. The
 term 'source reduction' does not include any practice which alters the physical, chemical, or
 biological characteristics or the volume of a hazardous substance, pollutant, or contaminant
 through a process or activity which itself is not integral to or necessary for the production of a
 product or the providing of a service." In effect, source reduction means reducing the amount of
 a pollutant that enters a waste stream or that is otherwise released into the environment prior to
 out-of-process recycling, treatment, or disposal.

 Although the PPA does not explicitly address storm water discharges or discharges from
 construction sites, the principles of the PPA are implicit in many of the practices used to reduce
 pollutant discharges from construction sites.  These include controls that minimize the potential
 for erosion such as stabilization of disturbed areas as soon as practicable.  These controls are
 described in section 5 of the Development Document.

 2.4    STATE REGULATIONS

 States and municipalities have been regulating discharges of runoff from construction and land
 development industry to varying degrees for some time.  A compilation of state and selected
 municipal regulatory approaches was  prepared to help establish the baseline for national and
 regional levels of control. Data were  collected by reviewing state and municipal web sites,
 summary references, state and municipal regulations and storm water guidance manuals. All
 states (and the selected municipalities) were contacted to confirm the data collected and to fill in
 data gaps, however, only  87 percent of the state agencies and a much smaller percentage of
 municipalities responded. The state and municipal regulatory data are summarized hi Section 3.3
 and the complete data sheets are included in Appendix A.

 2.5    REFERENCES

 US Senate, 1973. A Legislative History of the Federal Water Pollution Control Act Amendments
       of 1972.  U.S. Senate Committee of Public Works, Serial No. 93-1, January 1973.
       Washington, DC.
June 2002
                                                                                     2.6

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

 SECTION 3: DATA COLLECTION

 3.1    INTRODUCTION

 EPA gathered and evaluated technical and economic data from various sources in the course of
 developing the effluent limitation guidelines and standards for the construction and development
 industry. "EPA used existing data sources to profile the industry with respect to general industry
 description, industry trends, environmental impacts, and erosion and sediment control best
 management practices (BMPs) and cost. This chapter details the data sources used in the
 development of this proposal.

 3.2    LITERATURE SEARCH

 A literature search was performed to obtain information on various BMPs that pertain to the
 construction and land development industry.  Journal articles and professional conference
 proceedings were used to summarize the most recent BMP effectiveness data, design and
 installation criteria, applicability, advantages, limitations, and cost.

 3.3    COMPILATION OF STATE AND MUNICIPAL EXISTING CONTROL
       STRATEGIES, CRITERIA, AND STANDARDS

 A compilation of State and municipal regulations were prepared to determine national and
 regional approaches towards controlling construction site storm water. The data were collected
 by reviewing State and municipal web sites, summary references, and State and municipal
 regulations and storm water guidance manuals.  States and municipalities were contacted to
 confirm the data collected and to fill in data not available by these methods. Not all State and
 municipal contacts responded or were able to provide the missing information sought. While 87
 percent of the State agencies provided confirmation of the regulatory data collected for this study,
 a much smaller percentage of municipalities responded.

 A summary of criteria and standards that are implemented by States and municipalities as of
 August 2000 are presented in Tables 3-1 and 3-2, respectively. State requirements are generally
 equal to or less stringent than municipalities that are covered under the federal Clean Water Act
 NPDES Storm Water Program because State requirements apply to all development within their
 boundaries including single site development and low to high density developments. NPDES
 Storm Water Program designated municipalities generally have a population of 100,000 or more
 and can collect and fund the resources necessary to design, implement, and monitor separate and
potentially more stringent storm water management programs. Table 3-1 contains responses
 from 47 of the 54 State controlling agencies. The total is greater than 50 because Florida has 5
 intrastate regional authorities. Some State data were uncertain and repeated contacts to the   •
responsible State agencies to confirm the data were not returned.  For the same reason, some of
June 2002
                                                                                   3-1

-------
Development Document for Construction and Development Proposed Effluent Guidelines	

the data sought from municipal agencies also are not available for this report. Tables 3-1 and 3-2
are summaries of the regulatory controls used by States and municipalities as presented on Table
A-l: State regulations on the control of construction phase storm water.

Many data were not readily available. Appendix A presents Tables A-l which includes all of the
data that was collected.

The data collected reflect a cross section of the US geography but are representative primarily of
municipalities that have a population of 100,000 or greater and relatively few municipalities of
smaller population. Thirty-one municipalities are included in the summary tables, which is a
relatively small data set compared to the approximately 240 municipalities with NPDES
programs and nearly 3,000 municipalities nationwide.  Therefore, the data presented for the
States in Table 3-1 is fairly comprehensive while data for the municipalities presented in Table 3-
2 is not comprehensive but does reflect the diversity of management techniques used at the
municipal level.
June 2002
3-2

-------
 Development Document for Construction and Development Proposed Effluent Guidelines


               Table 3-1.  State or Regional Planning Authority Requirements
                                for Water Quality Protection
Standard
Solids or sediment
percent reduction
Numeric effluent limits
for TSS, settleable
solids, or turbidity
Numeric design depth or
volume for water quality
treatment
Habitat/biological
measures
Physical in-stream
condition controls
Water Quality or
Effluent Monitoring
Requirement
Number of
States with
Requirement9
• 11
2
22
3 •
8
3
Percent of
National
Developed
Acreage with
Requirement
24% '
11%
53%
7%
17%
6%
Percent of
National
Developed
Acreage
without
Requirement
61%
76%
28%
80%
70%
83%
Percent of
National
Developed
Acreage
without
Information
15%
13%
19%
13%
13%
11%
       a Florida has 5 Water Management Districts. If any of these Districts met a particular standard, the
       entire state annual developed acreage was counted.
June 2002
                                                                                            3-3

-------
Development Document for Construction and Development Proposed Effluent Guidelines
                 Table 3-2. Municipal Planning Authority Requirements
Standard
Design storm for peak
discharge control
Solids or sediment
percent reduction
Numeric design depth,
storm, or volume for
water quality treatment
Design storm for flood
control
Habitat/biological
measures
Physical in-stream
condition controls
Percent of
Municipalities
Reviewed with
Requirement
39%
7%
-
39%
3%
10%
Percent of
Municipalities
' Reviewed without
Requirement
45%
77%
-
16%
65%
58%
Percent of
Municipalities
without
Information
16%
16%
-
23%
32%
32%
        Note: This table reflects data collected from 31 municipalities

Tables 3-1 and 3-2 indicate that the following key control measures are being employed by States
and municipal/regional authorities to implement the NPDES Storm Water Program:

•  Storm water controls designed for peak discharge control
•  Storm water controls designed for water quality control
•  Storm water controls designed for flood control
•  Specified depths of runoff for water quality control
•  Percent reduction of loadings for water quality control (primarily solids and sediments)
•  Numeric effluent limits for water quality control (primarily total suspended solids, settleable
   solids, or turbidity)
•  Control measures for biological or habitat protection
•  Control measures for physical in-stream condition controls (primarily streambed and
   streambank erosion).

The water quantity control measures for peak discharge and runoff volume controls that apply to
the post-development conditions typically are  not applicable during the construction phase when
the site is disturbed.  Pollutant control measures are commonly required during the construction
phase, though the requirements for post-development storm water management are broader and
potentially more stringent.
June 2002
3-4

-------
Development Document for Construction and Development Proposed Effluent Guidelines	

3.4  OTHER DATA SOURCES

3.4.1   PHASE II STORM WATER RULE ECONOMIC ANALYSIS

The Economic Analysis of the Final Phase II Storm Water Rule (USEPA, 1999) estimated Phase
II Storm Water Rule compliance costs for two major categories of pollutant controls for
construction sites: erosion and sediment control BMPs and post-construction storm water
management controls. Total costs for implementing the Phase II Rule encompass expenditures
for installation of erosion and sediment control technologies, labor requirements for submitting a
Notice of Intent (NOI) to be covered by a general permit, a Notification to Municipalities, a
Storm Water Pollution Prevention Plan (SWPPP), and maintenance costs. Costs were derived on
a per-site basis and then aggregated to the State and national level based on the number of
building permits issued. As described in the Economic Analysis Report for the Phase II Rule,
census data were used to project the annual number of construction permits by Standard
Industrial Classification (SIC) Code and construction permit data from 14 municipalities were
used to categorize construction activities by site size.

3.4.2   1997 USDA NATIONAL RESOURCE INVENTORY

The 1997 National Resources Inventory (NRI) (USDA, 2000) is a statistically based survey that.
has been designed and implemented to assess conditions and trends of soil, water, and related
resources on non-Federal lands in the United States. The NRI is conducted every 5 years by the.
U.S. Department of Agriculture's (USDA) Natural Resources Conservation Service (NRCS), in
cooperation with the Iowa State University Statistical Laboratory. The inventory provides
scientifically valid, timely, and relevant information that is used to formulate effective
agricultural and environmental policies and legislation, implement resource conservation
programs, and enhance the public's understanding of natural resources arid environmental
conditions.

The NRI is a compilation of natural resource information on non-Federal land in the United
States-nearly 75 percent of the country's land base. The inventory captures data on land cover
and use, soil erosion,  prime farmland, wetlands, habitat diversity, selected conservation practices,
and related resource attributes at more than 800,000 scientifically selected sample sites. The NRI
can be accessed at http://www.nrcs.usda.gov/technical/NRI/.

3.4.3   NATIONAL  STORM WATER BMP DATABASE

The National Stormwater BMP Database, developed by the American Society of Civil Engineers
(ASCE), is designed to be a source of reliable data to help improve water quality nationwide by
sharing consistent and transferable information on the performance of storm water best
management practices. The database helps water quality professionals across the United States
learn about successful BMPs and apply proven methods to local water quality projects. The
June 2002
                                                                                    3-5

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

 database is based on extensive screening of a bibliography of more than 800 existing BMP
 studies and was designed by national storm water experts on ASCE's Urban Water Resources
 Research Council.  As of June 2002, the database contains data on 198 BMPs. Representative
 information provided for BMPs includes test site location, researcher contact data, watershed
 characteristics, regional climate statistics, BMP design parameters, monitoring equipment types,
 and monitoring data such as precipitation, flow, and water quality. The database can be accessed
 online at http://www.bmpdatabase.org.
 3.4.4
BMP DESIGN MANUALS AND GUIDANCE DOCUMENTS DEVELOPED
BY GOVERNMENTAL AND OTHER ORGANIZATIONS
 A variety of manuals and documents were used to obtain information on design and effectiveness
 of various BMPs. Examples include: (1) State design manuals such as the Virginia Erosion and
 Sediment Control Handbook (http://www.dcr.state.va.us/sw/e&s-ftp.htmX the Maryland Storm
 Water Design Manual rhtrp://www.mde.state.md.us/environment/wma/stormwatermanualX and
 the Denver Urban Drainage Criteria Manual (http://www.udfcd.org): (2) Guidance documents
 such as the Texas Nonpoint Source Book http://www.txnpsbook.org) and EPA's National Menu
 of BMPs (http://www.epa.gov/npdes/menuofbmps/menu.htm): and (3) Consensus design
 manuals such as manuals of practice on storm water design developed by ASCE and the Water
 Environment Federation (ASCE and WEF, 1992 and!998) were used to determine various
 management strategies. Links to on-line manuals and guidance documents are provided on
 EPA's website at http://www.epa.gov/waterscience/guide/construction/.

 3.5    REFERENCES

 ASCE and WEF. 1992. Design and Construction of Urban Stormwater Management Systems.
       ASCE Manual and Report on Engineering Practice No. 77; WEF Manual of Practice No.
       FD-20. American Society of Civil Engineers, New York, NY.  Water Environment
       Federation, Alexandria, VA.  http://www.asce.org and http://www.wef.org .

 ASCE and WEF. 1998. Urban Runoff Quality Management.  ASCE Manual and Report on
       Engineering Practice No. 87; WEF Manual of Practice No. 23.  American Society of Civil
       Engineers, Reston, VA. Water Environment Federation, Alexandria, VA.
       http://www.asce.org and http://www.wef.org .'

 USEPA. 1999. Economic Analysis of the Final Phase II Storm Water Rule. U.S. Environmental
       Protection Agency, Office of Wastewater Management. Washington, DC.

USDA. 2000.1997 National Resources Inventory. U.S. Department of Agriculture, National
      Resources Conservation Service, Washington, DC.
      http://www.nrcs.usda.gov/technical/NRI/.
June 2002
                                                                    3-6

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	.

 SECTION 4: INDUSTRY PROFILE

 4.1    INTRODUCTION

 The construction sector is among the largest and most important sectors in the national
 economy, accounting for approximately 4 percent of the U.S. gross domestic product. During
 1997, approximately 262,000 construction companies with payroll in the United States employed
 nearly 2.4 million workers nationwide. Another 1.6 million workers associated with construction
 activities were self-employed. The construction industry is divided into three major subsectors:
 general building contractors, heavy construction contractors, and special trade contractors.
 General contractors build residential, industrial, commercial, and other buildings. Heavy
 construction contractors build sewers, roads, highways, bridges, and tunnels. Special trade
 contractors typically provide carpentry, painting, plumbing, and electrical services.

 Because the proposed effluent guidelines are being developed to address water quality issues,
 this document focuses on the construction subsectors most closely associated with land-
 disturbing activities. General contractors and heavy construction establishments are by definition
 the most likely to conduct activities that could affect water resources. It should be noted,
 however, that for individual projects responsibility for land-disturbing activities and potential
 impacts on water quality might not be obvious because general contractors often subcontract all
 or some of the actual construction work. Hence, the following subsections describe the subsector
 categories most likely to be responsible for land-disturbing activities at the national level.

 4.2    INDUSTRY DESCRIPTION

 4.2.1  INDUSTRY DEFINITION AND CLASSIFICATION OF SUBSECTORS BY NAIC
       AND SIC CODES

 The construction and land development industry is classified in the 1997 North American
 Industry Classification System (NAICS, 1997) under Sector 23, Construction.  NAICS 1997 is
 the system currently used for' classifying industry establishments by type of economic activity.  It
 replaced the U.S.  Standard Industrial Classification (SIC) system.

 Construction work includes new construction, additions, alterations, and repairs. Establishments
 identified as construction-management firms are also included. The construction sector is divided
 into three types of activities or subsectors:

     • Subsector 233-Building, Developing, and General Contracting

       This subsector is made up of establishments responsible for the construction of building
       projects. Builders, developers, and general contractors, as well as land subdividers and
       land developers, are included in the subsector.  The construction work may be done for
June 2002
                                                                                     4-1

-------
Development Document for Construction and Development Proposed Effluent Guidelines            	

       others and performed by custom builders, general contractors, design builders, or turnkey
       contractors. This construction activity may be for sale as performed by speculative or
       operative builders.

     •  Subsector 234—Heavy Construction

       This subsector comprises establishments engaged in the construction of heavy
       engineering and industrial projects (except buildings), such as highways, power plants,
       and pipelines. Establishments in this subsector usually assume responsibility for entire
       nonbuilding projects, but they may hire subcontractors for some or all of the actual
       construction work.  Special trade contractors are included in this group if they are
       engaged in activities primarily related to heavy construction, such as grading for
       highways.  The kinds of establishments in this group include heavy-construction general
       contractors and design builders.

     •  Subsector 235-Special Trade Contractors

       This subsector comprises establishments engaged in specialized construction activities,
       such as plumbing, painting, and electrical work.. The activities in this subsector may be
       subcontracted from builders or general contractors, or the work may be performed
       directly for project owners. Special trade contractors usually perform most of their work
       at the job site..

Table 4-1 provides a list of the 3-digit subsectors, 4-digit industry groups and 5-digit NAICS
industries in the construction sector.
June 2002
                                                                                       4-2

-------
 Development Document for Construction and Development Proposed Effluent Guidelines

                   Table 4-1.  1997 NAICS, Subsectors, Industry Groups,
                    and Industries Performing Construction Activities
                                That Might Disturb Land
1997 NAICS Sector 23 - Construction
233
2331
23311
2332
23321
23322
2333
23331
23332
234
2341
23411
23412
2349
23491
23492
23493
23499
235
2357
23571
2359
23593
Building, Developing, and General Contracting
Land Subdivision and Land Development
Land Subdivision and Land Development
Residential Building Construction
Single-family Housing Construction
Multifamily Housing Construction
Nonresidential Building Construction
Manufacturing and Industrial Building Construction
Commercial and Institutional Building Construction
Heavy Construction
Highway, Street, Bridge, and Tunnel Construction
Highway and Street Construction
Bridge and Tunnel Construction
Other Heavy Construction
Water, Sewer, and Pipeline Construction
Power and Communication Transmission Line Construction
Industrial Nonbuilding Structure Construction
All Other Heavy Construction
Special Trade Contractors
Concrete Contractors
Concrete Contractors
Other Special Trade Contractors
Excavation Contractors
Before the creation of the NAICS, construction and land development industries were classified
using the SIC system. Any data collected before January 1997 might still be classified under that
system. SIC classifications are relevant to the effluent guidelines, because certain U.S. Bureau of
the Census (BOC) data for the construction industry were collected until 1994 and therefore
classified under the SIC system rather than the NAICS. Under the SIC system, industries that
might perform land-disturbing activities were classified under Division C-Construction, and
June 2002
4-3

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	.	,	

 Division H-Finance, Insurance, and Real Estate. These divisions include the following SIC
 major groups:

      •  SIC Major Group 15-Building Construction General Contractors and Operative Builders

        This group includes general contractors and operative builders primarily engaged in the
        construction of residential, farm, commercial, or other buildings. General building
        contractors who combine a special trade with their contracting are also included,

      •  SIC Major Group 16-Heavy Construction Other Than Building Construction Contractors

        This group includes general contractors primarily engaged in heavy construction other
        than building construction, such as highways and streets, bridges, sewers, railroads,
        irrigation projects, flood control projects, and marine construction, as well as special
        trade contractors primarily engaged in activities of a type clearly specialized in such
        heavy construction and not normally performed on buildings or building-related projects.

      •  SIC Major Group  17-Construction Special Trade Contractors
                                                                                      ^
        This group includes special trade contractors who undertake activities of a type that are
        specialized either in building construction or in both building and nonbuilding projects.

      •  SIC Major Group 65-Real  Estate

        This group includes real estate operators and the owners and lessors of real property,  as
        well as buyers, sellers, developers, agents, and brokers.

Major groups 15 and 16 are further defined by the type of construction performed.  Table 4-2
provides a list of the more specific industry groups and industries that might perform land-
disturbing activities.
June 2002
4-4

-------
Development Document for Construction and Development Proposed Effluent Guidelines
              Table 4-2.  1987 SIC Industry Groups Performing Construction
                              Activities That May Disturb Land
                                      SIC Major Group 15
           Industry Group 152: General Building Contractors - Residential
            1521
General Contractors - Single-family Houses
            1522
General Contractors - Residential Buildings, Other Than Single-family
           Industry Group 153: Operative Builders
            1531   Operative Builders
           Industry Group 154: General Building Contractors - Nonresidential
            1541
General Contractors - Industrial Buildings and Warehouses
            1542   General Contractors - Nonresidential Buildings, Other Than Industrial
                                      SIC Major Group 16
           Industry Group 161: Highway and Street Construction, Except Elevated Highways
            1611   Highway and Street Construction, Except Elevated Highways
           Industry Group 162: Heavy Construction, Except Highway and Street
            1622
Bridge, Tunnel, and Elevated Highway Construction
            1623
Water, Sewer, Pipeline, and Communications and Power Line
            1629   Heavy Construction Not Elsewhere Classified
                                      SIC Major Group 17
            Industry Group 179: Miscellaneous Special Trade Contractors
            1771
Concrete Work
            1794  Excavation Work
                                       SIC Major Group 65
            Industry Group 655: Land Subdividers and Developers
            6552
Land Subdividers and Developers, Except Cemeteries
The focus of this Development Document is on construction activities carried out by firms
covered by NAICS codes 233 and 234 or SIC codes 15 and 16.  (As discussed in Section VIA in
the preamble of the proposed rule, Special Trade Contractors, NAICS 235 or SIC 17, are
typically subcontractors and not identified as NPDES permittees.) Furthermore, the residential,
 June 2002
                                                                                            4-5

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

 non-residential, and heavy construction subsectors receive the greatest emphasis, because they
 account for the vast majority of construction projects and are responsible for most of the land
 disturbance in the United States. The following subsections describe these subsectors in terms of
 size, distribution, and recent growth trends.
 4.2.2
RESIDENTIAL BUILDING CONSTRUCTION GROUP
 Residential Construction Industry Description.  The U.S. Bureau of the Census (BOC), a
 division of the Department of Commerce (DOC), divides the residential construction industry
 into two categories. The first encompasses single-family housing construction and includes
 mobile homes, prefabricated houses, row houses, town houses, and single-family detached
 houses.  The second encompasses, multifamily housing construction and includes high-rise
 apartments, garden apartments, and town house apartments in which units are not separated by
 ground-to-roof walls.

 Historic Trends. The DOC began collecting detailed information on housing starts in 1963.
 Data on housing permits and starts are published monthly by the DOC and are viewed by
 economists as leading indicators of economic activity. More detailed industry information is
 collected through the Census of Construction Industries (CCI), which is conducted every 5 years
 (in years ending in a 2 or a 7) as part of the Census Bureau's Economic Census program. These
 data provide the most detailed snapshot of the status of the construction industry. The CCI
 covers all employer establishments primarily engaged in construction as defined by the NAICS
 and includes nonresidential construction activities. Table 4-3 summarizes housing starts for the
 period from 1979 to 1999.

 In Table 4-3, the number of construction starts is shown by regional location and type 'of
 structure. The table also provides national totals for both single- and multifamily housing starts
 (BOC, 2001). As shown in the table, single-family housing starts account for the majority of
 housing construction starts. Figure 4-1 combines single- and multifamily  housing starts and
 graphically depicts annual changes during the 1997-1999 period. The number of construction
 starts for privately owned housing units has decreased from approximately 1.7 million starts in
 1979 to roughly 1.6 million starts in 1999 (BOC, 2001).
June 2002
                                                                                    4-6

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

            Table 4-3. Annual Housing Construction Starts by Type and Region
                                 (Starts are in thousands)
Year
1979
1980
1981
1982
1983
1984
1985
1986
1987
1988
1989
1990
1991
1992
1993
1994
1995
1996
1997
1998
1999
United
States
1,745
1,292
1,084
1,062
1,703
1,750
1,742 "
1,805
1,621
.1,488
1,376
1,193
1,014
1,200
1,288
1,457 ,
1,354
1,447
1,474
1,617
1,641
Northeast
Single-
family
123
87
84
79
123
158
182
228
204
181
132
104
99
112
116
123
102
112
111
122
126
Multi-
family
55
38
33
37
45.
46
70
66
65
54
47
27
14
15
11
16
16
20
26 .
26
29
Midwest
Single-
family
243
142
no.
99
153
167
148
188
203
194
190
193
191
236
251
268
233
254
238
223
289
Multi-
family
106
76
55
50
65'
76
92
108
95
80
76
60
•42
52
47
61
57
68
66
58
59
South
Single-
family
522
428
363
357
557
528
504
504
485
443
, 409
371
353
439
498
522
485
524
507
573
580
Multi-
family
225
215
198
234
378
338
278
229
149
132
127
108
62
58
63
117
130
138
164
169
167
West
Single-
family
306
196
148
127
234
230
239
261
255
264
272
226
197
244
261
286
256
271
278
303
308
Multi-
family
165
110
' 92
78
148
206
230
222
165
140
124
103
57
45
41
65
76
90
. 86
92
84
  Source: BOC, 2001.
June 2002
                                                                                      4-7

-------
Development Document for Construction and Development Proposed Effluent Guidelines	

Housing start data tend to reflect the health of the U.S. economy. Therefore, as shown in Figure
4-1, the number of housing starts dropped significantly from 1986 to 1991 as the national
economy fell into a recession. Conversely, the robust economy over the past several years has
been accompanied by a strong growth in housing starts.
2,000 T-
1,800 • -
1,600 • -


800--













1979













1981























































































































-





















1983 1985 1987 1989 1991 1993 ' 1995 1997 1999
Year
                          Figure 4-1. Annual Housing Starts

Industry Size. As a result of the recent strong growth in demand for new housing, the number of
workers employed in residential construction has also increased. According to the BOC (1999b),
the total number of employees in the housing construction industry rose from 452,257 in 1992 to
628,886 in 1997, an increase of almost 40 percent. Table 4-4 shows the number of workers
employed, the payroll for those workers, and the value of completed construction for 1997. As
shown in the table, the number of establishments and workers associated with construction of
single-family housing greatly exceeds that for multifamily housing construction. It should also
be noted that although construction of single-family homes is performed by both small and large
firms, most multifamily housing construction is performed by large firms.  Specifically, a special
study by the Census Bureau (BOC, 2000a) found that about 39 percent of single-family homes
are built by small builders (fewer than 25 units in the year); 21 percent by medium builders (25-
99 units);  and 40 percent by large builders (more than 100 units).  In contrast, construction of
multifamily housing is performed primarily by larger builders. During 1997, large builders
constructed 77 percent of multifamily housing units.

The value of construction is defined as work done by general contractors, heavy construction
contractors, and special trade contractors.  Included in these estimates are new construction.;
June 2002
                                                                                     4-8

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

 additions, alterations or reconstruction, and maintenance and repair; the costs of industrial and
 other special equipment not an integral part of a structure are excluded. According to the 1997
 Construction Census, the value of completed construction exceeded $161 billion.  Single-family
 housing construction accounted for almost $147 billion, or more than 90 percent of the total.

               Table 4-4. Residential Construction Industry Profile for 1997

Total number of employees
Number of construction
establishments during the
year
Payroll (thousands)
Value of construction
completed nationwide
State with the highest dollar
value of construction work
for establishments with
payroll
Single-Family Housing
Construction
570,990
138,849
$14,964,583
$146,798,768
California
($18,137,680)
Multifamily Housing
Construction
58,896
7,543
$1,794,143
$14,487,308
Florida
($2,403,233)
             Source: BOC, 1999b, 1999cT

"Single-Family Housing Construction Trends.  As noted earlier, housing construction starts
increased significantly during the second half of the 1990s.  In 1999, single-family home
construction starts totaled more than 1.3 million, a level not reached since 1978 (BOC 2001).

As indicated in Table 4-5 by the number of permits issued, Atlanta, Georgia, led all U.S. major
markets for single-family housing construction activity in 19991.  The other leading market areas
for single-family construction were Phoenix, Arizona; Dallas-Ft. Worth, Texas; Chicago,
Illinois; and Washington, D.C.  Table 4-5 also shows the percent change in construction permits
issued from 1998 to 1999 (U.S. Housing Markets, 1999a).

Multifamily Housing Construction Trends.  Construction of structures with multiple housing
units also increased significantly during the 1990s.  For example,  construction starts of these
       1 Permits issued do not necessarily translate into housing starts, since a permit issued in
one year may not lead to actual construction until the next year. Furthermore, some permits
issued never lead to actual construction.  Nonetheless, permit counts can serve as a good
indicator of construction activity in the near future.
June 2002
                                                                                       4-9

-------
Development Document for Construction and Development Proposed Effluent Guidelines	     .

buildings increased from about 173,500 in 1991 to more than 338,500 in 1999, an increase of
about 95 percent.
                      Table 4-5. Busiest Markets for Single-Family
                                Housing Permits for 1999
Market Area
Atlanta
Phoenix
Dallas-Ft. Worth
Chicago
Washington, DC
Single-family Housing
Permits (1999)
25,066
21,290
17,434
14,954
14,703
Percent Change
From 1998
. +11%
+13%
+6%
+7%
0.07
Source: U.S. Housing Markets, 1999a.
Much of the growth in multifamily housing was in the construction of facilities with more than
five units. According to U.S. Housing Markets (1999b), the top five busiest markets for
multifamily construction permits for 1999 were Dallas-Ft. Worth, Texas; Orlando, Florida; New
York-Long Island; Puget Sound, Washington; and Houston, Texas. Table 4-6 shows the number
of multifamily permits and the percent change in permits issued from 1998 to 1999.

Regional Housing Start Trends (Single-family and Multifamily Structures). The Census Bureau
estimates housing starts at the regional level through statistical analysis of its survey data.

                       Table 4-6. Busiest Markets for Multifamily
                               Housing Permits for 1999
Market Area
Dallas-Ft. Worth
Orlando
New York-Long Island
Puget Sound
Houston
Multifamily Housing
Permits (1999)
8,488
7,303
6,255
6,122
5,900
Percent Change
From 1998
-15%
+46%
+55% t ,
+19%
-50%
              Source: U.S. Housing Markets, 1999b.
June 2002
4-10

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

 As shown in Figure 4-2, the Census Bureau divides the United States into four regions:
 Northeast2, Midwest3, South4, and West5.  Table 4-7 summarizes changes in construction starts
 at the regional level for the years 1989 and 1999.
                                                                                         Rl
                                                                                  Regions
                                                                                         Midwest
                                                                                  HI Northeast
                                                                                  E%^j South
                                                                                  |    ~[ West    .
                         Figure 4-2. Bureau of Census Housing Regions
As noted earlier, national housing starts have increased significantly over the past decade.  At the
regional level, however, growth rates have varied to a large degree. As shown in Figure 4-3 and
         The Northeast includes the following states: Connecticut, Maine, Massachusetts, New Hampshire, New Jersey, New
York, Pennsylvania, Rhode Island, and Vermont.

         The Midwest includes the following states: Illinois, Indiana, Iowa, Kansas, Michigan, Minnesota, Missouri, Nebraska,
North Dakota, Ohio, South Dakota, and Wisconsin.
        4
         The South includes the following states: Alabama, Arkansas, Delaware, District of Columbia, Florida, Georgia,
Kentucky, Louisiana, Maryland, Mississippi, North Carolina, Oklahoma,.South Carolina, Tennessee, Texas, Virginia, and West
Virginia.

         The West includes the following states: Alaska, Arizona, California, Colorado, Hawaii, Idaho, Montana, Nevada,
New Mexico, Oregon, Utah, Washington, and Wyoming.
June 2002
4-11

-------
Development Document for Construction and Development Proposed Effluent Guidelines
summarized in Table 4-7, construction of housing increased by nearly 40 percent in the South,
whereas construction starts in the Northeast actually decreased by almost 13 percent
from 1989 levels.  Housing starts in the Midwest also increased significantly over 1989 levels
while housing starts in the West remained at about the same level as a decade earlier.
Table 4-7. Changes in Housing Starts by Region (1989 and 1999)
Region
Northeast
Midwest
South
West
Total
1989 Housing Starts
(in thousands)
178.5
265.8
536.2
395.7
1,376.1
1999 Housing Starts
(in thousands)
155.7
347.3
746.0
391.9
1,640.9
Percent Change from
1989 to 1999
-12.77
30.66 ,
39.13
-0.96
19.24
Source: BOC, 1999a, 2001
June 2002
4-12

-------
Development Document for Construction and Development Proposed Effluent Guidelines
        800
          1988
1990
1992
1996
         1994

         Year

Northeast-*- Midwest* South ^ West
1998
2000
                      Figure 4-3. Annual Housing Starts by Region
4.2.3   NONRESIDENTIAL BUILDING CONSTRUCTION GROUP

The NAICS Nonresidential Building Construction group comprises establishments classified
either as Manufacturing and Industrial Building Construction or Commercial and Institutional
Building Construction. The following buildings are considered nonresidential by the U.S.
Census Bureau and fall under either the manufacturing or the commercial classification:
manufacturing and light industrial buildings; manufacturing and light industrial warehouses;
hotels and motels; office buildings; all other commercial buildings not elsewhere classified, such
as stores, restaurants, and automobile service stations; commercial warehouses; religious
buildings;  educational buildings; health care and institutional buildings; public safety buildings;
nonresidential farm buildings; amusement, social, and recreational buildings; and all other
nonresidential buildings. Because of the transition from the SIC system used in the 1992
Economic Census to theJNAICS for the 1997 census, a valid comparison of data between the two
censuses is not feasible, and therefore no historical data are shown.
June 2002
                                                              4-13

-------
Development Document for Construction and Development Proposed Effluent Guidelines	

Manufacturing and Industrial Building Construction. This industry type comprises
establishments primarily responsible for the entire construction of manufacturing and industrial
establishments, such as plants, mills, and factories.  Establishments identified as management
firms for manufacturing and industrial building construction are also part of this industry. They
include manufacturing and industrial building general contractors, design builders, engineer-
constructors, joint-venture contractors, and turnkey  contractors (BOC, 1999e).

In 1997, there were 7,280 manufacturing and industrial building construction establishments with
payroll (BOC, 1999e). These establishments employed 143,066 people for a total payroll of
more than $5.1 billion. The total value of manufacturing and industrial building construction
work in 1997 was more than $33.5 billion (BOC, 1999e). The value of construction work in
1997 by construction type is shown in Table 4-8 and includes new construction, additions,
alterations  or reconstruction, maintenance and repair, and any construction work done by the
reporting establishments for themselves.

Table 4-9 shows the value U.S. of construction work for establishments with payroll by work
location. States are grouped into four geographic regions: Northeast, Midwest, South, and West.
The South and the Midwest each accounted for approximately one-third of total 1997
construction value (southern region, approximately 32.4 percent; Midwest, nearly 30.1 percent).
The West and Northeast made up the remaining third (West, 23.4 percent; Northeast, 11.1
percent). Of the 50 states, California had the highest value of construction work at $3.4 billion,
10.1 percent of the total for the entire United States.  Michigan had the second-highest amount at
$2.9 billion (8.7 percent), followed by Texas at $1.9 billion (5.8 percent), and Ohio at $1.8
billion (5.3 percent).  The remaining states and Washington, D.C., each had less than 5 percent of
the total value of manufacturing and industrial building construction work in the United States in
1997.
June 2002
4-14

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

     Table 4-8. Value of Construction Work for Manufacturing and Industrial Building
         Construction Establishments With Payroll by Type of Construction, 1997
Type of Construction
Manufacturing and Light
Industrial Buildings
Manufacturing and Light
Industrial Warehouses
Hotels and Motels
Office Buildings
All Other Commercial Buildings,
Not Elsewhere Classified
Commercial Warehouses
Educational Buildings
Health Care and Institutional
Buildings
All Other Nonresidential
Buildings
Building Construction, Total
Nonbuilding Construction, Total1
Construction Work, Not Specified
by Kind
Manufacturing and Industrial
Building Construction, Total2
Value of Construction Work (thousands of dollars)
Total
$17,590,062
7,058,148
432,789
2,478,594
1,141,600
1,040,691 '
823,028
862,907
1,580,244
33,008,063
503,956
2,324
$33,514,342
New
Construction
$10,914,455
5,421,819
373,322
1,570,275
799,522
883,412
541,081
464,788
1,073,758
22,042,431
316,697
Not
Applicable
$22,359,127
Additions,
Alterations, or
Reconstruction
$4,280,143
1,358,864
49,580
810,808
298,166
131,005
255,540
355,116
436,029
7,975,252
123,832
Not
Applicable
$8,099,084
Maintenance
and Repair
$2,395,463
277,466
9,887
97,511
43,912
26,275
26,407
43,003
70,457
2,990,381
63,427
Not
Applicable
$3,053,807
    1. This information is shown for the breakdown of total industrial building construction values.
    2. Detail may not add to total because of rounding.
    Source: BOC, 1999e.
June 2002
4-15

-------
Development Document for Construction and Development Proposed Effluent Guidelines
Table 4-9. Value of Manufacturing and Industrial Building
Construction Work for Establishments with Payroll
by Location of Construction Work, 1997
(thousands of dollars)
Northeast
CT $260,593
ME 170,314
MA 403,700
NH 68,159
NJ 755,854
NY 920,425
PA 1,114,271
RI D
VT 14,812'








Total: $3,708,1282
Midwest
IL $1,208,663
IN 1,207,426
IA 381,922
KS 281,419
MI 2,908,857
MN 593,542
MO 745,632
NE 221,626
ND 89,251
OH 1,772,426
SD D
WI 669,575





Total: $10,080,339^
South
AL $1,080,420
AR 182,142
DE 169,305
DC 3,685
FL 920,179
GA 1,090,761
KY 861,206
LA 521,420
MD 253,778
MS 284,626
NC 921,364
OK 190,593
SC 689,581
TN 946,818
TX 1,934,909
VA 677,103
WV 144,481
Total: $10,872,3712
West
AK $62,907
AZ 561,785
CA 3,440,637
CO 330,551
HI S
ID 776,661
MT 26,176
NV 86,998
NM 377,538
OR 895,078
UT 314,621
WA 915,678
WY 52,326




Total: $7,840,9562
Total Value of Construction for United States: 33,514,3422
D: Withheld to avoid disclosing data of individual companies; data are included in
United States total.
S: Withheld because estimates did not meet publication standards.
1. Sampling error exceeds 40 percent.
2. Totals for regions do not include states with "S" and "D" criteria.
Source: BOC, 1999e. •
June 2002
4-16

-------
Development Document for Construction and Development Proposed Effluent Guidelines	

Commercial and Institutional Building Construction. This industry type comprises
.establishments primarily responsible for the entire construction of commercial and institutional
buildings, such as stores, schools, hospitals, office buildings, and public warehouses (BOC,
1999d). Establishments identified as management firms for commercial and institutional
building construction are also part of this industry type, which includes commercial and
institutional building general contractors, design builders, engineer-constructors, joint-venture
contractors, and turnkey contractors (BOC, 1999d).

In 1997, there were 37,430 commercial and institutional building construction establishments in
the United States employing a total of 528,173 people, with a payroll of $19.2 billion (BOC,
1999d). The value of construction work in 1997 by construction type is shown in Table 4-10.
Value includes new construction, additions, alterations or reconstruction, maintenance and repair,
and any construction work done by the reporting establishments for themselves (BOC, 1999d).

Table 4-11 shows the value of commercial and institutional building construction work by
location. The  data are reported by state, by region (Northeast, Midwest, South, and West), and
for the entire United States.  The South had the highest dollar value of construction activity,
accounting for $47.9 billion (27.7 percent) of commercial and institutional building construction
in the entire U.S. The West accounted for 20.6 percent of the total, followed by the Midwest at
16.8 percent, and then the Northeast at 9.7 percent.  Of the 50 states, California had the highest
value of commercial and institutional construction work at $18 billion, or 10.4 percent of the
total for the entire United States. Texas had the second highest value of construction at
approximately $13 billion  (7.5 percent), followed by Illinois at $7.9 billion (4.5 percent), and
then Georgia at $7.1 billion (4.1 percent).  The remaining states and Washington, D.C. each
accounted for  less than 4 percent of the total value of commercial and institutional building
construction work in the United States in 1997.
June 2002
                                                                                     4-17

-------
Development Document for Construction and Development Proposed Effluent Guidelines
        Table 4-10. Value of Construction Work for Commercial and Institutional
                  Building Construction Establishments With Payroll
                             by Type of Construction, 1997
Type of Construction
Single-Family Houses, Detached
and Attached
Apartment Buildings,
Apartment-Type Condominiums
and Cooperatives
Manufacturing and Light
Industrial Buildings
Manufacturing and Light
Industrial Warehouses
Hotels and Motels
Office Buildings
All Other Commercial Buildings,
Not Elsewhere Classified
Commercial Warehouses
Religious Buildings
Educational Buildings
Health Care & Institutional
Buildings
Public Safety Buildings
Farm Buildings, Nonresidential
Amusement, Social, and
Recreational Buildings
Other Building Construction
Building Construction, Total
Value of Construction Work (thousands of dollars)
Total
$2,690,846
4,081,493
8,083,739
3,325,768
8,313,559
36,147,979
32,715,012
6,929,460
4,324,007
23,974,844
17,446,710
5,345,602
1,904,128
6,529,907
3,429,673
166,818,246
New
Construction
$1,473,065
2,905,159
5,201,932
2,428,651
6,433,138
21,235,715
21,866,915
5,465,600
2,870,724.
15,587,110
11,187,636
4,183,179
1,508,380
5,141,460
1,984,749
110,618,170
Additions,
Alterations, or
Reconstruction
$1,000,110
1,016,097
2,425,390
776,335
1,679,856
13,524,406
9,631,103
1,215,709
1,342,559
7,893,507
5,917,408
1,064,693
272,836
1,275,033
895,522
50,325,006
Maintenance
and Repair •
$217,672
160,237.
456,417
120,783
200,564
1,387,858
1,216,994
248,151
110,724
494,>227
361,666
97,730
122,912
113,414
549,401
5,875,070
June 2002
4-18

-------
Development Document for Construction and Development Proposed Effluent Guidelines


Type of Construction
Nonbuilding Construction1
Construction W.ork, Not
Specified by Kind
Commercial and Institutional
Building Construction, Total2
Value of Construction Work (thousands of dollars)
Total
4,091,548
2,295,888
$173,205,680
New
Construction
2,697,377
Not
Applicable
$113,315,547
Additions,
Alterations, or
Reconstruction
1,205,513
Not Applicable
$51,530,519
Maintenance
and Repair
188,658
Not
Applicable
$6,063,728

   1. This information is shown for the breakdown of total industrial building construction values.
   2. Detail may not add to total because of rounding.
   Source: BOC, 1999d.
June 2002
                                                                                                  4-19

-------
  Development Document for Construction and Development Proposed Effluent Guidelines
Table 4-11. Value of Commercial and Institutional Building
Construction Work for Establishments With Payroll
by Location of Construction Work, 1997
(thousands of dollars)
Northeast
CT D
ME 385,818
MA 4,518,815
NH 697,186
NJ 4,973,021
NY D
PA 5,966,516
RI D
VT 303,481








Total: $16,844,837'
Midwest
IL 7,860,551
IN 3,132,116
IA 1,610,654
KS 1,609,747
MI 4,791,024
MN 3,361,074
MO D
NE 895,824
ND 297,619
OH 5,620,984
SD D
WI D





Total: $29,179,593'
South
AL D
AR D
DE 891,394
DC 1,724,839
FL D
GA 7,134,326
KY 1,961,212
LA 1,855,800
MD 3,693,531
MS D
NC 5,949,386
OK D
SC 2,417,316
TN 3,751,331
TX 12,953,464
VA 5,076,575
WV 529,092
Total: $47,938,266'
West
AK 509,429
AZ 3,287,644
CA 18,093,906
CO 3,728,688
HI D
ID D
MT 342,606
NV D
NM 913,252
OR 2,599,182
UT 1,796,639
WA 4,155,050
WY 211,989




Total: $35,638,385'
Total Value of Construction for United States: $173,205,6802
D: Withheld to avoid disclosing data of individual companies; data are included in
United States total.
1. Totals for regions do not include states with "D" criteria.
2. Detail may not add to total because of rounding, and because of "D" criteria.
Source: BOC, 1999d.
June 2002
                                                                                              4-20

-------
Development Document for Construction and Development Proposed Effluent Guidelines
4.2.4
HEAVY CONSTRUCTION SUBSECTOR
Industry Overview.  The heavy construction industry encompasses broad types of activities with
highway and street construction; bridge and tunnel construction; and water, sewer, and pipeline
construction as the three main types of heavy construction. The U.S. Census Bureau administers
a separate economic census for each of these three types of construction activities.

In general, most of the heavy construction industry indicators (e.g., value of completed work;
employment) have increased over the past two decades, although the health of the industry, like
that of the housing subsector, is closely tied to the overall state of the U.S.  economy. This
subsector has experienced both upturns and downturns over the past 20-year period.

The period encompassing the two most recent census years, 1992  and 1997, saw modest growth
in the heavy construction subsector. By 1997, the value of construction completed by the three
main types of heavy construction reached about $80 billion. As shown in  Table 4-12, the .
highway construction category of the heavy construction subsector accounted for about 60
percent of the total value of heavy construction. Highway construction employed the majority of
workers in the heavy construction subsector, accounting for about 278,000 of a total of 488,000
employees for all three categories of heavy contructio'n (BOC, 1999g).  Of the three heavy
construction categories, only the water, sewer, and pipeline category has experienced a decline in
number of establishments and number of employees.

Regional Distribution of Heavy Construction Activities.  The U.S. Bureau  of Census reports data
for the heavy construction industries at the state and regional levels. As in the case of the
housing subsector, the Census Bureau divides the United States into four major regions,
Northeast, Midwest, South, and West, each contributing to the total value of construction work.
As shown in Figure 4-4, the South and Midwest accounted for the majority of the establishments
and value of heavy construction work in 1997. In particular, these two regions accounted for 55
percent of the construction firms and 61 percent of the value of construction.

Of the three major types of heavy construction activities, highway.and street construction
accounted for almost 60 percent of the total value of heavy construction activities in 1997. The
distribution of highway construction establishments and the value  of completed work among the
different regions of the country are similar to those of the other heavy construction categories
For example, the South contributed more than $16 billion, or 34 percent, to the total value of
highway construction work in the United States. It should be reiterated, however, that the census
provides only a snapshot and that construction activities such as highway construction are
dependent on government funding and can change significantly in magnitude and location over
time.
June 2002
                                                                      4-21

-------
Development Document for Construction and Development Proposed Effluent Guidelines
                       Table 4-12. Overview of Heavy Construction
                                 Industry, 1992 and 1997
Year
Highway
'Bridges
Water, Sewer,
and Pipeline
Value of Construction (thousands of dollars)
1992
1997
35,331,607
48,472,284
7,198,275
9,539,041
20,205,048
22,204,058
Number of Establishments
1992
1997
10,090
11,270
1,041
1,177
10,233
8,042
Number of Employees *
1992
1997
257,356
277,979
43,701
47,764
194,252
162,566
1. Number of employees is the sum of all employees during
the pay periods that include the 12th of March, May, August,
and November, divided by four.
Source: BOC, 1992a, 1992b, 1992c, 1999f, 1999g, 1999h.
June 2002
                                                                                      4-22

-------
  Development Document for Construction and Development Proposed Effluent Guidelines
                NKteast
                                                                            Total
                                                  , Sfewr, atdPjidte
              Figure 4-4. Value of Heavy Construction Work by Region, 1997
 4.3     INDUSTRY PRACTICES AND TRENDS
 4.3.1
OVERVIEW OF CONSTRUCTION LAND-DISTURBING ACTIVITIES
 Constructing a building or facility involves a variety of activities, including the use of equipment
 that alters the site's environmental conditions. These changes include vegetation and top soil
 removal, regrading, and drainage pattern alteration.  The following provides a brief description of
 typical land-disturbing activities at construction sites and the types of equipment employed.

 Construction Site Preparation. Construction activities generally begin with the planning and
 engineering of the site and site preparation. During this stage, mobile offices, which are usually
 housed in trailers, are established on the construction site. The construction company uses these
 temporary structures to handle vital activities such as preparing and submitting applicable
 permits, hiring employees and subcontractors, and ensuring that proper environmental
 requirements are met.  The entire construction yard is delineated with erosion and sediment
 controls installed and security measures established.  The latter includes installing fences and.
 signs to warn against trespassing and to mark dangerous areas. After the site is secured,
 equipment is brought to the site (and is stored there throughout the construction period).
June 2002
                                                                                    4-23

-------
Development Document for Construction and Development Proposed Effluent Guidelines	
Clearing, Excavating, and Grading.  Construction on any size parcel of land almost always calls
for a remodeling of the earth (Lynch and Hack, 1984). Therefore, actual site construction begins
with site clearing and grading. Organic material cannot support the weight of buildings and
should be removed from the top layer of ground. (Some developers stockpile the organic material
for use during the landscaping phase of construction rather than paying for it to be hauled from
the site.) Construction contractors are to ensure that earthwork activities meet local, State, and
Federal regulations for soil and erosion control, runoff, and other environmental controls. The
size of the site, extent of water present, soil types, topography, and weather determine the kinds
of equipment used in site clearing and grading (Peurifoy and Oberlender,  1989). Material that
will not be used on the site should be hauled away by tractor-pulled wagons, dump tracks, or
articulated trucks (Peurifoy and Oberlender, 1989).

Equipment used for lifting excavated and cleared materials include aerial-work platforms,
forwarders, cranes, rough-terrain forklifts, and truck-mounted cranes.  In addition, track loaders
are used for digging and dumping earth (Caterpillar, 2000; Construction Equipment On-Line,
1996-1998; Lynch and Hack, 1984; and Peurifoy and Oberlender, 1989).

Excavation and grading are performed by several different types of machines. These tasks can
also be done by hand, but this is generally more expensive (Lynch and Hack, 1984).  When
grading a site, builders typically ensure that new grades are as close to the original as possible, to
avoid erosion and storm water runoff (Lynch and Hack, 1984).  Proper grading also ensures a flat
surface for development and drains water away from constructed buildings.

Excavation and grading equipment includes backhoes, bulldozers (including the versatile tracked
bulldozer), loaders, directional drilling rigs, hydraulic excavators, motor graders,  scrapers,
skid-steer loaders, soil stabilizers, tool carriers, trenchers, wheel loaders, and pipeliners.
Equipment selection depends on functions to be performed and specific site conditions
(Caterpillar, 2000; Construction Equipment On-Line, 1996-1998; Lynch and Hack, ,1984; and
Peurifoy and Oberlender, 1989).  Therefore, multiple types of equipment  are used throughout the
clearing and grading process.

Self-transporting trenching machines, wheelrtype trenching machines, and ladder-type trenching
machines are also used during site excavation. Self-transporting trenching machines are used to
create shallow trenches, such as for underground wire and cables. This type of machine has a
bulldozer blade attached to the front, is highly maneuverable, and can be used to dig narrow,
shallow trenches. Wheel-type trenching machines also dig narrow trenches, most often for water
mains and gas and oil pipelines.  Ladder-type trenching machines are used to dig deep trenches,
such as  for sewer pipes.  These machines might have a boom mounted at the rear. Along the
boom are cutter teeth and buckets that are attached to chains. As the machine moves, it digs dirt
and moves it to the sides of the newly formed trench (Peurifoy and Oberlender, 1989).
June 2002
                                                                                     4-24

-------
 Development Document for Construction and Development Proposed Effluent Guidelines   	

 Power shovels can also be used for excavating soils. They are used on all classes of earth that
 have not-been loosened. For solid rock, prior loosening is often necessary. As materials are
 excavated, they are immediately loaded onto trucks or tractor-pulled wagons and hauled from the
 site (Peurifoy and Oberlender, 1989). Hydraulic excavators, with either a front or a back shovel,
 are also used to dig into the earth and to load a hauling vehicle. There are several categories of
 hydraulic excavators, including backhoes, back shovels, hoes, and pull shovels. Hydraulic
 excavators are one of the most widely used types of excavating equipment because of their ease
 of use and their ability to remove the earth that caves as it is moved. They are effective
 excavating machines, and they are easy to use in terms of loading some sort of hauling vehicle
 (Peurifoy and Oberlender, 1989).

 Draglines, used to dig ditches or build levees, can transport soil within casting limits, thus
 eliminating the need for hauling equipment (Peurifoy and Oberlender, 1989).  Draglines have a
 bucket that hangs from a cable.  The bucket is brought through the dirt and toward the operator
 (Lynch and Hack, 1984). Draglines can be used on both wet and dry ground and can dig earth
 out of pits that contain water (Peurifoy and Oberlender, 1989). They are most useful for making
 large cuts and channels betow the level of the machine as well as for making valleys, mounds,
 slopes, and banks (Lynch and Hack, 1984).  Draglines have a lower output than power shovels,
 and do  not excavate rock as well as power shovels (Peurifoy and Oberlender, 1989).

 Draglines can be converted to clamshells by replacing the dragline bucket with a clamshell
 bucket. A clamshell is typically used for handling sand, gravel, crushed stone, sandy loam, and
 other loose materials; it is not efficient in handling compacted earth, clay, or other dense
 materials.  A clamshell is lowered into a material, and the bucket closes on the .material. It is
 then raised over a hauling vehicle and the materials deposited (Peurifoy and Oberlender, 1989).

 Scrapers, either self-powered or drawn by tractors, dig and compact materials by taking up earth
 from its underside with toothed scoops and loading it into hauling vehicles.  Scrapers are useful
 in removing earth and weak or broken rock, and for excavating hills and rock faces. Some
 scrapers are designed for long hauls; others with good traction are used on steep slopes (Lynch
 and Hack, 1984).

 A crawler tractor, which pulls a rubber-tired self-loading scraper, is often used for short-haul
 distances. The crawler tractor uses a drawbar pull to load the scraper. It has good traction and
 can operate on muddy roads.  It is, however, a slower vehicle and thus is more  appropriate for
 shorter hauls.

 Wheel-type tractor-pulled scrapers, which come in two- and four-wheel tractors, are used for
 longer hauling distances. Unlike the crawler tractor-pulled scrapers, the wheel-type tractor-
pulled scrapers do not maintain good traction.  Under such conditions, a helper tractor,, such as a
bulldozer, might be used (Peurifoy and Oberlender, 1989).
June 2002
                                                                                     4-25

-------
Development Document for Construction and Development Proposed Effluent Guidelines	._	
All these machines shape and compact the earth, a crucial site preparation step.  In addition,
earthwork activities might suggest that fill be brought in.  In such cases, the fill should be spread
in uniform, thick layers and compacted to a specified density with an optimum moisture content.
Graders and bulldozers are the most common earth-spreading machines.  Machines that compact
include tractor-pulled sheep's foot rollers, smooth-wheel rollers, pneumatic rollers, and vibrating
rollers, among other equipment (Peurifoy and Oberlender, 1989). Rollers and scarifiers are used
either to compact or to break up the ground (Lynch and Hack, 1984).

In order to remove rock,  it should first be loosened and broken up, usually through drilling or
blasting. Drilling equipment includes jackhammers, wagon drills, drifters, churn rills, and rotary
drills; each is designed to work on a specific size and type of rock. Dynamite and other
explosives are used to loosen rock (Peurifoy and Oberlender, 1989).

Once materials have been excavated and removed and ground cleared and graded, the site is
ready for construction.
4.3.2
CONSTRUCTION SITE SIZE CATEGORIES AND ESTIMATES OF
AMOUNT OF DISTURBED LAND
The proposed effluent guidelines would apply to construction sites of all types (i.e., residential,
commercial, and industrial) of more than one acre (5 acres, in the case of the guideline's Option
2). Because the costs of best management practices (BMPs) for erosion and sediment control are
largely driven by site size, EPA estimated the distribution of construction sites-by size category,
land use type, and geographic region in order to estimate the total cost of the proposed rule. (In
addition, estimating distribution of sites by type allows EPA to estimate the cost to each
construction sector.)

The method used to estimate the number of construction sites by size category, and therefore the
total area disturbed, is based on a number of data sources, including U.S. Census data and data
collected during  the Phase II Storm Water rulemaking.
4.3.2.1
National Estimates of Disturbed Acreage
EPA used the U.S. Department of Agriculture's (USDA's) 1997 National Resources Inventory
(NRI) to estimate the level of new U.S. development each year. (NRI is designed to track
changes in land cover and land use over time.) The inventory, conducted every five years, covers
all non-federal lands in the U.S. (75 percent of the U.S. total).  The program captures land use
data from some 800,000 statistically selected locations.  From 1992 to 1997, an average of 2.2
million acres per year were converted from non-developed to developed status. Table 4-13
shows the allocation of this converted land area by type of land or land cover.
June 2002
                                                                                    4-26

-------
Development Document for Construction and Development Proposed Effluent Guidelines





































Table 4-13. Acres Converted from Undeveloped to Developed State",
1992-1997

Type of Land


Cropland
Conservation Reserve
Program land
Pastureland
Rangeland
Forest land
Other rural area
Water areas and federal land
Total
Acres Converted to
Development 1992-1997
(thousands)
Annual Average
574.8
1.5

391.2
. 245.9
939.0
89.1
1.8
2,243.4

Percent Contribution by
Type of Land

26.6%
0.1%

17.4%
11.0%
41.9%
4.0%
0.1%
100.0%
a. NRI defines developed land as a combination of the following land cover/use categories
large urban and built-up areas, small built-up areas, and rural transportation land.
These are defined as follows:
Large urban and built-up areas. A land cover/use category composed of developed tracts
of at least 10 acres and meeting the definition of urban and built-up areas.
Small built-up areas. A land cover/use category consisting of developed land units of 0.25
to 10 acres, which meet the definition of urban and built-up areas.
Rural transportation land. A land cover/use category which consists of all highways,
roads, railroads and associated right-of-ways outside urban and built-up areas; also
includes private roads to farmsteads or ranch headquarters, logging roads, and other
private roads (field lanes are not included).
Urban and built up areas are in turn defined as:
Urban and built-up areas. A land cover/use category consisting of residential, industrial,
commercial, and institutional land; construction sites; public administrative sites; railroad
yards; cemeteries; airports; golf courses; sanitary landfills; sewage treatment plants; water
control structures and spillways; other land used for such purposes; small parks (less than
10 acres) within urban and built-up areas; and highways, railroads , and other
ransportation facilities if they are surrounded by urban areas. Also included are tracts of
ess than 10 acres that do not meet the above definition but are completely surrounded by
Urban and built-up land. Two size categories are recognized in the NRI: areas of 0.25 acre
o 10 acres, and areas of at least 10 acres.
Source: USDA, 2000.





































June 2002
                                                                                               4-27

-------
Development Document for Construction and Development Proposed Effluent Guidelines
4.3.2.2
Distribution of Acreage by Project Type
To allocate the NRI acreage among the various segments of the industry, EPA has estimated the
distribution of acres-developed by type of project in the following way. First, EPA multiplied
the number of building permits issued annually by estimates of the average site size for each
project type. Thus for single-family residential construction, EPA multiplied the number of new
single-family home building permits by the average lot size for new single-family construction.
Estimates for other types of construction were based on extrapolations from the U.S. Census
permit data and EPA estimates of average project size. Second, EPA adjusted the estimates of
acres converted to reconcile any differences between the total number of acres accounted for
using this approach and the total acres developed as estimated in the NRI.

Single-family Residential

Census data indicate that in recent years the number of new single-family housing units
authorized  has averaged just over 1.0 million units per year (see Table 4-14). The average lot
size for new single-family housing units is 13,553 square feet, or 0.31 acres (1 acre = 43,560
square feet). Using the average lot size (see Table 4-15), however, will underestimate the total
acreage converted for single-family residential projects because it does not include common
areas of developments not counted as part of an owner's lot. These  areas include streets,
sidewalks,  parking areas, storm water management structures, and open spaces.

                Table 4-14. New Single-Family and Multifamily Housing
                               Units Authorized, 1995-1997
Year
1995
1996
1997
1995-1997 avg
All Housing Units
1,332,549
1,425,616
1,441,136
1,399,767
Single-Family
Housing Units
997,268
1,069,472
1,062,396
1,043,045
Multifamily
Housing Units
335,281
356,144
378,740
356,722
                 Source: BOC, 2000b.
June 2002
                                                                                     4-28

-------
Development Document for Construction and Development Proposed Effluent Guidelines

                        Table 4-15.  Average and Median Lot Size
                                 for New Single-Family
                             Housing Units Sold, 1995-1997
Year
1995
1996
1997
1995-1997 avg
Average Lot Size
(Square Feet)
13,665
13,705
13,290
13,553
Median Lot Size
(Square Feet)
9,375
9,100
9,000
9,158
                      Source: BOC, 1995, 1996, 1997.

To account for these differences, EPA examined data obtained from a survey of municipalities
conducted in support of the Phase II Storm Water rule (EPA 1999). This survey identified 14
communities that consistently collected project type and size data as part of their construction
permitting programs.6 EPA's review of permitting data from these communities covered 852
single-family developments encompassing 18,134 housing units. The combined area of these
developments was 11,460 acres. This means that each housing unit accounted for 0.63 acres
(11,460 acres- 18,1.34 units = 0.63 acres per unit).  This estimate, essentially double the average
lot size, appears to more than account for the common areas and undeveloped areas in a typical
single-family residential development. For this reason, EPA averaged the Census estimate of the
national average lot size (0.31 acres) and the Phase II estimate of 0.63 acres per unit to arrive at
an estimate of 0.47 acres per unit.  This number was multiplied by the average number of single-
family housing units authorized by building permit, 1.04 million, to arrive at an estimate of
490,231 acres (see Table 4-18).

Multifamily Residential

For residential construction other than single-family housing,  EPA divided the average number
of units authorized during 1995-1997 (356,722, from Table 4-14) by the average number of units
per new multifamily building.  The average  number of units per building was obtained by
examining the distribution of units by unit size class in Census data (BOC 2000b). EPA
estimated the number of buildings in each size class by dividing the number of units in each  class
by the average number of units. The total number of units was then divided into the estimated
number of buildings to arrive at an average number of approximately 10 units per building across
       6 The communities were: Austin, TX; Baltimore County, MD; Gary, NC; Ft. Collins, CO; Lacey, WA;
 Loudoun County, VA; New Britain, CTf Olympia, WA; Prince George's County, MD; Raleigh, NC; South Bend,
 IN; Tallahassee, FL; Tuscon, AZ; and Waukesha, WI.
 June 2002
                                                                                     4-29

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

 all building size classes. Dividing 356,722 units authorized (Table 4-14) by 10 units per building
 project yields 35,672 individual development projects.

 EPA next examined data on the average site size for multifamily residential developments. The
 Center for Watershed Protection reports survey results showing that an average building footprint
 occupies 15.6 percent of the total site (CWP 2001). EPA assumed that the average-sized
 multifamily building (10.8 units) would have two floors and that each unit would occupy the
 national average of 1,095  square feet (NAHB 2002).  The total square footage accounted for by
 living space is thus 11,826 square feet. Multiplying by a factor of 1.2 to account for common
 areas and other non-living space (utility rooms, hallways, stairways), and dividing by 2 to reflect
 the assumption of a 2-story structure, EPA obtained a typical building footprint of 7,096-square
 feet (11,826 x 1.2 * 2 = 7,096). Combining this with the CWP estimate of the building footprint
 share of total site size (15.6 percent), the average site size was estimated to be 45,487 square feet
 (7,096'* 0.156 = 45,487), or just over 1 acre (1.04 acres).

 EPA compared the average site size obtained using this approach with data from the 14
 community survey referenced above under the Phase II Storm Water rule.  That study's review of
 permitting data identified 286 multifamily developments covering a total of 3,476 acres.  The
 average site size, 12.1 acres, is considerably higher than that calculated above.  EPA has no
 indication that the permits reviewed in these communities are for projects of a larger-than-
 average size. Therefore, for purposes of this analysis, EPA has taken the midpoint of the
 estimates, 6.5 acres, as the average size of multifamily projects. This number was multiplied by
 the average number of multifamily housing developments authorized by building permit, 35,672,
 to arrive at an estimate of 231,868 acres (see Table 4-18).
Nonresidential Construction

EPA lacked current data on the number of nonresidential construction and development projects
authorized annually because the Census Bureau ceased to collect data on the number of permits
issued for such projects in 1995.  EPA used regression analysis to forecast the number of
nonresidential building permits issued in 1997, based on the historical relationship between
residential and nonresidential construction activity. Using this approach, EPA estimates that a
total of 426,024 nonresidential permits were issued in 1997. These represent a variety of project
types, including commercial and industrial, institutional, recreational, as well as nonresidential,
nonbuilding projects such as parks and road or highway projects.

EPA first combined a number of project types into a larger "commercial" category, which
included hotels and motels and retail and office projects, as well as religious, public works, and
June 2002
                                                                                     4-30

-------
 Development Document for Construction and Development Proposed Effluent Guidelines      	

 educational projects.7 EPA's reasoning for including the latter categories under the commercial
 category is based on engineering judgment that erosion and sediment control practices would be
 similar across each project type. The total estimated number of commercial permits in 1997 was
 254,566 (59.7 percent of the nonresidential total). (EPA calculated a estimate for the industrial,
 category, which totaled 12,140 permits (2.8 percent), separately.) The residual 159,318 permits
 (37.4 percent), are nonbuilding, nonresidential projects that include parks, bridges, roads, and
 highways.  EPA accounts for these projects in the steps described below.

 For the industrial and commercial categories, EPA reviewed the project size data collected from
 the 14-community Phase II rule survey referenced earlier (EPA, 1999). This study identified 817
 commercial sites occupying 5,514 acres and 115 industrial sites occupying 689 acres.  The
 average site sizes are 6.7 and 6.0 acres, respectively.

 EPA also reviewed estimates from CWP (2001) on the average percent of commercial and
 industrial sites taken up by the building footprint. These percentages, 19.1 and 19.6 respectively,
 were multiplied across the model project site sizes of 1, 3, 7.5, 25, 70, and 200 acres to estimate
 building size on each site, assuming single-story buildings in each case.  These estimates are
 shown in Table 4-16.
                             Table 4-16.  Average Building Area
                                         (square feet)
Project Size
(Acres)
1
3
7.5
25
70
200
Commercial
8,320
24,960 .
62,400
207,999
582,397
1,663,992
Industrial
8,555
25,666
64,164
213,880
598,863
1,711,037
                      Estimates were obtained by multiplying the site size in square
                      feet by the percentage of the site estimated to be occupied by
                      the building "footprint," based on data from CWP (2001).

As seen in the table, the average building size corresponding to the 6- to7- acre sites estimated
from the 14-community study are in the 60,000 square feet range. EPA next examined R.S.
        The commercial category included: hotels/motels, amusement, religious, parking garages, service
stations, hospitals, offices, public works, educational, stores, and other nonresidential buildings.
June 2002
                                                                                       4-31

-------
Development Document for Construction and Development Proposed Effluent Guidelines	
Means' Building Construction Cost Data (2000), which provides cost data for "typical"
commercial and industrial buildings.  As part of the cost data, R.S. Means identifies the typical
range of building sizes based on a database of actual projects. Table 4-17 shows the typical size
and size range for a variety of building types that would fall into either the commercial or
industrial category. While some of the building types correspond with the estimated average of
60,000 square feet, these appear high for other categories, such as low-rise office and
supermarkets, warehouses, and elementary schools. EPA believes generally that there are more
small projects than large ones. As a result, EPA inferred that this approach would suggest an
average building size of 25,000 square feet, which implies an average site size of 3 acres, based
on Table 4-16.

To reconcile the estimates obtained from the two approaches, EPA has taken the midpoint of the
estimates. For commercial development, EPA assumes an average site size of 4.85 acres (the
average of 6.7 and 3.0 acres) and for industrial development EPA assumes an average site size of
4.5 acres (the average of 6.0 and 3.0 acres).

              liable 4-17. Typical Building Sizes and Size Ranges by Type of
                                        Building
Building Category/Type
Commercial - Supermarkets
Commercial - Department
Store
Commercial - Low-Rise Office
Commercial - Mid-Rise Office
Commercial - Elementary2
Industrial - Warehouse
Typical Size
(Gross Square
Feet)
20,000
90,000
8,600
52,000
41,000
25,000
Typical Range
(Gross Square Feet)
Low
12,000
44,000
4,700
31,300
24,500
8,000
High
30,000
122,000
19,000
83,100
55,000
72,000
            a. For purposes of this analysis EPA combines a number of building types, including
            educational, under the commercial category.
            Source: R.S. Means, 2000.  •  '

The resulting average project sizes were then multiplied by the estimated number of commercial
and industrial permits to obtain an estimate of the total acreage developed (and thus land acreage
disturbed) for these project categories. Table 4-18 shows the results of this "bottom-up"
approach to estimating the number of acres of land developed. The overall estimate of the
amount of land developed is 2.01 million acres per year. Residential single-family development
June 2002
                                                                                     4-32

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	
 accounts for 24.4 percent of the total, multifamily development for 11.5 percent of the total,
 commercial for 61.4 percent, and industrial for 2.7 percent.
             Table 4-18. National Estimates of Land Area Developed Per Year,
                              Based on Building Permit Data
Type of Construction
Residential
Nonresidential
Single-family
Multifamily
Commercial11
Industrial
Total
Permits
Number
1,043,045
35,672
254,566
12,140
1,345,423
Pet. of
Total
77.5%
2.7%
18.9%.
0.9%
100.0%
Average
Site Size"
0.47
6.5
4.9
4.5
--
Acres Disturbed
Number
490,231
231,868
1,234,645
54,630
2,011,374
Pet. of total
24.4%
11.5%
61.4%
2.7%
100.0%
    a. For single-family residential, this is the average of the average lot size for new construction in 1999
    (BOC, 2000b) and the average obtained in EPA (1999). For all other categories, the site sizes are EPA
    assumptions based on representative project profiles contained in R.S. Means (2000) and the 14-
    community survey conducted in support of the Phase IINPDES storm water rule (EPA, 1999).
    b. A number of project types were grouped together to form the "commercial" category, including:
    hotels/motels, amusement, religious, parking garages, service stations, hospitals, offices, public works,
    educational, stores, other nonresidential buildings.

The estimate of 2.01 million acres (Table 4-18) of annual construction is close to the estimated
2.24 million acres of annual new urban land obtained from 1997 NRI.  Areas not accounted for
in EPA's estimates include those converted as a result of road, highway, bridge, park, monument,
and other non-building construction  projects. EPA has not developed engineering costs
applicable to these types of projects, but assumes that the builders and developers of these areas
will face similar compliance costs per acre to the residential, commercial and industrial sectors,
and therefore, the acreage should be included in EPA's analysis. For the purpose of developing
national compliance costs, therefore, EPA has allocated the entire annual new urban acreage
from the 1997 NRI into the four land use categories using the distribution  shown in the final
column of Table 4-18. The third column in Table 4-19 summarizes the results of this allocation.
EPA next adjusted the annual developed acreage to account for sites that would not be required
to obtain a permit due to the low rainfall erosivity waiver contained in the  Phase  II rule, as well
as to eliminate sites less than 1 acre.  EPA estimated based on the Phase II economic analysis
that 33,517 acres would qualify for a low soil loss waiver, and analysis of the 14 community
survey data indicates that 33,828 acres would be in sites less than 1 acre.  This yields 67,345
acres of annual new development that would not be within the scope of the proposal. EPA
allocated this acreage among the four land uses based on an analysis of the number of permits
June 2002
                                                                                       4-33

-------
Development Document for Construction and Development Proposed Effluent Guidelines   	     •	
less than five acres contained within each respective segment.. The results of this allocation are
contained in the fourth column of Table 4-19, and the revised NRI acreage accounting for
waivers and  sites less than 1 acre is presented in the last column Table 4-19. EPA further
estimated acreage that would be eliminated from coverage given the 5 acre cutoff contained in
Option 2. A discussion of this analysis is included in the Economic Analysis supporting
document.
           Table 4-19. National Estimates of Land Area Disturbed Based on
                          National Resources Inventory Totals
Type of Construction
Residential
Nonresidential
Single-
family
Multifamily
Commercial0
Industrial
Total
Total NRI
Acreage"
546,783
258,616
1,377,070
60,932
2,243,400
Acres Waived or
not Covered
12,905
6,434
44,594
3,412
67,345
Adjusted NRI
Acreage1"
533,878
252,182
1,332,476
57,523
2,176,058
         a. This column distributes the total acreage estimated in NRI to be converted on an
         annual basis (adjusted for waivers) according to the distribution by type of development
         estimated through analysis of permits data contained in Table 4-18.
         b.  This column presents the total national acreage estimated after adjusting for rainfall
         erosivity waivers and sites less than 1 acre.
         c. A number of project types .were grouped together to form the "commercial" category,
         including: hotels/motels, amusement, religious, parking garages, service stations,
         hospitals, offices, public works, educational, stores, other nonresidential buildings.
4.3.2.3 Distribution of Developed Acreage by Project Size and Geography

For each of the four land use categories in Table 4-19, EPA developed a distribution to allocate
developed acre estimates among six different project size categories. The project size
distribution is based on a survey of construction permits issued in 14 communities conducted in
support of the Phase II storm water rule.  Table 4-20 shows the distribution of the 14 community
survey data by project size for each land use category.  The percentages shown in Table 4-20   •
were used to allocate the total estimated development within each of the four land use sectors in.
Table 4-19 into six site size categories. The results of this analysis are presented in Table 4-21.
June 2002
                                                                                         4-34

-------
Development Document for Construction and Development Proposed Effluent Guidelines   	
In addition, EPA developed procedures to spatially distribute land development regionally, using
19 eccregions covering the contiguous states. A description of this methodology is presented in
the Environmental Assessment supporting document.
June 2002
4-35

-------
 Development Document for Construction and Development Proposed Effluent Guidelines
                     Table 4-20. Distribution of 14 Community Survey
                                    Permits by Site Size
Site Size (Acres)
No. of
Permits
Acres by Size
Pet. Acres by Size
Single-Family Residential
1
3
7.5
25
70
200
Total
266
228
138
175
30
15
852
Multifamily Residential
1
3
7.5
25
70
200
Total
43
100
61
71
10
1
286
Commercial
1
3
7.5
25
70
200
Total
266
356
86
91
16
0
815
266
684
1,035
4,375
2,100
3,000
• 11,460

43
300
458
1,775
700
200
3,476

266
"1,068
. 645
2,275
1,260
0
5,514
2.3%
6.0%
9.0%
38.2%
18.3%
26.2%
100.0%

1.2%
8.6%
13.2%
'51.1%
20.1%
5.8%
100.0%

- 4.8%
19.4%
11.7%
41.3%
22.9%
0.0%
100.0%
June 2002
4-36

-------
 Development Document for Construction and Development Proposed Effluent Guidelines

                      Table 4-20. Distribution of 14 Community Survey
                                      Permits by Site Size
Site Size (Acres)
No. of
Permits
Acres by Size
Pet. Acres by Size
Industrial
1
'3
7.5
25
70
200
Total
39
55
10
8
3
0
115
39
165
75
200
210
0
689
5.7%
23.9%
10.9%
29.0%
30.5%
0.0%
100.0%
Total
1
3
7.5
25
70
200
Total
614
739
295'
345
59
16
2,068
614
2,217
2,213
8,625
4,270
3,200
21,139
2.9%
10.5%
10.5%
40.8%
20.2%
15.1%
100.0%
                 Based on permitting data from the following municipalities or counties:
                 Austin, TX; Baltimore County, MD; Gary, NC; Ft. Collins, CO; Lacey,
                 WA; Loudoun County, VA; New Britain, CT; Olympia, WA; Prince
                 George's County, MD; Raleigh, NC; South Bend, IN; Tallahassee, FL;
                 Tucson, AZ; and Waukesha, WI.
                 Source: USEPA, 1999
June 2002
4-37

-------
Development Document for Construction and Development Proposed Effluent Guidelines
                 Table 4-21. Distribution of National Construction by Site
                               Size and Development Type
Site Size (Acres)
No. of
Permits
Acres by Size
Pet. Acres by Size
Single-Family Residential
1
3
7.5
25
70
200
Total
12,392
10,622
6,429
8,153
1,398
699
3.9,691
12,392
31,865
48,217
203,815
97,831
139,759
533,878
2.3%
6.0%
9.0%
38.2%
18.3%
26.2%
100.0%
Multifamily Residential
1
3
7.5
25
70
200
Total
Commercial
1
3
7.5
25
70
200
Total
3,120
7,256
4,426
5,152
726
73
20,752

64,280 .
86,029
20,782
21,990
4,350
0
197,431
3,120
21,768
33,196
128,794
50,792
14,512
252,182

64,280
258,086
155,866
,549,761
304,483
0
1,332,476
1.2%
8.6%
13.2%
51.1%
20.1%
5.8%
100.0%

4.8%
19.4%
11.7%
. 41.3%
22.9%
0.0%
100.0%
June 2002
                                                                                      4-38

-------
Development Document for Construction and Development Proposed Effluent Guidelines	

                  Table 4-21. Distribution of National Construction by Site
                                 Size and Development Type
Site Size (Acres)
No. of
Permits
Industrial
1
3
7.5
25
70
200
Total
Totals
1
3
7.5
25
70
200
Grand Total
3,256
4,592
835
668
250
0
9,601

83,048
108,498
32,472
35,963
6,723
771
267,475
•Acres by Size

3,256
13,775
6,262
16,698
17,532
0
57,523

83,048
325,494
243,541
899,067
470,638
154,271
2,176,059
Pet. Acres by Size

5.7%
23.9%
10.9%
29.0%
30.5%
0.0%
100.0%

3.8%
15.0%
1-1.2%
41.3%
21.6%
7.1%
100.0%
                 Based on permitting data from the following municipalities or counties: Austin,
                 TX; Baltimore County, MD; Gary, NC; Ft. Collins, CO; Lacey, WA; Loudoun
                 County, VA; New Britain, CT; Olympia, WA; Prince George's County, MD;
                 Raleigh, NC; South Bend, IN; Tallahassee, FL; Tuscon, AZ; and Waukesha, WI.
                 Source: USEPA, 1999.
4.4    REFERENCES

BOC. 1992a.  1992 Census of Construction Industries: Highway and Street Construction
       Contractors, Except Elevated Highways. U.S. Bureau of the Census, Washington, DC.
       http://www.census.gOV/prod/l/constr/92ind/cci06f.pdf. Accessed May 21, 2002.
June 2002
4-39

-------
Development Document for Construction and Development Proposed Effluent Guidelines	
BOC. 1992b.  1992 Census of Construction Industries: Bridge, Tunnel and Elevated Highway
       Construction Contractors.  U.S. Bureau of the Census, Washington, DC.
       http://www.census.gOV/prod/l/constr/92ind/cci07f.pdf. Accessed May 21, 2002.

BOC. 1992c.  1992 Census of Construction Industries: Water, Sewer, Pipeline and
       Communications andPowerline Construction. U.S. Bureau of the Census, Washington,
       DC. http://www.census.gOV/prod/l/constry92ind/cci08f.pdf. Accessed May 21, 2002.

BOC. 1995. Characteristics of New Housing: 1995, Current Construction Reports.  U.S. Bureau
       of the Census, Washington, DC. http://www.census.gov/prod/l/constr/c25/c25_95a.pdf.
       Accessed May 28, 2002.

BOC. 1996. Characteristics of New Housing: 1996, Current Construction Reports.  U.S. Bureau
       of the Census, Washington, DC. http://www.census.gOV/prod/l/constr/c25/c25-96a.pdf.
       Accessed May 28, 2002.

BOC. 1997. Characteristics of New Housing: 1997, Current Construction Reports.  U.S. Bureau
       of the Census, Washington, DC. http://www.census.gOV/prod/3/98pubs/c25-97a.pdf.
       Accessed May 28, 2002.

BOC. 1999a. Housing Starts, Current Construction Reports. U.S. Bureau of the Census,
       Washington, DC. http://www.census.gov/prod/99pubs/c20-9901.pdf. Accessed October
       1,2000.

BOC. 1999b. Single-family Housing Construction, 1997 Economic Census,
       Construction Industry Series. U.S. Bureau of the Census, Washington, DC.
       http://www.census.gov/prod/ec97/97c2332a.pdf! Accessed October 2, 2000.

BOC. 1999c. Multifamily Housing Construction, 1997 Economic Census, Construction Industry
       Series. U.S. Bureau of the Census, Washington, DC.
       http://www.census.gov/prod/ec97/97c2332b.pdf. Accessed October 15, 2000.

BOC. 1999d. Commercial and Institutional Building Construction, 1997 Economic Census,
       Construction Industry Series. U.S. Bureau of the Census, Washington, DC.
       http://www.census.gov/prod/ec97/97c2333b.pdf. Accessed October 25, 2000.

BOC. 1999e. Manufacturing and Industrial Building Construction, 1997 Economic Census,
       Construction Industry Series. U.S. Bureau of the Census, Washington, DC.
       http://www.census.gov/prod/ec97/97c2333a.pdf. Accessed October 25, 2000.
June 2002
4-40

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	
 BOC. 1999f. Water, Sewer, Pipeline Construction,  1997 Economic Census, Construction
       Industry Series. U.S. Bureau of the Census, Washington, DC.
       http://www.census.gov/prod/ec97/97c2349a.pdf. Accessed October 25, 2000.

 BOC. 1999g. Highway and Street Construction Contractors. U.S. Bureau of the Census,
       Washington, DC. http://www.census.gov/prod/ec97/97c2341a.pdf. Accessed October 25,
     '  2000.

 BOC. 1999h. Bridge and Tunnel Construction. U.S. Bureau of the Census, Washington, DC.
       http://www.census.gov/prod/ec97/97c2341b.pdf. Accessed October 25, 2000.

 BOC. 2000a. 1997 Economic Census-construction Sector Special Study: Housing Start
       Statistics, a Profile of the Homebuilding Industry.  Issued July 2000. U.S. Bureau of the
       Census, Washington, DC.

 BOC. 2000b. New Privately Owned Housing Units Authorized by Building Permits in Permit-
       issuing Places, Annual Data. U.S. Bureau of the Census, Washington, DC.

 BOC. 2001. New Privately Owned Housing Units Started: Annual Data, 2001. U.S. Bureau of
       the Census, Washington, DC. http://www.census.gov/const/startsan.pdf. Accessed May
       23,2002.

 Caterpillar. 2000. Caterpillar, Inc. http://www.cat.com

 Construction Equipment On-line. 2000. Reed Business Information, U.S. www.coneq.com

 CWP. 2001.  Impervious Cover and Land Use in the Chesapeake Bay Watershed. Ellicott  City,
       MD: Center for Watershed Protection, January. Additional data table, "Chesapeake bay
       watershed impervious cover results by land use polygon," received via a facsimile from
       Tetra Tech, Inc., September 20, 2001.

Lynch, Kevin and Hack, Gary. 1984. Site Planning (3rd ed.). Cambridge, MA: MIT Press.

NAHB. 2002. Characteristics of New Multifamily Buildings 1987-1999. National Association
       of Home Builders, http://www.nahb.com/multifamily/characteristics.htm. Accessed May
       29,2001.

NAICS. 1997. North American Industry Classification System-U.S.  U.S. Department of
       Commerce,  National Technical Information  Service, Washington, DC.

Peurifoy, Robert L. and Oberlender, Garold D. (198.9). Estimating Construction Costs (4th ed.).
       New York: McGraw Hill Book Company.
June 2002
                                                                                   4-41

-------
Development Document for Construction and Development Proposed Effluent Guidelines	
R.S. Means.  2000. Building Construction Cost Data 58th Annual Edition. R.S. Means Co.,
       Kingston, Massachusetts.

R.S. Means.  2001. Heavy Construction Cost Data 15th Annual Edition. R.S. Means Co.,
       Kingston, Massachusetts.

USDA. 2000.1997 National Resources Inventory. U.S. Department of Agriculture, National
       Resources Conservation Service, Washington, DC. www.nrcs.usda.gov/technical/NRI/.

USEPA. 1999. Economic Analysis of the Final Phase II Storm Water Rule.  U.S. Environmental
       Protection Agency, Office of Wastewater Management. Washington, DC.

U.S. Housing Markets.  1998. Homebuilders Face a Steep Climb Chasing a 20-year-old-famify
       Mark. Meyers Real Estate Information, Inc.
       htt]j://www.housingusa.com/ushm/pr/ushm998.pr.html. Accessed October 25, 2000.

U.S. Housing Markets.  1999a.  Busiest Markets, Single-family Building. Meyers Real Estate
       Information, Inc.
       http://www.housingusa.com/ushm/pmts/mulbuzz.html. Accessed October 25, 2000.

U.S. Housing Markets. 1999b. Busiest Markets, Multifamily Construction. Meyers Real Estate
       Information, Inc.
       http://www.housingusa.com/ushm/pmts/mulbuzz.html Accessed October 25, 2000.
 June 2002
                                                                                   4-42

-------
 Development Document for Construction and Development Proposed Effluent Guidelines
 SECTION 5: TECHNOLOGY ASSESSMENT

 This'technology assessment of available data sources is intended to determine the depth and
 breadth of effectiveness data for various erosion and sediment controls, and to identify the
 amount and quality of data available to describe the performance of all currently used and
 innovative runoff control practices, the ability of each practice to effectively control impacts due
 to runoff, and the design criteria or standards currently used to size each practice to ensure
 effective control of runoff.

 5.1    CONSTRUCTION EROSION AND SEDIMENT CONTROLS

 5.1.1         INTRODUCTION

 Part 1, reported in this sub-section, addresses the erosion and sediment control BMPs for the
 construction phase of development. Prior to initiating this aspect of the work, EPA reviewed the
 findings of information sources and literature assessments to identify the appropriate definition
 of "performance" or the various definitions or "levels" of performance that are considered in
 evaluating and defining the levels of performance for these BMPs. A scientific-based approach
 to describe the performance of erosion and sediment control BMPs was devised similar to the
 approach developed by Barfield and Clar (1985) in the evaluation of the Maryland Erosion and
 Sediment Control Standards, as well as the one recently developed in the American Society of
 Civil Engineers BMP Database (ASCE, 1999).  The approach used in this assessment has been
 designed to provide the information needed to address several important issues, including
 whether to use a design-based approach, or an effluent-based concentration, or a loading
 approach in reporting on the current status of the technology. This sub-section identifies the
 following:

 •  The amount and quantity of data available to describe the performance of all currently used
   and innovative runoff control practices.

 •  The ability of each practice to effectively control impacts due to runoff.

 •  The design criteria or standards currently used to size each practice to ensure effective
   control of runoff.

Before a detailed evaluation of the BMPs can be provided, some background information is
necessary. Sub-section 5.2 describes the procedure for assessing the technology.  Sub-section
5.3 provides a historical background on the subject. Next, sub-section 5.4 presents a discussion
of goals, control strategies, criteria, and standards in general, and sub-section 5.5 provides a
detailed description and discussion of each BMP.
June 2002
                                                                                     5-1

-------
Development Document for Construction and Development Proposed Effluent Guidelines	


In the discussion of BMPs in sub-section 5.5, the major focus will be on sediment. This does not
imply that there are no other impacts; however, construction BMPs have focused on erosion aiad
sediment control rather than on other impacts.

In the assessment of BMPs> considerable attention is focused on whether to use a design-based
approach, an effluent-based concentration, or a loading approach in reporting on the current
status of the technology. Attention is also given to the recent emphasis in the literature on the
use of an integrated approach to evaluate impacts to the receiving waters and downstream areas.

5.1.2          PROCEDURE FOR TECHNOLOGY ASSESSMENT

5.1.2.1           IDENTIFICATION OF PERFORMANCE GOALS

In assessing the literature, particular consideration was given to definitions of performance of
BMPs and how they addressed the range of receiving water impacts identified.  It is important to
point out that the overarching performance goal of all the BMPs is to minimize the impact of
construction site runoff on receiving  waters and downstream areas.

Control strategies that have been identified for construction BMPs can be divided into three
categories.

Strategy 1. Control Based on Design Standards—-Control at this level is based on standard
designs that may include such things as volume requirements for reservoirs, detention time, and
trapping efficiency that do not directly limit an allowable discharge to receiving waters or limit a
downstream impact.

Strategy 2. Control Based on Effluent Standards—Control at this level is based on limiting the
quantity of one or more substances such as peak discharge, runoff volume, TSS, and settleable
solids. This directly addresses effluent, but does not directly address downstream impacts.

Strategy 3. Control Based on an Integrated Approach—Control at this level uses an integrated
approach (Snodgrass et al., 1998), including biological, chemical, and physical criteria, to define
BMP performance. A combination of water quality, biohabitat, and geomorphic criteria is used
to evaluate .whether a receiving stream is at the targeted goal pf fishable and swimmable, or the
extent of departure from this goal.

The majority of BMPs address Strategies. 1 or 2.  Although Strategy 3 is being discussed in the
literature, it has not been adopted in practice.  There is an analog in the surface mining industry,
where a cumulative hydrologic impact analysis on a watershed basis is required by the U.S.
Surface Mining and Reclamation Act of 1977 (PL95-87).  When moving from Strategy 2 to
Strategy 3, a number of other parameters are added to the performance criteria in Strategy 2,
June 2002
                                                                                     5-2

-------
 Development Document for Construction and Development Proposed Effluent Guidelines
 including (1) stream buffer retention and thermal impacts considerations, (2) volume control
 considerations such as are presented in the Low Impact Development concept approach, which
 are added to the peak discharge and ground water recharge criteria to achieve maintenance of
 hydrologic function at a site-specific level, and (3) geomorphic criteria as described by Lane
 (1955), Leopold et al. (1964), Rosgen (1996), and others.

 An important point must be made about controlling sediment. From a practical standpoint, a
 reasonably sized structure should not necessarily be expected to meet an effluent TSS standard
 unless the TSS specified in the standard is set at a very high value or unless some form of
 chemical treatment is used to enhance flpcculatidn. The settling velocity for primary clay
 particles is in the range of feet per month for all but the largest particles. Since these size
 particles are frequently encountered in large percentages in sediment from construction sites, the
 expected trapping efficiencies will not approach 100 percent, nor will the effluent TSS be hi the
 range of 100 mg/L or lower (Haan et al., 1994).
5.1.2.2
GOALS, ENVIRONMENTAL IMPACT AREAS, AND ASSESSMENT
SCALES
For the purposes of this report, impact areas are divided into three categories, local area,
receiving water, and downstream areas.

Local Area. This is the area between the construction site and the receiving stream. Typically,
these areas have ephemeral streams with low baseflows and highly variable flow rates, hi these
areas, the flows fluctuate widely, with geomorphology and habitat being very susceptible to
changes in hydrologic regime (Klaine, 2000). hi some developments, there would essentially be
no local area, and flows would exit directly into receiving waters.

Receiving Waters. This is the point at which flows enter a well-defined stream. Depending on
the local geology, flows may primarily be ephemeral, there may be a well-established baseflow,
or there may be something intermediate between the two extremes.  The degree to which flows,
sediment,  and chemicals impact the receiving waters depends largely on the type of receiving
water. For example, if the receiving waters have a low baseflow and highly variable flow rates,
the habitat and geomorphology will be very sensitive to significant changes in the hydrologic
regime.  However, if the receiving waters have a high baseflow, the sensitivity to changes in .
flow rate will be much less and the primary problems will likely be chemical in nature. Thus, it
is important to address impacts on a site-specific basis.

Downstream Areas.  A definition of the downstream area can be somewhat nebulous. (A
definition  of the aerial extent of "downstream areas" is something that needs to be developed in
follow-up  studies.)  However, consideration of this area is important. For example, use of peak
discharge criteria may directly control the local area impacts and impacts to the point at which
June 2002
                                                                                    5-3

-------
Development Document for Construction and Development Proposed Effluent Guidelines	 '	


flow enters the receiving waters.  If the watershed being considered is combined with other
downstream watersheds and all use peak discharge control without controlling runoff volume,
there can be an increase in flooding due to superposition of long duration peak flows exiting the
numerous reservoirs (Smiley and Haan, 1976). This increased discharge can negatively impact
channel geomorphology, habitat,  and riparian areas.

Another important issue related to construction is the fraction of the watershed under
construction at any one time. One argument about the relative importance of the construction
phase versus the post-construction phase is that the construction phase is short-lived and the
impact may be reversible after the site has stabilized. While this argument may  have some
validity on the local area, it is invalid when considering the downstream areas. On a larger
watershed under development, major construction may occur in the watershed for a long time,
with a potential long-term major cumulative impact.  When considering the entire watershed, it
may be desirable to  limit the area under construction at any one time to prevent exceeding some
threshold that would result in an irreversible impact.  This indicates the need to conduct a
cumulative impact analysis on a river basin scale to evaluate the potential for such an impact to
occur.                                     .                                   •

When considering area impacts, the following comments can be made about the strategies listed
above.

Strategy 1. No guarantees can be made that impacts would be controlled at any level unless the
design standards are highly conservative. This would result in overdesign for most situations so
that the standard would be adequate for all situations.

Strategy 2. This strategy should ensure control at the local level. Downstream, the impacts may
be positive or negative as a result of the control.  Examples include the control of peak discharge
only in storm water runoff.  Control of peak discharge on all construction areas at the local level
can result in increased peak discharge downstream (Smiley and Haan, 1976).  These increases
result from detaining increased volumes of runoff resulting from urbanization and releasing them
at the predisturbed peak rate over a long period of time.                              ,

Strategy 3. This approach should ensure control in both the local area and downstream areas.

Scale is very important to BMP effectiveness analyses.  A given BMP may be quite effective in
controlling impacts nearby but have a significant negative impact when applied over a large area.
In the final analysis, effectiveness should be  evaluated at multiple scales before a decision is
made.  This will require both local and watershed level analyses.                  .
June 2002
                                                                                      5-4

-------
 Development Document for Construction and Development Proposed Effluent Guidelines
 5.1.2.3
QUALITATIVE VERSUS QUANTITATIVE ASSESSMENT
 In the assessments, the issue may be addressed on a qualitative or a quantitative basis. The
 difference can be explained in the following manner, using water temperature as an example.  It
 is well known that turbidity impacts the depth of penetration of solar energy into a waterbody;
 hence, turbidity impacts temperature. When evaluating the impact of standards on water
 temperature, it is obvious that a TSS standard directly addresses water temperature because of
 the impact of TSS on turbidity. Thus, a qualitative analysis would simply state that TSS
 standards may impact water temperature, but give no degree to which the standard does impact
 temperature. A quantitative analysis, however, would define the degree to which a given TSS
 standard increased or decreased the impact of storm water TSS on temperature.

 5.1.3      REVIEW OF HISTORICAL APPROACHES TO EROSION AND SEDIMENT
           CONTROL

 Most early sediment control was related to agriculture and was installed as a way to maintain our
 natural resource base. On-site control was the primary emphasis, attempting to prevent erosion
 rather than trap sediment. Strategies were developed to minimize exposure of bare soil to the
 erosive power of rainfall  and runoff, using aboveground cover management, residue
 management, strip cropping, and terracing to limit the length of overland flow.  Impacts to
 receiving streams and downstream areas had not yet been identified as an issue, hi the 1960s,
 concern began to be expressed about the quantities of sediment in streams and reservoirs, and
 sediment was first identified as a pollutant Initially, the major focus of sediment control  was on
 the surface mining industry, with the passage of the Clean Water Act and then the Surface
 Mining, Reclamation, and Control Act (SMRCA) (PL 95-87) (U.S. Congress, 1977).  The first  '
 approach taken to sediment control was a design standard, requiring a sediment detention basin
 with a 24-hour detention time; TSS standards of 35 mg/L average and 70 mg/L peak were also
 promulgated, but were not typically enforced. The U.S. Environmental Protection Agency
 (USEPA) later evaluated  the TSS standard and moved to a settleable solids standard of 0.5 ml/L,
 based on a modeling effort that showed that it was not possible to trap fine sediments, but that  a
 0.5 ml/L settleable solids  standard could be met  with a reasonably sized sediment basin (Ettinger
 and Lichty, 1979).

 hi the late 1960s and early 1970s, sediment in streams and waterways originating from urban
 construction sites became an issue, which was then addressed in the Clean Water Act. EPA
 developed a list of BMPs  and standards for their construction. (USEPA, 1971).  In general, these
 standards were adopted from those of other agencies and were not based on studies related to
 urban runoff.

 hi 1987, the Clean  Water  Act was amended to include storm water discharges from urban areas.
 The Phase INPDES Stormwater regulations were published in 1990, requiring all municipalities
'June 2002
                                                                                    5-5

-------
Development Document for Construction and Development Proposed Effluent Guidelines	   ;	


with Municipal Separate Storm Sewer System (MS4) serving populations over 100,000,
construction sites 5 acres and larger, and certain industrial sites to obtain a permit. The permit
required the development of a stormwater pollution prevention plan (SWPPP) that typically
included a storm water and sediment control plan. In 1999, the Phase IINPDES stormwater
regulations were published, extending permit coverage to construction sites of 1 acre or larger
and municipalities to 50,000 or 10,000 population if the density is more than 1,000 per square
mile. The regulations allow use of general permits in lieu of individual site or facility permits.
The degree of oversight of construction varies widely among the states.

In the last two decades, increased concern at the local level has been focused on sediment
pollution of streams and waterways, particularly originating from construction, while less
concern has been focused on the impacts of increased construction on storm water and chemical
production. Much of this government concern originated from the Phase I and Phase II NPDES
stormwater regulations. A number of states and their local agencies have developed standards
and BMPs for sediment control, most of which do not have a scientific basis, but were adopted
from other agencies.  Some states, however, did conduct studies that gave their standards some
scientific basis. For example, Maryland evaluated its BMP standards in the 1980s by using
modeling techniques and the state changed its sediment basin standards to account for the
impacts of surface area on the trapping efficiency in sediment ponds. Based on typical soils in
the region and modeling studies, the state adopted a surface area to peak discharge ratio of 0.01
cfs/acre as a criterion (Barfield and Clar, 1985; McBurnie, 1990). Maryland was thus the first
state to use a design criterion that was related to the overflow rate.  Other states also used some
of Maryland's results (Smolen et al, 1988).

Recent efforts have moved closer to an effluent standard approach.  South Carolina conducted a
detailed analysis and published regulations that required a trapping efficiency or settleable solids
standard (SCDHEC,  1995). In addition, results from a detailed model were used to develop
simplified design aids (Hayes and Barfield, 1995; Holbrook et al., 1998). Some municipalities
are following suit to develop scientifically based standards of their own. For example, in 1998
Louisville, Kentucky (Hayes et al., 2001) developed standards and design aids for their storm
water and sediment control, following the example of South Carolina.

There are no analogs in which the integrated approach to storm water and sediment control have
been used on construction sites. The closest analog is the Cumulative Hydrologic Impact
Analysis (CHIA) required in surface mining by the SMRCA. SMRCA requires each applicant
for a surface mining permit to conduct a hydrologic impact analysis. Subsequently, the
regulatory authority is required to conduct a CHIA for the entire watershed. It should be pointed
out that although a CHIA is required, it is seldom undertaken on a scale that is useful.

Many of the advances in sediment control have been based on the capability to predict, a priori.,
the ability of a given design to meet a standard. For example, when the  settleable  solids standard
June 2002
                                                                                      5-6

-------
  Development Document for Construction and Development Proposed Effluent Guidelines       	


  was developed for surface mining, most regulatory authorities adopted it, with the requirement
  that permit applicants would demonstrate through the use of widely accepted computer models,
  that the proposed design would meet the settleable solids standard.

  Most of the early work in modeling sediment production stemmed from efforts in the 1950s to
  develop a soil loss equation that would apply to the entire nation and allow evaluation of
  alternative erosion control practices. This led to the relationship known as the Universal Soil
  Loss Equation (USLE) (Wischmeier and Smith, 1965) and its .subsequent derivative, the Revised
  USLE (RUSLE)  (Renard et al, 1994).  These efforts focus on erosion control; thus, the
  relationships do not predict sediment yield. A flurry of efforts were addressed in the late 1970s
  and early 1980s leading to the development of sediment yield relationships such as yielding the
  Modified USLE (MUSLE) by Williams (Williams, No Date), the CREAMS model (Knisel
  1980), and SEDIMOt II (Wilson et al., 1982), and its derivatives. The MUSLE and CREAMS
 models did not include methods to evaluate the impact of sediment trapping structures, but
  SEDIMOT H contained relationships developed at the University of Kentucky to predict the
 impact of reservoirs (Ward et al., 1977; Wilson et al, 1984), check dams (Hirschi, 1981), and
 vegetative filter strips (Hayes et al, 1984). The MUSLE and SEDIMOT II models were based
 on single storms while the CREAMS model was based on continuous simulation modeling.
 Details on these models can be found in Haanetal. (1994).      .          .

 More recently, modeling has improved, resulting in several new relationships. The WEPP
 watershed model  is one example of a continuous simulation approach. It includes computational
 procedures for a wide variety of sediment control structures (Lindley et al, 1998).  Another
 example of a single storm-based model is SEDIMOT HI (Barfield et al, 1996), which modifies
 the earlier SEDIMOT II model to include channel erosion routines and a wide variety of
 sediment control techniques. A significant drawback in the SEDIMOT m and WEPP models is
 that they do not have a good technique for predicting the impact of filter fence, which is the most
 common technique used today for sediment control.

 Concerns for changes in geomorphology resulting from flow changes have resulted in several
 modeling approaches. Early efforts were focused on what is known as the regime theory, in
 which changes in  channel property are linked, qualitatively, to changes in flow. Examples
 include models of Lane (1955) and Schumm (1977). In addition, some statistically based
 models were developed, but they are not universally applicable (Blench, 1970; Simons and
 Albertson, 1960).  More recently, models have been developed using physically based concepts
 to predict changes in geomorphology as related to changes in flow. The models of Chang (1988)
 are good examples. It is possible to predict, to a limited extent, the change in channel properties
 as impacted by changes in flow.

 The impact of changes in flow and geomorphology on habitat is one major area where
 information is lacking. Although this deficiency can be addressed in a qualitative manner, it is
June 2002
                                                                                   5-7

-------
Development Document for Construction and Development Proposed Effluent Guidelines	•	


not possible to predict quantitatively how a given change in geomorphology will impact habitat.
Additional information is needed to develop a strategy based on the integrated assessment
approach.

5.1.4         GOALS, CONTROL STRATEGIES, CRITERIA, AND STANDARDS

5.1.4.1           GOALS, CONTROL STRATEGIES, CRITERIA, AND STANDARDS:
                 HOW THEY RELATE

The relationship between goals, control strategies, criteria, and standards can sometimes be
confusing. For the purposes of the discussion on construction BMPs, the following definitions
will be used.

Goal. The overarching objective of having a storm water, sediment, and pollution control
program is known as the goal. It is what the program is trying to achieve. All BMPs should
relate to that goal. As stated earlier, the goal of this program is to minimize the impact of
construction on receiving water and downstream areas. The impacts of concern are identified in
the Environmental Assessment.

Control Strategies.  The methods by which the regulatory agency tries to achieve the goal are
called control strategies.

Criteria. The particular variables that are targeted by a given strategy are known as the criteria.
For example, if the strategy is to control impacts by limiting the discharge of sediment generated
to the receiving waters, then sediment becomes the criterion.                        ;

Standard. The specific variable chosen for the criteria and its numeric value is referred to as the
standard. For example, if the control strategy is to limit sediment discharge to the receiving
waters, the criterion is sediment, and the particular limiting variable and numeric value chosen is
a peak settleable solids concentration of 0.5 mg/L, then the standard would be a peak settleable
solids concentration of 0.5 mg/L.

The relationship among goals, control strategies, criteria, and standards is shown graphically in
Figure 5-1.
 June 2002
                                                                                     5-8

-------
Development Document for Construction and Development Proposed Effluent Guidelines

Criteria 1
Flow Variables*
Sediment Variables*
Geomorphology.


Criteria 2
To Be
Developed
Figure 5-1. Flow Diagram Showing Relationship Among Goals, Strategies, Criteria, and Standards
June 2002
5-9

-------
 Development Document for Construction and .Development Proposed Effluent Guidelines
 5.1.4.2
LEVELS OF PERFORMANCE OR "HOW WELL DO THE
STRATEGIES WORK?"
 Table 5-1 provides a description on the level of performance for the three strategies discussed in
 sub-section 5.2.1.

        Table 5-1. Description of Levels of Performance of Three Control Strategies
Level
0
1
2
3
4
5
Description of Performance
No consideration of impact.
Performance defined by a design standard. No guarantee that the design will control the impact to a
desired level on the specific watershed. Example: reservoir volume standard for runoff control.
Effluent standard based on controlling a single entity entering receiving waters. Control of the single
parameter will not guarantee that the desired protection will occur for receiving waters or
downstream impact. Example: controlling peak storm water discharge or peak TSS.
Effluent standard based on controlling two or more entities entering receiving waters, but not all
entities causing environmental impact. Example: controlling peak discharge and sediment, but not
storage volume or runoff volume.
Effluent standards for all entities entering receiving waters and causing environmental impact. Even
Qontrolling all quantities entering receiving waters will not guarantee that there are no undesired
downstream impacts. Example: Controlling runoff rate, runoff volume, peak discharge, and TSS in
receiving streams does not guarantee that there will be no undesirable biological impacts. i
Control based on integrated evaluation of impacts on receiving stream and downstream.
5.1.4.3 STRATEGIES, CRITERIA, STANDARDS, AND ENFORCEMENT
The effectiveness of a given strategy, criterion, or standard is directly related to the ability of an
enforcement agency to enforce the rules. Thus, a given standard may theoretically provide
excellent protection to the environment, but be so difficult to enforce that it is less effective than.
a less stringent standard that is enforceable. In general, the difficulty in enforcement increases as
the level of desired performance increases. An estimate of relative difficulty in enforcement is
given in Table 5-2 for the various levels of performance from Table 5-1. For example, it is
easiest to enforce the design standard, since enforcement is based entirely on reviewing plans
and inspection of the site to ensure that the plans are put into action properly.

Important issues related to enforcement include the following:

•   A priori demonstration by the best computational technology that the proposed design can
    meet the standard.
June 2002
                                                                                     5-10

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	
                                                En


 •   As-built inspections to verify that the installed practices match the approved plan.

 •   Self-monitoring of effluent in the case of effluent standards, with spot checks by the
    regulatory authority to make sure that evaluations are being done properly.

 •   Evaluation of downstream impacts.

 •   Clearly defined rules for monitoring the effectiveness of a practice.

              Table 5-2. Descriptions of Levels of Difficulty in Enforcement
Level of
Performance
from Table 1-1
0
1
2
3
4
5
Difficulty in
Enforcing
(Relative)
0
1
2
2.5
2.5
5
Description of Difficulty
Nothing to enforce.
Enforcement consists of reviewing plans and ensuring proper installation
and maintenance.
Enforcement requires some monitoring and typically requires a
preconstruction review of plans and submission of calculations showing
that the standard can be met.
Same as above except multiple variables.
Same as above.
Enforcement required some a priori demonstration of the expected flow
and concentration changes and their impact of the receiving waters and
downstream variables. In addition, routine monitoring of downstream
variables such as geomorphology, aquatic life, aesthetics, and riparian
zones would be required.
A Priori Demonstration of Performance:
A priori demonstration that a given design can meet the standard is very important. Experience
with the surface mining industry indicates that a sediment control plan is no better than its
design. If the best computational technology indicates that the design will not meet the standard,
then field monitoring of the BMP is not likely to show that the, standards are being achieved.
Thus, it will be important to have scientifically based and verified computational technologies to
predict the performance of BMPs relative to meeting a specified standard.     '<

In recognition of this need the USEPA funded the development of the National Stormwater BMP
Database project by the Urban Water Resources Research Council of the American Society of
Civil Engineers (ASCE, 1999) in order to establish the state of the art of BMP performance with
respect to pollutant removal and peak discharge control (level 3).  The database can be found at:
June 2002
5-11

-------
Development Document for Construction and Development Proposed Effluent Guidelines
http://www.bmpdatabase.org/. The ASCE project team prepared a report that contains several
different methods of evaluating BMP efficiency data. This report presents statistically based
approaches that involve conducting a statistical analysis to characterize inflow and outflow
EMCs, and then evaluates whether or not there is a statistically significant difference between
the two. The application of this approach in evaluating the data contained in the database has led
the study team to conclude that evaluating effluent quality is a good indicator of performance of
BMPs with respect to pollutant removal. A brief summary of the approach is provided in
Appendix A.

As-built Inspections
Another important issue related to enforcement is as-built inspections of installed practices.  •
Although the rules may call for certification by an appropriately licensed professional, it is
important that the regulatory authority conduct routine inspections to ensure that the licensed
professionals are doing their job properly.

Monitoring
Finally, there are issues related to self-monitoring versus monitoring conducted by the regulatory
authority.  The use of effluent standards would require some type of monitoring to ensure that
performance meets the standards.  However, storm water and sediment control structures that
control flows are highly variable and temporally stochastic.  This means that it is not possible to
plan ahead when the monitoring will occur. It will be necessary to have trained professionals to
conduct the monitoring.

A monitoring methodology for BMPs should meet three  criteria: (1) provide scientifically based
numbers to evaluate effectiveness, (2) be executable and sufficiently simple to allow the use of
trained technicians who would reasonably be available to do the monitoring, and (3) be  adequate
to ensure that the desired standards are met without excessive sampling or analysis.  The,flrst
criterion could be met by providing clear documentation on the monitoring methodology that
specifies times, frequency, and location of sampling relative to storms, as well as clearly
articulated protocols for handling samples. The second criteria can be met by being sure that the
techniques proposed have actually been field applied by technicians in the monitoring business.
The third criterion can be evaluated by an error analysis that determines the expected accuracy of
measurement as a function of number and frequency of sampling.

Several possible criteria or standards have special measurement problems that should be
mentioned. These include criteria or standards based on trapping efficiency, and/or effluent TSS
and settleable solids (average or peak). The issues associated with these criteria are discussed
below.
June 2002
                                                                                     5-12

-------
 Development Document for Construction and Development Proposed Effluent Guidelines
 Trapping Efficiency. Literature citations frequently include studies that attempt to measure
 trapping efficiency by sampling one or more inflow and outflow concentrations (Barrett et al.,
 1995). While this simplicity seems attractive, it is a grossly erroneous measure of trapping
 efficiency. A correct definition of trapping efficiency is given in Equation 1:
       Equation 1:
              where: Mi is inflow total mass
                    M0 is outflow total mass
                    Mt is given by integrating the product of inflow concentration and inflow
                    rate over the duration of a hydrograph
                           or
Equation 2:
                                 'D
                           Mt = J
              where: C; is inflow concentration
                    q} is inflow flow rate
                    t is time
                    tD is the duration of the storm.

Outflow total mass M0 is calculated by substituting the subscript o for i in Equation 2.  Thus, to
monitor trapping efficiency correctly, it is necessary to measure both flow and concentration as a
function of time over the duration of both inflow and outflow. Such measurement is quite
difficult and time-consuming, requiring many samples.

Statistical Evaluation of Inflow/Outflow Data (mean, median, standard deviation,
coefficient of variance). To measure average or peak TSS, it is necessary to measure TSS in the
effluent over the duration of the outflow hydrograph as well as the flow rate. This requires that
multiple samples be taken and that the samples be centered around the peak discharge.  The
ACSE database data analysis document has the ability, depending upon the number of samples
collected, to show a difference between various samples. Again, this is time-consuming and
difficult since the timing of an event and the timing of the peak discharge are not known a priori.
The average concentration is a weighted concentration, using flow rate as a weighting function.
June 2002
                                                                            5-13

-------
Development Document for Construction and Development Proposed Effluent Guidelines
5.1.5
CONTROL TECHNIQUES, BMP SYSTEMS
5.1.5.1
   EROSION CONTROL AND PREVENTION
5.1.5.1.1
       PLANNING, STAGING, SCHEDULING
General Description

A construction sequence schedule is a specified work schedule that coordinates the timing of
land-disturbing activities and the installation of erosion and sediment control measures. The
goal of a construction sequence schedule is to reduce on-site erosion and off-site sedimentation'
by performing land-disturbing activities and installing erosion and sediment control practices in
accordance with a planned schedule (Smolen et al., 1988).

Construction site phasing involves disturbing only part of a site at a time to prevent erosion from
dormant parts (Claytor, 1997).  Grading activities and construction are completed and soils are
effectively stabilized on one part of the site before grading  and construction commence at
another part. This differs from the more traditional practice of construction site sequencing, in
which construction occurs at only one part of the site at the time, but site grading and other
site-disturbing activities typically occur simultaneously, leaving portions of the disturbed site
vulnerable to erosion. Construction site phasing must be incorporated into the overall site plan
early on. Elements to consider when phasing construction activities include the following
(Claytor, 1997):

•  Managing runoff separately in each phase.
•  Determining whether water and sewer connections and extensions can be accommodated.
•  Determining the fate of already completed downhill phases.
•  Providing separate construction and residential accesses to prevent conflicts between
   residents living in completed stages of the site and construction equipment working on later
   stages (USEPA, 2000).

Applicability

Construction sequencing can be used to plan earthwork and erosion and sediment control
activities at sites where land disturbances might affect water quality in a receiving waterbody.
June 2002
                                                                       5-14

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	_^

 Design and Installation Criteria

 Construction sequencing schedules should, at a minimum, include the following (NCDNR  1988'
 MDE, 1994):                                                     '

 •   The erosion and sediment control practices that are to be installed

 •   The principal development activities

 •   The measures that should be installed before other activities are started -

 •   The compatibility with the general contract construction schedule
June 2002
                                                                                      5-15

-------
Development Document for Construction and Development Proposed Effluent Guidelines
Table 5-3 summarizes other important scheduling considerations in addition to those listed
above.
Table 5-3. Scheduling Considerations for Construction Activities
Construction Activity
Construction survey stakeout
Pre-construction meeting between owner,
contractor and regulatory agency
Construction access — entrance to site,
construction routes, areas designated for
equipment parking
Clearing and grading required for the
installation of controls •
Sediment traps and barriers — basin traps,
silt fences, outlet protection
Runoff control — diversions, perimeter
dikes, water bars, outlet protection
Runoff conveyance system— stabilize
streambanks, storm drains, channels, inlet
and outlet protection, slope drains
Land clearing and grading — site
preparation (cutting, filling, and grading;
sediment traps; barriers; diversions; drains;
surface roughening)
Surface stabilization— temporary and
permanent seeding, mulching, sodding,
riprap
Building construction — buildings, utilities,
paving
Landscaping and final
stabilization — adding top soil, trees, and
shrubs; permanent seeding; mulching;
sodding; riprap
Schedule Consideration
Prior to initiating any construction activity a construction survey stakeout
should be conducted. The stakeout should identify the limits of disturbance,
' and location of control structures, especially perimeter controls
This meeting should take place before any construction activity begins at the
site. The survey stakeout is reviewed, especially the limits of disturbance and
location of controls
This is the first land-disturbing activity. As soon as construction takes place,
stabilize any bare areas with gravel and temporary vegetation.
In conjunction with the construction access, the clearing and grading required
for the installation of E&S controls should take place.
After construction site has been accessed, install principal basins, with the
addition of more traps and barriers as needed during grading.
Install key practices after the installation of principal sediment traps and
before land grading. Additional runoff control measures may be installed
during grading.
If necessary, stabilize streambanks as soon as possible, and install principal
runoff conveyance system with runoff control measures. The remainder of
the svstems may be installed after grading. •
Implement major clearing and grading after installation of principal sediment
and key runoff control measures, and install additional control measures as
grading continues. Clear borrow and disposal areas as needed, and mark trees
and buffer areas for preservation.
Immediately apply temporary or permanent stabilizing measures to any
disturbed areas where work has been either completed or delayed.
During construction, install any erosion and sedimentation control measures
that are needed.
This is the last construction phase. Stabilize all open areas, including borrow
and spoil areas, and remove and stabilize all temporary control measures.
Effectiveness

Construction sequencing can be an effective tool for erosion and sediment control because it
ensures that management practices are installed where necessary and when appropriate. A
comparison of sediment loss from a typical development and from a comparable phased project
showed a 42 percent reduction in sediment export in the phased project (Claytor, 1997).
June 2002
                                                                                     5-16

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

 Limitations

 Weather and other unpredictable variables may affect construction sequence schedules. The
 proposed schedule and a protocol for making changes resulting from unforseen problems should
 be plainly stated in an applicable erosion and sediment control plan.

 Maintenance

 The construction sequence should be followed throughout the project, and the written erosion
 and sediment control plan should be modified before any changes in construction activities are
 executed. The plan can be updated if a site inspection indicates the need for additional erosion
 and sediment control as determined by contractors, engineers, or developers.

 Cost

 Construction sequencing is a low-cost BMP because it requires a limited amount of a
 contractor's time to provide a written plan for the coordination of construction activities and
 management practices. Additional time might be needed to update the sequencing plan if the
 current plan is not providing sufficient erosion and sediment control.

 Although little research has been done to assess the costs of phasing versus conventional
 construction costs,  it is known that it will be to implement successful phasing for a larger project
 (Claytor, 1997).
 5.1.5.1.2
VEGETATIVE STABILIZATION
Vegetation can be used during construction to stabilize and protect soil exposed to the erosive
forces of water, as well as during post-construction to provide a filtration mechanism for storm
water runoff pollutants. The following discussion refers to vegetative stabilization as a
construction BMP that stabilizes and protects soil from erosion.

General Description

Vegetative stabilization measures employ plant material to protect soil exposed to the erosive
forces of water and wind. Selected vegetation can reduce erosion by more than 90 percent
(Fifield,  1999). Natural plant communities that are adapted to the site provide a self-maintaining
cover that is less expensive than structural alternatives. Plants provide erosion protection to
vulnerable surfaces by the following (Heyer, n.d.):

•   Protecting soil surface from the impact of raindrops.

•   Holding soil particles in place.
June 2002
                                                                                      5-17

-------
Development Document for Construction and Development Proposed Effluent Guidelines           •,	

•   Maintaining the soil's capacity to absorb water.

•   Using living root systems to hold soil in place, increasing overall bank stability.

•   Directing flow velocity away from the streambank.

•   Acting as a buffer against abrasive transported materials.

•   Causing sediment deposition, which reduces sediment load and reestablishes the streambank.

The designer should be aware of and respond to local conditions that may influence the
development of vegetative stabilization measures. As with any planting design, climate,
maintenance practices, the availability of plant material (including native species), and many
other factors will influence such considerations as plant or seed mix selection, installation
methods, and project scheduling.

Slope Stabilization. On slopes, the goal of vegetative stabilization is not only to redupe surface
erosion but also to prevent slope failure. Vegetation should provide dense coverage to protect
soils from the direct impact of precipitation and help intercept runoff. A variety of plants should
be used to provide root systems that are distributed throughout all levels of the soil, increasing
slope shear strength and giving plants a greater ability to remove soil moisture. Uniform mats of
shallow rooting plants should be avoided because, while such plants may increase runoff
infiltration, they cannot remove soil moisture beyond the surface level, leaving slopes potentially
saturated and prone to slippage. Shallow, interlocking root systems may also increase the size of
a soil slippage by holding together and pulling down a larger area of slope after a small section
has given way. Large trees that have become unstable may also pull down slopes and should be
removed.  Using plants with low water requirements can reduce the potential for soil saturation
from irrigation.

Swale Stabilization.  On swales, the goal of vegetative stabilization is to prevent erosion within
the swale, where runoff is concentrated and flows at higher velocities. If natural stream channels
are involved, vegetation with deep root systems should be preserved, or if absent, planted above
the channel to help maintain the channel banks.  More information is provided in the subsequent
section dealing with grass-lined swales.

Surface Stabilization. On large, flat areas, the goal of vegetative stabilization is to reduce the
loss of surface soil from sheet erosion. Vegetation should provide complete coverage to reduce
the force of precipitation, which can shift soil  particles to seal openings in the soil, reducing
infiltration and increasing runoff. Vegetation should also provide many stem penetrations to
slow runoff and increase infiltration. Deep rooting plants are less critical for erosion control in-
flat areas than on slopes because soils are not  subject to the same forces that may cause slippage
on a slope. However, trees and shrubs can increase infiltration, lessening the buildup of runoff,
and transpire large volumes of water, reducing soil saturation.
June 2002
                                                                                      5-18

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

 In areas susceptible to wind erosion, the goal of vegetative stabilization is to establish direct
 protection of the soil. Vegetation should provide dense and continuous surface cover. Binding
 the soil deeply is generally not a requirement. The ideal vegetation for this purpose is grass,
 which forms a mat of protection. In areas where the vegetation is developed, the grass generally
 has high maintenance requirements. In less developed, open areas, unmown grass, including
 perennial native species, can be used to provide protection. Trees and shrubs also can provide
 protection from the wind.

 Shoreline Stabilization.  In lakes and ponds, the goal of vegetative stabilization is to prevent
 erosion of the shoreline. Wetland plants anchor the bottom of the lake or pond adjacent to the
 shore and help dissipate the erosive energy of waves. An important consideration in planting
 along shorelines is the need to establish favorable conditions for plant establishment and growth.
 These include the proper grading of side slopes and the control of upland erosion to prevent the
 buildup of silt and associated pollutants in the water.  Designers should maintain awareness of
 regulatory requirements that may influence vegetation projects  in a wetland environment
 (USAF,1998)..

 Vegetation used for shoreline stabilization work should be native material selected on the basis.
 of strength, resiliency,  vigor, and ability to withstand periodic inundation.  Woody vegetation
 with short, dense, flexible tops and large root systems works well. Other important factors
 include rapid initial growth, ability to reproduce, and resistance to disease and insects.

 According to Heyer, n.d., most strearribank stabilization plantings have used various willows,
 including black willow (Salix nigrd), sandbar willow (S. interior), meadow willow (S.
petiolaris), heartleaf willow (S. rigida), and Ward willow (S.  caroliniand). The size used
 depends on the severity of the erosion and the type of bank to be stabilized. Whatever the size, it
 is important to use dormant cuttings and to remove all lateral branches.  Most tree revetment
 projects used either eastern red cedar (Juniperus virginiand) or hardwoods such as northern pin
 oak (Quercus ellipsoidalis). Important suggestions include the  following:

 •  Choose trees with many limbs and branches to trap as much sediment as  possible.

 •  Select decay-resistant trees.

 •  Use recently cut trees—dead trees are more brittle and likely to break apart.

 •  The tree size-diameter of the tree crown should be about two-thirds of the height of the
   eroding bank.

 •  Gut off any trunk without limbs.
 •  Place the tree revetments overlapping, butt end pointing upstream.

 •  Begin and end revetments at stable points along the bank.
June 2002
                                                                                      5-19

-------
Development Document for Construction and Development Proposed Effluent Guidelines	


•  Choose an anchoring system according to the bank material to be stabilized and the weight of
   the object to be anchored.

Vegetative measures for streambank stabilization offer an alternative to structural measures and
are becoming well known as bioengineering techniques for streambanks. Utilizing vegetative
material for streambank stabilization could be the first step in the reestablishment of the riparian
forest, which is essential for long-term stability of the streamside and floodplain areas. Each site
must be evaluated separately as to the feasibility of using natural material (Heyer, n.d.).

Vegetative streambank stabilization, with the goal to protect streambanks from the erosive forces
of flowing water, is generally applicable where bankfull flow velocity does not exceed 6 ft/sec
and soils are erosion resistant (Smolen, 1988).  Table 5-4 includes general guidelines for
maximum allowable velocities in streams to be protected by vegetation.
Table 5-4. Conditions Where Vegetative Streambank Stabilization Is Acceptable
Frequency of Bankfull Flow
> 4 times/yr
1 to 4 times/yr

Maximum Allowable Velocity for
Highly Erodible Soil
4 ft/sec
5 ft/sec
6 ft/sec
Maximum Allowable Velocity for
Erosion-Resistant Soil ,
5 ft/sec
6 ft/sec ,
6 ft/sec
Source: Smolen, 1988.  '

Temporary Vegetative Stabilization.  Temporary vegetative cover such as rapidly growing
annuals and legumes can be used to establish a temporary vegetative cover.  Such covers are
recommended for areas that (Fifield, 1999):

•   Will not be brought to final grade within 30 days or are likely to be redisturbed.

•   Require seeding of cut and fill slopes under construction.

•   Require stabilization of soil storage areas and stockpiles.

•   Require stabilization of temporary dikes, dams, and sediment containment systems.  .

•   Require development of cover or nursery crops to assist with establishing perennial grasses.

Examples of temporary vegetation include wheat, oats, barley, millet, and sudan.  Temporary
seeding may not be effective in arid or semi-arid regions where seasonal conditions (lack of
moisture) prevent germination. It may be necessary to use a mixture of warm and cool season
grasses to ensure germination. Mulching and geotextiles can be used to help provide temporary
stabilization with vegetation, particularly hi situations where establishing cover may be difficult.
June 2002
                                                                                      5-20

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

 Permanent Vegetative Stabilization. Permanent vegetative cover such as a perennial grass or a
 legume cover can be used to establish a permanent vegetative cover. Permanent vegetation is
 recommended for (Fifield, 1999)

 •   Final graded or cleared areas where permanent vegetative cover is needed to stabilize the soil

 •   Slopes designated to be treated with erosion control blankets

 •   Grass-lined channels or waterways designed to be channel liners

 The following sub-sections discuss the various types or means of providing vegetative
 stabilization.
 5.1.5.1.2.1
 General Description
GRASS-LINED CHANNELS
 Grass-lined channels, or swales, convey storm water runoff through a stable conduit.  Vegetation
 lining the channel reduces the flow velocity of concentrated runoff.  Grassed channels are
 usually not designed to control peak runoff loads by themselves and are often used in
 combination with other BMPs such as subsurface drains and riprap stabilization.

 Applicability

 Grassed channels should be used in areas where erosion-resistant conveyances are needed, such
 as in areas with highly erodible  soils and slopes of less than 5 percent. They should be installed
 only where space is available for a relatively large cross-section.  Grassed channels have a
 limited ability to control runoff from large storms and should not be used in areas where velocity
 exceeds 5 feet per second unless they are on erosion-resistant soils with dense groundcover at the
 soil surface.

 Design and Installation Criteria
                      )

 Because of their ease of construction and low cost, vegetated-lined waterways are frequently
 used on diversion and collection ditches. USDA's Soil Conservation Service's (SCS)
Engineering Field Manual (1979) recommends the following maximum permissible velocities
 for individual site conditions shown in Table 5-5.
June 2002
                                                                                     5-21

-------
Development Document for Construction and Development Proposed Effluent Guidelines

              Table 5-5. Maximum Permissible Velocities for Individual Site
                               Conditions for Grass Swales
Site Location
Areas where only a sparse cover can be established or
maintained because of shale, soils, or climate
If the vegetation is to be established by seeding
Areas where a dense, vigorous sod is obtained quickly
or where the runoff can diverted out of the waterway
while the vegetation is being established
Velocity
3.00 ft/sec (0.91 m/sec)
3.00 to 4.00 ft/sec (0.91 to 1.22 m/sec)
4.00 to 5.00 ft/sec (1.22 to 1.52 m/sec)
Source: USDA, 1979

Grassed waterways typically begin eroding in the invert of the channel if the velocity exceeds
the sheer strength of the vegetation soil interface. Once the erosion process has started, it will
continue until an erosion-resistant layer is encountered. If erosion of a channel bottom is \
occurring, rock or stone should be placed in the eroded area or the design should be changed
(UNEP, 1994).

Grassed waterways on construction land must be able to carry peak runoff events from snpwmelt
and rainstorms (in some areas limited to up to 1 cubic meter of water per second). The size of the
waterway depends on the size of the area to be drained. A typical grassed waterway cross-section
is parabolic-shaped with a nearly flat-bottomed channel, a bottom width of 3 m and channel
depth of at least 30 cm. Side slopes usually rise about 1 m for every 10 m horizontal distance but
may be as steep as a 1 m rise for every 2 m of horizontal distance. The waterway should follow
the natural drainage path if possible (Vanderwel, 1998). The design should be site-specific and
use available, well-established procedures.      .

Lined channels are a means of dropping water to lower elevations along steep parts of a
waterway. Those portions of the waterway are precisely shaped and carefully lined with heavy-
duty erosion control matting, a type of geotextile product. The lining is covered with a layer of
soil and seeded to grass. The resulting -channel is highly resistant to erosion. Lined channels are
appropriate  for waterways that only carry water occasionally and have slopes of up to 10 percent.
Companies  that sell geotextile products provide detailed information on installation of their
products (Vanderwel and Abday, 1998). The design should be site-specific, using well-
established procedures.  No standard procedure is available for evaluating the effectiveness of
geotextile liners for pollutant removal.

Grass-lined channels should be sited in accordance with the natural drainage system and should
not cross ridges.  The channel design should not have sharp curves or significant changes in
slope. The  channel should not receive direct sedimentation from disturbed areas and should be
sited only on the perimeter of a construction site to convey relatively clean storm water runoff.
They should be separated from disturbed areas by a vegetated buffer or other BMP to reduce
sediment loads.
 June 2002

-------
  Development Document for Construction and Development Proposed Effluent Guidelines 	

  Although exact design criteria should be based on local conditions, basic design
  recommendations for grassed channels include the following:

  •   Construction and vegetation of the channel should occur before grading and paying activities
     begin.

  •  - Design velocities should be less than 5 ft/sec.

  •   Geotextiles can be used to stabilize vegetation until it is fully established.

  •   Covering the bare soil with sod or geotextiles can provide reinforced storm water
     conveyance immediately.

 •   Triangular-shaped channels  should be used with low velocities and small quantities of
     runoff; parabolic grass channels are used for larger flows and where space is available;
     trapezoidal channels are used with large flows of low velocity (low gradient).

 •   Outlet stabilization structures might be needed if the runoff volume or velocity has the
     potential to exceed the capacity of the receiving area.

 •   Channels should be designed to convey runoff from a 10-year storm without erosion.

 •   The sides of the channel should be sloped less than 3:1, with V-shaped channels along roads
     sloped 6:1 or less for safety.

 •  All trees, bushes, stumps, and other debris should be removed during construction.

 Effectiveness

 Grass-lined channels can effectively transport storm water from construction areas if they are
 designed for expected flow volumes and velocities and if they do not receive sediment directly
 from disturbed areas. The primary function is to carry the flow at a higher velocity without
 eroding or overtopping the channel.

 Limitations

 Grassed channels, if improperly installed, can alter the natural flow of surface water and have
 adverse impacts  on downstream waters. Additionally, if the design capacity is exceeded by a
 large storm event, the vegetation might not be sufficient to prevent erosion and the channel
 might be destroyed. Clogging with sediment and debris reduces the effectiveness of grass-lined
 channels for storm water conveyance.
June 2002
                                                                                      5-23

-------
Development Document for Construction and Development Proposed Effluent Guidelines	__	

Maintenance                                                            ,

Maintenance requirements for grass channels are relatively minimal. During the vegetation
establishment period, the channels should be inspected after every rainfall.  Other maintenance
activities that should be carried out after vegetation is established are mowing, litter removal,
and spot vegetation replacement. The most important objective in the maintenance of grassed
channels is the maintaining of a dense and vigorous growth of turf.  Periodic cleaning of
vegetation and soil buildup in curb cuts is required so that water flow into the channel is;
unobstructed. During the growing season, channel grass should be cut no shorter than the level
of design flow, and the cuttings  should be removed promptly.

Cost

Costs of grassed channels range according to depth, with a 1.5-foot-deep, 10-foot-wide grassed
channel estimated at between $6,395 and $17,075 per trench, while a 3.0-foot-deep, 21-foot-
wide grassed channel is estimated at $12,909 to $33,404 per trench (SWRPC, 1991).

As an alternative cost approximation, grassed channel construction costs can be developed using
unit cost values. Shallow trenching (1 to 4 feet deep) with a backhoe in areas not requiring
dewatering can be performed for $4 to $5 per cubic yard of removed material (R. S. Means,
2000). Assuming no disposal costs (i.e., excavated material is placed on either side of the
trench), only the cost of fine grading, soil treatment, and grassing (approximately $2 per square
yard of earth surface area) should be added to the trenching cost to approximate the total
construction cost. Site-specific hydrologic analysis of the construction site is necessary to
estimate the channel conveyance requirement, however, it is not unusual to have flows on the
order of 2 to 4 cfs per acre served. For channel velocities between 1 and 3 feet per secorid, the
resulting range in the channel cross-section area can be as low as 0.67 square foot per acre
drained to as high as 4  square feet per acre. If the average channel flow depth is 1 foot, then the
low estimate for grassed channel installation is .$0.27 per square foot of channel bottom per acre
served per foot of channel length. The high estimate is $ 1.63 per square-foot of channel bottom
per acre served per foot of channel length.

5.1.5.1.2.2           SEEDING

General Description

Permanent seeding, is used to control runoff and erosion on disturbed areas by establishing
perennial vegetative cover from seed.  It is used to reduce erosion, decrease sediment yields from
disturbed areas, and provide permanent stabilization.  This practice is both economical and
adaptable to different site conditions, and it allows selection of the most appropriate plant
materials. Seeding is a best management practice that is particularly susceptible to local
conditions such as the  climatic  conditions, physical and chemical characteristics of the soil,
topography, and time of year.
 June 2002
                                                                                      5-24

-------
Development Document for Construction and Development Proposed Effluent Guidelines	]	

Applicability

Permanent seeding is well-suited in areas where permanent, long-lived vegetative cover is the
most practical or most effective method of stabilizing the soil. Permanent seeding can be used
on roughly graded areas that will not be regraded for at least a year. Vegetation controls erosion
by protecting bare soil surfaces from displacement by raindrop impacts and by reducing the
velocity and quantity of overland flow.  The advantages of seeding over other means of
establishing plants include lower initial costs and labor inputs.

Design and Installation Criteria

Areas to be stabilized with permanent vegetation must be seeded or planted 1 to 4 months after
the final grade is achieved unless temporary stabilization measures are in place. Successful plant
establishment can be maximized with proper planning; consideration of soil characteristics;
selection of plant materials that are suitable for the site; adequate seedbed preparation, liming,
and fertilization; timely planting; and regular maintenance. Climate, soils, and topography are
major factors that dictate the suitability of plants for a particular site. The soil on a disturbed site
might require amendments to provide sufficient nutrients for seed germination and seedling
growth. The surface soil must be loose enough for water infiltration and root penetration. Soil
pH should be between 6.0 and 6.5 and can be increased with liming if soils are too acidic. Seeds
can be protected with mulch to retain moisture, regulate soil temperatures, and prevent  erosion
during seedling establishment.

Seedbed preparation is critical in established vegetation. Spraying seeds on a scraped slope,will
generally not provide satisfactory results. Typical seedbed preparation will begin with a soil test
to determine title amount of lime or fertilizer that should be added. In addition, tillage should be
performed that will break up clods so that seed contact can be established. When the seed is
applied, it should be covered and lightly compacted. An appropriate natural or synthetic mulch is
recommended to provide surface stabilization until the vegetation is established. In addition to
providing surface stabilization, the mulch will also retard evaporation and encourage rapid   ~
growth. A suitable tack to hold the mulch may be necessary if the mulch is not otherwise
anchored. Mulches are covered in a subsequent sub-section.

Depending on the amount of use permanently seeded areas receive, they can be considered high-
or low-maintenance areas. High-maintenance areas are mowed frequently, limed and fertilized
regularly, and either (1) receive intense use (for example, athletic fields) or (2) require
maintenance to an aesthetic standard (for example, home lawns). Grasses used for high-
maintenance areas are long-lived perennials that form a tight sod and are fine-leaved.
High-maintenance vegetative cover is used for homes, industrial parks, schools, churches, and
recreational areas.

Low-maintenance areas are mowed infrequently or not at all and do not receive lime or fertilizer
on a regular basis. Plants must be able to persist with minimal maintenance over long periods of
June 2002
                                                                                       5-25

-------
 Development Document for Construction and Development Proposed Effluent Guidelines    	|	

 time.  Grass and legume mixtures are favored for these sites because legumes fix nitrogen from
 the atmosphere. Sites suitable for low-maintenance vegetation include steep slopes, streambanks
 or channel banks, some commercial properties, and "utility" turf areas such as road-banks.

 Effectiveness

 Seeding that results in a successful stand of grass has been shown to remove between 50 and 100
 percent of total suspended solids from storm water runoff, with an average removal of 90 percent
 (USEPA, 1993).

 Limitations                .

 The effectiveness of permanent seeding can be limited because of the high erosion potential
 during establishment, the need to reseed areas that fail to establish, limited seeding times
 depending on the season, and the need for stable soil temperature and soil moisture content
 during germination and early growth. Permanent seeding does not immediately stabilize
 soils—temporary erosion and sediment control measures should be in place to prevent off-site
 transport of pollutants frofn disturbed areas.  Use of mulches and/or geotextiles may improve the
 likelihood of'successfully establishing vegetation.

 Maintenance  .

 Grasses should emerge within 4 to 28 days and legumes 5 to 28 days after seeding, with legumes
 following grasses.  A successful stand should exhibit the following:                     ,

 •   Vigorous dark green or bluish green seedlings—not yellow

 •   Uniform density, with nurse plants,  legumes, and grasses well intermixed

 •   Green leaves—perennials remaining throughout the summer, at least at the plant bases',

 Seeded areas should be inspected for failure, and necessary repairs and reseeding should be
 made as soon as possible. If a stand has inadequate cover, the choice of plant materials and
 quantities of Lime and fertilizer should be reevaluated. Depending on the condition of the stand,
 areas can be repaired by overseeding or reseeding after complete seedbed preparation.  If the
 timing is bad, an annual grass seed can be overseeded to temporarily thicken the stand until a
 suitable time for seeding perennials. Consider seeding temporary, annual species if the season is
 not appropriate for permanent seeding.  If vegetation fails to grow, the soil should be tested to
 determine whether low pH or nutrient imbalances are responsible.  Local NRCS or county
 extension agents can also be contacted for seeding and soil testing recommendations.

 On a typical disturbed site, full plant establishment usually requires refertilization in the second
growing season.  Soil tests should be used to determine whether more fertilizer needs to be
June 2002
                                                                                     5-26

-------
Development Document for Construction and Development Proposed Effluent Guidelines	

added.  Do not fertilize cool season grasses in late May through July. Grass that looks yellow
may be nitrogen deficient. Nitrogen fertilizer should not be used if the stand contains more than
20 percent legumes.                                    .            .

Cost

Seeding costs range from $200 to $1,000 per acre and average $400 per acre.  Maintenance costs
range from 15 to 25 percent of initial costs and average 20 percent (USEPA, 1993). R. S. Means
(2000) indicates the cost of mechanical seeding to be approximately $900 per acre, and
demonstrates that the coverage cost varies with the seed type, seeding approach and scale (total
acreage to be seeded).  For example, hydro or water-based seeding for grass is estimated to be,
$700 per acre but seeding of "field" grass species is only $540 per acre (Costs include materials,
labor, and equipment, with profit and overhead).  If surface preparation is required;, then the
installation costs increase. R. S. Means suggests the cost of fine grading, soil treatment, and
grassing is approximately $2 per square yard of earth surface area.
5.1.5.1.2.3
SODDING
General Description

Sodding is a permanent erosion control practice that involves laying a continuous cover of grass
sod on exposed soils. In addition to stabilizing soils, sodding can reduce the velocity of storm
water runoff. Sodding can provide immediate vegetative cover for critical areas and stabilize
areas that cannot be vegetated by seed. It can also stabilize channels or swales that convey
concentrated flows and reduce flow velocities.  While sodding is not as dependent as seeding on
local conditions, it does depend on soil and climatic conditions to be successful. Capability to
water immediately after installation and occasionally until establishment is generally beneficial.

Applicability

Sodding is appropriate for any graded or cleared area that might erode, requiring immediate
vegetative cover.  Locations particularly well-suited to sod stabilization are:

•  Waterways and channels carrying intermittent flow

•  Areas around drop inlets that require stabilization

•  Residential or commercial lawns and golf courses where prompt use and aesthetics are
   important

•  Steeply sloped areas
June 2002
                                                                 5-27

-------
Development Document for Construction and Development Proposed Effluent Guidelines	

Design and Installation Criteria

Sodding eliminates the need for seeding and mulching and produces more reliable results with
less maintenance. Sod can be laid during times of the year when seeded grasses can faili The
sod must be watered frequently within the first few weeks of installation.. Some seedbed
preparation is recommended, including smoothing to provide contact between the sod and the
soil surface and soil testing to determine liming and fertilizer application rates. Since sod
provides instantaneous cover, mulches are not typically recommended, but anchoring may be
appropriate on steep slopes.

The type of sod selected should be composed of plants adapted to site conditions. Sod
composition should reflect environmental conditions as well as the function of the area where
the sod will be laid. The sod should be of known genetic origin and be free of noxious weeds,
diseases, and insects.  The sod should be machine cut at a uniform soil thickness of 15 to 25 mm
at the tune of establishment (this does not include top growth or thatch).  Soil preparation and
addition of lime and fertilizer may be needed—soils should be tested to determine whether
amendments are needed. Sod should be laid in strips perpendicular to the direction of water flow
and staggered in a brick-like pattern.  The corners and middle of each strip should be stapled
firmly. Jute or plastic netting may be pegged over the sod for further protection against washout
during establishment.

Areas to be sodded should be cleared of trash, debris, roots, branches, stones, and clods larger
than 2 niches in diameter.  Sod should be harvested, delivered, and installed within a period of
36 hours.  Sod not transplanted within this period should be inspected and approved prior to its
installation.

Limitations

Compared to seed, sod is more expensive and more difficult to obtain, transport, and store. Care
must be taken'to prepare the soil and provide adequate moisture before, during, and after
installation to ensure successful establishment.  If sod is laid on poorly prepared soil or
unsuitable surface, the grass will die quickly because it is unable to root.  Sod that is not
adequately irrigated after installation may cause root dieback because grass does not root rapidly
and is subject to drying out.

Effectiveness

Sod has been shown to remove between 98 and 99 percent of total suspended solids in runoff
(USEPA, 1993). It is therefore a highly effective management practice for erosion and sediment
control.
June 2002
                                                                                     5-28

-------
Development Document for Construction and Development Proposed Effluent Guidelines	

Maintenance

Watering is very important to maintain adequate moisture in the root zone and to prevent
dormancy, especially within the first few weeks of installation, until it is fully rooted. Mowing
should not result in the removal of more than one-third of the shoot. Grass height should be
maintained at between 2 and 3 inches. After the first growing season,  sod might require
fertilization or liming. Permanent, fine turf areas require yearly maintenance fertilization.
Warm-season grass should be fertilized in late spring to early summer, and cool-season grass in
late winter and again in early fall.

Cost

Average construction costs of sod average $0.20 per square foot and range from $0.10 to $1.10
per square foot; maintenance costs are approximately 5 percent of installation costs (USEPA,
1993). R. S. Means (2000) indicates the sodding ranges between $250 and $750 per 1000 square
feet for 1" deep bluegrass sod on level ground, depending on the size of the area treated (unit
costs value are for orders over 8,000 square feet and less than 1000 square feet, respectively).
Bent grass sod values range between $350 and $500 per 1000 square feet, again the lower value
is more likely for most construction sites because it is for large area applications. (Costs include
materials, labor, and equipment, with profit and overhead).
5.1.5.1.2.4
MULCHING
General Description

Mulching is a temporary erosion control practice in which materials such as grass, hay, wood
chips, wood fibers, straw, or gravel are placed on exposed or recently planted soil surfaces.
Mulching is highly recommended as a stabilization method and is most effective when anchored
in place until vegetation is well established. In addition to stabilizing soils, mulching can reduce
the velocity of storm water runoff.  When used in combination with seeding or planting,
mulching can aid plant growth by holding seeds, fertilizers, and topsoil in place; by preventing
birds from eating seeds;  by retaining moisture; and by insulating plant roots against extreme
temperatures.

Mulch mattings are materials such as jute or other wood fibers that are formed into sheets and
are more stable than loose mulch.  They can also be easily unrolled during the installation
process and are particularly useful in steeper areas  or in channels. Netting can be used to
stabilize soils while plants are growing, although netting does not retain moisture or insulate
against extreme temperatures.  Mulch binders consist of asphalt or synthetic materials that are
sometimes used instead of netting to bind loose mulches but have been found to have limited
usefulness.
June 2002
                                                                                      5-29

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

 Applicability

 Mulching is often used in .areas where temporary seeding cannot be used because of
 environmental constraints. Mulching can provide immediate, effective, and inexpensive erosion
 control. On steep slopes and critical areas such as waterways, mulch matting is used with
 netting or anchoring to hold it in place. Mulches can be used on seeded and planted areas where
 slopes are steeper than 2:1  or where sensitive seedlings require insulation from extreme
 temperatures.

 Design and Installation Criteria

 When possible, organic mulches should be used for erosion control and plant material
 establishment. Suggested materials include loose straw, netting, wood cellulose, or agricultural
 silage. All materials should be free of seed, and loose hay or .straw should be anchored by
 applying tackifier, stapling netting over the top, or crimping with a mulch crimping tool.
 Materials that are heavy enough to stay in place do not need anchoring (for example, gravel).
 Steepness of the slope will also affect the extent of anchoring the mulch.  Other examples
 include hydraulic mulch products with 100 percent post-consumer paper content, yard trimming
 composts, and wood mulch from recycled stumps and tree parts. Inorganic mulches such as pea
 gravel or crushed granite can be  used in unvegetated areas.

 Mulches may or may not require a binder, netting, or tacking. All straw and loose materials
 must have a binder to hold them hi place.  Mulch materials that float away during storms can
•clog drainage ways and lead to flooding.  The extent of binding depends on the type of mulch
 applied. Effective use of netting and matting material requires firm, continuous contact between
 the materials and the soil.  If there is no contact, the material will not  hold the soil and erosion
 will occur underneath the material.  Grading is not necessary before mulching.

 There must be adequate coverage, or erosion, washout, and poor plant establishment will result.
 If an appropriate tacking agent is not applied, or if it is applied in an insufficient amount, mulch
 will not withstand wind and runoff. The channel grade and liner must be appropriate for the
 amount of runoff, or the channel bottom will erode.  Also, hydromulch should be applied in
 spring, summer, or fall to prevent deterioration of the mulch before plants can become
 established. Table 5-6 presents guidelines for installing mulches, but local conditions may
 warrant additional requirements.
June 2002
5-30

-------
 Development Document for Construction and Development Proposed Effluent Guidelines

               Table 5-6. Typical Mulching Materials and Application Rates
Material
Rate per
Acre
Requirements
Notes
Organic Mulches .
Straw
Wood fiber or wood
cellulose
Wood chips
Bark
1-2 tons
0.5-1 ton
5-6 tons
35yd3
Dry, unchopped, unweathered; avoid
weeds.

Air dry. Add fertilizer N, 12 Ib/ton.
Air dry, shredded or hammermilled,
or chips.
Spread by hand or machine; must be
tacked or tied down.
Use with hydroseeder; may be used to
tack straw. Do not use in hot, dry
weather.
Apply with blower, chip handler, or
by hand. Not for fine turf areas.
Apply with mulch blower, chip
handler, or by hand. Do not use
asphalt tack.
Nets and Mats
Jute net
Excelsior
(wood fiber) mat
Fiberglass roving
Cover area
Cover area
0.5-1 ton
Heavy, uniform; woven of single jute
yarn. Used with organic mulch.

Continuous fibers of drawn glass
bound together with a non-toxic
agent.
Withstands water flow.

Apply with compressed air ejector.
Tack with emulsified asphalt at a rate
of 25-35 gal/1,000 ft2.
 Effectiveness

 Mulching effectiveness varies with the type of mulch used and local conditions such as rainfall
 and runoff amounts. Percent soil loss reduction for different mulches ranges from 53 to 99.8
 percent used and associated water velocity reductions range from 24 to 78 percent (Harding,
 1990).  Table 5-7 shows soil loss and water velocity reductions for different mulch treatments.
• June 2002
5-31

-------
Development Document for Construction and Development Proposed Effluent Guidelines	     j

       Table 5-7. Measured Reductions in Soil Loss for Different Mulch Treatments
Mulch characteristics
100% wheat straw/top net
100% wheat straw/two nets
70% wheat straw/30% coconut
fiber
70% wheat straw/30% coconut
fiber
100% coconut fiber
Nylon monofilament/two nets
Nylon
monofilament/rigid/bonded
Vinyl
monofilament/flexible/bonded
Curled wood fibers/top net
Curled wood fibers/two nets
Antiwash nettingft'ute)
Interwoven paper and thread
Uncrimped wheat straw-2,242
kg/ha
Uncrimped wheat straw— 4,484
kg/ha
Soil loss
reduction
(%)
97.5
98.6
98.7
99.5
98.4
99.8
53.0
89.6
90.4
93.5
91.8
93.0
84.0
89.3
Water velocity reduction
(%)relative to bare soil
73
56
71
78
77
74
24
32
47
59
59
53
45
59
Source: Harding, 1990, as cited in USEPA, 1993.

Limitations                                         '                              ,

Mulching, matting, and netting might delay seed germination because the cover changes soil
surface temperatures. The mulches themselves are subject to erosion and may be washed away
in a large storm if not sufficiently anchored with netting or tacking. Maintenance is necessary to
ensure that mulches provide effective erosion control.
June 2002
5-32

-------
  Development Document for Construction and Development Proposed Effluent Guidelines	

  Maintenance

  Mulches must be anchored to resist wind displacement.  Netting should be removed when
  protection is no longer needed and disposed of in a landfill or composted. Mulched areas should
  be inspected frequently to identify areas where mulch has loosened or been removed, especially
  after rain storms.  Such areas should be reseeded (if necessary) and the mulch cover replaced
  immediately. Mulch binders should be applied at rates recommended by the manufacturer.  If
  washout, breakage, or erosion occurs, surfaces should be repaired, reseeded, and remulched, and
  new netting should be installed. Inspections should be continued until vegetation is firmly '
  established.

  Cost

  The costs of seed and mulch average $1,500 per acre and range from $800 to $3,500 per acre
  (USEPA, 1993). R. S. Means (2000) estimates the cost of power mulching to be $22.50 per
  1000 square feet, for large volume applications. In addition, hydro- and mechanical seeding are
 approximately $700 to $900 per acre. Coverage cost varies with the seed type, seeding approach,
 and scale (total acreage to be seeded). For example, hydro or water-based seeding for grass is
 estimated to be $700 per acre, but seeding of "field" grass species is only $540 per acre. (Costs
 include materials, labor, and equipment, with profit and overhead.) If surface preparation is
 required, then the installation costs increase. R. S. Means (2000) suggests the cost of fine
 grading, soil treatment, and grassing, is approximately $2 per square yard of earth surface area.
 5.1.5.1.2.5
GEOTEXTILES
 General Description

 Geotextiles are porous fabrics also known as filter fabrics, road rugs, synthetic fabrics,
 construction fabrics, or simply fabrics.  Geotextiles are manufactured by weaving or bonding
 fibers made from synthetic materials such as polypropylene, polyester, polyethylene, nylon,
 polyvinyl chloride, glass, and various mixtures of these materials. As a synthetic construction
 material, geotextiles are used for a variety of purposes such as separators, reinforcement,
 filtration and drainage, and erosion control (USEPA, 1992).  Some geotextiles are made of
 biodegradable materials such as mulch matting and netting. Mulch mattings are jute or other
 wood fibers that have been formed into sheets and are more stable than normal mulch. Netting
 is typically made from jute, wood fiber, plastic,, paper, or cotton and can be used to hold the
 mulching and matting to the ground. Netting can also be used alone to stabilize soils while the
 plants are growing; however, it does not retain moisture or temperature well.

 Geotextiles can aid in plant growth by holding seeds, fertilizers, and topsoil in place. Fabrics are
 relatively inexpensive for certain applications—a wide variety of geotextiles exist to match the
 specific needs of the site.
June 2002
                                                                                     5-33

-------
Development Document for Construction and Development Proposed Effluent Guidelines	;	

Applicability

Geotextiles can be used for erosion control by using it alone. Geotextiles can be used as matting,
which is used to stabilize the flow of channels or swales or to protect seedlings on recently
planted slopes until they become established. Matting may be used on tidal or streambanks
where moving water is likely to wash out new plantings.  They can also be used to protect
expo'sed soils immediately and temporarily, such as when active piles of soil  are left overnight.

Geotextiles are also used as separators.  An example of such a use is geotextile as a separator
between riprap and soil. This "sandwiching" prevents the soil from being eroded from beneath
the riprap and maintaining the riprap's base.

Design and Installation Criteria

Many types of geotextiles are available.  Therefore, the selected fabric should match its purpose.
State or local requirements, design procedures, and any other applicable requirements should be
considered.  In the field, important concerns include regular inspections to determine whether
cracks, tears, or breaches are present in the fabric and appropriate repairs should be made.
Effective netting and matting require firm, continuous contact between the materials and the soil.
If there is no contact, the material will not hold the soil and erosion will occur underneath the
material.

Effectiveness

A geotextile's effectiveness depends upon the strength of the fabric and proper installation. For
example, when protecting a cut slope with a geotextile, it is important to properly anchor the
fabric using appropriate length and spacing of wire staples. This will ensure that it will not be
undermined by a storm event.

Limitations

Geotextiles (primarily synthetic types) have the potential disadvantage of being sensitive to light
and must be protected prior to installation. Some geotextiles might promote  increased runoff
and might blow away if not firmly anchored. Depending on the type of material used,
geotextiles might need to be disposed of in a landfill, making them less desirable than vegetative
stabilization.  If the fabric  is not properly selected, designed, or installed, the effectiveness may
be reduced drastically.

Maintenance

Regular inspections should be made to determine whether cracks, tears, or breaches have formed
in the fabric—it should be repaired or replaced immediately. It is necessary to maintain contact
between the ground and the geotextile at all times.

June 2002                                  ~           T"                                5'34

-------
Development Document for Construction and Development Proposed Effluent Guidelines	•	.

Cost

Costs for geotextiles range from $0.50 to $10.00 per square yard depending on the type chosen
(SWRCP, 1991).  Geosynthetic turf reinforcement mattings (TRMs) are widely used for
immediate erosion protection and long-term vegetative reinforcement, usually for steeply sloped
areas or areas exposed to runoff flows. The Erosion Control Technology Council (a geotextile
industry support association) estimates TRMs cost approximately $7.00 per square yard
(installed) for channel protection (ECTC, 2002a). Channel protection is one of the most
demanding of installations (much more demanding than general coverage of denuded area). The
ECTC estimates the cost to install a simple soil blanket (or rolled erosion control product), seed,
and fertilizer to be $ 1.00 per square yard (ECTC, 2002b).
5.1.5.1.2.6
VEGETATED BUFFER STRIPS
General Description

Vegetated buffers are areas of either natural or established vegetation that are maintained to
protect the water quality of neighboring areas. Buffer zones reduce the velocity of storm water
runoff, provide an area for the runoff to permeate the soil, allow groundwater recharge, and act
as filters to catch sediment. The reduction in velocity also helps to prevent soil erosion.

Applicability

Vegetated buffers can be used in any area that is able to support vegetation, but they are most
effective and beneficial on floodplains, near wetlands, along streambanks, and on steep, unstable
slopes.  They are also effective in separating land use areas that are not compatible and hi
protecting wetlands or waterbodies by displacing activities that might be potential sources of
nonpoint source pollution.

Design  and Installation Criteria

To establish an effective vegetative buffer, the following guidelines should be followed:

•  Soils should not be compacted.

•  Slopes should be  less than 5 percent.

•  Buffer widths should be determined after careful consideration of slope, vegetation, soils,
   depth to impermeable layers, runoff sediment characteristics, type and quantity of storm
   water pollutants, and annual rainfall.

•  Buffer widths should increase as slope increases.
June 2002
                                                                 5-35

-------
Development Document for Construction and Development Proposed Effluent Guidelines	;	

•   Zones of vegetation (native vegetation in particular), including grasses, deciduous and
    evergreen shrubs, and understory and overstory trees, should be intermixed.

•   In areas where flows are concentrated and velocities are high, buffer zones should be
    combined with other structural or nonstructural BMPs as a pretreatment.            '•
                                                                                 \
Vegetated strips have been studied extensively, with emphasis placed on their effectiveness in
removing sediment and other pollutants. Vegetated strips are most appropriate at sites where
sediment loads are relatively low, as high sediment loads will cause large quantities of
deposition along the leading edge of the vegetation. This deposition will cause the flow to divert
around the vegetation hi a concentrated flow pattern, which will cause short-circuiting and
greatly reduce removal efficiency. Variability in vegetation density and uniformity often :causes
similar problems. Removal efficiency depends on a combination of slope, length, and width of
the filter; density of the vegetation; sediment characteristics, hydraulics of the flow; and!
infiltration. The interaction of these variables is complex and prevents the process from being
reduced to a simple relationship except on a local basis. For site-specific local conditions,
methods have been developed that allow trapping to be related to strip length and slope.'

Effectiveness

Considerable data have been collected on the effectiveness of buffer strips for specific
conditions. Numerous factors such as infiltration rate, flow depth, slope, dimensions of the
buffer, density and type of vegetation, sediment size, and sediment density impact removal rates.
Recent studies show that even short vegetative buffers can trap high percentages of sediment and
certain chemicals. A significant concern is whether flow is allowed to concentrate, which will
greatly reduce the travel tune through the buffer and prevent the removal of pollutants.

Several researchers have measured greater than 90 percent reductions in sediment and nitrate
concentrations; buffer/filter strips do a reasonably good job of removing phosphorus attached to
sediment, but are relatively ineffective hi removing dissolved phosphorus (Gillman, 1994).
However, since the hydraulics of flow through buffers strips are not well defined and can vary
considerably based on site conditions, it is difficult to  consistently estimate the effectiveness of
buffers strips.

Limitations

Vegetated buffers require plant growth before they can be effective, and land must be available
on which to plant the vegetation. If the cost of the land is very high, buffer zones might hot be
cost-effective. Although vegetated buffers help to protect water quality, they usually do not
effectively counteract concentrated storm water flows  to neighboring or downstream wetlands.
June 2002
                                                                                     5-36

-------
  Development Document for Construction and Development Proposed Effluent Guidelines	

  Maintenance                                                             '    .

  Keeping vegetation in vegetated buffers healthy requires routine maintenance, which (depending
  on species, soil types, and climatic conditions) can include weed and pest control, mowing,
  fertilizing, liming, irrigating, and pruning.  Inspection and maintenance are most important when
  buffer areas are first installed.  Once established, vegetated buffers do not require much
  maintenance beyond the routine procedures listed earlier and periodic inspections of the areas,
  especially after any heavy rainfall and at least once a year. Inspections should focus on
  encroachment, gully erosion, density of vegetation, evidence of concentrated flows through the
  areas, and any damage from foot or vehicular traffic. If there is more than 6 inches of sediment
  in one place, it should be removed.

  Cost

 Conceptual cost estimates for grassed buffer strips can be made based on square footage using
 unit cost values. R. S. Means (2000) estimates the cost of fine grading, soil treatment, and
 grassing to be $2 per square yard of earth surface area. This cost estimate is based on
 application of traditional lawn seed. The cost for field seed is lower than lawn seed, reducing the
 coverage price.  Where gently sloping areas just need to be grassed with acceptable species, the
 cost can be as low as $0.38 per square yard.
 5.1.5.1.2.7
EROSION CONTROL MATTING
 General Description

 Erosion control mats can be either organic or made from a synthetic material.  A wide variety of
 products exist to match the specific needs of the site. Organic mats are made from such materials
 as wood fiber, jute net, and coconut coir fiber.. Unlike organic matter, synthetic mats are
 constructed from non-biodegradable materials and remain in place for many years. These
 organic mats are classified as Turf Reinforcement Mats (TRMs) and Erosion Control and
 Revegetation Mats (ECRMs) (USDOT, 1995).

 Erosion control matting aids in plant growth-by holding seeds, fertilizers, and topsoil in place.
 Matting can be used to stabilize the flow of channels or swales or to protect seedlings on recently
 planted slopes until they become established. Matting can be used on tidal or streambanks  *
 where moving water is likely to wash out new plantings. It can also be used to protect exposed
 soils immediately and temporarily, such as when active piles of soil are left overnight.

 Applicability

 Mulch mattings, netting, and filter fabrics are particularly useful in steep areas and drainage
 swales where loose seed is vulnerable to being washed away or failing to survive dry soil
 (UNEP, 1992).
June 2002
                                                                                     5-37

-------
Development Document for Construction and Development Proposed Effluent Guidelines         	;	,	

Erosion control mats can also be used to separate riprap and soil. This results in a
"sandwiching" effect, maintaining the riprap's base and preventing the soil beneath from being
eroded.

Design and Installation Criteria

Matting is especially recommended for steep slopes and channels (UNEP, 1992).

Many types of erosion control mats are available. Therefore, the selected product should match
its purpose. Effective netting and matting require firm, continuous contact between the materials
and the soil. If there is no contact, the material will not hold the soil and erosion will occur
underneath the material.

Wood fiber or curled wood mat consists of curled wood with fibers, 80 percent of which are 150
mm or longer, with a consistent thickness and even distribution of fiber over the entire mat. The
top side of the mat is covered with a biodegradable plastic mesh. The mat is placed in thp
channel or on the slope parallel to the direction of flow and secured with staples and check slots.
This is applied immediately after seeding operations (USDOT, 1995).

Jute net consists of jute yarn, approximately 5 mm in diameter, woven into a net with openings
that are approximately 10 by 20 mm (or 0.40 to 0.79 inches). The jute net is loosely laid in the
channel parallel to the direction of flow.  The net is secured with staples and check slots at
intervals along the channel. Placement of die jute net is done immediately after seeding
operations (USDOT, 1995).

Coconut blankets are constructed of biodegradable coconut fibers that resist decay for 5 to 10
years to provide long, temporary erosion control protection. The materials are often encased in
ultraviolet stabilized nets and sometimes have a composite, polypropylene structure to provide
permanent turf reinforcement.  These materials are best used for waterway stabilization and
slopes that require longer periods to stabilize (USDOT, 1995).

Under the synthetic mat category there are TRMs and.ECRMs. Turf reinforcement mats\SK
three-dimensional polymer nettings or monofilaments formed into a mat. They have sufficient
thickness (>13 mm or 0.5 inch) and void space (>90 percent) to allow for soil filling and
retention. The mat acts as a traditional mat to protect the seed and increase germination.  As the
turf establishes,  the mat remains in place as part of the root structure.  This gives the established
turf a higher strength and resistance to erosion (USDOT, 1995).

Erosion control and revegetation mats are composed of continuous monofilaments bound by
heat fusion or stitched between nettings. They are thinner than TRMs and do not have the void
space to allow for filling of soil.  They act as a permanent mulch and allow vegetation to grow
through the mat (USDOT, 1995).
 June 2002
                                                                                      5-38

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

 Effectiveness

 The effectiveness of erosion control matting depends upon the strength of the material and
 proper installation. For example, when protecting a cut slope with an erosion control mat, it is
 important to anchor the mat properly.  This will ensure that it will not be undermined by a storm
 event.

 While erosion control blankets can be effective, their performance varies. Some general trends
 are that organic materials tend to be the most effective (Harding, 1990) and that thicker materials
 are typically superior (Fifield, 1992), but there are exceptions to both of these trends.
 Information about product testing of blankets is generally lacking. One notable exception is the
 Texas Department of Transportation, which publishes the findings of their testing program in the
 form of a list of acceptable and unacceptable materials for specific uses.

 Limitations

 Erosion control mats (primarily synthetic types) are sensitive to light and for this reason must be
 protected prior to installation. Some erosion control mats might cause an increase in runoff or
 blow away if not firmly anchored. Erosion control mats might need to be properly disposed of in
 a landfill, depending on the type of material. Effectiveness may be reduced if the fabric is not
 properly selected, designed, or installed.

 Maintenance

 Regular inspections are necessary to determine whether cracks, tears or breaches have formed in
 the fabric. Contact between the ground and erosion control mat should be maintained at all
 times and trapped sediment removed after each storm event.

 Cost

 Costs for erosion control mats range from $0.50 to $10.00 per square yard depending on the type
 chosen (SWRCP, ,1991). Geosynthetic turf reinforcement mattings (TRMs) are widely used for
 immediate erosion protection and long-term vegetative reinforcement, usually for steeply sloped
 areas or areas exposed to runoff flows. The Erosion Control Technology Council (a geotextile
 industry support association) estimates TRMs cost approximately $7.00 per square yard
 (installed) for channel protection (ECTC, 2002a).  Channel protection is one of the most
 demanding of installations (much more demanding than general coverage of denuded area). The
ECTC estimates the cost to install a simple soil blanket (or rolled erosion control product), seed,
and fertilizer to be $ 1.00 per square yard (ECTC, 2002b).
June 2002
                                                                                     5-39

-------
Development Document for Construction and Development Proposed Effluent Guidelines
5.1.5.1.2.8
TOPSOILING
General Description

Topsoiling is the placement of a surface layer of soil enriched in organic matter over a prepared
subsoil to provide a suitable soil medium for vegetative growth on areas with poor moisture, low
nutrient levels, undesirable pH, and/or the presence of other materials that would inhibit the
establishment of vegetation. Advantages of topsoil include its high organic-matter content and
friable consistency and its available water-holding capacity and nutrient content. The texture
and friability of topsoil are usually more conducive to seedling emergence root growth. In
addition to being a better growth medium, topsoil is often less credible than subsoils, and the
coarser texture of topsoil increases infiltration capacity and reduces runoff. During construction,
topsoil is often removed from the project area and stockpiled. It is replaced over areas to be
grassed  or landscaped during the final stages  of the project.                            ;

Applicability

Conditions where topsoiling apply include the following:

•  Where a sufficient supply of quality topsoil is available.

•  Where the subsoil or areas of existing surface soil present the following problems:
   -   The structure, pH, or nutrient balance of the available soil cannot be amended by
       reasonable means to provide an adequate growth medium for the desired vegetation.
   -   The soil is too shallow to provide adequate rooting depth or will not supply necessary
       moisture and nutrients for growth of desired vegetation.
   -   The soil contains substances toxic to the desired vegetation.

•  Where high quality turf or ornamental plants are desired.

•  Where slopes are 2:1 or flatter.                                                  :

Design  and Installation Criteria

The topsoil should be uniformly distributed over the subsoil to a minimum compacted depth of
50 mm (2 inches) on slopes steeper than 3 horizontal to 1 vertical  and  100 mm (4 inches) on
flatter slopes. Thicknesses of 100 to 150 mm is preferred for vegetation establishment via
seeding. The topsoil should not be placed while in a frozen or muddy condition or when the
subsoil is excessively wet, frozen, or in a condition that is detrimental  to proper grading or
seedbed preparation. The final surface should be prepared so that any irregularities are corrected
and depressions and water pockets do hot form. • If the topsoil has been treated with soil
sterilants, it should not be placed until the toxic substances have dissipated (USDOT, 1995).
Table 5-8 summarizes the cubic yards of topsoil required for application to various depths.
June 2002
                                                                                      5-40

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

       Table 5-8.  Cubic Yards of Topsoil Required for Application to Various Depths
Depth (inches)
1
2
3
4
5
6
Per 1,000 Sq Ft
3.1
6.2
9.3
,. 12.4
15.5
18.6
Per Acre
134
268
403
536
670
804
        Source: Smolen et al., 1988.

 On slopes and areas that will not be mowed, the surface may be left rough after spreading
 topsoil. A disk may be used to promote bonding at the interface between the topsoil and subsoil
 (Smolen et al., 1988).

 Effectiveness

 No information is available describing the effectiveness of applying topsoil as a BMP.

 Limitations

 Limitations of applying topsoil can include to following:

 • i  Topsoil spread when conditions were too wet, resulting in severe compaction.

 •   Topsoil mixed with too much unsuitable subsoil material, resulting in poor vegetation
    establishment.

 •   Topsoil contaminated with soil sterilants or chemicals, resulting in poor or no vegetation
    establishment.

 •   Topsoil not adequately incorporated or bonded with the subsoil, resulting in poor vegetation
    establishment and soil slippage on sloping areas.

 •   Topsoiled areas not protected, resulting in excessive erosion.

 Maintenance

Newly topsoiled areas should be inspected frequently until the vegetation is established. Eroded
 or damaged areas should be repaired and revegetated.
June 2002
                                                                                      5-41

-------
• Development Document for Construction and Development Proposed Effluent Guidelines            ,	

 Cost

 Top soiling costs are a function of the price of topsoil, the hauling distance, and the method of
 application. R. S. Means (2000) report unit cost values of $3 and $4 per square yard, for 4 and 6
 inches of top soil cover, respectively.  This price is for furnishing and placing of top soil, and
 includes materials, labor, and equipment, with profit and overhead.

 5.1.5.2       WATER HANDLING PRACTICES

 5.1.5.2.1             EARTH DIKE

 General Description

 An earth dike is a temporary or permanent ridge of soil designed to channel water to a desired
 location. Dikes are used to divert the flow of runoff by constructing a ridge of soil that  ,
 intercepts and directs the runoff to the desired outlet or alternative management practice, such as
 a pond. This practice serves to reduce the length of a slope for erosion control and protect
 downslope areas. An earth dike can be used to prevent runoff from going over the top of a cut
 and eroding the slope, directing runoff away from a construction site or building; to divert clean
 water from a disturbed area; or to reduce a large drainage area into a more manageable size.
 Dikes should be stabilized with vegetation after construction (NAHB, n.d.).

Applicability

Earth dikes are applicable to all areas; the. size of the dike is correlated to the size of the drainage
 area (NAHB, n.d.).

Design and Installation Criteria         -

The location of dikes should take into consideration outlet conditions, existing land use, '
topography, length of slope, soils, and development plans. The capacity of earth dikes and
diversions should be suitable for the area that is being protected, including adequate freeboard,
or extra depth that is added as a safety margin.  For homes, schools, and industrial buildings, the
recommended design frequency storm is 50 years and the freeboard is 0.5 feet (NAHB, n.'d.).

Earth dikes can be employed as a perimeter control. For small sites, a compacted 2-foot-tall dike
is usually suitable, if hydroseeded. Larger dikes will actually divert runoff to another portion of
the site, usually to a downstream sediment trap or basin. Therefore, the designer should ensure
that they have the capacity for the 10-year storm event, and that the channel created behind the
dike is properly stabilized to percent erosion (Brown et al., 1997). In addition, the downstream
structure must be sized to handle the flow from the dike.  Dikes should be designed using
standard hydrologic and hydraulic calculations and certified by a professional hydrologist, or
engineer.
June 2002
5-42

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

 Diversion dikes should be installed prior to the majority of the soil-disturbing activity. As soon
 as the dike form is completed, it should be machine compacted, fertilized, and either seeded and
 mulched or sodded.  Excavated materials should be properly stockpiled for future use or
 disposed of properly. Dikes should have an outlet that functions with a minimum of erosion.
 Depending on site conditions and outlet structures,  the runoff directed by dikes may have to be
 conveyed to a sediment-trapping device, such as a sediment basin or detention pond. As grades
 increase over 4 percent, geotextile material or sod may be required to control erosion.  Slopes
 greater than 8 percent may require riprap. Dikes may be removed when stabilization of the
 drainage area and outlet are complete (NAHB, n.d.).  Dike design criteria must incorporate site-
 specific  conditions, as dimensions depend on expected flows, soil types, and climatic conditions.
 All of these inputs vary tremendously over different sections of the country.

 Effectiveness

 No information has been found on the effectiveness of earth dikes used as BMPs, although
 terraces often have sediment removal rates of up to  90 percent.

 Limitations

 An erosion-resistant lining in the channel may be needed to prevent erosion in the channel
 caused by excessive grade.  In addition, the channel should be deepened and the grade realigned
 if there is overtopping caused by sediment in the channel where the grade decreases or reverses.
 If overtopping occurs at low points in the ridge where the diversion crosses the shallow draw, the
 ridge should be reconstructed with a positive grade  toward the outlet at all points. Finally, if
 there is erosion at the outlet, an outlet stabilization structure  should be installed and if
 sedimentation occurs at the diversion outlet, a temporary sediment trap should be installed.

 Maintenance

 An earth dike should be inspected for signs of erosion after every major rain event.  Any repairs
 and/or revegetation should be completed promptly (NAHB, n.d.).  The following actions can be
 taken to properly maintain an earth dike:

 •  Remove debris and sediment from the channel immediately after the storm event.

 •  Repair the dike to its original height.

 •  Check outlets and make necessary repairs to prevent gully formation.

 •  Clean out sediment traps when they are 50 percent full.
June 2002
                                                                                     5-43

-------
Development Document for Construction and Development Proposed Effluent Guidelines  	'.	
                                              f
•  Once the work area has been stabilized, remove the diversion ridge, fill and compact the
   channel to blend with the surrounding area, and remove sediment traps, disposing of unstable
   sediment in a designated area.

Cost

The cost of an earth dike depends on the design and materials used. Small dikes can cost,
approximately $2.00 per linear foot, while larger dikes can cost approximately $2.00 per cubic
yard.  EPA states that an earth dike can cost approximately $4.50 per linear foot (NAHB, n.d.).

An alternative means to estimate conceptual costs for earthen dikes is to use unit cost values and
a rough estimate of the quantities needed. Shallow trenching (1 to 4 feet deep) with a backhoe in
areas not requiring dewatering can be performed for $4 to $5 per cubic yard of removed material
(R. S. Means, 2000). Based on this value, $2 per linear foot provides for 11 square feet of flow
area and $4.50 per linear foot provides for 24 square feet of flow area. This suggests that the
size of the dike  is required prior to specifying a cost, which requires a site-specific hydrologic
evaluation. Based on standards for Virginia (1992), most small drainage areas (made up of 5
acre or less), diversion dikes are approximately 18" tall, with a 4.5' base.  Assuming the
excavation volume equals the volume of the dike, the resulting excavation volume is
approximately 7 cubic feet per linear foot, which (conservatively) equates to $1.03 to $1.30 per
linear foot for construction costs.                                                   ;

If the earthen dikes are to be permanent, then additional costs are incurred to vegetate the dike.
R. S. Means (2000) estimates the cost of fine grading, soil treatment, and grassing is    •
approximately $2 per square yard of earth surface area. This adds approximately $6 per linear
foot of dike. Where gently sloping areas just need to be grassed with acceptable species, the cost
can be as low as $0.3 8 per square yard.        ,                                    •
5.1.5.2.2
TEMPORARY SWALE
General Description

The term swale (grassed channel, dry swale, wet swale, biofilter) refers to a series of vegetated,
open channel management practices designed specifically to treat and attenuate storm water
runoff for a specified water quality volume. As storm water runoff flows through these
channels, it is treated by filtering through the vegetation in the channel, filtering through a
subsoil matrix, and/or infiltration into the underlying soils.  Variations of the grassed swale
include the grassed channel, dry swale, and wet swale.  The specific design features and methods
of treatment differ in each of these designs, but all are improvements on the traditional drainage
ditch and incorporate modified geometry and other features for use of the swale as a treatment
and conveyance practice.
June 2002
                                                                                     5-44

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

 Applicability

 Grassed swales can be applied in most situations with some restrictions and are very well suited
 for treating highway or residential road runoff because they are linear practices. Perimeter
 dikes/swales should.be limited to a drainage area of no more than 0.8 hectare and usually work
 best on gently sloping terrain. Perimeter dikes may not work well on moderate slopes, and they
 should never be established on slopes exceeding 20 percent (UNEP, 1994).

• Regional Applicability.  Grassed swales can be applied in most regions of the country. In arid
 and semi-arid climates, however, the value of these practices needs to be weighed against the
 water needed to irrigate them.

 Ultra-Urban Areas. Ultra-urban areas are densely developed urban areas in which little
 pervious surface exists. Grassed swales are generally not well suited to ultra-urban areas because
 they require a relatively large area of pervious surfaces.

 Storm Water Hot Spots.  Storm water hot spots are areas where land use or activities generate
 highly contaminated runoff, with concentrations of pollutants in excess of those commonly
 found in storm water. A typical example is a gas station or convenience store.  With the
 exception of the dry swale design, hot spot runoff should not be directed toward grassed
 channels. These practices  either infiltrate storm water or intersect the groundwater, making use
 of the practices for hot spot runoff a threat to groundwater quality.

 Storm Water Retrofit. A storm water retrofit is a storm water management practice (usually
 structural),  put into place after development has occurred, to improve water quality, protect
 downstream channels, reduce flooding, or meet other specific objectives. One retrofit
 opportunity using grassed swales modifies existing drainage ditches. Ditches have traditionally
 been designed only to convey storm water away from roads as quickly as possible.  In some
 cases, it may be possible to incorporate features to enhance pollutant removal or infiltration such
 as check dams (for example, small dams along the ditch that trap sediment, slow runoff, and
 reduce the longitudinal slope).  Since grassed swales cannot treat a large area, using this practice
 to retrofit an entire watershed would be expensive because of the number of practices needed to
 manage runoff from a significant amount of the watershed's land area.

 Cold Water (Trout) Streams.  Grassed channels are a good treatment option within watersheds
 that drain to cold water streams. These practices do  not retain water for a long period of time
 and often induce infiltration.  As a result, standing water will not typically be subjected to
 warming by the sun in these practices.
June 2002
                                                                                      5-45

-------
Development Document for Construction and Development Proposed Effluent Guidelines	;	

Design and Installation Criteria

Temporary swales should be designed using standard hydrologic and hydraulic calculations.
Designs should be certified by a professional hydrologist, engineer, or other appropriate
professional.

Perimeter dikes/swales should be established before any major soil-disturbing activity takes
place. Dikes should be compacted with construction equipment to the design height plus 10
percent to allow for settlement. If they are to remain in place for longer than 10 days, they
should be stabilized using vegetation, filter fabric, or other material. Diverted water should be
directed to a sediment trap or other sediment treatment area (UNEP, 1994).

In addition to the broad applicability concerns described above, designers need to consider
conditions at the site level. In addition, they need to incorporate design features to improve the
longevity and performance of the practice, while minimizing the maintenance  burden.   :

Siting Considerations
                                                                                 t
In addition to considering the restrictions and adaptations of grassed swales to different regions
and land uses, designers must ensure that this management practice is feasible at the site in
question. Depending on the design option, grassed channels can be highly restricted practices
based on site characteristics.

Drainage Area. Grassed swales generally should treat small drainage areas of less than 5 acres.
If the practices are used to treat larger areas, the flows and. volumes through the swale become
too large to design the practice to treat storm water runoff through infiltration and filtration.

Slope. Grassed swales should be used on sites with relatively flat slopes (less than 4 percent).
Runoff velocities within the channel become too high on steeper slopes. This  can cause erosion
and does not allow for infiltration or filtering in the swale.

Soils /Topography. Grassed swales can be used on most soils, with some restrictions on the
most impermeable soils. In the dry swale, a fabricated soil bed replaces on-site soils to ensure
that runoff is filtered as it travels through the soils of the swale.                        :

Groundwater. The depth to groundwater depends on the type of swale used.  In the dry swale
and grassed channel options, designers should separate the bottom of the swale from the
groundwater by at least 2 feet to prevent a moist swale bottom or contamination of the  '
groundwater.  In the wet swale option, treatment is enhanced by a wet pool, which is maintained
by intersecting the groundwater.
June 2002
                                                                                     5-46

-------
  Development Document for Construction and Development Proposed Effluent Guidelines      	_^^^

  Design Considerations

  Although the grass swale has different design variations, including the grassed channel, dry
  swale, and wet swale, some design considerations are common to all three. One overriding
  similarity is the cross-sectional geometry of all three options. Swales  should generally have a
  trapezoidal or parabolic cross-section with relatively flat side slopes (flatter than 3:1). Designing
  the channel with flat side slopes maximizes the wetted perimeter. The wetted perimeter is the
  length along the edge of the swale's cross-section where runoff flowing through the swale is in
  contact with the vegetated sides and bottom of the swale.  Increasing the wetted perimeter slows
  runoff velocities and provides more contact with vegetation to encourage filtering and
  infiltration. Another advantage to flat side slopes is that runoff entering the grassed swale from
  the side receives some pretreatment along the side slope. The flat bottom of all three should be
  between 2 and 8 feet wide.  The minimum width ensures an adequate filtering surface for water
  quality treatment, and the maximum width prevents braiding, that is, the formation of small
  channels within the swale bottom.

 Another similarity among all three designs is the type of pretreatment needed. In all three  design
 options, a small forebay should be used at the beginning of the front of the  swale to trap
 incoming sediments.  A pea gravel diaphragm, a small trench filled with river run gravel, should
 be used to pretreat runoff entering the sides of the swale.

 Two other features designed to enhance the treatment ability of grassed swales are a flat
 longitudinal slope (generally between 1 and 2 percent) and a dense vegetative cover in the
 channel. The flat slope helps to reduce the velocity of flow in the channel.  The dense vegetation
 also, helps reduce velocities, protect the channel from erosion, and act as a filter to treat storm
 water runoff. During construction, it is important to stabilize the channel before the turf has
 been established, either with a temporary grass cover or with the use of natural or synthetic
 erosion control products.

 In addition to treating runoff for water quality, grassed swales need to convey larger storms
 safely.  Typical designs allow the runoff from the 2-year storm (for example, the storm that
 occurs, on average, once every 2 years) to flow through the swale without causing erosion.
 Swales should also have the capacity to pass larger storms (typically a 10-year storm) safely.

 The length of the swale necessary to infiltrate runoff waters can be calculated by using a mass
 balance of runoff waters and infiltration waters for a triangular-shaped cross-sectional area.

 Design Variations

 The following discussion identifies three different variations of open channel practices,
 including the grassed channel, the dry swale, and the wet swale.
June 2002
                                                                                     5-47

-------
Development Document for Construction and Development Proposed Effluent Guidelines            :	

Grassed Channel. (Discussed in more length in sub-section 5.5.1.2.1) Of the three grassed
swale designs, grassed channels are the most similar to a conventional drainage ditch, with the
major differences being flatter side slopes and longitudinal slopes and a slower design velocity
for water quality treatment of small storm events.  Of all of the grassed swale options, grossed
channels are the least expensive, but they also provide the least reliable pollutant removal. The
best application of a grassed channel is as pretreatment to other structural storm water practices.

One major difference between the grassed channel and most of the other structural practices is
the method used to size the practice. Most water quality practices for storm water management
are sized by volume. This method sets the volume available in the practice equal to the Water
quality volume, or the volume of water to be treated in the practice. The grassed channel, on the
other hand, is a flow rate-based design. Based on the peak flow from the water quality storm
(this varies from region to region but a typical value is the 1-inch storm), the channel should be
designed so that runoff takes, on average, 10 minutes to flow from the top to the bottom|of the
channel. A procedure for this design can be found in Design of Storm Water Filtering Systems
(CWP, 1996).                                                     .

Dry Swales.  Dry swales are similar in design to bioretention areas. These designs incorporate a
fabricated soil bed into their design. The existing soil is replaced with a sand/soil mix that meets
minimum permeability requirements. An underdrain system is used under the soil bed. This
system is a gravel layer that encases a perforated pipe. Storm water treated in the soil bed flows
through the bottom into the underdrain, which conveys this treated storm water to the storm
drain system.  Dry swales are a relatively new design, but studies of swales with a native soil
similar to the man-made soil bed of dry swales suggest high pollutant removal.        '

Wet Swales. Wet swales intersect the groundwater and behave similarly to a linear wetland cell.
This design variation incorporates a shallow permanent pool and wetland vegetation to provide
storm water treatment.  This design also has potentially high pollutant removal. One
disadvantage of the wet swale is that its use in residential or commercial settings is unpopular
because the shallow standing water in the swale is often viewed as a potential nuisance by
homeowners.

Regional Variations

Cold Climates. In cold or snowy climates, swales may serve a dual purpose by acting as both a
snow storage/treatment and a storm water management practice.  This dual purpose is  j
particularly relevant when swales are used to treat road runoff. If used for this purpose, Iswales
should incorporate salt-tolerant vegetation, such as creeping bentgrass.

Arid Climates. In arid or semi-arid climates, swales should be designed with drought-tolerant
vegetation, such as buffalo grass. As pointed out in the Applicability discussion, the value of
vegetated practices for water quality needs to be weighed against the cost of water needed to
maintain them in arid and semi-arid regions.                                       '-•
 June 2002
                                                                                      5-48

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

 Effectiveness

 Swales act to control peak discharges in two ways.  First, the grass reduces runoff velocity,
 depending on the length and slope of the swale. Second, a portion of the storm water runoff
 volume passes through the swale and infiltrates into the soil. Table 5-9 summarizes grassed
 swale pollutant removal efficiencies.

                Table 5-9. Grassed Swale Pollutant Removal Efficiency Data
Grassed Swale Removal Efficiencies
Study
Goldberg, 1993
Seattle Metro and Washington
Department of Ecology, 1992
Seattle Metro and Washington
Department of Ecology, 1992
Wang etal., 1981
Dormanetal., 1989
Harper, 1988
Kercher, Landon, and Massarelli,
1983
Harper, 1988
Koon, 1995
Occoquan Watershed Monitoring
Lab, 1983
Yousef et al., 1985
Occoquan Watershed Monitoring
Lab, 1983
Yousef etal., 1985
Occoquan Watershed Monitoring
Lab, 1983
Welborn and Veenhuis, 1987
Yu, Barnes, and Gerde, 1993
Dormanetal., 1989 -
Pitt and McLean, 1986
Oakland, 1983
Dorman et al.. 1989
TSS
67.8
60
83
80
98
87
99
81
67
-100
• -
-50
-
31
0
68
65,
0
33
-85
TP
4.5
45
29
-
18
83
99
17
39
-100
8
' -9.1
-19.5
-23
-25
60
41
-
-25
12
TN
-
-
-
-
-
84
99
40
-
-100
13
-18.2
8
36.5
-25
-
-
0
-
-
NO,
31.4
-25
'-25
-
45.
80
99
52
9
-
11 .
-
2
.-
-25
-
11
-
-
-100
Metals
42-62
2-16
46-73
70-80
37-81
88-90
99
37-69
-35 to 6
-100
14-29
-100
41-90
-100 to 33
0
74
14-55
0
20-58
14-88
Bacteria
-100
—25
-25 .
-
-
-
-
-
-
- '
-
-
-
-
-
.
,
0
0
-
Type
Grassed channel
Grassed channel
Grassed channel
Dry swale
Dry swale
Dry swale
Dry swale
Wet swale
Wet swale
Drainage channel
Drainage channel
Drainage channel
Drainage channel
Drainage channel
Drainage channel
Drainage channel
Drainage channel
Drainage channel
Drainage channel
Drainage channel
Limitations

Common problems associated with swales include excessive erosion along unlined channels
(usually because of excessive grade), erosion or sedimentation at the outlet point, or overtopping
of the dike at low points (UNEP, 1994).

Additional limitations of the grass swale include the following:

•  Grassed swales cannot treat a very large drainage area.

•  Swales do not appear to be effective at reducing bacteria.
June 2002
                                                                                     5-49

-------
Development Document for Construction and Development Proposed Effluent Guidelines	____^	

•   Wet swales may become a nuisance because of mosquito breeding.

•   If designed improperly (for example, proper slope is not achieved), grassed channels will
    have very little pollutant removal.

•   A thick vegetative cover is needed for these practices to function properly.

Maintenance

As with any BMP, swales must be maintained to continue functioning as effective pollutant
removal methods.  Maintenance may include occasional mowing, fertilizing, and liming., In
addition, any areas that become damaged by erosion should be immediately repaired and
replanted. The swales should be protected from concentrated flows and checked for downstream
obstructions.

Cost                                                                            [

To produce a conceptual cost approximation, grassed channel construction costs can be
developed using unit cost values. Shallow trenching (1 to 4 feet deep) with a backhoe in areas
not requiring dewatering can be performed for $4 to $5 per cubic yard of removed material
(R. S. Means, 2000). Assuming no disposal costs (i.e., excavated material  is placed on either side
of the trench), only the cost of fine grading, soil treatment, and grassing (approximately $2 per
square yard of earth surface area) should be added to the trenching cost to  approximate the total
construction cost. Site-specific hydrologic analysis of the construction site is necessary to
estimate the channel conveyance requirement and the desired retention time in the swale. It is
not unusual to have flows on the order of 2 to 4'cfs per acre served.                    ;

For a design channel velocity of 1 foot per second, the resulting range in the channel cross-
section area can be as low as 2 but as high as 4 square feet per acre drained. If the average
channel flow depth is 1 foot, then the low estimate for grassed channel installation is $0.74 per
square foot of channel bottom per acre served per foot of channel length. The high estimate is
$ 1.48 per square foot of channel bottom per acre served per foot of channel length.      j

Table 5-10 summarizes additional costs of grass swales.
June 2002
5-50

-------
Development Document for Construction and Development Proposed Effluent Guidelines

     Table 5-10.  Average Annual Operation and Maintenance Costs for a Grass Swale
Component
Mowing
General grass
care
Debris/litter
removal
Reseeding/
fertilization
Inspection and
general
administration
TOTAL
Estimated
Unit Cost ($)
0.89/100 m2
8.8/100 m2
0.5 1/m2
0.35/m2
0.74/m2

$ for Swale Size:
0.5 m Deep
0.3m Bottom Width .
3m Top Width
145.0
162.98
93.0
5.9
231.0
638.0
$ for Swale Size:
1 m Deep
1m Bottom Width
7 m Top Width
241.0
274.0
93.0
10.37
231.0
850.0
Comments
Mow 2-3 times per year
Grass maintenance area is (top
width + 3 m) x length

Area revegetated is 1% of
maintenance area per year
Inspection once per year

Source: Ellis, 1998.
5.1.5.2.3
TEMPORARY STORM DRAIN DIVERSION
General Description

A temporary storm drain diversion is a pipe that reroutes an existing drainage system to
discharge flow into a sediment trap Or basin. This practice reduces the amount of sediment-
laden runoff from construction sites that enters waterbodies without treatment. Temporary storm
drain diversions can be used when a permanent storm water drainage system has not yet been
installed. It should be recognized that diversion channels can also be installed but are not
considered in the following discussion.

Applicability

A temporary storm drain diversion should be used to temporarily redirect discharge to a
permanent outfall and should remain in place until the area draining to the storm sewer is no
longer disturbed.  Temporary storm drain diversions can also be combined with other structures
and used as a sediment-trapping device when the completion of a permanent outfall has been
delayed; alternatively, a sediment trap can be placed below a permanent outfall to remove
sediment before the final flow discharge.

Design and Installation Criteria

Since the diversion is only temporary, the layout of piping and the overall impact of the
diversion's installation on post-construction drainage patterns must be considered. Once
construction is completed,  the temporary diversion should be moved to restore the original
system.  The following activities should be done at this time:
June 2002
                                                                5-51

-------
Development Document for Construction and Development Proposed Effluent Guidelines             	

•   The storm drain should be flushed before the sediment trap is removed.

•   The outfall should be stabilized.

•   Graded areas should be restored.

•   State or local specifications should be checked for more detailed requirements arid ah
    appropriate professional should certify that the design meets local hydrologic and hydraulic
    requirements.

Effectiveness

If installed properly to capture the bulk of runoff from a construction site, temporary storm drain
diversions can be effective in reducing the discharge of sediment-laden, untreated water to
waterbodies. When used in combination with other erosion and sediment control practices such
as minhriized clearing or vegetative and chemical stabilization, the level of pollution frorn a
construction site can be substantially reduced or eliminated.

Limitations
                                   V .                                             !
Installation of a temporary storm drain diversion may result in the disturbance of existing storm
drainage patterns.  Care must be taken to ensure that the original system is properly restored
once the temporary system is removed. The most common source of problems is excessive
velocity at the outlet.  Installation of an outlet stabilization structure is typically required and
may be constructed of riprap, reinforced concrete, geotextile linings, or a combination.  ;

Maintenance

Once installed, temporary storm drain diversions require very little maintenance. Frequent
inspection and maintenance  of temporary storm drain systems, especially after large storms,
should ensure that pipe clogging does not occur and that runoff from the site is being
successfully diverted.  After removal of the temporary diversion, the permanent storm drain
system should be carefully inspected to ensure that drainage patterns have not been altered by
the temporary system.                                             '

Cost

Depending on the size of the construction site, a temporary storm drain diversion can be bostly..
Costs include those associated with materials needed to construct the diversion and sediment trap
or basin (mainly piping, concrete, and gravel), and also labor costs for installation and removal
of the system, all of which may involve excavation, regrading, and inspections. Based on the
variety of conditions that can affect storm drain diversion designs, typical costs per installation
are not presented here. However, site-specific cost estimates can be produced using unit cost

June 2002                                              '                           :    T52

-------
Development Document for Construction and Development Proposed Effluent Guidelines	

values along with site-specific quantity estimates. R. S. Means (2000) indicates a range of pipe
costs for surface placement, between $5.00 per linear foot for 4" diameter PVC piping, and $9.20
per linear foot for 10" diameter PVC piping. On construction sites, temporary inlets and outlets
are usually formed by small rock-lined depressions. Assuming 4 cubic yards of crushed rock
(1.5" mean diameter) per opening, an inlet and outlet combine to add approximately $200 per
pipe installation, based on $25 per cubic yard of stone (R. S. Means, 2000).
5.1.5.2.4
PIPE SLOPE DRAIN
General Description

Pipe slope drains are used to reduce the risk of erosion on slopes by discharging runoff to
stabilized areas.  Consisting of a metal or plastic flexible pipe if temporary, or pipes or paved
chutes if permanent, these drains are placed from the top to the bottom of a slope to carry surface
runoff from the top to the bottom of a slope that has already been damaged by erosion or is at
high risk for erosion.  These drains are also used to drain saturated slopes that have the potential
for soil slides.

Applicability

Temporary slope drains can be used on most disturbed slopes to eliminate gully erosion
problems resulting from concentrated flows discharged at a diversion outlet.  Slope drains should
be used as a temporary measure for as long as the drainage area remains disturbed. They will
need to be moved once construction is complete and a permanent storm drainage system is
established. Appropriate restoration measures will then need to be taken, such as adjusting
grades and flushing sediment from the pipe before it is removed (UNEP, 1994).

Design and Installation Criteria

Pipe slope drains can be placed directly on the ground or buried under the surface.  The inlet
should be located at the top of the slope and should be fitted with an apron, attached with a water
tight connection.  Filter cloth should be placed under the inlet to prevent erosion. Flexible pipes,
which are positioned on top of the ground, should be securely anchored with grommets placed
10 feet on center. The outlet at the bottom of the slope should also be stabilized with riprap.
The riprap should be placed along the bottom of a swale that leads to a sediment-trapping
structure or another stabilized structure.

Slope drain pipe sizes are based on drainage area and the size of the design storm.  Pipes should
be connected to a diversion ridge at the top of the slope by covering with compacted fill material
where it passes through the ridge. Discharge from a slope drain should be to a sediment trap,
sediment basin, or other stabilized outlet (UNEP, 1994).
June 2002
                                                                                      5-53

-------
 Development Document for Construction and Development Proposed Effluent Guidelines   	    '      	

 Pipe slope drains should be installed perpendicular to the contour down the slope, and the design
 should be able to handle the peak runoff for the 10-year storm. Recommendations of slope drain
 diameter are summarized in Table 5-11 (NAHB, n.d).

              Table 5-11. Recommended Pipe/Tubing Sizes for Slope Drains
Maximum Drainage
Area (acres)
0-0.5
0.5
0.75
1.0
1.5
2.5
3.5
5.0
Pipe/Tubing
Diameter*(inches)

12


18
21
24
30
Pipe/Tubing
Diameter11 (inches)

12


18

24

Pipe/Tubing
' Diameter0 (inches)

8
10
12
Individually designed



•UNEP, 1994.
bUSDOT, 1995.
°IDNR,1992.

Recently graded slopes that do not have permanent drainage measures installed should have a
temporary slope drain and a temporary diversion installed. A temporary slope drain used^in
conjunction with a diversion conveys storm water flows and reduces erosion until permanent
drainage structures are installed.

The following are design recommendations for temporary slope drains:

•   The drain should consist of heavy-duty material manufactured for the purpose and have
    grommets for anchoring at a spacing of 10 feet or less.

•   Minimum slope drain diameters should be observed for varying drainage areas.

•   The entrance to the pipe should consist of a standard flare end section of corrugated metal.
    The corrugated metal pipe should have watertight joints at the ends. The rest of the pipe is
    typically corrugated plastic or flexible tubing, although for flatter, shorter slopes, a
    polyethylene-lined channel is sometimes used.

•   The height of the diversion at the pipe should be the diameter of the pipe plus 0.5 foot.

•   The outlet should be located at a reinforced or erosion-resistant location.
June 2002
5-54

-------
Development Document for Construction and Development Proposed Effluent Guidelines
    Temporary slope drains should be designed to adequately convey runoff for a desired
    frequency storm, typically either 2 years or 10 years depending on local regulations.  Both
    the size and the spacing can be determined based on the contributing drainage area. Drains
    are spaced at intervals corresponding to the specified drainage areas. For larger drainage
    areas and critical locations, the drains should be sized on an individual basis (USDOT,
    1995).

    Slope drains should be constructed in conjunction with diversion berms such that the berms
    are not overtopped. At the pipe inlet, the top of the berm should be a minimum of 300
    millimeters (11.81 inches) higher than the top of the pipe. The entrance should be
    constructed of a standard flared end section or a Tee section if designed properly. The
    entrance should be placed in a 150 millimeters (5.90 inches) minimum depressed sump
    (USDOT, 1995).

    The outlet of the slope drain must be protected with a riprap apron.  If the slope drain is
    draining a disturbed area and sufficient right-of-way is available, the drain may empty into a
    sediment trap (USDOT, 1995). Table 5-12 summarizes slope drain characteristics.

                         Table 5-12. Slope Drain Characteristics
Capacity
Material
Inlet section
Connection to ridge at top of
slope
Outlet
2-yr frequency, 24-hr-duration storm event
Strong, flexible pipe, such as heavy duty, nonperforated, corrugated plastic
Standard "T" or "L" flared-end section with metal toe plate
Compacted fill over pipe with minimum dimensions, 1.5 ft depth, 4 ft top width, and 6
in higher than ridge
Pipe extends beyond toe of slope and discharges into a sediment trap or basin unless
contributing drainage area is stable
Source: IDNR, 1992.
Effectiveness                '.

There is currently no information on the effectiveness of pipe slope drains.

Limitations

The area drained by a temporary slope drain should not exceed 5 acres. Physical obstructions
substantially reduce the effectiveness of the drain. A common slope drain problem is
overtopping of the inlet due to an undersized or blocked pipe, or erosion at the outlet point due to
insufficient protection (UNEP, 1994). Other concerns are failures from overtopping because of
inadequate pipe inlet capacity and reduced diversion channel capacity and ridge height
June 2002
5-55

-------
Development Document for Construction and Development Proposed Effluent Guidelines	

Solutions to common problems include the following (IDNR, 1992):     .

•   Washout - A washout along a pipe due to seepage and piping may be caused by inadequate
    compaction, insufficient fill, or installation that may be too close to the edge of the slope.

•   Overtopping caused by undersized or blocked pipe - The drainage area may be too large.

•   Overtopping caused by improper grade of channel and ridge - A positive grade should be
    maintained.

•   Overtopping caused by poor entrance conditions and trash buildup at the pipe inlet - Deepen
    and widen the channel at the pipe entrance and frequently inspect and clear the inlet.  ;

•   Erosion at outlet - The pipe should be extended to a stable grade or an outlet stabilization
    structure is needed.

•   Displacement or separation of pipe - The pipe should be tied down and the joints secured.

Maintenance

Pipe slope drains must be inspected after each significant runoff event for evidence of erosion
and uncontrolled runoff. Any repairs to the drain should be made immediately.  Significant
amounts of sediment trapped at the outfall should also be removed in a  timely manner and
disposed of properly (NAHB, n.d.).

The following actions should be taken to properly maintain a pipe slope drain (IDNR, 1992):

•   Inspect slope drains and supporting diversions once a week and after every storm event.

•   Check the inlet for sediment or trash accumulation; clear and restore to proper entrance
    condition.

•   Check the fill over the pipe for settlement, cracking, or piping holes; repair immediately.

•   Check for holes where the pipe emerges from the dike; repair immediately:

•   Check the conduit for evidence of leaks or inadequate anchoring; repair immediately.

•   Check the outlet for erosion or sedimentation; clean and repair, or extend if necessary.

•   Once slopes have been stabilized, remove the temporary diversions  and slope drains, and
    stabilize all disturbed areas.
June 2002
5-56

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

 Cost

 The cost of pipe slope drains and their installation varies with the design and materials used.
 Site-specific cost estimates can be produced using unit cost values with site-specific quantity
 estimates. R. S. Means (2000) indicates a range of pipe costs for surface placement between
 $5.00 per linear foot for 4" diameter PVC piping, and $9.20 per linear foot for 10" diameter PVC
 piping. On construction sites, temporary inlets and outlets are usually formed by small rock-lined
 depressions. Assuming 4 cubic yards of crushed rock (1.5" mean diameter) per opening, 'an inlet
 and outlet combine to add approximately $200 per pipe installation, based on $25 per cubic yard
 of stone (R. S. Means, 2000).
 5.1.5.2.5
STONE CHECK DAM
 General Description

 A check dam is a small temporary barrier or dam constructed across a drainage channel or swale
 to reduce the velocity of the flow. By reducing the flow velocity, the erosion potential is
 reduced, detention times are lengthened, and more sediments are able to drop out of the water
 column. Check dams can be constructed of stone, gabions, treated lumber, or logs
 (NAHB, n.d.).

 Check dams are inexpensive and easy to install.  They may be used permanently if designed
 properly to allow a high proportion of sediment in the runoff to settle out and reduce velocity
 and may provide aeration of the water (NAHB, n.d.). However, the use of check dams in a
 channel should not be a substitute for the use of other sediment-trapping and erosion control
 measures. As with most other temporary structures, check dams are most effective when used in
 combination with other storm water and erosion and sediment control measures.

 Applicability

 Check dams are commonly used (1) in channels that are degrading but where permanent
 stabilization is impractical because of their short period of usefulness and (2) in eroding channels
 where construction delays or weather conditions prevent timely installation of erosion-resistant
 linings (IDNR,  1992).
                                                                                   V
 Check dams are also useful in steeply sloped swales, in small channels, in swales where
 adequate vegetative protection cannot be established, or in swales  or channels that will be used
 for a short period of time where it is not practical to line the channel or implement other flow
 control practices (USEPA, 1993). In addition, check dams are appropriate where temporary
 seeding has been recently implemented but has not had time to fully develop and take root. The
 contributing drainage area should range from 2 to 10 acres. Check dams should be used only in
 small open channels that will not be overtopped by flow once the dams are built and should not
be built in steam channels, either intermittent or perennial (UNEP, 1994).
June 2002
                                                                                    5-57

-------
Development Document for Construction and Development Proposed Effluent Guidelines            	

Design and Installation Criteria

Check dams can be constructed from a number of different materials. Most commonly, they are
made of rock, logs, sandbags, or straw bales.  Rock or stone is often preferred because of its
cost-effectiveness and longevity. Logs and straw bales will decay with time and are not
recommended as they may cause waterway blockage if they fail. When using rock or stone, the
material diameter should be 2 to 15 inches. The stones should be extended 18 inches beyond the
banks, and the side slopes should be 2:1 or flatter.  Lining the upstream side of the dam with a
foot of 1- to 2-inch gravel may improve the efficiency of the dam (NAHB, n.d.). Logs should
have a diameter of 6 to 8 inches. Regardless of the material used, careful construction of a check
dam is necessary to ensure its effectiveness.                                        j

The distance between rock check dams will vary depending on the slope of the ditch, with closer
spacing when the slope is steeper. The size of stone used in the check dam should also vary with
the expected design velocity and discharge. As velocity and discharge increase, the rock size
should also increase. For most rock check dams, 3  inches to 12 inches is a suitable stone size.
To improve the sediment-trapping efficiency of check dams, a filter stone can be applied to the
upstream face. A well-graded coarse aggregate that is less than 1 inch in size can be used as a
filter stone.

All check dams should have a maximum height of 3 feet. The center of the dam should be at
least 6 inches lower than the edges. This design creates a weir effect that helps to channel flows
away from the banks and prevent further erosion. Additional stability can be achieved by
implanting the dam material approximately 6 inches into the sides and bottom of the channel
(VDCR, 1995).

When installing more than one check dam in a channel, outlet stabilization measures should be
installed below the final dam in the series. Because this area is likely to be vulnerable to further
erosion, riprap or some other stabilization measure is highly recommended.

Effectiveness

Field experience has shown that rock check dams are more effective than silt fences or straw
bales to stabilize wet-weather ditches (VDCR, 1995).  Straw bales have been shown to have vary
low trapping efficiencies and should not be used for check dams. For long channels, check dams
are most effective when used in a series, creating multiple barriers to sediment-laden runoff.
June 2002
                                                                                    5-58

-------
Development Document for Construction and Development Proposed Effluent Guidelines	

Limitations

Check dams should not be used in perennial streams unless approved by an appropriate
regulatory agency (USEPA, 1992; VDCR, 1995). Because the primary function of check dams
is to slow runoff in a channel, they should not be used as a stand-alone substitute for other
sediment-trapping devices. Also, leaves have been shown to be a significant problem as they
clog check dams; therefore, increased inspection and maintenance might be necessary in the fall.

Common problems with check dams include channel bypass and severe erosion when
overtopped and ineffectiveness due to accumulated sediment and debris. When designing check
dams, the fact that they will reduce the capacity of a channel to transmit storm water runoff and
thus will need to be sized appropriately should be taken into account (UNEP, 1994). The check
dam may also kill grass linings in the channel if the water level remains high after it rains or if
there is significant sedimentation. In addition, a check dam may reduce the hydraulic capacity of
the channel and create turbulence, which erodes the channel banks (NAHB, n.d.).

Maintenance

Check dams should be inspected periodically to ensure that they have not been repositioned as a
result of storm water flow. In addition, the center of a check dam should always be lower than
its edges. Additional stone may have to be added to maintain the correct height.  Sediment
should not be allowed to accumulate to more than half the original dam height. Any required
maintenance should be performed immediately. When check dams are removed, care must be
taken to remove all dam materials to ensure proper flow within the channel. The channel should
subsequently be seeded for stabilization (NAHB, n.d.).

Cost

The cost of check dams varies based on the material used for construction and the width of the
channel to be dammed. In general, it is estimated that  check dams constructed of rock cost about
$100 per dam (USEPA, 1992). Brown (1997) estimated rock check dam would cost
approximately $62 per installation, including the cost for filter fabric bedding. Other materials,
such as logs and sandbags, may be a less expensive alternative, but they might require higher
maintenance costs.

5.1.5.2.6            LINED WATERWAYS

General Description

Lined channels convey storm water runoff through a stable conduit.  Vegetation lining the
channel reduces the flow velocity of concentrated runoff. Lined channels usually are not
designed to control peak runoff loads by themselves and are often used in combination with
other BMPs such as subsurface drains and riprap stabilization. Where moderately steep slopes

June 2002                           '.         ~         •':      5^59

-------
Development Document for Construction and Development Proposed Effluent Guidelines   	;  •

require drainage, lined channels can include excavated depressions or check dams to enhance
runoff storage, decrease flow rates, and enhance pollutant removal. Peak discharges can be
reduced through temporary detention hi the channel.  Pollutants can be removed from storm
water by filtration through vegetation, by deposition, or in some cases by infiltration of soluble
nutrients into the soil. The degree of pollutant removal in a channel depends on the residence
time of the water hi the channel and the amount of contact with vegetation and the soil surface,
but pollutant removal is not generally the major design criterion.

Often construction increases the velocity and volume of runoff, which causes erosion in newly
constructed or existing urban runoff conveyance channels. If the runoff during or after
construction will cause erosion hi a channel, the channel should be lined or flow control
practices instituted. The first choice of lining should be grass or sod since this reduces runoff
velocities and provides water quality benefits through filtration and infiltration.  If the velocity in
the channel would erode the grass or sod, riprap, concrete, or gabions can be used (USEPA,
2000). Geotextile materials can be used hi conjunction with either grass or riprap linings ;to
provide additional protection at the soil-lining interface.                              ;

Applicability

Lined channels typically are used in residential  developments, along highway medians, or as an
alternative to curb and gutter systems. Grass-lined channels should be used to convey runoff
only where slopes are 5 percent or less. These channels require periodic mowing, occasional
spot-seeding, and weed control to ensure adequate grass cover (UNEP, 1994).

Lined channels should be used hi areas where erosion-resistant conveyances are needed, such as
hi areas with highly erodible soils and slopes of less than 5 percent. They should be installed
only where space is available for a relatively large cross-section.  Grassed channels have a
limited ability to control runoff from large storms and should be used with the recommended
allowable velocities for the specific soil types and vegetative cover.

Design and Installation Criteria

The design of a lined waterway requires proper determination of the channel dimensions. It must
ensure that (1) the velocity of the flowing water will not wash out the waterway and that (2) the
capacity of the waterway is sufficient to carry the surface flow from the watershed without
overtopping.

Vegetative-Lined Channels. Grass-lined channels have been previously discussed in detail and
are only summarized hi this section. The allowable velocity of water in the waterway depends
upon the type, condition, and density of the vegetation, as well as the erosive characteristics of
the soil. Uniformity of vegetative cover is important because the stability of the most sparsely
covered area determines the stability of the channel. Grasses are a better vegetative coverthan
legumes because grasses resist water velocity more effectively.                        <
June 2002
                                                                                      5-60

-------
  Development Document for Construction and Development Proposed Effluent Guidelines	

  Vegetative-lined channels may have triangular, parabolic, or trapezoidal cross-sections. Side
  slopes should not exceed 3:1 to facilitate the establishment, maintenance, and mowing of
  vegetation. A dense cover of hardy, erosion-resistant grass should be established as soon as
  possible following grading. This may necessitate the use of straw mulch and the installation of
  protective netting until the grass becomes established. If the intent is to create opportunities for
  runoff to infiltrate into the soil, the channel gradient should be kept near zero, the channel
  bottom must be well above the seasonal water table, and the underlying soils should be relatively
  permeable (generally, with an infiltration rate greater than 2 centimeters [0.78 niches] per hour).

  Rock-Lined Channels.  Riprap-lined channels may be installed on somewhat steeper slopes
  than grass-lined channels. They require a foundation of filter fabric or gravel under the riprap.
  Generally, side slopes should not exceed 2:1, and riprap thickness should be 1.5 times the
  maximum stone diameter. Riprap should form a dense, uniform, well-graded mass
  (UNEP, 1994).

  Lined channels should be sited in accordance with the natural drainage system and should not
  cross ridges.  The channel design should not have sharp curves or significant changes in slope.
  Channels should not receive direct  sedimentation from disturbed areas and should be established
  only on the perimeter of a construction site to convey relatively clean storm water runoff and
  separated from disturbed areas by a vegetated buffer or other BMP to reduce sediment loads.

 Basic design recommendations for lined channels include the following:

 •  Construction and vegetation of the channel should occur before grading and paving activities
 .  begin.                                                                             ,

 •  Design velocities should be below 5 feet per second.

 •  Geotextiles can be used to stabilize vegetation until it is fully established.

 •  Covering the bare soil with sod or geotextiles can provide reinforced storm water
    conveyance immediately.

 •  Triangular-shaped channels should be used with low velocities and small 'quantities of
    runoff; parabolic grass channels are used for larger flows and where space is available;
    trapezoidal channels are used with large flows of low velocity (low slope).

 •   Outlet stabilization structures might be needed if the runoff volume or velocity has the
    potential to exceed the capacity of the receiving area.

 •   Channels should be designed to convey runoff from a 10-year storm without erosion.
June 2002
                                                                                      5-61

-------
Develc
lent Document for Construction and Development Proposed Effluent Guidelines
•  The sides of the channel should be sloped less than 3:1, with V-shaped channels along roads
   sloped 6:1 or less for safety.

•  All trees, bushes, stamps, and other debris should be removed during construction.

Effectiveness                                                                   :

Lined channels can effectively transport storm water from construction areas if they are designed
for expected flow volumes and velocities and if they do not receive sediment directly from
disturbed areas.

Limitations

Lined channels, if improperly installed, can alter the natural flow of surface water and have
adverse impacts on downstream waters. Additionally, if the design capacity is exceeded by a
large storm event, the vegetation might not be sufficient to prevent erosion and the channel
might be destroyed. Clogging with sediment and debris reduces the effectiveness of grass-lined
channels for storm water conveyance.                                             :

Common problems in lined channels include erosion of the channel before vegetation is; fully
established and gullying or head cutting in the channel if the grade is too steep. In addition, trees
and brush tend to invade lined channels, causing maintenance problems.              ;

Riprap-lined channels can be designed to safely convey greater runoff volumes on steeper
slopes. However, they should generally be avoided on slopes exceeding 10 percent because stone
displacement, erosion of the foundation, or channel overflow and erosion resulting from a
channel that is too small can occur. Thus, channels established on slopes greater than 10 percent
will usually require protection with rock gabions, concrete, or other highly stable and protective
surfaces (UNEP, 1994).                                                          ;

 Maintenance

 Maintenance requirements for lined channels are relatively minimal.  During the vegetation
 establishment period, the channels should be inspected after every rainfall. Other maintenance
 activities that should be carried out after vegetation is established are mowing, litter removal,
 and spot vegetation repair. The most important objective in the maintenance of lined channels is
 maintaining a dense and vigorous growth of turf. Periodic cleaning of vegetation and soil
 buildup in curb cuts is required so that water flow into the channel is unobstructed. During the
 growing season, channel grass should be cut no shorter than the level of design flow, and the
 cuttings should be removed promptly.                                             \
 June 2002
                                                                                      5-62

-------
 Development Document for Construction and Development Proposed Effluent Guidelines    	

 Cost

 Costs of grassed channels range according to depth, with a 1.5-foot-deep, 10-foot-wide grassed
 channel estimated at $6,395 to $17,075 per trench, while a 3.0-foot-deep, 21-foot-wide grassed
 channel is estimated at $12,909 to $33,404 per trench (SWRPC, 1991)

 Readers are also referred to the discussion of costs for grass-lined channels, which contains
 many of the design and cost elements required for installing lined waterways. Designers have a
 range of options for lining new channels. Geosynthetic turf reinforcement mattings (TRMs) can
 be used for immediate erosion protection in channels exposed to runoff flows. The Erosion
 Control Technology Council (a geotextile industry support association) suggests TRMs cost
 approximately $7.00 per square yard (installed) for channel protection (ECTC, 2002a). R. S.
 Means indicates machine-placed riprap costs of approximately $40 per cubic yard. The riprap
 maximum size is typically between 6 and 12 inches, depending on the channel design velocity. A
 cubic yard of riprap will cover between 36 and 18 square feet of channel bed for these riprap
 sizes (assuming depth of riprap is 1.5 times the maximum size). These estimates suggests that
 riprap lining will be between $10 and $20 per square foot of channel (Costs include materials,
 labor, and equipment, with overhead and profit).
5.1.5.3
SEDIMENT TRAPPING DEVICES
The devices listed under this group of BMPs trap sediment primarily through impounding water
and allowing for settling to occur (Haan et al., 1994).  Silt fence, super silt fence, straw bale
dikes, sediment traps, and sediment basins all control flow through a porous flow control system
such as filter fabric or straw bales or they use a dam to impound water with a pipe, open channel,
or rock fill outlet.  The filtering capacity of silt fence (filter fabric) contributes only a small
amount of trapping, but serves to make the fence less porous and hence increases ponding . For
steady-state flows, the trapping that occurs behind the flow control device can be shown to be
directly proportional to the surface area and indirectly proportional to flow through the system
(Haan et al., 1994). The ratio of the surface area to flow is known as the overflow rate, and
trapping in such systems is predicted by the ratio of overflow rate to particle settling velocity.
Although flows in nature are inherently non-steady state and more complex than steady-state
systems, studies have shown that the best predictor of trapping in such systems is still the ratio of
settling velocity to overflow rate (Hayes et al., 1984).  In the case of non-steady state, the
overflow rate is best defined by the ratio of peak discharge from the system to a surface area
(Hayes et al., 1984; McBurnie et al., 1990).

The amount of trapping in these structures depends on the size of the structure, flow rates into
the system, hydraulics of the flow control system, the size distribution of the sediment flowing
into the structure, and the chemistry of the sediment-water system (Haan et al., 1994).  Trapping
can be enhanced by chemical treatment of flows into the structure, but the impacts have not been
widely defined for varying mineralogy and chemistry of the sediment-water system (Haan et al.,
1994; Tapp and Barfield, 1986).  Recent studies have been conducted on the application of
June 2002
                                                                                    5-63

-------
Development Document for Construction and Development Proposed Effluent Guidelines	    ;	

polyacrilamides (PACs) to disturbed areas for enhancing settling (Benik et'al., 1998; Masters et
al., 2000; Roa-Espinosa et al., 2000), but results have not been definitive. No known studies
have evaluated the impacts of PAC application to disturbed areas on settling in sediment
trapping devices.                                                                 !

Sediment flowing into sediment trapping devices is composed of primary particles and  I  -
aggregated particles.  Aggregates are formed when clays, silts, and sands are cemented together
to form larger particles that have settling velocities far greater than those of any individual
particles alone although the degree of aggregation depends on the amount of cementing material
present (typically clays and organic matter).  Since the aggregates have higher settling velocities
than primary particles, the degree of aggregation that is present has a large impact on the;
trapping that occurs.  Procedures are available to measure the combined size distribution! of
aggregate and primary particle size distribution (Barfield et al., 1979; Haan et al., 1994).!
Procedures are also available to predict particle size distributions of aggregates and primary
particles (Foster et al., 1985) but have not been found to be very accurate for subsoils exposed
during construction in at least one study (Barfield et al., 1983).

In the absence of chemical treatment, the sediment that can be captured in sediment trapping
devices is typically the settleable solids. To trap the smaller size clay particles, structures with
surface areas larger than the construction site itself would have to built in many cases (Barfield,
2000). Chemical treatment can be used to reduce the size, but it has not been adopted on a wide
scale because of the cost and complexity of the operation (Tapp et al., 1981).

Sediment trapping devices also provide some storm water detention by virtue of detaining flows
long enough to allow sediment to settle out and be deposited. However, to operate as a storm  -
water detention structure, the design should include storm water detention as well.      ;

Virtually all of the available information on sediment trapping structures, both theoretical and
experimental, is on impacts to receiving waters and not downstream effects, m a very limited
analysis, Barfield (2000) combined the SEDIMOTII computer model together with the
FLUVIAL model to theoretically evaluate the impact of sediment trapping structures on -,
downstream geomorphology in a Puerto Rican watershed.                            ;
5.1.5.3.1
General Description
SILT FENCE
Silt fences are used as temporary sediment barriers consisting of filter fabric anchored across and
supported by posts. Their purpose is to retain sediment from small disturbed areas by reducing
the velocity of sediment-laden runoff and promoting sediment deposition (Smolen et al., 1998).
Silt fences capture sediment by ponding water and allowing for deposition, not by filtration. Silt
fence fabric first screens silt and sand from runoff, resulting in clogging of the lower part of the
June 2002
                                                                                     5-64

-------
Development Document for Construction and Development Proposed Effluent Guidelines	•   	

fence. The pooling water allows sediments to settle out of the runoff.  Silt fences work best in
conjunction with temporary basins, traps, or diversions.

Applicability

Silt fences are generally placed at the toe of fills, along the edge of waterways, and along the site
perimeter.  The fences should not be used in drainage areas with concentrated and high flows, in
large areas, or in ditches and swales where  concentrated flow is present.

The drainage area for the fence should be selected based on design storms and local hydrologic
conditions so that the silt fence is not expected to overtop. A typical design calls for no greater
than l/4 acre per 100 feet offence, but this is highly variable depending on climate.  The fence
should be stable enough to withstand runoff from a 10-year peak storm. Table 5-13 lists the
maximum slope length specified by the USDOT.  These slope lengths should be based on
sediment load and flow rates. This would mean that the values given below should be adjusted
for climatic conditions  instead of "one size fits all" for a silt fence to ensure maximum
effectiveness.

                   Table 5-13. Maximum Slope Lengths for Silt Fences
Slope (%)
<;2
5
10
20
25
30
35
40
45
50
18- inch (460 mm) Fence
250 ft (75 m)
100 ft (30m)
50 ft (15 m)
25 ft (8 m)
6 m (20 ft)
15 ft (5 m)
15ft(5m)-
15 ft (5m)
10 ft (3m)
10 ft (3m)
30- inch (760 mm) Fence
500 ft (150m)
250 ft (75m)
150 ft (45m)
70 ft (21 m)
55 ft (17 m)
45 ft (14 m)
40 ft (12 m)
35 ft (10 m)
30 ft (9 m)
25 ft (8m)
       Source: USDOT, 1995.

Typical standards and specifications call for the silt fence to be located on fairly level ground
and follow the land contour. However, field evaluations by Barfield and Hayes (1992,1999) in
South Carolina and Kentucky indicate that installations on the contour as well as along a slope
have problems with undercutting,  hi either case, the installations are such that a slight slope may
occur along the fence hi spite of the best installation practices.  Runoff can move down the
contour until a weak spot occurs in the buried toe and undercuts the fence. Alternatively, flow
may move to a low spot where it accumulates and causes an overtopping,  hi either case,
trapping by the silt fence is essentially zero, and flows have then been concentrated at a point
causing downslope channel erosion.
June 2002
5-65

-------
Development Document for Construction and Development Proposed Effluent Guidelines              	

Design and Installation Criteria

Design criteria are of two types:                                                      ,

Hydrologic design for a required trapping of sediment and flow rate to pass the design storm.
Selection of appropriate installation criteria such that the silt fence will perform as designed.

Hydrologic Design

Hydrologic design should result in a design that passes the design storm without causing damage
while trapping the required amount of sediment. It is necessary to use either a database or; some
type of model to develop the appropriate hydrologic design. Efforts to model the sediment
trapping that occurs through the use of a silt fence have resulted in models that predict the:
settling in the ponded area upstream from the fence {Barfield et al., 1996; Lindley et al., 1998).
The results from model simulations  show that trapping depends primarily on the surface area of
the impounded water and the flow rate through the filter. The models utilize a clear water slurry
flow rate, typically specified by the manufacturer, to predict discharge.  However, numerous
studies have shown that sediment laden flows cause clogging of the geotextiles used to construct
the fence, dependent on the opening size and size of the sediment (Britton et al., 2001; Wyant,
1980; Barrett et al., 1995; Fisher and Jarret, 1984). Thus, results from model studies to date are
suspect and need to be modified to account for the impacts of clogging on flow rate. Barfield et
al., (2000) developed a model of flow rate using conditional probability concepts, but the results
have not been experimentally verified.

Design aids have been developed for silt fence, using simulations from the SEDIMOT III model
(Hayes and Barfield, 1995). In the model, predictions are made about trapping efficiency using
the ratio of settling velocity for the dlsl of the eroded sediment, divided by the ratio of discharge
to ponded surface area. The design aids yield conservative estimates as compared to the  •
SEDIMOT IK model,-but the database used for generating the design aid is based on the
assumption that clogging does not impact flow rates. The discussion above shows that
assumption to be erroneous.

The bottom line on the discussion above is that it is not possible to predict with any expected
accuracy the trapping efficiency of silt fence under a given set of conditions.

Installation Criteria                                                                 ;

General installation criteria for the silt fence should incorporate the following factors:     ;
       'dls:15 percent by weight of suspended solids are smaller than those that are trapped by this device;
Similarly djo indicates that 50 percent by weight of suspended solids are smaller than those trapped.

June 2002
5-66

-------
Development Document for Construction and Development Proposed Effluent Guidelines	•

•   The fabric must have sufficient strength to counter forces created by contained water and
    sediment (Sprague, 1999).

•   The posts must have sufficient strength to counter the forces transferred to them by the fabric
    (Sprague, 1999).

•   The fabric must be installed to ensure that the loads are all adequately transferred through the
    fabric to the posts or the ground without overstressing (Sprague, 1999).

•   The fence must be designed based on site-specific hydrologic and soil conditions such that it
    will not overtop during design events.

•   The fence must be installed (anchored) with a buried toe of sufficient depth so that it does
    not become detached from the soil surface.

•   In general, the fence requires a metal wire backing to provide  sufficient strength to prevent
    failure from the weight of trapped sediment and to prevent the.toe of the fabric from being
    removed from the ground.

•   Maximum drainage area behind the fence should be determined based on the local rainfall
    and the infiltration characteristics of the soil and cover.

Silt fence material is typically synthetic filter fabric or a pervious  sheet of polypropylene, nylon,
polyester, or polyethylene yarn. The fabric should have ultraviolet ray inhibitors and stabilizers
to provide for a minimum useful construction life of 6 months or the duration of construction,
whichever is  greater.  The height of the fence fabric should not exceed 3 feet. If standard
strength filter fabric is used, it should be reinforced with  a wire fence, extending down into the
trench that buries the toe. The wire should be of sufficient strength to support the weight of the
deposited sediment and water. In general, a minimum 14 gauge and a maximum mesh spacing
of 6 inches is called for (Smolen et al., 1988). Typical requirements for the silt fence physical
properties, as specified in selected local BMP standards and specifications, are included in Table
5-14.
June 2002
5-67

-------
Development Document for Construction and Development Proposed Effluent Guidelines

                  Table 5-14. Typical Requirements for Silt Fence Fabric
Physical Property
Filtering Efficiency
Tensile Strength at
20% (maximum)
Elongation
Slurry Flow Rate
Water Flow Rate
UV Resistance
Requirements >
Woven Fabric
85%
Standard Strength — 30
pound/linear inch
Extra Strength — 50 pound/linear
inch
0.3 gallon/square feet/minute
15 gallon/square feet/minute
70%
Non- Woven Fabric '.
85% ;
Standard Strength — 50 pound/linear
inch |
Extra Strength — 70 pound/linear inch
4.5 gallon/square feet/minute
220 gallon/square feet/minute ;
85%
    Source: NCDNR, 1988; IDNR 1992.                                                  i

It should be pointed out that these numbers, particularly the flow rates, could vary widely
depending on the local soil condition due to possible clogging of the filter material.

Material for the posts used to anchor the filter fabric can be constructed of either wood on steel.
Wooden stakes should be buried at a depth sufficient to keep the fence, when loaded with
sediment and water, from falling over. The depth of burial should depend on soil strength
characteristics when saturated and post diameter.  Many standards and specifications set a
minimum length of the post of 5 feet long, and a diameter of 4 inches for posts composed of
softwood (e.g, pine), and 2 niches for posts composed of hardwood (e.g., oak)(Smolen et al.,
1988).  Steel posts should also be designed based on local soil strength characteristics when wet.
Some standards and specifications for these posts set a minimum weight of 1.33 pound/linear
feet with a minimum length of 4 feet. Steel posts should also have projections to adhere filter
fabric to the post (Smolen et al., 1988).

A silt fence should be erected in a continuous fashion from a single roll of fabric so as to ;
eliminate unwanted gaps in the fence. If a continuous roll of fabric is not available, the fabric
should overlap from both directions  only at posts with a mmirnum overlap of 6 inches and be
rolled together with a special flexible rod to keep the ends from separating. Fence posts should
be spaced at a  distance based on wet soil strength characteristics and post size and strength;
generally, the posts are spaced approximately 4 to 6 feet apart.  If standard strength fabric is
used in combination with wire mesh, the spacing can be larger. Typically, the standards and
specifications call for the posts to be no more than 10 feet apart. If extra-strength fabric is used
without wire mesh reinforcement, some standards call for the support posts to be spaced no more
than 6 feet apart (VDCR, 1995). Again, this spacing should depend on wet soil strength
characteristics and post size.

A silt fence must provide sufficient storage capacity or be stabilized over flow outlets such that
the storage volume of water will not overtop the fence.  The return period event (size of the
rainfall event managed) used for design is typically a prerogative of the regulatory agency, For
temporary fences, a 2-year storm event is typically used as a design standard. Fences that will be
hi place for 6 months or longer are commonly designed based on a 10-year storm event
June 2002
                                                                                     5-68

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

 (Sprague, 1999).  The space behind the fence used for impoundment volume must be sufficient
 to adequately contain the sediment that will be deposited. Each storm will deposit sediment
 behind the fence, and after a period of time the amount of sediment accumulated will render the
 fence useless.  Frequency offence management is a function of its sizing (i.e. whether the fence
 was installed for a 2-year or a 10-year storm event) (Sprague, 1999) and the amount of erosion
 that occurs in the area draining to the fence.

 Effectiveness

 The performance of silt fences has not been well defined. Laboratory studies using carefully
 controlled conditions have shown trapping efficiencies in the range of 40 to 100 percent,
 depending on the type  of fabric, overflow rate, and detention time (Barrett et al., 1995; Wyant,
 1980; Wishowski et al., 1998). Field studies have been limited and quite inadequate;  however,
 the results show that field-trapping efficiencies are very low. In fact, Barrett et al.  (1995)
 obtained a value of zero percent trapping averaged over several samples with a standard error of
 26 percent. Barrett et al. (1995) cite the following reasons for the field tests not showing the
 expected results:                   '

 •   Inadequate fabric splices                                                     .

 •   Sustained failure to correct fence damage resulting from overtopping

 •   Large holes in the fabric

 •   Uhder-runs due to inadequate "toe-ins"

 •   Silt fence damaged and partially covered by the temporary placement of stockpiles of
    materials

 Field inspections conducted by Barfield and Hayes (1992) were made in which more than 50
 construction sites in South Carolina and Kentucky were visited. Inspections found that silt fence
 was seldom installed and, when installed, was rarely set up according to specifications. In areas
 where installations did meet  standards, it was obvious that flows sought the weakest spot on the
 fence and either flowed through cuts in the fabric, or undercut or overtopped the fence. This
 flow was thus changed  from  the overland flow coming into the  site to concentrated flow, causing
 significant erosion.                                                                  •

 Silt fences are effective at removing large particle sediment, primarily aggregates, sands, and
larger silts.  Sediment is removed through impounding of water to slow velocity. It is argued
that the silt fence will not contribute to a reduction hi small particle sediment and is not effective
against other pollutants (WYDEQ, 1999). EPA (1993) reports the following effectiveness
ranges, for silt fences constructed of filter fabric: average total suspended solids removal of 70
percent, sand removal of 80 to 90 percent, silt-loam removal of 50 to 80 percent, and
June 2002
                                                                                     5-69

-------
Development Document for Construction and Development Proposed Effluent Guidelines  	    •	

silt-clay-loam removal of 0 to 20 percent. However, the EPA numbers from the Nationwide
Urban Runoff Program should not be considered to apply to every location.           i
The actual trapping will vary widely for a given design because of differences in hydrolpgic
regimes and soil types.

The advantages of using silt fences include: minimal labor requirement for installation,: low
cost, high efficiency in removing sediment, durability, and sometimes reuse (Sprague, 1:999).
Silt fences are the most readily available and cost-effective control options where options like
diversion are not possible. Silt fences are also a popular choice; because contractors have used
them extensively, the familiarity makes silt fence use more likely for future construction
activities. The visibility of a silt fence is also an advantage, for the fence is "advertising" the use
of erosion and sediment control structures. In addition, the silt fence visibility makes site
inspection easier for contractors and government inspectors (CWP, 1996).            •

Limitations

Silt fences should not be installed along areas where rocks or other hard surfaces will prevent
uniform anchoring offence posts and entrenching of the filter fabric because an insufficient
anchor-will greatly reduce the effectiveness of silt fencing and may create runoff channels
leading off-site.  In addition, open areas where wind velocity is high may present a maintenance
challenge, as high winds may accelerate deterioration of the filter fabric (Smolen et al., 1988).
When the pores of the silt fence fabric become clogged with sediment, pools of water ate likely
to form on the uphill side offence. Siting and design of the silt fence should account for this
problem and care should be taken to avoid unnecessary diversion of storm water from these
pools which might cause further erosion damage. Silt fences can act as a diversion if placed
slightly off-contour and can control shallow, uniform flows from small, disturbed areas^and
deliver sediment-laden water to deposition areas.                                   :

Silt fences will sag or collapse if a site is too large, if too much sediment accumulates, if the
approach slope is too  steep, or if the fence was not adequately supported.   If the fence bottom is
not properly installed or the flow velocity is too fast, fence undercuts or blowouts can occur
because of excess runoff.  Erosion around the end of the fence can occur if the fence ends do not
extend upslope to prevent flow around the fence (IDNR, 1992).

Maintenance

Site operators should inspect silt fences after each rainfall event to ensure they are intact and that
there are no gaps at the fence-ground interface or tears along the length of the fence. If gaps or
tears are found, they should be repaired or the fabric should be replaced immediately.  |
Accumulated sediments should be removed from the fence base when the sediment reaches
one-third to halfway up the height of the fence. Sediment removal should occur more frequently
if accumulated sediment is creating a noticeable strain on the fabric and there is the possibility
that the fence might fail from a sudden storm event.
 June 2002
                                                                                      5-70

-------
 Development Document for Construction and Development Proposed Effluent Guidelines
 Cost

 There is a wide range of data on installation costs for silt fences.  EPA estimates these costs at
 approximately $6.00 per linear foot (USEPA, 1992) while SWRPC estimates unit costs between
 $2.30 and $4.50 per linear foot (SWRPC, 1991).  Silt fences have an annual maintenance cost
 that is 100 percent of installation cost (Brown et al., 1997). These values are significantly greater
 than that reported by R. S. Means (2000), which indicates a 3 foot tall silt fence installation cost
 between $0.68 and $0.92 per linear foot (for favorable and challenging installations). It should
 be noted that the R. S. Means value covers just a single installation, without the expected costs
 of maintenance (e.g., removal of collected sediment). In addition, the type of silt fence fabric
 employed will also affect the total installation costs.
 5.1.5.3.2
SUPER SILT FENCE
 General Description

 Super silt fence is a modification of a standard silt fence.  The two central differences between
 the standard silt fence and the super silt fence is that the super silt fence has toe that is buried
 more deeply and the backing material is chain .link fence held in place by steel posts-a concept
 that originated in Maryland. The Maryland super silt fence requires a Geotextile Class F fabric
 over a chain link fence to intercept sediment-laden runoff from small drainage areas.  The super
 silt fence provides a barrier that can collect and hold debris and soil more effectively than a
 standard silt fence, preventing material from entering critical areas. It is best used where the
 installation of a dike would destroy sensitive areas, woods, and wetlands.

 Applicability

 Super silt fences can be used in the same conditions as a silt fence. Fences should follow the
 contour of the land. Table 5-15 lists the distance a super silt fence should be from a slope to
 ensure maximum effectiveness (MDE,  1994).

                      Table 5-15. Slope Lengths for Super Silt Fences
Slope (%)
0-10
10-20
20-33
33-50
50+
Slope Length
Minimum
Unlimited
200 feet
100 feet
100 feet
50 feet
Maximum
Unlimited
1,500 feet
1,000 feet
500 feet
250 feet
June 2002
                                                                                      5-71

-------
Development Document for Construction and Development Proposed Effluent Guidelines	    ,	

Design and Installation Criteria

As with the standard silt fence, design criteria are of two types, hydrologic design for a required
trapping of sediment and flow rate to pass the design storm and selection of appropriate  |
installation criteria such that the" silt fence will perform as designed.                    :

Hydrologic Design                                                                \

Hydrologic design criteria are the same as the criteria for the standard silt fence.         i

Installation Criteria                          ,                                     •

The criteria used for the Maryland super silt fence indicate the following, although they have not
been tested with field data:

•   The fence should be placed as close to the contour as possible, with no section of the silt
    fence exceeding a grade of 5 percent for a distance of more than 50 feet.

•   Fabric should be no more than 42 inches in height and should be held in place with a 6-foot
    chain link fence.

•   Fabric should be attached to the steel pole using wire ties or staples.  Fabric should be
    securely fastened to the chain link fence with ties spaced every 24 inches at the top arid
    midsection.                                                                   ;

•   Fabric should be embedded into the ground at a minimum of 8 inches.

•   Edges of fabric should overlap by 6 inches.

Table 5-16 describes the physical properties of Geotextile class F fabric (MDE, 1994).

    Table 5-16.  Minimum Requirements for Super Silt Fence Geotextile Class F Fabric
Physical Properties
Tension Strength
Tensile Modulus
Flow Rate
Filtering Efficiency
Requirements
50 pound/inch
20 pound/inch
0.3 gallon/tf/minute
75%
 June 2002
                                                                                      5-72

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

 Effectiveness

 Performance data have not been collected for super silt fences. The fences have been proposed
 for locations within a sensitive watershed, or where site conditions prohibit the use of a standard
 silt fence. However, until performance data are collected under field conditions, effectiveness is
 speculative.

 Limitations

 Super silt fences are not as likely to fail structurally as are standard silt fences, but they are more
 expensive than standard silt fences.

 Maintenance

 Maintenance requirements for super silt fences are generally the same as for standard silt fences.

 Cost

 The cost of the super silt fence is more than the standard silt fence because of deeper burial at the
'toe and the cost of chain linked fencing. R. S. Means (2000) indicates a rental price of $10 to $11
 per linear foot of chain linked fence for periods up to 1 year. Overall,- rental is expected for most
 construction site installation because rental rates are approximately half the price of permanent
 chain link fencing.
5.1.5.3.3
STRAW BALE DIKE
General Description

The straw bale dike is a temporary measure used to trap sediment from small, sloping disturbed
areas.  It is constructed of straw bales (not hay bales) wedged tightly together and placed along
the contour downslope of disturbed areas. The bales are placed in a shallow excavation, and the
upslope side is sealed with soil. Stakes are driven through the bales into the soil to help hold the
bales in place. The dike works by impounding water, which allows sediment to settle out in the
upslope area (Haan et al., 1994). Straw bale dikes are recommended for short duration
application and are usually effective for less than 3 months because of rapid decomposition
(USDOT, 1995).

Applicability                                        .

Straw bale dikes are generally placed at the toe of fills to provide for a broad shallow sediment
pool. The dikes should not be used in drainage areas with concentrated and high flows, hi large
areas, or in ditches and swales. The location of the straw bale dike should be fairly level, at least
June 2002
                                                                 •5-73

-------
 Development Document for Construction and Development Proposed Effluent Guidelines            ;	

 10 feet from the toe, and should follow the land contour. Table 5-17 lists the distance a istraw
 bale dike should be placed from a slope to ensure maximum effectiveness.            !

         Table 5-17. Maximum Land Slope and Distances Above a Straw Bale Dike
Land Slope (%)
Less than 2
2-5
5-10
10-20
More than 20
Maximum Distance Above Dam (ft)
100
75
50
25
15
              Source: USDOT, 1995.                                                !

Design and Implementation Criteria

Hydrologic Design                                                      '        ;'.

Hydrologic design dictates the structure necessary to withstand a storm without causing damage
while trapping the required amount of sediment. Either a database or some type of model are
needed to find the appropriate design. Efforts to model the sediment trapping that occurs in
straw bale dikes have resulted in models that predict the settling in the ponded area upstream
from the fence (Barfield et al., 1996; Lindley et.al.,  1998). The results from model simulations
show that trapping depends primarily on the surface area of the impounded water and flqw rate
through the filter. The models utilize a clear water slurry flow rate to predict discharge. It is
anticipated, based on visual observations, that sediment will clog the straw bale barrier, reducing
the slurry flow rate.  Thus, results from model studies to date are suspect and need to be',
modified to account for the impact of clogging on flow rate.

Installation Criteria

The USDOT's BMP Manual and the Indiana BMP Manual (IN Manual) calls for bales to be:

•  Anchored by driving two 36-inch long (minimum) steel rebars or 2 x 2-inch hardwood stakes
   through each bale;

•  Sized according to the standard bale size of 14 inches x 18 inches x 35 inches;

•  Placed in an excavated trench at least 4 inches deep, a bale's width, and long enough that the
   end bales are somewhat upslope of the sediment pool;

•  Abutted tightly against each other; and,

•  Sized such that impounded water depth should not exceed 1.5 feet.
June 2002
5-74

-------
Development Document for Construction and Development Proposed Effluent Guidelines	

The USDOT BMP Manual does not require that straw bale dikes be designed; however, the
Indiana Manual limits the drainage area to % acre per 100 feet of dam and the total drainage area
draining to a straw bale dike to 2 acres.

Effectiveness

The information on performance of straw bale dikes is very limited. In laboratory studies of
bales at varying orientations, Kouwen (1990) found that trapping efficiencies ranged from 60 to
100 percent. Field data on trapping have not been collected; however, visual inspection of sites
indicate that straw bales are not properly installed to prevent flows from undercutting of flowing
between bales (Barfield and Hayes, 1992, 1999). In addition, bales deteriorate rapidly and need
to be replaced frequently. Because of these problems, the use of straw bale dikes as a perimeter
control is not recommended, except in special circumstances.  Only 27 percent of Erosion and
Sediment Control (ESC) experts rated the straw bale dike as an effective ESC practice, although
its use was still allowed in half of the communities surveyed (Brown and Caraco, 1997).

Limitations

Straw bale dikes should not be used as a diversion, in streams, in channels, or in areas with
concentrated flow. The bales are not recommended for paved areas because of the inability to
anchor the bales (IDNR, 1992).

Care must be taken to ensure that the bales are not installed in an area where there is a
concentrated flow of runoff, in a drainage area that is too large, or on an excessive slope (IDNR,
1992). Under these conditions, erosion around the end of the bales, overtopping and
undercutting of the bales, and bale collapsing and dislodging are likely to occur. Overtopping
will also occur if the storage capacity is underestimated and where provisions are not made for ._
safe bypass of storm flow (IDNR, 1992). Undercutting will occur if the bales are not
entrenched at least 4 inches and backfilled with compacted soil or were not abutted or chinked
properly.  Straw bale dikes are likely to collapse or dislodge if the bales are not adequately
staked, or if too much sediment is allowed to accumulate before cleanout (IDNR, 1992).

Maintenance

For the straw bale dike to be most effective, it is important to replace deteriorated bales when
appropriate.

Cost

The cost of straw bale dikes are relatively low, making their use relatively attractive. R. S.
Means (2000) indicates a staked straw bale unit cost of $2.61 per linear foot (Costs include
materials, labor, and equipment, with profit and overhead).
June 2002
                                                                                     5-75

-------
Development Document for Construction and Development Proposed Effluent Guidelines
5.1.5.3.4
SEDIMENT TRAP
General Description

A sediment trap is a temporary control device used to intercept sediment-laden runoff and to trap
sediment to prevent or reduce off-site sedimentation. It is normally a more temporary type of
structure than a sediment pond and is constructed to control sediment on the construction area
during a selected phase of the construction operation. A sediment trap can be formed by;
excavation and/or embankments constructed at designated locations accessible for cleanqut. The
outlet for a sediment trap is typically a porous rock fill structure, which serves to detain the flow,
but a pipe structure can also be used. A temporary sediment trap may be located in a    ]
drainageway, at a storm drain inlet, or at other points of discharge from a disturbed area. They
may be constructed independently or in conjunction with diversions and may be used in most
drainage situations to prevent excessive siltation of pipe  structures (USEPA, 1992).     !

Applicability

Sediment traps can simplify the storm water control plan design process by trapping sediment at
specific spots at a construction site (USEPA,  1992). They should be installed as early in |the
construction process as possible and are primarily effective as a short-term solution to trapping
sediment from construction sites (WYDEQ 1999).  Natural drainage patterns should be nbted,
and sites where runoff from potential erosion can be directed into the traps should be selected.
Traps are most effective when capturing runoff from areas where 2 to 5 acres drain to one
location. Sediment traps should not be located in areas where their failure resulting from excess
storm water-runoff can lead to further erosive damage of the landscape.  Alternative diversion
pathways should be designed to accommodate these potential overflows. Traps should be
accessible for clean-out and located so that they do not interfere with construction activity, hi
addition, the traps are easily adaptable to most conditions.

Design and Implementation Criteria

Hydrologic Design                                                                \

A sediment trap should be designed to maximize surface area and sediment settling. This will
increase the effectiveness of the trap  and decrease the likeliness of backup  during and after
periods of high runoff intensity. The design of a trap includes determining the storage volume,
surface area, dimensions of spillway or outlet, and elevations of embankment (USDOT, 1995).
Sediment traps should be designed to meet a 2- year, 24-hour duration storm event, but the
selection of a return period varies among regulatory agencies (EDNR, 1992).

Storage volume is created by a combination of excavation of land and construction of an
embankment to detain runoff (USDOT, 1995). Trap storage volume and length of spillway are
determined as a function of the runoff volume and rate for the design storm. These parameters
June 2002
                                                                5-76

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

 will vary depending on return period rainfall and watershed hydrologic characteristics. Some
 standards specify a storage volume per acre disturbed. For example, Smolen et al. (1998)
 specified that approximate storage capacity of each trap should be at least 67 cubic yards per
 acre disturbed draining into the trap, but more recent guidelines suggest 134 cubic yards per acre
 of drainage area (VDCR, 2001). Any national standard, however, should be based on runoff
 volume and peak discharge in order to be generally applicable. Local regulations can translate
 this into applicable volume and area standards.

 A more important criterion than storage volume relates to sediment trapping. If a trapping
 efficiency is specified, as in the case of South Carolina (SCDHEC, 1995), it is necessary to
 design for trapping efficiency. If a TSS or settleable solids effluent criterion is adopted
 (SCDHEC, 1995), settleable solids must be estimated. In both cases, a national standard should
 address how to estimate trapping efficiency or settleable solids. Efforts to model the sediment
 trapping that occurs in sediment traps have resulted in models that predict the settling in the '
 ponded area (Barfield et al., 1996; Lindley et al., 1998). The results from model simulations
 show that trapping depends primarily on surface area of the impounded water and flow rate
 through the rock fill outlet. In fact, the ratio of peak outflow rate to surface area is the best
 simple predictor of trapping.  The models utilize a modification of the Herrera and Felton (1991)
 relationship developed by Haan et al. (1994) to predict discharge rates. The predicted flow rates
 do not take into  account clogging that can occur in rock fill.. No models or procedures are
 available to estimate this clogging or its impact on flow criteria.

 Design aids have also been developed for sediment traps, using simulations  from the
 SEDIMOT HI (Barfield et al., 2001; Hayes et al., 2001). In the model, predictions are made of
 trapping efficiency using the ratio of settling velocity for the d15 of the eroded sediment, divided
 by the ratio of discharge to ponded surface area.  The design aid yields conservative estimates,
 but the database used for generating the design aid is based on the assumption that flow rates are
 not impacted by clogging. This latter assumption is not likely to be a critical issue, but should be
 addressed in future research.

 Installation Specifications

 USDOT standards call for the embankment to be constructed of compacted earth, at a maximum
 height of 5 feet (1.5 meters), a width of 4 to 5 feet (1.2 meters), and side slopes of 2:lor flatter.
 These values may change as a result of local criteria and with changing soil characteristics.
 Temporary vegetation should be applied to the embankment (USDOT).
June 2002
                                                                                     5-77

-------
Development Document for Construction and Development Proposed Effluent Guidelines    	    ;	

Two types of outlet structures are typically used for sediment traps, a rock outlet and a pipe
outlet. Spillways of large stones or aggregate are the most common type of outlet designed for
sediment traps. The crest of the spillway should be constructed 1 foot below the top of the
embankment and the spillway depth 1.5 feet below the top of the embankment Weir length of
the spillway is determined based on the contributing drainage area (Table 5-18) (USDOT, 1995),
The outlet apron should be a minimum of 5 feet long, and situated on level ground with a filter
fabric foundation to ensure exit velocity of drainage to receiving stream is nonerosive (IDNR,
1992).

The length of the rock outlet should be determined based on peak discharge required and rock
characteristics, typically rock diameter. Flow rate calculations can be made with the relationship
of Herrera and Felton (1991) as modified by Haan et al. (1994). Alternatively, the USDOT has
specified the weir length for a given drainage area as shown in Table 5-18. However, the values
should be adjusted for each climatologic area to account for local hydrologic and return period
rainfall.

                      Table  5-18. Weir Length for Sediment Traps
Contributing.
Drainage Area
1
2
3
4
5
Weir Length (ft)
4
5
6
10
12
              Source: USDOT, 1995.
The pipe outlet, constructed of corrugated metal or PVC pipe riser, is an alternative to the rock
outlet. Pipe diameter is based on the peak discharge rate required.  To obtain appropriate;
freeboard, the top of pipe should be placed 1.5 feet below embankment elevation. Perforated
pipe is sometimes used. USDOT suggests perforations of 1- inch (25 mm) diameter holes or
0.5 x 6 inch (13x15 mm) slits in the upper two-thirds of the pipe; however, the discharge should
be calculated for this pipe specification to ensure that it matches the required peak discharge.

The pipe should be placed vertically and horizontally above wet storage elevation
(USDOT, 1995).  Riprap should be used as an outlet protection and placed at the outlet of the
barrel to prevent scour from occurring (USDOT, 1995).  A stable channel should be provided to
convey discharge to the receiving channel(USDOT, 1995).                           |

Effectiveness

If it is assumed that the flow can be accurately controlled by the rock fill outlet, sediment traps
should operate as effectively as sediment basins, with trapping efficiencies reduced as a result of
smaller surface areas. The NURP study (USEPA, 1993), Stahre and Urbonas .(1990), and
June 2002
                                                                                     5-78

-------
 Development Document fdr Construction and Development Proposed Effluent Guidelines    	

 Haan, et al., (1994), report that sediment basins effectively trapped sediment and chemical as
 shown in Table 5-19.                           .

             Table 5-19. Range of Measured Long-Term Pollutant Removal for
                                Sediment Detention Basins
Item
Total suspended solids (TSS)
Total phosphorus (TP)
Nitrogen
Organic matter
Lead
Zinc
Hydrocarbons
Bacteria
Removable Percentage
50-70
10-20
10-20
20-40
75-90
30-60
50-70
50-90
               Source: Stahre and Urbonas, 1990.

 Information on the actual effectiveness of sediment trapsis limited. The discussion should start
 first with the flow hydraulics of the rock fill outlet typically employed as a principal spillway for
 sediment traps. Procedures for estimating flow through rock fill have been developed by Herra
 and Felton (1991) to estimate flow as a function of average rock diameter, standard deviation of
 rock size., and flow length.  If these parameters could be controlled in an actual situation, the
 flow could be accurately predicted. However, given that standard construction practices consist
 of end-dumping the rock fill in place, one would expect little correlation between design and
 construction and the actual discharge and trapping efficiency would be expected to be     '
 dramatically different from the design.  This analysis does not mean that sediment traps are
 ineffective, but that a given design could not be guaranteed to meet the effluent criteria, even
 though the predictions indicate compliance. Sediment trapping efficiency is a function of
 surface area and inflow rate (Smolen, 1988). Those traps that provide pools with large length-to-
 width ratios have a greater chance of success.

 Sediment traps remove larger size sediment, primarily sized from silt to sands, by slowing water
 velocity and allowing for sediment settling in ponded water (Haan et al., 1994). Although
 sediment traps allow for settling of eroded soils, because of their short detention periods for
 storm water they typically do not remove fine particles such as  silts and clays without chemical
 treatment.  Sediment settling ability is related to the square of the particle  size; halving particle
 sizes quadruples the time needed to achieve  settlement (WYDEQ 1999). To increase overall
 effectiveness, traps should be constructed in smaller areas with low slopes.

 Sediment traps are typically designed to remove only sediment  from surface water, but some
 non-sediment pollutants are trapped as well (Haan et al., 1994).
June 2002
                                                                                     5-79

-------
Development Document for Construction and Development Proposed Effluent Guidelines

Limitations

Common concerns associated with sediment traps are included in Table 5-20.

             Table 5-20.  Common Concerns Associated with Sediment Traps
Common Concern
Inadequate spillway size
Omission or improper installation of geotextile fabric
Low point in embankment caused by inadequate
Stone outlet apron does not extend to stable grade

Inadequate vegetative protection
Inadequate storage capacity
Contact slope between stone spillway and earth
embankment too steep
Outlet pipe installed in vertical side of trench
Corrugated tubing used as outlet pipe 	
Result
Results in overtopping of the dam and possible failure
of the structure
Results in piping under the sides or bottom of the stone
and outlet section
Results in overtopping and possible failure ;
Results in erosion below the dam
Results in stone displacement
Results in erosion of embankment
Caused by sediment not being removed from the basin
enough
Results in piping failure ;
Results in piping failure of embankment
Results in crushed pipe and inadequate outlet capacity
Source: IDMR, 1992.

Maintenance

The primary maintenance consideration for temporary sediment traps is the removal of  j
accumulated sediment from the basin, which must be done periodically to ensure the continued
effectiveness of the sediment trap. Sediments should be removed when the basin reaches
approximately 50 percent sediment capacity.

A sediment trap should be inspected after each rainfall event to ensure the trap is draining
properly.  Inspectors should also check the structure for damage from erosion or piping. 'The
depth of the spillway should be checked and maintained at a minimum of 1.5 feet below the low
point of the trap embankment

Cost

The cost of installing temporary sediment traps ranges from $0.20 to $2.00 per cubic foot of
storage (about $1,100 per acre of drainage). For a recent national assessment, USEPA (1999)
estimated the following costs for sediment traps, which vary as a function of the volume ;of
storage: $513 for 1,800 cubic yards, $1,670 for 3,600 cubic yards, and $2,660 for 5,400 cubic
yards. In addition, it has been reported that a sediment trap has an annual maintenance cdst of 20
percent of installation cost (Brown et al., 1997).
 June 2002
                                                                                    5-80

-------
 Development Document for Construction and Development Proposed Effluent Guidelines
 5.1.5.3.5
SEDIMENT BASINS
 General Description

 A sediment basin is a storm water detention structure formed by constructing a dam across a
 drainageway or excavating a storage volume at other suitable locations and using it to intercept
 sediment-laden runoff. Sediment basins are generally larger and more effective in retaining
 sediment than temporary sediment traps and typically remain active throughout the construction
 period. Jurisdictions that require postdevelopment flow to be less than or equal to
 predevelopment flow during construction may employ the designed detention facilities as a
 temporary sediment basin during construction.

 When sediment basins are designed properly, they can control sediment pollution through the
 following functions (Faircloth, 1999):

 •   Sediment-laden runoff is caught to form an impoundment of water and create conditions
    where sediment will settle to the bottom of the basin.

 •   Treated runoff is released with less sediment concentration than when it entered the basin.

 •   Storage is provided for accumulated sediment, and resuspension by subsequent storms is
    limited.

 Applicability

 Sediment basins should be located at a convenient concentration point for sediment-laden flows
 (NCDNR,  1988). Ideal sites are areas where natural topography allows a pond to be formed by
 constructing a dam  across a natural swale; such sites are preferred to those that require
 excavation (Smolen et al., 1998).                                                   -      .

 Sediment basins are also applicable in drainage areas where it is anticipated that other erosion
 controls, such as sediment traps, will not be sufficient to prevent off-site transport of sediment.
 Choosing to construct a sediment basin with either an earthen embankment or a stone/rock dam
 will depend on the materials available, location of the basin, and desired capacity for storm water
 runoff and  settling of sediments.

 Rock dams are suitable where earthen embankments would be difficult to construct or where
 riprap is readily available. Rock structures are also desirable where the top of the dam structure
 is to be used as an emergency overflow outlet. These riprap dams are best for drainage areas of
 less than 50 acres. Earthen damming structures are appropriate where failure  of the dam will not
result in substantial damage or loss of property or life. If properly constructed, sediment basins
with earthen dams can handle storm water runoff from drainage basins as large as 100 acres.
June 2002
                                                                                     5-81

-------
 Development Document for Construction and Development Proposed Effluent Guidelines-  	    '.	;	

 Design and Implementation Criteria

 Hydrologic Design                                                              !

 A sediment basin can be constructed by excavation or by erecting an earthen embankment across
 a low area or drainage swale. Sediment basins can be designed to drain completely during dry
 periods, or they can be constructed so that a shallow, permanent pool of water remains between
 storm events. Depending on the size of the basin constructed, the basin may be subject to
 additional regulation, particularly state and federal regulations related to dam safety.

 Sediment basins can be used for any size watershed, but the U.S. Department of Transportation
 recommends a drainage area range of 5 to 100 acres (USDOT,  1995). Components of a sediment
 basin that must be considered in the hydrologic design include the following (Haan et al., 1994):

 •  A sediment storage volume sized to contain the sediment trapped during the life of the
    structure or between cleanouts.

 •  A permanent pool volume (if included) above the sediment storage to protect trapped
    sediment and prevent resuspension as well as providing a first flush of discharge that has
    been subjected to an extended detention period.                                 !

 •  A detention volume that contains storm runoff for a period sufficient to trap the necessary
    quantity of suspended solids.

 •  A principal spillway that can be a drop-inlet pipe and barrel, a trickle tube, or other type of
    controlled release structure.

 •  An emergency spillway that is designed to handle excessive runoff from the rarer events and
    prevent overtopping.                                                         \

• The following recommended procedures for conducting the hydrologic design are summarized
 from Haan etal. (1994).                                                          ;

 Sediment Storage Volume. This volume should be sufficient to store the sediment trapped
 during the life of the structure or between cleanouts.  Sediment storage volume can be calculated
 based on 'sediment yield using relationships such as the Revised Universal Soil Loss Equation
 with an appropriate delivery ratio (Renard et al., 1994) or a computer model such as    !
 SEDMOT III (Barfield et al., 1996). Many design specifications, however,  base the sediment
 storage volume on a volume per acre disturbed.  This volume is highly site-specific, depending
 on rainfall distributions, soil types, and construction techniques. It is recommended that care be
 exercised in developing appropriate values to be sure that existing variations in rainfall
 throughout a state or region are incorporated in the statutory requirements.
 June 2002
                                                                                     5-82

-------
Development Document for Construction and Development Proposed Effluent Guidelines	

Permanent Pool Volume. Providing a first flush of discharge that has been subjected to an
extended detention period can help to minimize degradation of water quality and justify some
permanent pool. The recommended capacity of the permanent pool varies with the regulatory
agency. The U.S. Department of Transportation, for example, recommends 67 cubic yards per
acre (126 mVha) (USDOT, 1995). If an effluent criterion such as allowable peak TSS or peak
settleable solids is used, the final design of both permanent pool and detention volume should be
selected only after using a computer model to predict the expected peak effluent concentrations.

Detention Volume. Storm runoff must be contained for a period of time sufficient to trap the
necessary quantity of suspended solids. Since inflow is occurring simultaneously with outflow,
the detention time for each plug of flow is different and should be considered individually. The
size of the detention volume, as stated above, should also be developed in concert with
determining the size of the permanent pool volume as well as the size of the principal spillway.
When effluent TSS and settleable solids criteria are used, the size of the detention volume and
permanent pool volume should be determined through, on a computer model calculation of
expected effluent concentrations for a given design. The return period used to size the detention
volume depends on the regulatory agency, but a return period of 10 years is typical.

Principal Spillway. The principal spillway is a hydraulic outlet structure sized to provide the
appropriate outflow rate to meet the effluent or trapping efficiency criteria. The principal
spillway should have a dewatering device that slowly releases water contained in the detention
storage over an extended period of time and at a rate determined to trap the required amount of
sediment and/or provide for the appropriate effluent concentration in the design storm.  The
more common outlet structures are the drop-inlet structure and the trickle tube. Sizing of the
principal spillway should follow standard hydrologic and sedimentology design procedures but
sizing the structure to simply pass the design storm is inappropriate and will not result in
meeting an effluent or trapping efficiency standard.  The size to be used in a given structure
should be determined based on the effluent or trapping efficiency standard being targeted and
site-specific hydrologic and soil conditions. Appropriate design will require the use of a
computer model such as SEDIMOT in (Barfield et al., 1996) or design aids such as those
developed for South Carolina (Hayes and Barfield, 1995). In general, the design is developed to
maximize surface area, which will minimize peak discharge. Since failure of the dam could
result in downstream damage, the design should be done and certified by a licensed engineer
with expertise in hydrologic computation.

It has been proposed that a surface skimmer made of PVC, aluminum, or stainless steel and
designed to prevent trash from clogging and can also be used to replace conventional principal
spillways. The skimmer puts the basin drain just below the  water surface, allowing for a
constant head rather than variable head from the bottom.  It  is proposed that the skimmer allows
water to be released from the top of the basin, which would be the cleanest water, and that the
skimmer properly regulates the fill and draining of the basin (Fairchild, 1999). The skimmer
floats on the surface of the basin and rises as water in the basin rises during the storm. After the
storm the skimmer slowly releases water from the basin.  As the basin drains, the skimmer
June 2002
                                                                                    5-83

-------
Development Document for Construction and Development Proposed Effluent Guidelines	L____	

settles to the bottom, draining the entire pool except for a pool directly under the skimmer.  The
skimmer can be attached directly to an outlet pipe that drains through the dam or can be attached
to an outlet pipe through a riser.  It is important to point out that use of the skimmer is    ,
controversial and not universally recognized as a good concept. Conventional hydraulic flow
theory would not concur with the statement that the flow would come only from the surface,
unless the pond had significant thermal gradients preventing flow from deeper levels. A single
hole placed just above the sediment cleanout level can also dewater the basin slowly.

Emergency Spillway. Since overtopping of the dam can cause failure and downstream damage,
an emergency spillway is necessary to handle excessive runoff from the rarer events and prevent
overtopping. The design storm for the emergency spillway will depend on the hazard    \
classification of the sediment basin. Typical return periods vary between 25 and 100 years, with
25 years recommended by the USDOT. Sizing of the emergency spillway is typically
accomplished to simply transmit the rare event without eroding the base of the spillway.
Procedures for making the hydrologic and hydraulic computations are summarized in Haan et al.
(1994).  Again, since failure of the dam could result in downstream damage, the design should
be done and certified by a licensed engineer with expertise in hydrologic computation.

Installation Criteria

The embankment for permanent  sediment basins should use standard geotechnical construction
techniques. The fill is typically constructed of earthen fill material placed and compacted in
continuous layers over the entire length of the fill.  USDOT recommends 6- to 8- inch layers
(USDOT, 1995). The embankment should be stabilized with vegetation after construction of the
basin. A cutoff trench should be excavated along the centerline of the dam to prevent excessive
seepage beneath the dam, and sized using standard geotechnical computations. USDOT
recommends that a minrmum depth of the cutoff trench should be about 2 feet (600 mm), the
height should be to the riser crest elevation, the minimum bottom width should be 4 feet (1.2 m)
or wide enough for compaction equipment, and slopes should be no steeper than 1:1.     ',

Sediment basins can also be constructed with rock dams in a design that is similar to a sediment
basin with an earthen embankment. It is important to remember that rock fill is highly   '
heterogeneous and that flow rates calculated with any available procedure are not likely to match
those that will actually occur.  Since sediment trapping is  inversely proportional to flow r^te, the
trapping efficiency will be impacted significantly.  No data are available to determine the
variability of rock fill in actual installations so that confidence intervals can be placed on .
predicted flow rates.  Such data should be collected and the confidence intervals calculated prior
to recommending the use of rock dams as outlets on any structures other than sediment traps.

Effectiveness

The effectiveness of a sediment basin depends primarily on the sediment particle size and the
ratio of basin surface area to inflow rate (Smolen et al., 1998; Haan et al., 1994). Basins with a
June 2002
                                                                                    5-84

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

 large surface area-to-volume ratio will be most effective.  Studies by Barfield and Clar (1985)
 showed that a surface area-to-peak discharge ratio of 0.01 acres per cubic square foot would trap
 more than 75 percent of the sediment coming from the Coastal Plain and Piedmont regions in
 Maryland. This efficiency might vary for other regions of the country and should not be used as
 a national standard.  Studies by Hayes et al. (1984) and Stevens et al. (2001), however, show that
 similar relationships can be developed for other locations.

 Laboratory data collected on pilot-scale facilities are available on the trapping efficiency of
 sediment basins, on effluent concentrations, on dead storage and flow patterns, and on the
 impacts of chemical fiocculants on sediment trapping (Tapp et al., 1981; Wilson et al., 1984;
 Griffin et al., 1985; Jarrett et al., 1999; Ward et al.,  1977, 1979).  In general, the laboratory
 studies show that pilot-scale ponds can be expected to trap from 70 to 90 percent of sediment,
 depending on the sediment characteristics, pond volume, and flow rate. The trapping efficiency
 and effluent concentration are, in general, related to the overflow rate and can be reasonably well
 predicted using a plug flow model (Ward et al., 1977, 1979) and a Continuously Stirred Tank
 Reactor (CSTR) model (Wilson et al., 1982; Wilson et al., 1984). Extensive field-scale data are
-available on long term trapping efficiency in storm water detention basins (Brune,  1953) in
 which the annual trapping efficiency is related to the annual capacity inflow ratio of the basin.
 These structures are not representative of those used for sediment ponds, but would be
 representative of those used for regional detention.  A more limited database is available on
 single storm sediment trapping in the larger structures (Ward, et al., 1979) and on a field
 laboratory structure at Pennsylvania State University (Jarret et al., 1999).

 For maximum trap efficiency, Smolen et al. (1988) recommend the following:

 •   Allow the largest surface area possible, maximize the length-to-width ratio of the basin to
    prevent short circuiting, and ensure use of the entire design settling area;

 •   Locate inlets for the basin at the maximum distance from the principal spillway outlet;

 •   Allow the maximum reasonable time to detain water before dewatering the basin;  and,

 •   Reduce the inflow rate into the basin and divert all sediment-free  runoff.

 Jarett (1999) has shown that the smaller the depth of the basin, the more sediment is discharged.
 A 0.15 m (0.49 ft) deep basin lost twice as much sediment as a 0.46 m (1.50 ft) deep basin.
 Jarrett also found that the performance of a sediment basin will increase with the use of a
 skimmer in the principal spillway. The sediment discharged was 1.8 times greater with just a
 perforated riser than with a skimmer in the principal spillway. In addition, increasing the de-
 watering time, which will allow for more sediment deposition, decreases the sediment loss from
 the basin (Jarett, 1999).
June 2002
5-85

-------
Development Document for Construction and Development Proposed Effluent Guidelines	    •	

Limitations

Neither a sediment basin with an earthen embankment nor a rock dam should be used in areas of
continuously running water (live streams). The use of sediment basins is not intended for areas
where failure of the earthen or rock dam will result in loss of life, or damage to homes or other
buildings. In addition, sediment basins should not be used in areas where failure will interfere
with the use of public roads or utilities.

Because sediment basins are usually temporary structures, they are often designed poorly and
rarely receive the adequate attention and maintenance. As a result, these basins will not achieve
the function for. which they were designed, especially when conventional outlets cannot property
meter outflow to create an impoundment, thus allowing rapid release of sediment laden water
from the bottom of the basin to escape (Faircloth, 1999).                              i

Common concerns associated with sediment basins are included in Table 5-21.         ;

             Table 5-21. Common Concerns Associated with Sediment Basins
Common Concern
Piping failure along conduit
Erosion of spillway or embankment slopes
Slumping or settling of embankment
Bank failure due to slumping
Erosion and caving below principal spillway
Basin not located properly for access
Sediment not properly removed
Lack of anti-flotation
Principal and emergency spillway on design
plans
Safety or health hazard from pond water
Principal spillway too small
Result
Caused by improper compaction, omission of anti-seep collar,
leaking pipe joints, or use of unsuitable soil i
Caused by inadequate vegetation or improper grading and
sloping
Caused by inadequate compaction or use of unsuitable soil
Caused by steep side slopes
Caused by inadequate outlet protection
Results in difficult, ineffective, and costly maintenance
Results in inadequate storage capacity and potential \
resuspension
Results in the riser and barrel being blocked with debris
Results in improper disposal of accumulated sediment .
Caused by gravel clogging the dewatering system :
Results infrequent operation of emergency spillway and ,
increased erosion potential
Source: IDNR, 1992.                                                                ;
                                                                                !

Maintenance                                                                    :

Routine inspection and maintenance of sediment basins is essential to their continued
effectiveness. Basins should be inspected after each storm event to ensure proper drainage from
me collection pool and determine the need for structural repairs. Erosion from the earthen
embankment or stones moved from rock dams should be replaced immediately.         ;

Sediment basins must be located in an area that is easily accessible to maintenance crews for
removal of accumulated sediment.  Sediment should be removed from the basin when its storage
June 2002
                                                                                    5-86

-------
  Development. Document for Construction and Development Proposed Effluent Guidelines	

  capacity has reached approximately 50 percent. Trash and debris from around dewatering
  devices should be removed promptly after rainfall events.

  Cost

  The sediment basin has a 25 percent annual maintenance cost as a percentage of installation
  (Brown et al., 1997).

  If constructing a sediment basin with less than 50,000 cubic feet of storage space, the cost of
  installing the basin ranges from $0.20 to $1.30 per cubic foot of storage (about $1,100 per acre
  of drainage).  The average cost for basins with less than 50,000 cubic feet of storage is
  approximately $0.60 per cubic foot of storage (USEPA, 1993).

  If constructing a sediment basin with more than 50,000 cubic feet of storage space, the cost of
  installing the basin ranges from $0.10 to $0.40 per cubic foot of storage (about $550 per acre of
  drainage). The average cost for basins with greater than 50,000 cubic feet of storage is
  approximately $0.30 per cubic foot of storage (USEPA, 1993).

 As an alternative costing method, designers can use cost curves developed for permanent basins
 used to manage storm water from urban areas.  However, since permanent storm water basins
• typically include design features that would not be included in temporary sediment basins, this
 approach is expected to greatly overestimate the actual costs to construct sediment basins. For
 many sites, sedimentation basins installed for erosion and sediment control during the
 construction phase are retained/modified to meet other runoff management requirements. For
 example, site flood prevention requirements for the 10-year rainfall event can be met with a
 pond made from a converted sedimentation basin.  As a result, sedimentation basins installation
 costs are partially offset by a later cost reduction or savings.  Work by the Center for Watershed
 Protection (CWP, 1996), provides capital cost equations for different types of sediment basins
 for permanent installations. For example,

    dry extended duration ponds

       CC = 8.16 (Vs) A 0.78

   and for all ponds regardless of type (including wet ponds)  -   .     •

       CC = 20.18 (Vs) A 0.70

   Where:
      . CC = base construction cost, not including design, engineering, and contingencies
       Vs = Storage volume below the crest of the emergency spillway, in cubic feet
June 2002
                                                                                     5-87

-------
Development Document for Construction and Development Proposed Effluent Guidelines             '	

Design, engineering, and contingency costs are given as approximately 32 percent of the base
construction costs. Base construction costs for permanent ponds are composed of approximately
48 percent excavation/grading cost, 36 percent control structure cost, and 16 percent    .
appurtenances cost. R. S. Means (2000) suggests the cost to remove the eroded sediment;
collected in a small basin during construction is approximately $4 per cubic yard (value includes
a 100 percent surcharge for wet excavation). Disposal of material on-site will be an additional
cost that can only be computed from site-specific conditions. The cheapest management of
dredge material is application to land areas adjacent to the basin, followed with application of a
vegetative cover.
5.1.5.4
5.1.5.4.1
Description
OTHER CONTROL PRACTICES

       STONE OUTLET STRUCTURE
A stone outlet structure is a temporary stone dike installed in conjunction with and as a part of an
earth dike. The purpose of the stone outlet structure is to impound sediment-laden runoff,
provide a protected outlet for an e'arth dike, provide for diffusion of concentrated flow, and allow
the area behind the dike to dewater slowly. The stone outlet structure can extend across the end
of the channel behind the dike or be placed in the dike itself. In some cases, more than one stone
outlet structure can be placed in a dike.

Applicability

Stone outlet structures  apply to any point of discharge where there is a need to discharge runoff
at a protected outlet or to diffuse concentrated flow for the duration of the period of construction.
The drainage area to this practice is typically limited to one-half acre or less to prevent excessive
flow rates. The stone outlet structure should be located so as to discharge onto an already
stabilized area or into a stable watercourse. Stabilization should consist of complete vegetative
cover and paving, sufficiently established to be erosion resistant.                      ;

Design and Installation Criteria

Design criteria are of two types, hydrologic design for a required trapping of sediment and/or
flow rate to pass the design storm;  and selection of appropriate installation criteria such that the
stone outlet will perform as designed                                               ;
 June 2002
                                                                                      5-88

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

 Hydrologic Design

 The hydrologic design should be based on the design storm and standard hydraulic calculations
 and should include the following considerations:

 •   Design Rainfall and Design Storm. The design storm should be specified by the regulatory
    authority. Typically a return period of 2 to 5 years is used. Runoff rates should be calculated
    with standard hydrologic procedures, as allowed by the regulatory authority.

 •   Drainage Area. The drainage area to this structure is typically limited to less than half an acre
    to ensure that the flow rates are not excessive.

 •   Length of Crest and Height of Stone Fill. The crest length and height of stone fill should be
    of sufficient size to transmit the design storm without overtopping. The volume of water
    stored behind the dike can be estimated, but would require a routing of the storm flow in the
    design storm. Flow through the stone outlet can be calculated using the relationships of
    Herrera and Felton (1991) as modified by Haan et.al. (1994).  The height of the fill should be
    small enough to prevent excessive flow velocities through the stone fill and. prevent
    undercutting!                                                           '                •

 •   Outlet Stabilization.  The discharge from the stone outlet should be stabilized with vegetated
    waterways or riprap until the flow reaches a stable channel. Design of the stabilized outlet
    should follow procedures presented earlier.

Installation Criteria Specifications

A stone outlet structure should conform to the following specifications:

•   The outlet should be composed of 2- to 3- inch stone or recycled concrete equivalent is
    preferred, but clean gravel may be used if stone is not available.

•   The crest of the stone dike should be at least 6 inches lower than the lowest elevation of the
    top of the earth dike and should be level.

•   The stone outlet structure should be embedded into the soil a minimum of 4 inches.

•   The minimum length of the crest of the stone outlet structure should be 6 feet.

•   The baffle board should extend 1 foot into the dike and 4 inches into the ground and be
    staked in place.

•   The drainage area to this structure should be less than half an acre.
June 2002
                                                                                      5-89

-------
Development Document for Construction and Development Proposed Effluent Guidelines
5.1.5.4.2
ROCK OUTLET PROTECTION
Description                                                                       :

Rock outlet structures are rocks that are placed at the outfall of channels or culverts to reduce the
velocity of flow in the receiving channel to nonerosive rates.

Applicability

This practice applies where discharge velocities and energies at the outlets of culverts are •
sufficient to erode the next downstream reach and it applicable to outlets of all types such as
sediment basins, storm water management ponds, and road culverts.

Design and Installation Criteria

Hydrologic Design

Hydrologic design consists primarily of selecting the design runoff rate and sizing the outlfet
protection. Standard hydrologic calculations should be used to make the calculation, using an
appropriate return period storm for the outlet being protected. Typical return periods range from
2 to 10 years.

Sizing the outlet protection consists of:

•   Selecting the Type of Outlet Protection. The outlet protection may consist of a plunge pool
    (scour hole), an apron-type arrangement, or an energy dissipation basin (Haan et al., 1994).
    The design of each differs. Plunge pools are typically used for outlet pipes that are eleyated
    above the water surface.  Aprons are used for other types  of outlets.                  i

•   Selecting the Geometry of the Outlet.  Plunge pool geometry is based on the flow rate, jpipe
    size and slope, tailwater depth, and size of the riprap lining (Haan et al., 1994). Apron
    dimensions are determined by the ratio of the tailwater depth to pipe diameter (Haan et al.,
    1994). Energy dissipation basins are used as an alternative to the plunge pool. Dimensjions
    are a function of the brink depth-in the pipe at the design flow, pipe diameter, and size of
    riprap (Haan etal, 1994).                                                        ;
                                                                                   !
•   Size of Rock Lining. The size of the rock lining is a function of the discharge, pipe size,
    tailwater depth, and  geometry selected. Details on sizing the rock are given in Haan et al.
    (1994).

The design method presented here applies to the sizing of rock riprap and gabions to protect a
downstream area. It does not apply to rock lining of channels or streams. The design of rock
outlet protection depends entirely on the location. Pipe outlets at the top of cuts or on slopes
June 2002
                                                                 5-90

-------
Development Document for Construction and Development Proposed Effluent Guidelines	

steeper than 10 percent cannot be protected by rock aprons or riprap sections due to
reconcentration of flows and high velocities encountered after the flow leaves the apron.

Installation Criteria

The following criteria should be considered.

•   Bottom Grade: The outlet protection apron should be constructed with no slope along its
    length. There should be no obstruction at the end of the apron.  The elevation of the
    downstream end of the apron should be equal to the elevation of the receiving channel or
    adjacent ground.

•   Alignment: The outer protection apron should be located so that there are no beds in the
    horizontal alignment.

•   Materials:  The outlet protection may be done using rock riprap, or gabions. Riprap should
    be composed of a well-graded mixture of stone sized so that 50 percent .of the pieces, by
    weight, should be larger than the size determined by using the charts. The minimum d50 size
    to be used should be 9 inches. A well-graded mixture is defined as a mixture composed
    primarily of larger stone sizes but with a sufficient mixture of other sizes to fill the smaller
    voids between the stones. The diameter of the largest stone in such a mixture should be
    2.0 times the size selected in Table 5-22 (MDE, 1994).

•   Thickness:  The SHA riprap specification values are summarized in Table 5-22.

Class I
Class II
Class III
D,n (inches)
9.5
16
23
D,nn (inches)
15
24
34
Thickness (inches)
19
32
46
    Stone Quality: Stone for riprap should consist of field stone or rough and hewn quarry stone.
    The stone should be hard and angular and of a quality that will not disintegrate on exposure
    to water or weathering. The specific gravity of the individual stones should be at least 2.5.
    Recycled concrete equivalent may be used provided it has a density of at least 150 pounds
    per cubic foot and does not have any exposed steel or reinforcing bars.

    Filters: A filter is a layer of material placed between the riprap and the underlying soil
    surface to prevent soil movement into and through the riprap to prevent piping, reduce uplift
    pressure, and collect water. Riprap should have a filter placed under it in all cases.  A filter
    can be of two general forms: a gravel layer or a geotextile.
June 2002
                                                                                      5-91

-------
 Development Document for Construction and Development Proposed Effluent Guidelines                   	

 •  Gabions:  Gabion baskets may be used as rock outlet protection, provided they are made of
    hexagonal triple twist mesh with heavily galvanized steel wire. The maximum lined;
    dimension of the mesh opening should not exceed 4.5 inches. The area of the mesh opening
    should not exceed 10 square inches. Gabions should be fabricated in such a manner that the
    sides, ends, and lid can be assembled at the construction site into a rectangular basket of the
    specified sizes. Gabions should be of a single unit construction and should be installed
    according to the manufacturer's specifications.  The area on which the gabion is to be
    installed should be graded as shown on the drawings. Foundation conditions should be the
    same as for placing rock riprap. Geotextiles should be placed under all gabions.  Gabions
    must be keyed in to prevent undermining of the main gabion structure.

 •  The subgrade for the filter, riprap, or gabion should be prepared to the required lines and
    grades.  Any fill required in the subgrade shall be compacted to a density of approximately
    that of the surrounding undisturbed material.                                     '

 •  The rock or gravel should conform to the specified grading limits when installed in the riprap
    or filter, respectively.                                                          !

 •  Geotextiles should be protected from punching, cutting, or tearing.  Any damage other than
    occasional small holes should be repaired by placing another piece of geotextile fabric over
    the damaged part or by completely replacing the geotextile fabric. All overlaps, whether for
    repairs or for joining two pieces of geotextile fabric, should be a minimum of 1 foot in
    length.                                                                       ;

 •   Stone for the riprap or gabion outlets may be placed by equipment.  They should be
    constructed to the full course thickness in one operation and in such a manner as to  avoid
    displacement of underlying materials. Care should be taken to ensure that the stone is not
    placed so that rolling will cause segregation of stone by size, i.e., the stone for riprap or.
    gabion outlets should be delivered and placed in a manner that will ensure that it is
    reasonably homogeneous with the smaller stones filling the voids between the larger stones.
    Riprap must be placed in a manner to prevent damage to the filter blanket or geotextile
    fabric. Hand placement will be required to the extent necessary to prevent damage to the
    permanent works.                                                             '

 •   Stone should be placed so that it blends in with the existing ground and the depth to the  stone
    surface is sufficient to transmit the flow without spilling over onto the unprotected surface.

Effectiveness

There is currently no information on the effectiveness of rock outlet structures.
June 2002
                                                                                     5-92

-------
 Development Document for Construction and Development Proposed Effluent Guidelines 	

 Limitations               '

 Common problems with rock outlet structures include the following:

 •  Foundation not excavated deep enough or wide enough—restricts the flow cross-section,
    resulting in erosion around the apron and sour holes at the outlet.

 •  Riprap apron should be placed on a suitable foundation to prevent downstream erosion.

 •  Riprap installed smaller than specified—results in rock displacement; selectively grouting
    over the rock materials may stabilize the situation.

 •  Riprap not extended enough to reach a stable section of the channel—results in downstream
    erosion.

 •  No filter installed under the riprap—results in stone displacement and erosion of the
    foundation.

 Maintenance

 Once a riprap outlet has been installed, the maintenance needs are very low. It should be
 inspected after high flows to see if scour has occurred beneath the riprap, if flows have occurred
 outside the boundaries  of the riprap and caused scour, or if any stones have been dislodged.
 Repairs should be made immediately.

 Cost

 R. S. Means indicates machine-placed riprap costs of approximately $40 per cubic yard. For a
 riprap maximum size between 15 and 24 inches, a cubic yard of riprap will cover between 13.5
 and 17 square feet for channel bed (assuming depth of riprap as given in Table 5-22). This
 suggests that riprap lining will be between $21 and $27 per square foot of outlet (includes
 materials, labor, and equipment, with overhead and profit). R. S. Means (2000) provides a cost
 range for gabions ($2.80 to $9 per square foot of coverage) for stone fill depths of 6" to 36",
 respectively. These costs include all costs of materials, labor, and installation.
5.1.5.4.3
SUMP PIT
Description

A sump pit is a temporary pit from which pumping is conducted to remove excess water while
minimizing sedimentation. The purpose of the sump pit is to filter water being pumped to
reduce sedimentation to receiving streams.
June 2002
                                                                                     5-93

-------
Development Document for Construction and Development Proposed Effluent Guidelines	     •	

Applicability
                                                                                |
Sump pits are constructed when water collects and must be pumped away during excavating,
cofferdam dewatering, maintenance or removal of sediment traps and basins, or other uses as
applicable, such as" for concrete wash out.

Design and Installation Criteria                                                 '.

Hydrologic Design                                                               :

The only hydrologic calculation is determining the expected flow rate and volume to be handled.
This should follow standard hydrologic computational procedures based on design rainfall,
surface and soil conditions, and the size of the pump.                                ;

Installation Criteria and Specifications                                             i

The number of sump pits and their locations should be determined by the designer and included
on the plans.  Contractors may relocate sump pits to  optimize use, but discharge location changes
should be coordinated with inspectors.                                             ;

A perforated vertical sandpipe is wrapped with Vi inch hardware cloth and geotextiles and then
placed in the center of an excavated pit which is then backfilled with filter material consisting of
anything from clean gravel to stone.  Water is then pumped from the center of the sandpipe to a
suitable discharge area such as into a sediment trap,  sediment basin, or stabilized area.   ,

A sump pit should conform to the following specifications:

•   Pit dimensions are variable, with the minimum diameter being twice the diameter of the
    sandpipe.                                                                    :

•   The sandpipe should be constructed by perforating a 12- to 36-inch diameter pipe, then
    wrapping it with ^-inch hardware cloth and geotextiles. The perforations should be H- x 6-
    inch slits or 1-inch diameter holes 6 inches on center.                             \

•   The sandpipe should extend 12 to 18 inches above the lip of the pit or riser crest elevation
    (basin dewatering), and filter material should extend 3 inches minimum above the anticipated
    standing water level.

Effectiveness

There is currently no information on the effectiveness of the sump pit.
June 2002
                                                                                     5-94

-------
 Development Document for Construction and Development Proposed Effluent Guidelines   	

 Limitations

 The sump pit must be properly maintained and pumped regularly to avoid clogging.

 Maintenance

 To maintain, sump pits must be removed and reconstructed when water can no longer be pumped
 out of the sandpipe.

 Cost

 R. S. Means (2000) provides information appropriate for assessment of a wide range of
 dewatering scenarios (i.e., different sump sizes, dewatering durations, and discharge conditions).
 In general, installation of earthen sump pits are listed as costing approximately $1.50 per cubic
 foot of sump volume.  Piping to and away from the sump ranges from $30 to $60 per linear foot.
 Pump rentals and operation range between $150 and $500 per day of pumping, depending on the
 rate of dewatering. All costs include material, labor, and equipment, with overhead and profit.
 5.1.5.4.4
 Description
SEDIMENT TANK
A sediment tank is a compartmented tank container through which sediment-laden water is
pumped to trap and retain the sediment.  The purpose of a sediment tank is to trap and retain
sediment prior to pumping the water to drainageways, adjoining properties, and rights-of-way
below the sediment tank site.

Applicability

A sediment tank should be used on sites where excavations are deep and space is limited, such as
urban construction, where direct discharge of sediment-laden water to streams and storm
drainage systems should be avoided.

Design and Installation Criteria

The location of sediment tanks should facilitate easy cleanout and disposal of the trapped
sediment to minimize interference with construction activities and pedestrian traffic. The tank
size should be determined according to the storage volume of the sediment tank, 1 cubic foot of
storage for each gallon per minute of pump discharge capacity.
June 2002
                                                                                   5-95

-------
Development Document for Construction and Development Proposed Effluent Guidelines             	

Effectiveness

There is currently no information on the effectiveness of sediment tanks.

Limitations                                                                      :

The sediment tank does not provide any natural infiltration; thus, the trapped sediment and storm
water must be disposed of properly.

Maintenance

To properly maintain the sediment tank, it needs to be in a location that is easy to access.;

COSt                                                                             ;
There is currently no information on the cost of sediment tanks.                       ;

5.1.5.4.5             STABILIZED CONSTRUCTION ENTRANCE

Description

The purpose of stabilizing entrances to a construction site is to minimize the amount of sediment
leaving the area as mud attached to motorized vehicles.. Installing a pad of gravel over filter
cloth where construction traffic leaves a site can help stabilize a construction entrance.  As a
vehicle drives over the gravel pad, mud and. other sediments are removed from the vehicle's
wheels (sometimes by washing) and offsite transport of soil is reduced. The gravel pad also
reduces erosion and rutting on the soil beneath the stabilization structure. The fabric reduces the
amount of rutting caused by vehicle tires by spreading the vehicle's weight over a larger soil area
than just the tire width. The filter fabric also separates the gravel from the soil below, preventing
the gravel from being ground into the soil.                                           *

Applicability

Typically, stabilized construction entrances are installed at locations where construction traffic
leaves or enters an existing paved road. However, the applicability of site entrance stabilization
should be extended'to any roadway or entrance where vehicles will access or leave the site.

From a public relations point of view, stabilizing construction site entrances can be a worthwhile
exercise. If the site entrance is the most publicly noticeable part of a construction site, stabilized
entrances can improve the appearance to passersby and improve public perception of the
construction project by reducing the amount of mud tracked onto adjacent streets.
June 2002
                                                                                     5-96

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

 Design and Installation Considerations

 Hydrologic Design

 Not applicable.

 Installation Criteria and Specifications

 All entrances to a site should be stabilized before construction begins and further disturbance of
 the site area occurs.  The stabilized site entrances should be long enough and wide enough so
 that the largest construction vehicle that will enter the site will fit in the entrance with room to
 spare. If many vehicles are expected to use an entrance hi any one day, the site entrance should
 be wide enough for the passage of two vehicles at the same time with room on either side of each
 vehicle. For optimum effectiveness, a rock construction entrance should be at least 50 feet long
 and at least 10 to 12 feet wide (USEPA, 1992). If a site entrance leads to a paved road, the end
 of entrance should be "flared" (made wider as hi the shape of a funnel) so that long vehicles do
 not go off the stabilized area when turning onto or off of the paved roadway.

 If a construction site entrance crosses a stream, swale, roadside channel, or other depression, a
 bridge or culvert should be provided to prevent erosion from unprotected banks.

 Stone and gravel used to stabilize the construction site entrance should be large enough so that
 they are not carried off-site with vehicle traffic. In addition, sharp-edged stone should be
 avoided to reduce the possibility of puncturing vehicle tires.  Stone or gravel should be installed
 at a depth of at least 6 inches for the  entire length and width of the stabilized construction
 entrance.

 Effectiveness

 Stabilizing construction entrances to prevent sediment transport off-site is effective only if all
 entrances to the site are stabilized and maintained. Also, stabilization of construction site
 entrances may not be very effective unless a wash rack is installed and routinely used (Corish,
 1995) but this can be problematic for sites with multiple entrances with high vehicle traffic.

Limitations

Although stabilizing a construction entrance is a good way to help reduce the amount of
sediment leaving a site, some soil may still be deposited from vehicle tires onto paved surfaces.
To further reduce the chance that these sediments will pollute storm water runoff, sweeping of
the paved area adjacent to the stabilized site entrance is recommended.

For sites using wash stations, a reliable water source to wash vehicles before leaving the site
might not be initially available, hi this case, water may have to be trucked to the site at
June 2002
                                                                                       5-97

-------
Development Document for Construction and Development Proposed Effluent Guidelines	     !      	

additional cost. Discharge from the wash station should be directed into an appropriate sediment
control structure.
                                                                                  i
Maintenance

Stabilization of site entrances should be maintained until the remainder of the construction site
has been fully stabilized. Stone and gravel might need to be periodically added to each
stabilized construction site entrance to keep the entrance effective.  Soil that is tracked offsite
should be swept up immediately for proper disposal.                                  •

For sites with wash racks at each site entrance, sediment traps will have to be constructed and
maintained for the life of the project. Maintenance will entail the periodic removal of sediment
from the traps to ensure their continued effectiveness.

Cost

Without a wash rack,  construction site entrance stabilization costs range from $1,000 to $4,000.
On average, the initial construction cost is around $2,000 per entrance. When maintenance costs
are included, the average total annual cost for a 2-year period, is approximately $ 1,500. ;

If a wash rack is included in the construction site entrance stabilization, the initial construction
costs range from $1,000 to $5,000, with an average initial cost of $3,000 per entrance.  Tptal
annual cost, including maintenance for an estimated 2-year life span, is approximately $2,200
per year (USEPA, 1993).                                                           !
5.1.5.4.6
Description
LAND GRADING
Land grading involves reshaping the ground surface to planned grades as determined by an
engineering survey, evaluation, and layout.  Land grading provides more suitable topography for
buildings, facilities, and other land uses and helps to control surface runoff, soil erosion, and
sedimentation both during and after construction.

Applicability                                                                     ;

Land grading is applicable to sites with steep topography or easily erodible soils because it
stabilizes slopes and decreases runoff velocity. Grading activities should maintain existing
drainage patterns as much as possible.       •                                        ;
June 2002
                                                                  5-98

-------
Development Document for Construction and Development Proposed Effluent Guidelines	

Design and Installation Criteria

Before grading activities begin, decisions should be made regarding the steepness of cut-and-fill
slopes and how the slopes will be protected from runoff, stabilized, and maintained. A grading
plan that establishes which areas of the site will be graded, how drainage patterns will.be
directed, and how runoff velocities will affect receiving waters should be prepared.  The grading
plan also includes information regarding when earthwork will start and stop, establishes  the
degree and length of finished slopes, and dictates where and how excess material will be
disposed of (or where borrow materials will be obtained if needed). Berms, diversions, and
other storm water practices that require excavation and filling should also be incorporated into
the grading plan.

A low-impact development BMP that can be incorporated into a grading plan is site
fingerprinting, which involves clearing arid grading only those areas necessary for building
activities and equipment traffic.  Adhering to strict limits of clearing and grading helps to
maintain undisturbed temporary or permanent buffer zones in the grading operation and provides
a low-cost sediment control measure that will help reduce runoff and off-site sedimentation.  The
lowest elevation of the site should remain undisturbed to provide a protected storm water outlet
before storm drains or other construction outlets are installed.

Effectiveness

Land grading is an effective means of reducing steep slopes and stabilizing highly credible soils
when implemented with storm water management and erosion and sediment control practices in
mind. Land grading is not effective when drainage patterns are altered or when vegetated areas
on the perimeter of the site are destroyed.

Limitations

Construction sites are routinely graded to prepare a site for buildings and other structures.
Improper grading practices that disrupt natural storm water patterns might lead to poor drainage,
high runoff velocities, and increased peak flows during storm events. Clearing and grading of
the entire site without vegetated buffers promotes off-site transport of sediments and other
pollutants. Grading plans should be designed with erosion and sediment control and storm water
management goals in mind; grading crews should be carefully supervised to ensure that the plan
is implemented as intended.
June 2002
5-99

-------
Development Document for Construction and Development Proposed Effluent Guidelines	___,	

Maintenance

All graded areas and supporting erosion and sediment control practices should be periodically
checked, especially after heavy rainfalls. All sediment should be promptly removed from
diversions or other storm water conveyances.  If washouts or breaks occur, they should be
repaired immediately. Prompt maintenance of small-scale eroded areas is essential to prevent
these areas from becoming significant gullies.

Cost

Land grading is practiced at virtually all construction sites—additional site planning to
incorporate storm water and erosion and sediment controls in grading plans can require several
hours of planning by a certified engineer or landscape architect. Extra time might be required to
excavate diversions and construct berms, and fill materials might be needed to build up low-
lying areas or fill depressions.

Where grading is performed to manage on-site storm water,  R. S. Means (2000) suggests 'the
cost of fine grading, soil treatment, and grassing to be approximately $2 per square yard of earth
surface area. Shallow excavation/trenching (1 to 4 feet deep) with a backhoe in areas not
requiring dewatering can be performed for $4 to $5 per cubic yard of removed material. Larger
scale grading requires a site-specific assessment of an alternative grading apparatus and a
detailed fill/excavation material balance to retain as much soil on site as possible.
5.1.5.4.7
Description
TEMPORARY ACCESS WATERWAY CROSSING
A temporary stream crossing is a structure erected to provide a safe and stable way for   !
construction vehicle traffic to cross a running watercourse. The primary purpose of such a
structure is to provide streambank stabilization, to reduce the risk of damaging the streambed or
channel, and to reduce the risk of sediment loading from construction traffic. A temporary
stream crossing may be a bridge, culvert, or ford.                                     :

Applicability

Temporary stream crossings are applicable wherever heavy construction equipment must be
moved from one side of a stream channel to the other or where lighter construction vehicles will
cross the stream a number of tunes during the construction period. In either case, an appropriate
method for ensuring the stability of the streambanks and preventing large-scale erosion is'
necessary.

A bridge or culvert is the best choice for most temporary stream crossings.  If properly de'signed,
each can support heavy loads, and materials used to construct most bridges  and culverts can be
June 2002
                                                                                    5-100

-------
  Development Document for Construction and Development Proposed Effluent Guidelines	

  salvaged after they are removed. Fords are appropriate in steep areas subject to flash flooding,
  where normal flow is shallow or intermittent across a wide channel. Fords should be used only
  where stream crossings are expected to be infrequent.

  Design and Installation Criteria

  Because of the potential for stream degradation, flooding, and safety hazards, stream crossings
  should be avoided on a construction site whenever possible. Consideration should be given to
  alternative site access routes before arrangements are made to erect a temporary stream crossing.
  If it is determined that a stream crossing is necessary, an area where the potential for erosion is
  low should be selected.  The stream crossing structure should be selected during a dry period if
  possible to reduce sediment transport into the stream.

  If needed, over-stream bridges are generally the preferred temporary stream crossing structure.
  The expected load and frequency of the stream crossing, however, will govern the selection of a
  bridge as the correct choice for a temporary stream crossing. These types of temporary bridges
  usually cause minimal disturbance to a stream's banks and cause the least obstruction to stream
  flow and fish migration. They should be constructed only under the supervision and approval  of
  a qualified engineer.

 As general guidelines for constructing temporary bridges,  clearing and excavation of the stream
 shores and bed should be kept to a minimum. Sufficient clearance should be provided for
 floating objects to pass under the bridge. Abutments should be parallel to the stream and on
 stable banks.  If the stream is less than 8 feet wide at the point where a crossing is needed, no
 additional in-stream supports should be used.  If the crossing is to extend across a channel wider
 than 8 feet (as measured from top of bank to top of bank),  the bridge should be designed with
 one in-water support for each 8 feet of stream width.

 A temporary bridge should be anchored by steel cable or chain on one side only to a stable
 structure on shore. Examples of anchoring structures include trees with a large diameter, large
 boulders, and steel anchors. By anchoring the bridge on one side only, there is a decreased risk
 of causing a downstream blockage or flow diversion if a bridge is washed out.

 When constructing a culvert, filter cloth should be used to  cover the streambed and streambanks
 to reduce settlement and improve the stability of the culvert structure. The filter cloth should
 extend a minimum of 6 inches and a maximum of 1 foot beyond the end of the culvert and
 bedding material. The culvert piping should not exceed 40 feet in length and should be of
 sufficient diameter to allow for complete passage of flow during peak flow periods.  The culvert
 pipes should be covered with a minimum of 1 foot of aggregate. If multiple  culverts are used, at
 least 1 foot of aggregate should separate the pipes.

Fords should be constructed of stabilizing material such as  large rocks.
June 2002
                                                                                    5-101

-------
Development Document for Construction and Development Proposed Effluent Guidelines              	.—

Effectiveness                                                                    ;

Both temporary bridges and culverts provide an adequate path for construction traffic crossing a
stream or watercourse.                                                             ;

Limitations

Bridges can be considered the greatest safety hazard of all temporary stream crossing structures
if not properly designed and constructed. Bridges might also prove to be more costly in terms of
repair costs and lost construction tune if they wash out or collapse (Smolen et al., 1988).

The construction and removal of culverts are usually very disturbing to the surrounding area, and
erosion and downstream movement of soils are often great.  Culverts can also create obstructions
to flow in a stream and inhibit fish migration. -Depending on their size, culverts can be blocked
by large debris in a stream and are therefore vulnerable to frequent blockage and washout.

If given a choice between building a bridge or a culvert as a temporary stream crossing, a bridge
is preferred because of the relative minimal disturbance to'streambanks and the opportunity for
unimpeded flow through the channel. The approaches to fords often have high erosion potential.
In addition, excavation of the streambed and approach to lay riprap or other stabilization imaterial
causes major stream disturbance.  Mud and other debris are transported directly into the stream
unless the crossing is used only during periods of low flow.                           ;
                                                                                 E
Maintenance

Temporary stream crossings should be inspected at least once a week and after all significant
rainfall events. If any structural damage is reported to a bridge or culvert, construction traffic
should stop using the structure until appropriate repairs are made. Evidence of streambahk
erosion should be repaired immediately.

Fords should be inspected closely after major storm events to ensure that stabilization materials
remain in place.  If the material has moved downstream during periods of peak flow, the|lost
material should be replaced immediately.                                           ;

 Cost

 In general, temporary bridges are more expensive to design and construct than culverts.  Bridges
 are also associated with higher maintenance and repair costs should they fail. Additional costs
 may accrue to the site team in terms of lost construction time if a temporary structure is washed
 out or otherwise fails.                                                             ;

 Temporary bridging costs range as a function of the width of the bridge span and the duration of
 application.  If the bridging is permanent, a mean cost of $50 per square foot for an 8-foot wide

 June 2002             -                                                           !    5"102

-------
 Development Document for Construction and Development Proposed Effluent Guidelines     	

 steel arch bridge (no foundation costs included) can be used for conceptual cost estimation
 (R. S. Means, 2000). If rental bridging is employed, then rates are probably on the order of 20 to
 50 percent of the bridge (permanent) cost, but will range based on the rental duration and
 mobilization distance.
 5.1.5.4.8
DUST CONTROL
 General Description

 Dust control measures are practices that help reduce ground surface and air movement of dust
 from disturbed soil surfaces. Construction sites are good candidates for dust control measures
 because land disturbance from clearing and excavation generates a large amount of soil
 disturbance and open space for wind to pick up dust particles. To illustrate this point, research at
 construction sites has established an average dust emission rate of 1.2 tons/acre/month for active
 construction (WA Dept. of Ecology, 1992).  These airborne particles pose a dual threat to the
 environment and human health.  First, dust can be carried off-site, thereby increasing soil loss
 from the construction area and increasing the likelihood of sedimentation and water pollution.
 Second, blowing dust particles can contribute to respiratory health problems and create an
 inhospitable working environment.

 Applicability

 Dust control measures are applicable to any construction site where dust is created and there is
 the potential for air and water pollution from dust traveling across the landscape or through the
 air.  Dust control measures are particularly important in arid or semiarid regions where soil can
 become extremely dry and vulnerable to transport by high winds.

 Also, dust control measures should be implemented on all construction sites where there will be
 major soil disturbances or heavy construction activity, such as clearing, excavation, demolition,
 or excessive vehicle traffic.  Earthmoving activities are the major source of dust from
 construction sites, but traffic and general disturbances can also be major contributors (WA Dept
 of Ecology, 1992).

 The specific dust control measures implemented at a site will depend on the topography, land
 cover, soil characteristics and amount of rainfall at the site.

 Design  and Installation Criteria

 When designing a dust control plan for a site, the amount of soil exposed will dictate the
 quantity of dust generation and transport. Therefore, construction sequencing and disturbing
 small areas at one time can greatly reduce problematic dust from a site. If land must be
 disturbed, additional temporary stabilization measures should be  considered prior to disturbance.
June 2002
                                                                                    5-103

-------
Development Document for Construction and Development Proposed Effluent Guidelines	

A number of methods can be used to control dust from a site. The following is a brief list of
control measures and their design criteria. Not all control measures will be applicable to a given
site. The owner, operator, and contractors responsible for dust control at a site should determine
which practices accommodate theu: needs based on specific site and weather conditions.  •

Sprinkling/Irrigation: Sprinkling the ground surface with water until it is moist is an effective
    dust control method for haul roads and other traffic routes (Smolen et al., 1988). This
    practice can be applied to almost any site.                                         >

Vegetative Cover.  In areas not expected to handle vehicle traffic, vegetative stabilization! of
    disturbed soil is often desirable.  Vegetative cover provides protection to surface soils, and
    slows wind velocity at the ground surface, thus reducing the potential for dust to become
    airborne.

Mulch:  Mulching can be a quick and effective means of dust control for a recently disturbed
    area (Smolen et al., 1988).
                                                                                  I
Wind Breaks:  Wind breaks are barriers (either natural or constructed) that reduce wind velocity
    through a site and therefore reduce the possibility of picking up suspended particles.  Wind
    breaks can be trees or shrubs left in place during site clearing or constructed barriers such as
    a wind fence, snow fence, tarp curtain, hay bale, crate wall, or sediment wall (USEPA,
    1992).                                                                         ;

Tillage: Deep tillage in large open areas brings soil clods to the surface where they rest on top of
    dust, preventing it from becoming airborne.

Stone: Stone can be an effective dust deterrent for construction roads and entrances.

Spray-on Chemical Soil Treatments (palliatives): Examples of chemical adhesives include
    anionic asphalt emulsion, latex emulsion, resin-water emulsions, and calcium chloride.
    Chemical palliatives should be used only on mineral soils. When considering chemical
    application to suppress dust, consideration should be taken as to whether the chemicaj is
    biodegradable or water-soluble and what effect its application could have on the surrounding
    environment, including waterbodies and wildlife.                                 ;
June 2002
                                                                                     5-104

-------
Development Document for Construction and Development Proposed Effluent Guidelines	

Table 5-23 shows application rates for some common spray-on adhesives as recommended by
Smolen et al. (1988).

                  Table 5-23. Application Rates for Spray-On Adhesives
Spray on Adhesive
Anionic Asphalt Emulsion
Latex Emulsion
Resin in Water
Water Dilution
7:1
12.5:1
4:1
Type of Nozzle
Coarse spray
Fine spray
Fine spray
Application (gal/acre)
1,200
235
300
Source: Smolen et al., 1988.

Effectiveness

Sprinkling/Irrigation: Not available.

Vegetative Cover: Not available.

Mulch: Can reduce wind erosion by 80 percent.

Wind Breaks/Barriers:  For each foot of vertical height, an 8- to 10-foot deposition zone develops
    on the lee.ward side of the barrier. The barrier density and spacing will change its
    effectiveness at capturing windborne sediment.

Tillage:.Roughening the soil can reduce soil losses by approximately 80 percent.

Stone:  The sizes of the  stone can affect the amount of erosion that will take place.  In areas of
    high wind, small stones are not as effective as a 20 cm stone.

Spray-on Chemical Soil Treatments (palliatives): Effectiveness of polymer stabilization
    methods ranges from 70 percent to 90 percent.

Limitations

In areas where evaporation rates are high, water application to exposed soils may require near
constant attention.  If water is applied in excess, runoff may result from the site and possibly
create conditions where vehicles could track mud onto public roads.

Chemical applications should be used sparingly and only on mineral soils (not high organic
content soils) because their misuse can create additional surface water pollution from runoff or
contaminate groundwater.  Chemical applications might also present a health risk if excessive
amounts are used.
June 2002
5-105

-------
 Development Document for Construction and Development Proposed Effluent Guidelines

 Maintenance                                                                     i

 Because dust controls are dependent on specific site conditions, including the weather,
 inspection and maintenance are unique for each site. Generally, however, dust control measures
 involving application of either water or chemicals require more monitoring than structural or
 vegetative controls to remain effective. If structural controls are used, they should be inspected
 for deterioration on a regular basis to ensure, they are still achieving their intended purpose.

 Cost                                                                             ;

 Chemical dust control measures can vary widely in cost depending on specific needs of the site
 and level of dust control desired.  One manufacturer of a chloride product estimated a cost of
 $1,089 per acre for application to road surfaces, but cautioned that cost estimates withoutja
 specific site evaluation are rather inaccurate.
 5.1.5.4.9
Description
STORM DRAIN INLET PROTECTION
 Storm drain inlet protection measures are controls that help prevent soil and debris from dn-site
 erosion from entering storm drain drop inlets." Typically, these measures are temporary cbntrols
 that are implemented prior to large-scale disturbance of the surrounding site.  These controls are
 advantageous because their implementation allows storm drains to be used during even thfe early
 stages of construction activities.  The early use of storm drains during project development
 significantly reduces the occurrence of future erosion problems (Smolen et al, 1988).

 Three temporary control measures to protect storm drain drop inlets are                 ;

 •   Excavation around the perimeter of the drop inlet                                 !

 •   Fabric barriers around inlet entrances

 •   Block and gravel protection

Excavation around a storm drain inlet creates a settling pool to remove sediments.  Weep holes
protected by gravel are used to drain the shallow pool of water that accumulates around the inlet.
A fabric barrier made of porous material erected around an inlet can create an effective shield to
sediment while allowing water to flow into the storm dram. This type of barrier can slow runoff
velocity while catching soil and other debris at the drain inlet. Block and gravel inlet protection
uses standard concrete blocks and gravel to form a barrier to sediments while permitting Water
runoff through select blocks that are laid sideways.
June 2002
                                                                                    5-106

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

 In addition to the materials listed above, limited temporary storm water drop inlet protection can
 also be achieved with the use of straw bales or sandbags to create barriers to sediment.

 For permanent storm drain drop inlet protection after the surrounding area has been stabilized,
 sod can be installed as a barrier to slow storm water entry to storm drain inlets and capture
 sediments from erosion. This final inlet protection measure can be used as an aesthetically
 pleasing way to slow storm water velocity near drop inlet entrances and remove sediments and
 other pollutants from runoff.

 A new technology that uses an insert trap into the inlet itself has been developed (Adams et al.,
 2000).  This technique showed good results on initial testSj trapping more than 50 percent of the
 incoming sediment in flows typical of those into urban storm drains. This technique is being
 further developed with a pending patent application.

 Applicability

 All temporary controls should have a drainage area no greater than 1 acre per inlet. It is also
 important for temporary controls to be constructed prior to disturbance of the surrounding
 landscape. Excavated drop inlet protection and block and gravel inlet protection are applicable
 to areas of high flow where overflow is anticipated into the storm drain. Fabric barriers are
 recommended for smaller, relatively flat drainage areas (slopes less than 5 percent leading to the
 storm drain).

 Temporary drop inlet control measures are often used in combination with each other and with
 other storm water control techniques.

 Design and Installation Considerations

Hydrologic Design

Hydrologic computations are not necessary with present technologies. A specified limitation of
 1 acre per inlet limits flow rates, dependent on local rainfall and runoff considerations.

Installation Criteria and Specifications

The following criteria should be followed until future research establishes better techniques:

•   With the exception of sod drop  inlet protection, these controls should be installed before any
    soil disturbance in the drainage  area.                                      ,  .          -

•   Excavation around drop inlets should be dug a minimum of 1 foot deep  (2 feet maximum)
    with a minimum excavated volume of 35 cubic yards per acre disturbed. Side slopes leading
    to the inlet should be no steeper than 2:1. The shape of the excavated area should be
June 2002
                                                                                     5-107

-------
Development Document for Construction and Development Proposed Effluent Guidelines -            :•	

   designed such that the dimensions fit the area from which storm water is anticipated to drain.
   For example, the longest side of an excavated area should be along the side of the inlet
   expected to dram the largest area.                                               '

•  Fabric inlet protection is essentially a filter fence placed around the inlet.  The fabric asures
   should not be used as stand-alone sediment control measures. To increase inlet protection
   effectiveness, these practices should be used in combination with other measures, such as
   small impoundments or sediment traps (USEPA, 1992). Temporary storm drain inlet:
   protection is not intended for use in drainage areas larger than 1 acre. Generally, storm water
   inlet protection measures are practical for relatively low sediment and low volume flows.

Frequent maintenance of storm drain controls  is  necessary to prevent clogging.  If sediment and
other debris clog the water intake, drop intake control measures can actually cause erosion in
unprotected areas.                                      .                          '

Maintenance

All temporary control measures must be checked after each storm event.  To maintain the
sediment capacity of the shallow settling pools created from these techniques, accumulated
sediment should be removed from the area around the drop inlet (excavated area, around fabric
barrier, or around block structure) when the sediment storage is reduced by approximately 50
percent. Additional debris should be removed from the shallow pools on a periodic basis.

Weep holes in excavated areas around inlets can become clogged and prevent water from
draining from the shallow pools that form. Should this happen, unclogging the water intake may
be difficult and costly.                                                            i       •

Cost                                                                           •

The cost of implementing storm drain drop inlet protection measures will vary depending on the
control measure chosen. Generally, initial installation costs range from $50 to $150 per inlet,
with an average cost of $100 (USEPA, 1993). Maintenance costs can be high (annually, up to
100 percent of the initial construction cost) because of frequent inspection and repair needs. The
Southeastern Wisconsin Regional Planning Commission has.estimated that the cost of
installation of inlet protection devices ranges from $106 to $154 per inlet (SEWRPC,' 1991).

5.1.5.4.10           POLYACRYLAMIDE (PAM)

General Description

The term polyacrylamide (PAM) is a generic term that refers to a broad class  of compounds.
There are hundreds of specific PAM formulations, and all have unique properties that depend on
polymer chain length and number and kinds of functional group substitutions along the chain.

June 2002      ~~                                                               •   ~108

-------
  Development Document for Construction and Development Proposed Effluent Guidelines	

  PAMs are classified according to their molecular weight and ionic charge and are available in
  solid, granular, liquid, or emulsion forms.

  PAM's effectiveness to prevent or reduce erosion is due to its affinity for soil particles, largely
  via coulombic and Van der Waals attraction. These surface attractions enhance particle
  cohesion, stabilizing soil structure against shear-induced detachment and transport in runoff. In
  a soil application, PAM aggregates soil particles, increasing pore space and infiltration capacity,
  resulting in reduced runoff. These larger particle aggregates are less susceptible to raindrop and
  scour erosion, thus reducing the potential to mobilize sediments.

  Applicability

  Because of ease in application, PAM is well suited as a short-term erosion prevention BMP,
  especially for areas with limited access or steep slopes that hinder personnel from applying other
  cover materials. PAM can be used to augment other cover practice BMPs, though it can be
  effective when applied alone.  Thus, the ease of application, low maintenance,  and relatively low
  cost associated with PAM make it a practical solution to soil stabilization during construction.

 Application Criteria

 PAM can be applied to soil through either a dry granular powder or a liquid spray form. Optimal
 application rates to prevent erosion on construction sites are generally less than 1 kg/ha (about 1
 Ib/ac) (Tobiason et al., 2000).  However, the concentration required can vary for specific soil
 properties and construction phases. WDOT (2002) suggests a dosage of 60 mg/L for roadway
 erosion and sediment control.  This is higher than the rate recommended by the University of
 Nebraska for an agricultural application (10 parts per million).  To put this into context, one half
 pound of PAM inlOOO gallons of water results in a PAM concentration of 60 mg/L, which treats
 1 acre of exposed soil to WDOT recommendations.
                                                   e
 Effectiveness

 A study performed in Dane County, Wisconsin, analyzed 15 small plots (1 meter x 1 meter) for
 runoff and sediment yield on a construction site. The study concluded that when a solution of
 PAM-mix with mulch/seeding was applied to dry soil and compared with  the control (no PAM-
 mix application to dry soil), an average reduction of 93 percent in sediment yield was found.  An
 average reduction of 77 percent in sediment yield was the worst performing PAM treatment and
 occurred when PAM-mix in solution was applied to moist soil. The application of dry PAM-mix
 to dry soil reduced sediment by 83 percent and decreased runoff by 16 percent when compared
 to the control. The results show that regardless of the application method, PAM-mix was
 effective in reducing sediment yield in the test plots (Roa-Espinosa et al., 2000).
June 2002
                                                                                    5-109

-------
Develc
nt Document for Construction and Development Proposed Effluent Guidelines
A second study performed in Washington analyzed the runoff from three different construction
sites: an erosion control test facility, a highway construction site, and an airport runway. Table
5-24 summarizes the 225 samples analyzed in Tobiason et al. (2000).                  ;
                    Table 5-24. Turbidity Reduction Values from PAM

Maximum
Median
Minimum
Volume, m3
350
285
133
Turbidity Reduction (%)
99.97
97.6
46
Limitations

Currently PAMs are most commonly produced as dry granules. They completely dissolve and
remain dissolved if mixed properly.  If added too quickly or if not stirred vigorously the granules
rapidly form nondissolvable gels on contact with water or collect in low turbulence areas as
syrupy concentrations that dissolve slowly in an uncontrolled pattern over a period of hours or
days (USDA, 1994).                                                        -      |

In addition, when spilled on hard surfaces, PAM solutions are extremely slippery and hazardous
to foot and vehicle traffic. PAM dust is highly hygroscopic and, if inhaled, could impair
breathing.  Certain neutral and cationic PAMs at very high exposure levels produce irritation in
humans and are somewhat toxic to certain aquatic organisms; therefore, PAM should be used in
strict compliance with state and federal label requirements.

Finally, although PAM is rather inexpensive, there are considerable infrastructure needs and
operating costs; thus, sophisticated onsite polymer treatment systems may not be appropriate for
certain projects.                                                                 ;
                                              1                                  !
Cost

The cost of PAM ranges from $1.25 per pound to $5.00 per pound (Entry etal., 1999).  the cost
of PAM application depends on the system employed. PAM can be used in a centralized
treatment system (e.g., at a sedimentation basin) to treat larger areas, or dispersed in granular or
liquid form, hi Tobiason et al. (2000), the startup costs for the batch treatment system amounted
to $90,000. Monthly expenses averaged $18,000 for operations and maintenance and $13,000
for materials and equipment.  The total costs for this phase totaled about $245,000, less than 1
percent of total construction costs. If dispersed through irrigation systems (for agriculture), the
seasonal cost of PAM treatment is $9 to $15 per acre (Kay-Shoemake, et. al., 2000), where a
season probably requires between 5 and 10 applications.

For construction sites, it is more likely that PAM would be applied as an additive to the.
hydroseed mix and applied when final grade is established and cover vegetation is installed.
 June 2002
                                                                                     5-110

-------
  Development Document for Construction and Development Proposed Effluent Guidelines     	

  Based on a recent scan of the Internet, there are numerous suppliers who provide PAM as a low
  cost additive for hydroseeding, suggesting PAM application costs can be incorporated into that
  of hydroseeding ($540 to $700 per acre depending on which seed is applied). An additional cost
  would be incurred to sample site soils to customize the dosage and delivery mechanisms for
  individual sites. In addition, re-application of PAM in granular or liquid form to areas with rill
  development (poor vegetation cover) would require additional funds.  Where re-application of
  granular PAM is used, R. S. Means (2000) suggests a cost of approximately $5 per 1000 square
  feet for spreading soil admixtures by hand.
June 2002
                                                                                     5-111

-------
Development Document for Construction and Development Proposed Effluent Guidelines
5.1.6  SUMMARY

The BMP information presented in sub-section 5.1 is summarized in Tables 5-25 through 5-28.

       Table 5-25. Summary of Information on Erosion Control and Prevention BMPs
                                        (Sub-section 5.1.5.1)	
     BMP Type
                    Receiving Water Quality
                                   Physical Impact Mitigation
                                                              Other Impacts
                                     Downstream Impacts
     lanning/
    Staging/
    Scheduling
 lould be low cost.
One data set shows 42% reduction
    in sediment yield due to
    planning/staging/scheduling.
Requires additional advance
    planning and management.
 mpact could be evaluated with
    models as well as
    experimentally since several
    computer models are
    available.
 Jould be low cost.
Database is poor.
  o validated urban runoff models
    available for theoretical
    analysis of downstream
    impacts.
  otential exists to modify existing
    models to make the analysis
    of downstream impacts on
    geomorphology.
^o good cause-effect
    relationships ',
    available.
Other impacts not
    evaluated.
    Vegetative
    Stabilization
 'ould be low cost
Can be very effective in some
    cases with advance planning.
Can be important on streambanks.
Limited applicability hi the active
    construction area.
 implements other practices.   ••'
Practice is seasonably dependent
    hi most of nation.
 inpact could be evaluated with
    models as well as
    experimentally since several
    computer models are
    available.
Could be low cost.
Database is poor.
 <[o validated urban runoff models
    available for theoretical
    analysis of downstream
    impacts.
Potential exists to modify existing
    models to  make the analysis
    of downstream impacts on
    geomorphology.
 

-------
 Development Document for Construction and Development Proposed Effluent Guidelines
       Table 5-25. Summary of Information on Erosion Control and Prevention BMPs
                                    (Sub-section 5.1.5.1)
BMP Type

Seeding

>









Sodding .














Mulching
--













Physical Impact Mitigation
Receiving Water Quality
Low-cost method for establishing
vegetation.
Occurs near the end of active
construction.
Requires significant time for
establishment.
Need a prepared seedbed.
Good database on impact on soil
, erosion.
Should be supported by other
BMPs.

High-cost method of establishing
vegetation.
Immediate stabilization.
Requires significant management
attention during
establishment.
Good database on impact on soil
erosion.
Very effective way of controlling
erosion.
Works well for grass waterways
and other significant problems
area.
Should be supported by other
BMPs.
Relatively low-cost method of
providing cover.
Can be highly effective in
reducing soil loss when
properly anchored.
Good database on impact on soil
erosion.
Variety of materials can be used.
nstallation is rapid.
fot a stand-alone practice.
Due to interference with
construction operations, the
times that it can be used
during active construction are
limited.
Downstream Impacts
Should not be evaluated as stand-
alone practice, but as part of a
system.
Database is poor.
No validated urban runoff models
available for theoretical
analysis of downstream
impacts.
Some potential exists to modify
existing models to make the
analysis of downstream
impacts on geomorphology.
Should not be evaluated as stand-
alone practice, but as part of a
system.
Database is poor.
No validated urban runoff models
available for theoretical
analysis of downstream
impacts.
Some potential exists to modify
existing models to make the
analysis of downstream
impacts on geomorphology.



Database is poor.
"io validated urban runoff models
available for theoretical
analysis of downstream
impacts.
ome potential exists to. modify
existing models to make the
analysis of downstream
impacts on geomorphology.






Other Impacts

No good cause-effect
relationships available.
Other impacts not
evaluated.








sk> good cause-effect
relationships
available.
Other impacts not
evaluated.










'•Jo good cause-effect
relationships
available.
Other impacts not
evaluated.










June 2002
                                                                                    5-113

-------
Development Document for Construction and Development Proposed Effluent Guidelines   	     :

      Table 5-25. Summary of Information on Erosion Control and Prevention BMPs
                                   (Snb-section 5.1.5.1)
BMP Type

Erosion
Control
Matting
/Geotextiles














Vegetative
Buffer
Strips















Physical Impact Mitigation
Receiving Water Quality
Cost is highly variable.
Effectiveness in controlling
sediment is variable
depending on type material.
Can provide immediate protection
to exposed soils.
Not a stand-alone practice.
Due to interference with
construction operations, the
times that it can be used
during active construction are
limited.
Disposal is a significant problem
and may require landfilling.
Can be used for channel linings as
stand alone or under riprap.
Fair database on effectiveness in
preventing erosion.
Can be highly effective in trapping
sediment.
Effectiveness is well established
and considerable data
collected.
Well-validated models are
available to predict the
impacts of constructed filter
strips on sediment trapping.
Models are included in watershed
stormwater and sediment
models.
Modifications needed for natural
riparian zones.
Require routine maintenance.
May be most appropriate where
sediment loads are relatively
low.
Downstream Impacts
Database is poor.
No validated urban runoff models
available for theoretical
analysis of downstream
impacts.
Some potential exists to modify
existing models to make the
analysis of downstream
impacts on geomorphology



i





Database is poor.
No validated urban runoff models
available for theoretical
analysis of downstream
impacts.
Some potential exists to modify
existing models to make the
analysis of downstream
impacts on geomorphology









Other Impacts (

No good cause-effect
relationships
available.
Other impacts not
evaluated.
•


1









No good cause-effect
relationships
available. ;
Other impacts not
evaluated. ;









.
;


June 2002
                                                                                    5-114

-------
Development Document for Construction and Development Proposed Effluent Guidelines
      Table 5-25. Summary of Information on Erosion Control and Prevention BMPs
                                   (Sub-section 5.1.5.1)
BMP Type

Top soiling
Physical Impact Mitigation
Receiving Water Quality
Important in vegetative
establishment.
No protection until cover is
established.
Not a stand-alone practice, but
must be supported by other
BMPs.
No known information to describe
effectiveness and cost not
currently available.
Downstream Impacts
Database is poor.
No validated urban runoff models
available for theoretical
analysis of downstream
impacts.
Some potential exists to modify
existing models to make the
analysis of downstream
impacts on geomorphology
Other Impacts
No good cause-effect
relationships
available.
Other impacts hot
evaluated.
June 2002
5-115

-------
 Development Document for Construction and Development Proposed Effluent Guidelines
        Table 5-26. Summary of Information on Erosion Control and Prevention BMPs
                  	(Sub-section 5.1.5.2)
   BMP Type
                                      Physical Impact Mitigation
   Receiving Water Quality
        Downstream Impacts
      Other Impacts
 Earth Dike
 Used to protect down slope areas.
 Should be stabilized prior to use.
 Requires maintenance after every
 major storm.
 Can be significant source of
 sediment if not properly
 constructed.
 Little data available on its
 effectiveness as a BMP.
 Can be relatively inexpensive,
 depending on design.
 Not a stand-alone procedure.
No known information available.
                                                                                      No known information
                                                                                      available.
 Temporary
 Swale
Effectively a grass-lined drainage
ditch with shallow side slopes.
Can be applied in many areas, but
use limited in arid areas.
Contaminants that will harm
vegetation, such as oils and
greases, cannot be discharged to
the system.
Continuous water flow cannot be
tolerated by the grass lining.
Effectiveness depends on
infiltration. Can be a problem of
groundwater pollution with high
water tables.
Some studies show that they export
bacteria.
Some studies show high removal
efficiency for TSS, fair for
nutrients, are variable for metals.
No general relationships available
to predict the impact under widely
varied climates and conditions,
hence the effectiveness cannot be
predicted for a given situation
beyond the limited database.
No  known information available.
No known information
available.       ;
June 2002
                                                                                                        5-116

-------
   Development Document for Construction and Development Proposed Effluent Guidelif
          Table 5-26. Summary of Information on Erosion Control and Prevention BMPs
                    	(Sub-section 5.1.5.2)
     BMP Type
     Receiving Water Quality
                                          'hysical Impact Mitigation
                                                             Downstream Impacts
     mporary
   Storm Drain
   Diversion
   (Pipe)
   Reroutes existing drainage systems.
   Primary benefit is to separate
   drainage water originating from
   undisturbed and construction and
   reduce the volume of water to be
   treated.
   Can be combined with other
   structures, such as sediment traps,
   and used for sediment trapping.
 . Require little maintenance.
  Requires outlet stabilization.  Can
  be a significant source of sediment
  without outlet stabilization.
  Can be costly, depending on size,
  installation, and removal.
 No  known information available.
                                    No known information
                                    available.
    ipe Slop<
  Drain
  Routes runoff from concentrated
  flow to stabilized areas.
  Can be very effective in
  eliminating gully erosion problems,
  if properly installed and
  maintained.
  Can be constructed from low-cost
  corrugated PVC, but must be
  anchored or buried along slope.
  Needs to be checked frequently for
  sedimentation and other
  maintenance problems.
No known information available.
                                    No known information
                                    available.
  Dams
 Reduces velocity of flow and
 prevents erosion.
 Stabilizes channel slope on steep
 sections by stairstepping.
 Can trap small percentages of
 sediment behind dam.
 Used for short periods of time
 where channel lining is impractical.
 Limited lab studies show high
 effectiveness, but very limited field
 studies show low trapping
 efficiency.
 Must be installed such that
 overtopping occurs over the rock
 fill and not around the perimeter.
 Should not be used in continuously
 flowing streams.
 Relatively expensive, if properly
 installed.
 Procedures  for predicting impact of
properly installed stone check dams
are available and incorporated into
watershed computer models.	
                                                    No known information available.
                                   No known information
                                   available.
June 2002
                                                                                                         5-117

-------
        :Dc
: for Construction and Development Proposed Effluent Guidelines
    Table 5-26. Summary of Information on Erosion Control and Prevention BMPs

Lined
Waterways

• Designed for stability and capacity.
• Local rainfall-runoff conditions
and linings will influence channel
dimensions.
• Require some maintenance during
vegetative establishment.
• Not designed as sediment removal
device, but to prevent channel


Other Impacts
available.
June 2002
                                                                                   5-118

-------
  Development Document for Construction and Development Proposed Effluent Guidelines
         Table 5-27. Summary of Information on Erosion Control and Prevention BMPs
                   	      (Sub-section 5.1.5.3)
      BMP Typ
                      Receiving Water Quality
                       Physical Impact Mitigation
                                            Downstream Impacts
                                                                                        Other Impacts
     Silt Fence,
Most widely recognized sediment
     control BMP.
Frequently poorly installed with little
     design consideration.
Maintenance is frequently poor,
     resulting in frequent failure.
     Frequent maintenance is required
     for proper operation.
^aboratory studies show fair to good
     sediment trapping by filter fence,
     but limited field studies do not
     show the same results.
Evaluations of installations show that
     failure is frequent,' coming from
     undercutting of the fabric and
     subsequent gully erosion.
ihould not be installed where rocks anc
     other hard surfaces prevent
     anchoring.
'•Jo validated procedures are available
    to predict the effectiveness of the
    filter fence in trapping sediment,
    primarily because of the lack of
    validated relationships for
    predicting flow through the filter
    fence.
'rocedures for evaluating the
    anchoring requirements and
    support post requirements have not
    adequately accounted for variable
    soil strength conditions, resulting
    in frequent failure of the fence
    under loading.
 Database is poor.
 No validated urban runoff models
     available for theoretical analysis
     of downstream impacts.
 Some potential exists to modify
     existing models to make the
     analysis of downstream impacts on
     geomorphology
No good cause-effect
     relationships
     available.
Other impacts not
     evaluated.
      uper Silt
     ence
Modification of standard silt-fence to
    improve it structurally.
"Jo validation information is available.
{.ecommended to be used where
    destruction of the silt fence will
    destroy critical areas.
dore expensive than standard silt
    fence.	
database is poor.
^o validated urban runoff models
    available for theoretical analysis
    of downstream impacts.
 ome potential exists to modify
    existing models to make the
    analysis of downstream impacts on
    geomorphology.	
sfo good cause-effect
    relationships
    available.
)ther impacts not
    evaluated.
June 2002
                                                                                                           5-119

-------
Development Document for Construction and Development Proposed Effluent Guidelines	     :

      Table 5-27. Summary of Information on Erosion Control and Prevention BMPs
                                   (Sub-section 5.1.5.3)
BMP Type

Straw Bale
Dike










Sediment
Traps

















Physical Impact Mitigation
Receiving Water Quality
Works by impounding water.
Primary trapping mechanism is by
settling behind straw bale dike.
Information on performance is very
limited with much variation in the
limited data.
Should not be used in waterways or as a
perimeter control due to
biodegradation.
Idealized models of performance are
available for systems that are
properly installed.
Formed by excavation and/or
embankment.
Can simplify stprmwater control by
trapping sediment at specific
spots.
Can be installed quickly and serve as
short-term solution to sediment
trapping in small areas.
May require cleahout.
Detailed models as well as simplified
design aids are available to predict
performance in trapping sediment.
Data on performance are available from
both laboratory studies and field
studies.
Will likely control only the settleable
solids unless enhanced settling is
developed with chemical

Downstream Impacts
Database is poor.
No validated urban runoff models
available for theoretical analysis
of downstream impacts.
Some potential exists to modify
existing models to make the
analysis of downstream impacts on
geomorphology




Database is poor.
No validated urban runoff models
available for theoretical analysis
of downstream impacts.
Some potential exists to modify
existing models to make the
analysis of downstream impacts on
geomorphology.











Other Impacts ,

No good cause-effect
relationships ;
available.
Other impacts not
evaluated. ;
f



-.


Data, for trapping
nutrients are :
available, but show
wide variation.;
General models of ;
nutrient trapping are
not available.
Other impacts not
evaluated. . ;
1









June 2002
                                                                                     5-120

-------
 Development Document for Construction and Development Proposed Effluent Guidelines
       Table 5-27. Summary of Information on Erosion Control and Prevention BMPs
                                    (Sub-section 5.1.5.3)
BMP Type
Sediment
Basins





















Physical Impact Mitigation
Receiving Water Quality
Normally formed by construction of a
dam.
Stormwater detention basin may serve
as sediment basin during
construction.
Can be used for any size watershed.
May require cleanout.
Data on performance are available both
from laboratory studies and field
studies.
Will likely control only the settleable
solids unless enhanced settling is
developed with chemical
flocculation.
Most reliable and stable structure for
obtaining high sediment trapping
efficiency under widely varying '
conditions.
Must consider dam safety issues since
dam failure is a reasonable
possibility.
Structures are relatively large and can
be expensive.
Downstream Impacts
Database is poor.
No validated urban runoff models
available for theoretical analysis
of downstream impacts.
Some potential exists to modify
existing models to make the
analysis of downstream impacts on
geomorphology.















Other Impacts
Data for trapping
nutrients are
available, but show
wide variation.
General models of
nutrient trapping are
not available.
Other impacts not
evaluated.

,












June 2002
                                                                                    5-121

-------
Development Document for Construction and Development Proposed Effluent Guidelines
      Table 5-28. Summary of Information on Erosion Control and Prevention BMPs
                                   (Sub-section 5.1.5.4)
BMP Type
Stone Outlet
Structures
Rock Outlet
Protection
Sump Pit
Storm Drain
Met
Protection
Sediment Tank
Physical Impact Mitigation
Receiving Water Quality
Porous outlet structure constructed of
dumped rock, used as the outlet
for earth dikes.
Requires a stabilized outlet channel
until the flow reaches a stable
channel.
Data on the effectiveness are limited to
visual observations of field installations
where failure was frequent due to poor
installation.
Models are available to predict the
performance of stone outlets, but field
data have not been collected to evaluate
the accuracy of the model.
Used to reduce velocity of flow in
receiving channel and prevent
scouring.
Very effective when properly installed.
Design procedures are well established.
Maintenance is low, if properly
installed.
Should be inspected after high flows.
No data on impact.
Used to dewater during excavation.
Effectiveness not evaluated.
Potential exists to theoretically evaluate
the BMP's effectiveness in
trapping sediment.
Could be used at times other than storm
flow, such as removal of
groundwater flow.
Used to trap sediment that would
otherwise flow into storm drain
inlet.
Should be installed prior to land
disturbance.
Effectiveness in removing sediment has
not been evaluated, but is thought
to be low during construction.
Potential exists to use computer models
to evaluate effectiveness.
Cost can be high for maintenance
requirements.
Should not be used as stand-alone
sediment control.
Portable sediment trap.
Flows are pumped in and out of the
tank.
Used where spaced is limited No
effectiveness data are available.
Expected to be relativelv expensive.
Downstream Impacts
No validated urban runoff models
available for theoretical analysis
of downstream impacts.
Some potential exists to modify
existing models to make the
analysis of downstream impacts on
geomorphology.
No data available.
Database is poor.
No validated urban runoff models
available for theoretical analysis
of downstream impacts.
Some potential exists to modify
existing models to make the
analysis of downstream impacts on
geomorphology.
Database is poor.
No validated urban runoff models
available for theoretical analysis
of downstream impacts.
Some potential exists to modify
existing models to make the
analysis of downstream impacts on
geomorphology.
No data available.
Other Impacts
General models of
nutrient trapping are
not available.
Other impacts not
evaluated.
i
No data available.
No data available.:
No data available.
!
No data available.;
June 2002
5-122

-------
Development Document for Construction and Development Proposed Effluent Guidelines
      Table 5-28. Summary of Information on Erosion Control and Prevention BMPs
                                   (Sub-section 5.1.5.4)
BMP Type
Stabilized
Construction
Entrance

Temp Access
Waterways
Crossing
Dust Control
Physical Impact Mitigation
Receiving Water Quality
Used to minimize mud and sediment
attached to motorized vehicles.
Consists of an area that is covered with
rocks over which all vehicles must
drive.
Can be combined with a wash station.
Effective only if all entrances are
maintained.
Relatively expensive.
Will not remove highly cohesive clays.
Stabilizes slopes and decreases runoff
velocity.
Can be incorporated into low-impact
development plans.
Not effective when drainage patterns
are altered.
Mot effective when vegetative areas on
perimeter are destroyed.
Practiced at virtually all construction
sites.
No data available on BMP
effectiveness.
Reduces risk to damaging streambed
from construction equipment
tracking.
Can be a bridge, culvert, or ford.
Bridges and culverts preferred, but
more expensive.
Data on effectiveness in reducing
sediment are not available.
•mportant in arid and semi-arid regions.
Applicable to any construction site.
Construction and sequencing and
limiting exposure area can reduce
problems.
Spray-on adhesives are recommended.
Water application may require near
. constant attention.
ixcess water may cause runoff or
tracking of mud.
Very limited effectiveness information
available.
Costs can vary widely, depending on
local conditions.
Downstream Impacts
No data available.
No data available.
No data available.
No data available.
Other Impacts
No data available.
No data available.
Mo data available.
No data available.
June 2002
5-123

-------
Development Document for Construction and Development Proposed Effluent Guidelines             '•	

5.2    REFERENCES                                                          !

Adams, J, J. Janatzen, D. Loudenslager.  2000. A Device to Alleviate Pollution from Urban
       Stormwater.  Design Report, Biosystems and Agricultural Engineering, Oklahoma State
       University, Stillwater, OK.                                                j

ASCE. 1999, National Stormwater Best Management Practices (BMP) Database, prepared by the
       Urban Water Resources Research Council of the American Society of Civil Engineers,
       prepared for the US EPA, Office  of Science and Technology, Washington, DC.  !
       http://www.bmpdatabase.org                                              ;

Barfield, B. J.  2000.  Presentation on Nonpoint Source Sediment and Stormwater. University of
       Puerto Rico, San Juan, PR.

Barfield, B. J., Bamhisel, R. I., Powell, J. C., Hirschi, M. C., and Moore, I. D. 1983.
       Erodibilities and Eroded Size Distribution of Western Kentucky Mine Spoil and
       Reconstructed Topsail Final Report for Title HI. Institute for Mining and Minerals
       Research, University of Kentucky, Lexington, KY.                            ;

Barfield, B J., and Clar, M.  1985. Development of New Design Criteria for Sediment Traps and
       Basins. Prepared for the Maryland Resource Administration. Annapolis, MD, pp. 33.

Barfield, B. J., Haan, C. T>, Stevens, E., and Hayes, J. C., K. F. Holbrook. 2001.  "Engineering
       Design Aids for Sediment Control Practices," Proceedings of the International
       Symposium on Soil Erosion Research for the 21st Century. January 3-5, 2001. Honolulu,
       Hawaii. Symposium Sponsored and Published by American Society of Agricultural
       Engineers. St. Joseph, MI.                                .                 ^

Barfield, B. J. and Hayes, J. C. 1992. Unpublished Results of Field Evaluation of Sediment
       Controls in South Carolina Construction Operations. Biosystems and Agricultural
       Engineering, Oklahoma State University, Stillwater, OK.

Barfield, B. J. and Hayes, J. C. 1999. Unpublished Results of Field Evaluation of Sediment
       Controls in Louisville, KY Construction Operations. Biosystems and
       Agricultural Engineering, Oklahoma State University, Stillwater, OK           I
                                                                               I
Barfield, B. J., Hayes, J. C., Fogle, A. W., and Kranzler, K. A. 1996. "The SEDIMOT III - Model
       of Watershed Hydrology and Sedimentology,"  Proceedings of Fifth Federal Interagency
       Sedimentation Conference, Las Vegas, NV. 1996.                            ;
June 2000
                                                                                  5-124

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	"

 Barfield, B. J., Moore, I. D., and Williams, R. G.  1979.  "Sediment Yield in Surface Mined
        Watersheds," Proceedings Symposium on Surface Mine Hydrology, Sedimentology, and
        Reclamation.  College of Engineering, University of Kentucky, Lexington, KY.

 Barrett, M. E., Kearney, J. E., McCoy, T. G., Malina, J. F.,Charbeneau, R. J. and Ward, G. H.
        1995.  An Evaluation of 'the Use and Effectiveness of Temporary Sediment Controls.
        Technical Report CRWR 261, Center for Research in Water Resources, The University of
        Texas at Austin, Austin, TX 78712.

 Benik, S. R., B. N. Wilson, D. D. Biesboer, B. Hanse, and D. Stenlund. 1998. The Efficacy of
       Erosion Control Products at a MN/DOT Construction Site. Paper No. 982156, American
    -_   Society of Agricultural Engineers, St. Joseph, MI

 Blench, T.  1970. Regime Theory Design of Canals with Sand Beds.  Proceedings of the
       American Society of Civil Engineers.  96(IR2): 205-213.

 Britton, S. L., Robinson, K. M., and Barfield, B. J.  2001. Modeling the Effectiveness of Silt
       Fence.  Proceedings 7th Federal Interagency Sedimentation Conference, Reno, NV.

 Brown, W. and D. Caraco.  1997.  "Muddy Water In - Muddy Water Out?". Article No. 52  hi
       The Practice of Watershed Protection. Center for Watershed Protection, Ellicott City,
       MD. 2000. http://www.stormwatercenter.net

 Brown, W. and T. Schueler.  1997.  The Economics ofStormwater BMPs in the Mid-Atlantic
       Region. Prepared for: Chesapeake Research Consortium. Edgewater, MD. Center for
     "  Watershed Protection, Ellicott City, MD.                                ,         -

 Brown, W., D. Caraco, R. Claytor, P. Hinkle, H.Y. Kwon, T. Schueler, C. Swann, and J.
       Zielinski.  1998. Better Site Design: A Handbook for Changing Development Rules in
       Your Community.  Prepared by Center for Watershed Protection, Ellicott City, MD, for
       the Site Planning Roundtable.

 Caraco, D. S.  1997. "Keeping Soil in Its Place." Article No.  55 in The Practice of Watershed
       Protection. Center for Watershed Protection, Ellicott City, MD. 2000.
       http://www.stormwatercenter.net

 CWP.  1996. Design of Stormwater Filtering Systems. Prepared for Chesapeake Research
       Consortium, Solomons, MD, and U.S. EPA Region 5, Chicago, EL, by Center for
       Watershed Protection, Ellicott City, MD.

Chang, H. H. 1988. Fluvial Processes in River Engineering. Wiley Interscience, New York,
       432 p.
June 2000
                                                                                   5-125

-------
Development Document for Construction and Development Proposed Effluent Guidelines	[	._


Claytor, R. 1997. "Practical Tips for Construction Site Phasing."  Article No. 54 in The Practice
       of Watershed Protection. Center for Watershed Protection, Ellicott City, MD. 2000.
       http://www.stormwatercenter.net

Corish, K. 1995.  Clearing and Grading Strategies for Urban Watersheds. Environmental Land
       Planning Series. Publ. No. 95704.  Metropolitan Washington Council of Governments,
       Washington, DC. http://www.mwcog.org.                                   i

Coyle, JJ. (USD A). 1979. "Grassed Waterways and Outlets". Engineering Field Manual for
       Conservation Practices. U.S. Department of Agriculture, Soil Conservation Service
       (SCS).Ch7.p.7-7.

Dorman, M.E., J. Hartigan, R.F. Steg and T. Quasebarth.  1989.  Retention, Detention and
       Overland Flow for Pollutant Removal from Highway Stormwater Runoff Vol. 1.
       Research Report. Federal Highway Administration. FHWA/RD 89/202.  Also in:
       "Performance of Grassed Swales along East Coast Highways." Article No.  114 in The
       Practice of Watershed Protection. Center for Watershed Protection, Ellicott City, MD.
       2000. http://www.stormwatercenter.net                                     i

Ellis, B.  1998.  A Further Guide to Swales: Costs of Operation and Maintenance.  PERMEATE.
       Scottish Environmental  Protection Agency, Stirling, Scotland.
       http://www.sepa.org.uk/guidance/scontrol/swales.htm

Entry, J.A., and R.E. Sojka.  1999. Polyacrylamide Application to Soil Reduces the Movement
       of Microorganisms in Water.  In 1999 Proceedings of the International Irrigation Show,
       Irrigation Association, Orlando, FL, November 9,1999, pp. 93-99.

Ettinger, C. E. and J. E. Lichty.  1979. Evaluation of Performance Capability of Surface Mine
       Sediment Ponds. Final Report on Contract Number 68-03-2677, Industrial
       Environmental Research Laboratory.

Faircloth, W. 1999. "Searching for a Practical, Efficient, Economical Sediment Basin." In
       Proceeding of the 30th Conference of the International Erosion Control Association,
       Nashville, TN, Feb. 22-26, pp. 272-282.                                     !

Fifield, J. 2000. Designing for Effective Sediment and Erosion Control on Construction Sites.
       Forester Press, Santa Barbara, CA.

Fisher, L. S. and Jarrett, A. R.  1984. "Sediment Retention Efficiency of Synthetic FilterFabric".
       Transactions of the American Society of Agricultural Engineers. 30(l):82-89.
June 2000
                                                                                   5-126

-------
  Development Document for Construction and Development Proposed Effluent Guidelines

  Foster, G. R., Young, R. A., and Neibling, W. H. 1985. "Sediment Composition for Nonpoint
        Source Pollution Analyses." Transaction of the American Society of Agricultural
        Engineers. 28(1): 133-146.

  Goldberg.  1993. Dayton Avenue Swale Biofiltration Study. Seattle Engineering Department
        Seattle, WA.                                                        F       '

  Griffin, M. L., Barfield, B. J., and Warner, R. C. 1985.  "Laboratory Studies of Dead Storage in
        Sediment Ponds". Transactions of the American Society of Agricultural Engineers
        28(3):799-804.                                           -

 Haan, C. T., Barfield, B. J., and Hayes, J. C. 1994. Design Hydrology and Sedimentologyfor
        Small Catchments. Academic Press, San Diego, CA.

 Harding, M.V. 1990. Erosion Control Effectiveness: Comparative Studies of Alternative
        Mulching Techniques. Environmental Restoration, pp. 149-156, as cited in USEPA.
        1993. Guidance Specifying Management Measures for Sources of Nonpoint Pollution in
        Coastal Waters.  EPA 840-B-92-002. U.S. Environmental Protection Agency, Office of
        Water, Washington, DC.

 Harper, H. 1988. Effects of Stormwater Management Systems on Groundwater Quality. Final
        Report. Environmental Research and Design, Inc. Prepared for Florida Department of
        Environmental Regulation. Also in: "Runoff and groundwater dynamics of two swales in
        Florida. Article No. 113 in The Practice of Watershed Protection. Center for Watershed
        Protection, Ellicott City, MD.  2000.  http://www.stormwatercenter.net

 Hayes,  J. C., B. J. Barfield and R. I. Barnhisel.  1984.  Performance of Grass Filters under
        Laboratory and Field Conditions. Transactions of the American Society of Agricultural
        Engineers 27(5):  1321-1331.

 Hayes, J. C. and Barfield, B. J.  1995.  "Engineering Aids and Design Guidelines for Control of
        Sediment in South Carolina". In South Carolina Stormwater and Sediment Control
       Handbook for Land Disturbance Activities. South Carolina Department of Health and
       Environmental Control.  Columbia, SO.

 Hayes, J. C., Akridge, A. L., Barfield, B. J. and Holbrook, K. F.  2001.  "Simplifying Design of
       Sediment Controls in Jefferson County, Kentucky". In Proceedings of Soil Erosion
       Research for the 21st Century: An International Symposium and Exhibition, Honolulu,
       HI, January 2001. Sponsored by the American Society of Agricultural Engineers.
June 2000
                                                                                  5-127

-------
Development Document for Construction and Development Proposed Effluent Guidelines	;

Herrera, N. M. and G. K. Felton. 1991. "Hydraulics of Flow through a Rockfill Dam Usuig
       Sediment-Free Water." Transactions of the American Society of Agricultural Engineers.
       34(3):871-875.                                                           j

Heyer, T. No Date.  Vegetative Measures for StreambankStabilization. U.S. Department of
       Agriculture, Forest Service, St. Paul, MN.                                  |

Hirschi, M. C. 1981. "Efficiency of Small Sediment Controls." Unpublished Agricultural
       Engineering File Report, Agricultural Engineering, University of Kentucky, Lexington,
       KY.  (Details of the report are given in Chapter 9 of Haan et al, 1994).

Holbrook, K. F.5 Hayes, J. C. and Barfield, B. J.,  and A.W. Fogle. 1998. "Engineering Aids and
       Design Guidelines for Control of Sediment in South Carolina," In Proceedings of the
       International Erosion Control Association Conference. February 1998.          '

IDNR. 1992. Indiana Handbook for Erosion Controls in Developing Areas, Guidelines for
       Protecting Water Quality through the Control of Soil and Erosion and Sedimentation on
       Construction Sites.  Division of Soil Conservation, Indiana Department of Natural
       Resources, Indianapolis, EN.                                              !

Jarrett, A. 1999. "Designing Sedimentation Basins for Better Sediment Capture". In Proceeding
       of the 30th Conference of the International Erosion Control Association, Nashville, TN,
       Feb. 22-26, pp. 218-233.                                                  :

Kay-Shoemake, J, M. Watwood, R. Sojka, and R. Lentz. 2000. "Soil Amidase Activity in
       Polyacrylamide (PAM) Treated Soils and Potential Activity toward Common Amide
       Containing Agrochemicals". Biology and Fertility of Soils 31(2): 183-186.      >

Kercher, W.C., J.C. Landon and R. Massarelli. 1983. "Grassy Swales Prove Cost-Effective for
       Water Pollution Control. Public Works 16: 53-55.

Knisel, W. G., ed. 1980. CREAMS: A Field-Scale Model for Chemicals, Runoff, and Erosion
       from Agricultural Management Systems.  U. S. Department of Agriculture, Conservation
       Research Report No. 26, 640pp.                                   .        '    •

Koon, J. 1995. Evaluation of Water Quality Ponds and Swales in the Issaquah/East Lake
       Sammamish Basins. King County Surface Water Management and Washington
       Department of Ecology, Seattle,  WA.                                      |

Kouwen, N. 1990. Silt Fences to Control Sediment on Construction Sites.  Technical
       Publication MAT-90-03, Research and Development Branch, Ontario Ministry of
       Transportation, 1201 Wilson Avenue, Downsview, Ontario, Canada M3M 1J8. ;
June 2000
                                                                                  5-128

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

 Lane, E. W.  1955.  "The Importance of Fluvial Geomorphology in Hydraulic Engineering".
       Proceedings oj'the American Society of'Civil Engineers, 81:  1-17.

 Leopold, Luna B., M. Gordon Wolman, and John P. Miller, 1964.  Fluvial Processes in
       Geomorphology.  Dover Publications, Inc. Mineola, NY.

 Lindley, M., B. J. Barfield, J. Ascough, B. N. Wilson, and E. Stevens. 1998.  "The Surface
       Impoundment Element for WEPP".  Transactions of the American Society of Agricultural
       Engineers. 41(3):555-564.

 Masters, A., K. A. Flahive, S. Mostaghimi, D. Vaughan, A. Mendez, M. Peterie, S. Radke.  2000:
       A Comparative Investigation of the Effectiveness ofPolyacrylamide (PAM)for Erosion
       Control in Urban Areas. Paper No.  002176, American Society of Agricultural Engineers,
       St. Joseph, MI, pp. 21.

 McBurnie, J. C., Barfield, B. J., Clar, M. L.  and Shaver, E.  1990. "Maryland Sediment
       Detention Pond Design Criteria and Performance". Applied Engineering in Agriculture.
       6(2):167-173.   .                                      "

 MDE. 1994. Maryland Standards and Specifications for Soil Erosion and Sediment Control.
       Water Management Administration,  Maryland Department of the Environment, Soil
       Conservation Service, and State Soil Conservation Committee, Baltimore, MD.

 NAHB. No Date. NAHB Research Center Storm Water Runoff & Nonpoint Source Pollution
       Control Guide for Builders and Developers. National Association of Home Builders,
       Washington, DC.

 NCDNR. 1988. North Carolina Erosion and Sediment Control Planning and Design Manual.
       North Carolina Sedimentation Control Commission, the North Carolina Department of
       Natural Resources and Community Development, and the North Carolina Agricultural
       Extension Service, Raleigh, NC.

 Oakland, P.H. 1983. "An Evaluation of Stormwater Pollutant Removal through Grassed Swale
       Treatment." In Proceedings of the International Symposium of Urban Hydrology,
       Hydraulics and Sediment Control, H. J. Sterling, Lexington, KY.

 Occoquan Watershed Monitoring Laboratory. 1983; Final Report: Metropolitan Washington
       Urban Runoff Project. Prepared for the Metropolitan Washington Council of
       Governments. Manassas, VA.
June 2000
                                                                                  5-129 '

-------
Development Document for Construction and Development Proposed Effluent Guidelines	        \     	

Pitt, R. and J. McLean.  1986.  Toronto Area Watershed Management Strategy Study: Number
       River Pilot Watershed Project. Ontario Ministry of Environment and Energy, Toronto,
       Ontario, Canada.

Roa-Espinosa, A., G.D. Bubenzer, and E.S.Miyashita. 2000. Sediment and Runoff Control on
       Construction Sites Using Four Application Methods of Polyacrylamide Mix. National
       Conference on Tools for Urban Water Resource Management and Protection, Chicago,
       IL, February 7-10,2000, pp. 278-.                                         ;

Renard, K. G., Foster G. R., Weesies, G. A., McCool, KD. K. and Yoder D. C. 1994. Predicting
       Soil Erosion by Water—A Guide to Conservation Planning with the Revised Universal
       Soil Loss Equation (RUSLE). Agricultural Research Service Publication, U.S.
       Department of Agriculture.

Rosgen, D. L., 1996. Applied River Morphology. Wildland Hydrology, Pagosa Springs, CO.

Seattle Metro and Washington Department of Ecology. 1992. Biqfiltration Swale Performance:
       Recommendations and Design Considerations.  Publication No. 657.  Water Pollution
       Control Department, Seattle, WA. Also in:  Watershed Protection Techniques.  Fall 1994
       1(3):  117-119.                                                           ;

SCDHEC. 1995. South Carolina Stormwater Management and Sediment Reductions
       Regulations. In South Carolina Stormwater Management and Sediment Control
       Sedimentology Resource, South Carolina Department of Health and Environmental
       Control. Columbia, SC.

Schumm, S. A.  1977.  The Fluvial System.  John Wiley and Sons, New York. 338pp.

Simons, D.B., and M.L. Albertson,  1960. "Uniform Water Conveyance Channels in Alluvial
       Material." Proceedings of the American Society of Civil Engineers. 86(HY5):33-71.

Smiley, J., and C. T. Haan. 1976. "The Dam Problem of Urban Hydrology". Proceedings of the
       National Symposium on Urban Hydrology,  Hydraulics, and Sediment Control. Cpllege
       of Engineering, University of Kentucky, Lexington, KY.

Smolen, M.D., D.W. Miller, L;C. Wyall, J. Lichthardt, and A.L. Lanier. 1988. Erosion and  •
       Sediment Control Planning and Design Manual. North Carolina Sedimentation Control
       Commission and North Carolina Department of Natural Resources and Community
       Development,  Raleigh, NC.
June 2000
                                                                                 5-130

-------
 Development Document for Construction and Development Proposed Effluent Guidelines    	

 Snodgrass, W.J., B.W. Kilgour, L. Leon, N. Eyles, J. Parish, and D.R. Barton.  1998.  "Applying
        Ecological Criteria for Stream Biota and an Impact Flow Model for Evaluation of
        Sustainable Urban Water Resources in Southern Ontario".  In:  Sustaining Urban Water
        Resources in the 21st Century.  Proceedings for an Engineering Foundation Conference.
        Edited by A.C. Rowney, P. Stahre, and L.A. Roesner. Malmo, Sweden. September 7 -
      .  12, 1997.

 Stahre, P., and B. Urbonas. 1990.  Stormwater Detention For Drainage, Water Quality, and
        CSO Management. Prentice Hall, Englewood Cliffs, NJ.

 SWRPC. 1991. Costs of Urban Nonpoint Source Water Pollution Control Measures. Technical
        report no. 31.  Southeastern Wisconsin Regional Planning Commission, Waukesha, WI.

 Tobiason, S., D. Jenkins, E. Molash, and S. Rush. 2000. "Polymer Use and Testing for Erosion
        and Sediment Control on Construction Sites: Recent experience in the Pacific
        Northwest". In: Proceedings of Conference 31. International Erosion Control
        Association. Palm Spring,  CA, February 21-25, 2000, pp. 41-52.

 Tapp, J. S., and Barfield, B. J. 1986. "Modeling the Flocculation Process in Sediment Ponds".
        Transactions American Society of 'Agricultural Engineers. 29(3):''41-747.

 Tapp, J. S., Barfield, B. J., and Griffin, M. L. 1981. Predicting Suspended Solids Removal in
       Pilot Scale Sediment Ponds Utilizing Chemical Flocculation.  Research Report IMMR
        81/063, Institute for Mining and Minerals Research, University of Kentucky, Lexington,
       KY.

 UNEP.  1994. Guidelines for Sediment Control Practices in the Insular Caribbean. CEP
       Technical Report No. 32. UNEP Caribbean Environment Programme, Kingston,
       Jamaica, http://www.cep.unep.org/pubs/techreports/tr32en/content.html

 USAF.  1998.  USAF Landscape Design: Erosion Control Measures. U.S. Air Force, St Paul
  '   .  MN.
       http://www.afcee.brooks.af.mil/ldg/sl8erosioncontrol/cQ3process.html

 USDOT. 1995. Best Management Practices for Erosion and Sediment Control. Report No.
       FHWA-FLP-94-005. Eastern Federal Lands Highway Design, U.S Department of
       Transportation, Sterling, VA.

USDA andNRCS, 1985. National Engineering Handbook. Natural Resources Conservation
       Service, Soil Conservation Services.
June 2000
                                                                                  5-131

-------
Development Document for Construction and Development Proposed Effluent Guidelines            •	

USEPA. 1,971. Control of Sediment Resulting from Construction of Highways and Land
      Development. U.S. Environmental Protection Agency, Washington, DC.

USEPA.  1992. Storm Water Management for Construction Activities: Developing Pollution
      Prevention Plans and Best Management Practices. EPA 832-R-92-005.  U.S.    I
      Environmental Protection Agency, Office of Water, Washington, DC.          '

USEPA.  1992. Storm Water Management for Industrial Activities: Developing Pollution
      Prevention Plans and Best Management Practices. EPA 832-R-92-006. U.S.
      Environmental Protection Agency, Office of Water, Washington, DC.

USEPA.  1993. Guidance Specifying Management Measures for Sources ofNonpoint Pollution
      in Coastal Waters. EPA 840-B-927002. U.S. Environmental Protection Agency, Office
      of Water, Washington, DC.                                               >

USEPA.  2000. Urban Nonpoint Source Management Measure Guidance -Draft. U.S.  j
      Environmental Protection Agency, Office of Water, Washington, DC.          ;

Vanderwel, D. and S.  Abday. 1998. An Introduction to Water Erosion Control. Alberta,
      Agriculture, Food and Rural Development, Alberta, Canada.                  ;

VDCR.  1995. Virginia Erosion & Sediment Control Field Manual. Second Edition. Virginia
      Department of Conservation and Recreation, Division of Soil and Water Conservation,
      Richmond, VA.

VDCR.  2001. Virginia Erosion and Sediment Control Law, Regulations, and Certification
      Regulations. Virginia Department of Conservation and Recreation, Division of Soil and
      Water Conservation, Richmond, VA.                                      |

Wang, T., D. Spyridakis, B. Mar and R. Horner. 1981.  Transport, Deposition and Control of
      Heavy Metals  in Highway Runoff. FHWA-WA-RD-39-10. Department of Civil
      Engineering. University of Washington, Seattle, WA.                         [

Ward, A. D., Haan, C. T., and Barfield, B. J. 1977. Simulation of the Sedimentology of Sediment
      Detention Basins. Research Report No. 103, Water Resources Research Institute,
      University of Kentucky, Lexington, KY.

Ward, A. D., Haan, C. T., and Tapp, J. S. 1979. The DEPOSITS Sedimentation Pond Design
      Manual. OISTL, Institute for Mining and Minerals Research, University of Kentucky,
      Lexington, KY.                                                         I
June 2000
                                                                                 5-132

-------
' Development Document for Construction and Development Proposed Effluent Guidelines     	

 Washington State Department of Ecology (WDEC).  1992.  Stormwater Management Manual for
        the Puget Sound Basin. Technical Manual. Washington State Department of Ecology,
        Olympia, WA.

 Welborn, C., and J. Veenhuis.  1987. Effects of Runoff Controls on the Quantity and Quality of
        Urban Runoff in Two Locations in Austin, TX. U.S. Geological Survey Water Resources
        Investigations Report. 87-4004, pp.  88.

 Williams, J. R. No Date. "Sediment Yield Prediction with Universal Equation Using Runoff
        Energy Factor". Present and Prospective Technology for Predicting Sediment Yields and
        Sources-. Publication ARS-S-40, Agricultural Research Service, U.S. Department of
        Agriculture, Washington, DC.

 Wilson, B. N., and Barfield, B. J.  1984.  "A Sediment Detention Pond Model Using CSTRS
        Mixing Theory." Transactions American Society of Agricultural Engineers.  27(5);1339-
   -   .  1344.

 Wilson, B. N., Barfield, B. J., and Moore, I. D.  1982. A Hydrology and Sedimentology
        Watershed Model. Department of Agricultural Engineering, University of Kentucky,
       Lexington, KY.

 Wischmeier, W. H. and D. D. Smith. 1965. Predicting Rainfall-Erosion Losses From Cropland
       East of the Rocky Mountains-Guide for Selection of Practices for Soil and Water
       Conservation. U. S. Department of Agriculture, Agricultural Handbook No. 282. 47 pp.

 Wishowski, J. M., Mamo, M., and Bubenzer, G. D. 1998. Trap Efficiencies of Filter Fabric
       Fence. Paper No 982158, American Society of Agricultural Engineers, St. Joseph, MI.

Wyant, D. C. 1980. Evaluation of Filter Fabric for Use as Silt Fences. VA Highway and
       Transportation Research Council, Richmond, VA.

WYDEQ. 1999. Urban Best Management Practices for Nonpoint Source Pollution. Point and
      Nonpoint Source Programs; Water Quality Division, Wyoming Department of
      Environmental Quality, Cheyenne, WY.
      http://deq.state.wv.us/wqd/watershed/92171.pdf

Yousef, Y., M. Wanielista, H. Harper, D.  Pearce and R. Tolbert. 1985. Best Management
      Practices: Removal of Highway Contaminants by Roadside Swales. Final report.
      University of Central Florida. Florida Department of Transportation. Orlando, FL. Also
      in: "Pollutant Removal Pathways in Florida Swales".  Watershed Protection Techniques
      Fall 1995. 2(1): 299-301.
June 2000
                                                                                 5-133

-------
Development Document for Construction and Development Proposed Effluent Guidelines	

Yu, S., S. Barnes and V. Gerde. 1993.  Testing of Best Management Practices for Controlling
       Highway Runoff. Virginia Transportation Research Council. FHWA/VA-93-R16, pp. 60.
       Also in: Performance of Grassed Swales Along East Coast Highways.  Article No, 114 in
       The Practice of Watershed Protection. Center for Watershed Protection, Ellicott City,
       MD. 2000.  http://www.stormwatercenter.net
June 2000
                                                                                     5-134

-------
 Development Document for Construction and Development Proposed Effluent Guidelines        	

 SECTION 6: REGULATORY DEVELOPMENT AND RATIONALE

 In this section, the methodology used by EPA to develop regulatory options for the construction
 and land development industry is described.  EPA methodology first evaluated the pollutants
 discharged from the industry and evaluated existing Federal, State and local control strategies
 designed to manage impacts. Based on this analysis, EPA was able to identify several key
 components of existing regulatory strategies that would be applicable for national effluent
 guidelines regulations and develop regulatory options around these existing strategies. Following
 development of regulatory options, EPA evaluated the costs and environmental benefits of
 several options and determined the appropriate option for proposal based on factors such as total
 costs, monetized and non-monetized environmental benefits, ease of implementation, industry
 financial impacts, and industry acceptance. The following sections describe the components of
 this process involving identification of impacts, evaluation of available control strategies, and
 formulation of regulatory options. Costs of regulatory options are discussed in Section 7 of this
 document while a description of the environmental benefits estimation and industry financial
 analyses, can be found in the other supporting documents of this regulation (USEPA 2002 and
 2002a).                             .

 6.1  IDENTIFICATION OF INDUSTRY IMPACTS

 In developing effluent guidelines for controlling storm water discharges associated with
 construction and land development activities, EPA identified pollutants that are attributable to
 the industry.  In addition to pollutants discharged from construction sites and from long-term
 storm water discharges, EPA also looked at the broader range of environmental impacts that the
 land development process influences and that could potentially be addressed under effluent
 guidelines regulations. These categories include physical impacts to receiving streams due to the
 increased frequency of high flow rates and associated discharge of sediment, as well as, thermal
 impacts to receiving waters due to the increased temperature of storm water discharges.

 These analyses helped EPA to develop regulatory options and associated estimates of costs and
 benefits for temporary erosion and sediment controls. This approach allowed for the evaluation
 of different combinations of regulatory options when developing an overall regulatory strategy
 for this industry, with different combinations addressing various impact areas.

6.1.1   Pollutant Indicators

When determining which pollutants to assess, EPA applied the following priorities for
construction storm water discharges:

    Focus on pollutants  directly attributable to the industry, using indicator pollutants where
    necessary;                                     .
June 2002
                                                                                     6-1

-------
Development Document for Construction and Development Proposed Effluent Guidelines	        :

•    Focus on pollutants most commonly encountered under most settings, (i.e., not to pre-
     construction site contamination issues or accidental discharges);
•,    Focus on pollutants that are most manageable given the current suite of available
     technologies; and
•    Focus on pollutants that can be addressed under the authority of effluent guidelines. '

EPA conducted an extensive evaluation of the literature to identify pollutants present in siorm
water discharges from construction and land development sites. While the literature contains
extensive information on pollutants present in storm water discharges from urban areas, there
were little data available on pollutants present in storm water discharges from construction sites
during the active construction phase other than for sediment, TSS and turbidity. This is not
surprising, since construction site storm water management is primarily concerned with the
control of solids from exposed soil areas. There is the potential for other pollutants to be
discharged from construction sites depending on factors such as prior land uses. For example, if
the prior land use was agriculture, there is the potential for discharge of pollutants such as
nutrients and pesticides. Likewise, areas of redevelopment that occur on sites where previous
land uses included industry could discharge pollutants such as organics and metals. In addition,
pollutants such as metals and nutrients can be present in native site soils, and could be discharged
from construction sites. However, EPA was not able to identify sufficient data in the literature to
warrant development of controls specific to pollutants other than sediment, TSS and turbidity in
storm water discharges from construction sites. Some literature suggests that pollutants adhere to
sediment so regulating TSS should also act as a control for other pollutants.

There are extensive data in the literature describing pollutants present in storm water discharges
from urban areas. The most comprehensive evaluation of urban storm water was the Nationwide
Urban Runoff Program (NURP) (USEPA, 1983). While somewhat dated, the NURP results are
still valid, and serve as a primary means  of characterizing urban runoff pollutants. In addition to
NURP, a variety of other analyses conducted over the past 20 years have contributed greatly to
the understanding of pollutants present in urban storm water runoff. Literally thousands of
references can be found in the literature summarizing hundreds of studies evaluating urban runoff
pollutant levels. As a result, there are sufficient data available to identify the major pollutants
expected to be discharged from new land development activities. Based on these data sources,
EPA identified sediments  (measured as TSS), nutrients and metals as pollutants of concern for
this industry. EPA also evaluated the inclusion of organics, pesticides, and bacteria as potential
pollutants of concern, but  the literature indicates that control of these pollutants through
conventional storm water management strategies is potentially much more difficult, and that
there are little data linking their presence in storm water discharges directly with new land
development activities.  Source control may factor greatly into controlling these pollutant
sources.

Although EPA identified a number of pollutants of concern for this industry, EPA did not
develop regulatory options specifically targeted at controlling each of these individual pollutants.
 June 2002
                                                                                       6-2

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

 Instead, EPA chose to develop regulatory options using an indicator pollutant, TSS. While TSS
 levels may not be directly correlated with all pollutants of concern, it is certainly the most widely
 reported parameter in the literature due to its relative ease of collection and low cost. In addition,
 design of management systems for the control of TSS will likely result in control of pollutants
 such as sediment, nutrients and metals that are present in the solid-phase (attached to sediments).
 The one pollutant of concern that may not have a strong correlation with TSS is turbidity, since
 particles that contribute to turbidity may not be removed through conventional storm water
 management practices that control TSS.  Particles' that contribute to turbidity may be of such a
 fine grain that they will not be removed by the mechanisms whereby most BMPs operate, mainly
 settling and filtration.

 EPA's assessment of pollutant loadings for the industry was based on mathematical models.
 These models were developed using analyses prepared by EPA for the  NPDES Phase n
 rulemaking (USEPA, 1999), established hydrologic principles and storm water monitoring data
 from the literature. EPA estimated annual loadings with and without effluent guidelines from
 construction site storm water discharges using 225 site models which varied based on location,
 site size and site slope. In its assessment of the industry, EPA elected to use the estimated land
 area constructed annually in the nation for the contiguous states, based on the National Resources
 Inventory (NRI)(USDA, 2000). EPA did not develop estimates of pollutant loadings for Alaska,
 Hawaii, and the U.S. territories, due to a several factors, such as a lack of rainfall data and lack of
 data on annual land development. However, due to' the small amount of development that occurs
 in these areas, the omission of these areas from the analysis  is not expected to contribute a
 significant error to EPA's national estimates.

 In developing pollutant loadings of the land development industries, a distinction was made
 between primary pollutant loadings (e.g., discharge of sediments from disturbed ground surfaces)
 and secondary pollutant loadings  (e.g., loadings resulting from accelerated erosion of streams
 caused by increased high flows from urbanized land uses). This distinction was made because
 studies focusing on the impacts of land development have sometimes neglected the secondary
 pollutant loadings that result when changes to hydrology cause downstream channels to become
 unstable. The secondary pollutant loadings that occur year after year from increased stream
 flows have not been well inventoried.1
       1 Not all pollutant indicators listed above are directly used by EPA in its benefits
assessment, or in developing the C&D effluent guidelines. Nevertheless, EPA has collected data
to estimate all of the measures for potential future consideration of this and other
industries/activities.

June 2002    "  '        '                         ~~                                   ^3

-------
 Development Document for Construction and Development Proposed Effluent Guidelines

 6.1.2  Physical/Habitat Indicators

 In addition to assessing impacts of the construction and land development industry due to
 discharge of pollutants in storm water, EPA also developed a methodology for assessing the
 physical and habitat impacts caused by changes in hydrology and stream flow. Land
 development activities cause significant alterations in the natural hydrologic regime of   '
 developing watersheds. The removal of vegetation., the compaction of soils by construction
 equipment and the construction of impervious surfaces such as roads, driveways and buildings
 causes a marked increase in the total volume and peak flow rate of storm water discharges as
 compared to forested, open and agricultural land uses. As a result, streams receiving storm water
 discharges will frequently undergo significant channel alterations in order to adjust to the altered
 hydrologic regime.  This alteration results in mobilization of high quantities of sediment and
 associated water quality problems.                                          .       ;  '

 EPA's assessment attempted to develop an impacts time line, predicting when certain impacts
 will occur. Due to its relatively short duration, construction impacts (or benefits) were assumed
 to occur within a single year. The assessment of long-term impacts was based on the 30 year
 period immediately following conversion into urban land use.  This includes characterization of
 physical/habitat impacts related to hydrologic changes (e.g., increased flooding and stream
 erosion) and changes in runoff characteristics (e.g., runoff thermal signature).  In its modeling
 effort, EPA made assumptions that simplify (spatial and temporally) land development,
 compressing the period required for land to reach "build-out." EPA performed sensitivity
 evaluations to verify that these simplifications do not distort or abrogate its assessment of
 potential environmental impacts.                                                   ;

 Physical/Habitat Measures Estimated by EPA2 include:                               '.

 •   Miles of stream urbanized (located within the area urbanized nationally in a single year)
 •   Number of new stream crossings expected to become fish migration barriers
 •   Acres of stream habitat lost to new stream crossings
 •   Acres of stream-side area flooded by the 100-year rainfall event
 •   Tons of stream bank/bed sediment removed as a result of increased high flow rate frequency

A detailed discussion of EPA's environmental assessment methodology and results is presented
in other supporting documents of this rule (USEPA, 2002 and 2002a).                  :
       2 Not all physical/habitat measures listed above are directly used by EPA in its benefits
assessment, or in developing the C&D effluent guidelines. Nevertheless, EPA has collected data
to estimate all of the measures for potential future consideration of this and other        !
industries/activities.
June 2002
                                                                                     6-4

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

 6.2  DEVELOPMENT OF REGULATORY OPTIONS

 In developing effluent guidelines for the construction and development industries, EPA evaluated
 a variety of state and local programs to identify various management strategies and regulatory
 components that would be applicable on a national basis. For erosion and sediment control and
 other temporary BMPs, EPA considered a series of regulatory options. These options are
 designed to control the discharge of sediment, storm water and other pollutants from sites when
 construction is taking place. EPA considered a range of options that incorporate varying levels
 of management and various control strategies. Because long-term storm water management is
 beyond the scope of the controls proposed by EPA, the following discussion only presents
 information related to options for controlling storm water during the active phase of construction.

 The following discussion presents various options that EPA considered.

 Codify the EPA Construction General Permit

 EPA considered an option that would essentially codify the provisions contained in EPA's
 construction general permit (CGP)  (USEPA, 1998) as minimum national standards for erosion
 and sediment control (i.e., for all states, not only those with EPA as permitting authority) for sites
 of 5 acres or more of disturbed land. Requirements include preparing a Storm Water Pollution
 Prevention Plan (SWPPP) or equivalent, provisions for installing and sizing sediment basins on
 sites with more than 10 acres of disturbed land, requirements for providing cover on exposed soil
 areas within 14 'days after construction activity has ceased, and installation and maintenance of
 other erosion and sediment control  practices and other temporary BMPs on all construction sites,
 such as silt fencing, seeding and mulching, diversion dikes and berms, sediment traps, storm
 drain inlet protection, channel liners, erosion control blankets and mats, stabilized construction
 entrances, litter, trash and debris control, discarded building material control,  and concrete truck
 wash water control.

 Numerical Design Requirements

 EPA considered an option that would establish numerical requirements for the design of
 sediment basins and traps based on local or regional rainfall patterns and site-specific soil types.
 This options could be similar to existing requirements designed for managing storm water
 discharges, where sediment controls are sized based on a specified rainfall return frequency (such
 as the 2-year; 24-hour storm), or a specified runoff frequency (such as the 90th percentile runoff
 event).

 Numerical Pollutant Removal Requirements

,EPA considered options that would contain numerical requirements for the removal of specific
 pollutants from construction site runoff. EPA initially considered targeting a variety of pollutants

 June 2002                '                   "    ~~   :                            '     £?

-------
Development Document for Construction and Development Proposed Effluent Guidelines	

including sediment, TSS, turbidity, nutrients, metals and other priority pollutants, however there
are little data available supporting the feasibility of controlling pollutants other than sediment (or
associated indicator parameters such asTJ3S, turbidity, total suspended sediment, or settleable
solids). This option could be expressed as either a percent removal through sediment controls
(such as sediment basins or traps), or as a total site reduction (incorporating consideration of
sheet flow and diffuse runoff in addition to discrete conveyances); In addition to establishing
numerical requirements for the control of sediment, EPA preliminarily considered establishing
requirements for removing fine-grained and slowly- or non-settleable particles contained in
construction-site runoff (such as turbidity). This option would likely have relied primarily on
chemical treatment of soils  or construction site runoff using polymers or coagulants such as alum
in order to prevent the non-settleable fractions of solids from being transported off-site.

Discharge Monitoring

EPA considered the inclusion of monitoring requirements for evaluating the effectiveness of
erosion and sediment controls.  Monitoring of storm water discharges from construction sites
could be used to evaluate the effectiveness of individual sediment controls (such as sediment
basins), or monitoring the receiving water above and below construction sites. Monitoring
requirements could be incorporated with any of the previously discussed regulatory options
considered.        .                                                               '
                                                                                  i
Inspection and Certification                                                        i

EPA considered an option that includes mandatory site inspection, maintenance and reporting
provisions by site owners and operators in order to improve confidence in the implementation
and performance of construction site erosion and sediment controls. These certification
provisions maybe accomplished either through self-inspection by a qualified employee of the
owner and  operator (such as a professional engineer or person trained in erosion and sediment
control techniques) or inspection by a third-party (such as a consulting firm). The certification
provisions would consist of a checklist-type certification form that the permittee would be
required to complete at various stages of the project to certify that the provisions contained in the
permittee's SWPPP are being implemented, hi addition, the permittees would be required to
conduct periodic inspections in order to confirm that the permittee is conducting the maintenance
necessary to maintain the functionality of BMPs. The specific activities requiring certification
include: SWPPP preparation; installation of perimeter controls and sediment controls; site
inspections every 14 days; final stabilization of exposed soils and removal of temporary erosion
& sediment controls. The certification and inspection forms would be retained on the site, and
made available to the permitting authority and the public upon request.                  | -
 June 2002
                                                                                       6-6

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

 6.3  REGULATORY OPTIONS DEVELOPED FOR THE PROPOSED RULE

 6.3.1  Option 1 - Inspection and Certification

 Option 1 proposed by EPA would establish the site inspection and certification provisions
 discussed above as minimum requirements for all construction sites subject to the NPDES storm
 water regulations. This includes sites from 1 up to 5 acres that will be required to obtain a permit
 once the Phase II regulations are implemented and sites 5 acres or greater that are required to
 obtain a permit under the Phase I regulations.  The permittee would be required to conduct
 periodic inspections and provide certifications as to certain activities (such as SWPPP
 preparation, BMP installation, periodic maintenance, etc.). Under this option, these inspections
 and certifications would be performed by a qualified professional, such as a registered
 professional engineer or person trained in erosion and sediment control. The permittee may
 provide self-certifications if qualified.

 The specific inspection and certification provisions can be found in the proposed rule language
 and are summarized below:

 Site log book. The permittee would be required to maintain a record of site activities in a site log
 book. The specific requirements and information contained in the log book consists of the
 following:

 (1)  A copy of the site  log book would be required to be maintained on site and be made
     available to the permitting authority upon request. EPA recommends that the permittee also
     make a copy of the site log book available to the public upon request within a reasonable
     period;

 (2)  In the site log book, the permittee shall certify, prior to the commencement of construction
     activities, that'any plans required by the permit meet all Federal, State, Tribal and local
     erosion and sediment control requirements and are available to the permitting authority;

 (3)  The permittee would be required to have a qualified professional, conduct an assessment of
     the site prior to groundbreaking and certify that the appropriate BMPs described in plans
     required by the permit have been adequately designed, sized and installed to ensure overall
     preparedness of the site for initiation of groundbreaking activities. The permittee would be
     required to record the date of initial groundbreaking in the site log book. The permittee
     would also be required to identify and conduct any soil stabilization and BMP maintenance
     requirements identified in the permit within 48 hours of their identification;

(4)  The permittee would be required to post at the site, in a publicly-accessible location, a
     summary of the site inspection activities on a monthly basis. EPA recommends that the
     permittee provide contact information for obtaining a copy of the site inspection log book;
June 2002
                                                                                    6-7

-------
Development Document for Construction and Development Proposed Effluent Guidelines             	.

Site Inspections. The permittee or designated agent of the permittee (such as a consultant,
subcontractor, or third-party inspection firm) would be required to conduct regular inspections of
the site and record the results of such inspection in the site log book.  Specific inspection
provisions include:

(1) After initial groundbreaking, permittees would be required to conduct site inspections at
    least every 14 calendar days and within 24 hours of the end of a storm event of 0.5 inches or
    greater.  These inspections would be required to be conducted by a qualified professional.
    During each inspection, the permittee or designated agent would be required to conduct the
    following activities and record the following information:

       (i)     Indicate the extent of all disturbed site areas and  drainage pathways. Indicate site
              areas that are expected to undergo initial  disturbance or significant site work
              within the next 14-day period;
       (ii)     Indicate all areas of the site that have undergone  temporary or permanent
              stabilization;                                     -          .        :
       (iii)    Indicate all disturbed site areas that have  not undergone active site work during
              the previous 14-day period;
       (iv)    Inspect all sediment control practices and note the approximate degree of
              sediment accumulation as a percentage of the sediment storage volume (for
              example 10 percent, 20 percent, 50 percent,  etc.). Note all sediment control
              practices in the site log book that have sediment  accumulation of 50 percent or
              more; and                                                          i
       (v)     Inspect all erosion and sediment control BMPs and note compliance with any
              maintenance requirements such as verifying the integrity of barrier or diversion
              systems (e.g., earthen berms  or silt fencing)  and containment systems (e.g.,
              sediment basins and sediment traps).  Identify any evidence of rill or gully erosion
              occurring on slopes and any loss of stabilizing vegetation or seeding/mulching.
              Document in the site log book any excessive deposition of sediment or ponding
              water along  barrier or diversion systems. Note the depth of sediment within
              containment structures, any erosion near  outlet and overflow structures, and verify
              the ability of rock filters around perforated riser pipes to pass water.

(2)  Prior to  filing of the Notice of Termination or the end of permit term, the permittee or
     designated agent would be required to  conduct a final  site erosion and sediment control
     inspection. The inspector would be required to certify that the site has undergone final
     stabilization as required by the permit and that all temporary erosion and sediment controls
     (such as silt fencing) not needed for long-term erosion control have been removed.
 June 2002
                                                                                       6-8

-------
 Development Document for Construction and Development Proposed Effluent Guidelines   	

 6.3.2  Option 2 - Codify EPA CGP Requirements with Site Inspection and Certification
       Provisions

 Option 2 proposed by EPA would require the permittee to prepare a storm water pollution
 prevention plan (SWPPP) and implement the erosion and sediment controls contained in the
 EPA CGP. In addition, the permittee would be required to conduct periodic site inspections and
 provide certifications in a site log book This option would only apply to sites with 5 or more
 acres of disturbed land., The details of this option can be found in the proposed rule language and
 are summarized below:

 General Erosion and Sediment Controls

 Each SWPPP would be required to include a description of appropriate controls designed to
 retain sediment on site to the extent practicable. These general erosion and sediment controls
 would be required to be included in the SWPPP described below. The SWPPP would be
 required to include a description of interim and permanent stabilization practices for the site,
 including a schedule of when the practices will be implemented. Stabilization practices may
 include:

 (1) Establishment of temporary or permanent vegetation;

 (2) Mulching, geotextiles, or sod stabilization;

 (3) Vegetative buffer strips;

 (4) Protection of trees and preservation of mature vegetation.

EPA recommends that all controls be properly selected and installed in accordance with sound
engineering practices and, when feasible, manufacturer's specifications.

Sediment Controls

Operators would be required to design and install structural controls to divert flows from exposed
soils, store flows or otherwise limit runoff and the discharge of pollutants from exposed areas
and to describe controls in the SWPPP. These controls are as follows:

(1) For common drainage locations that serve an area with 10 or more acres disturbed at one
   time, the operator would be required to provide a temporary (or permanent) sediment basin
   that provides storage for a calculated volume of runoff from a 2 year, 24-hour storm from
   each disturbed acre drained, or equivalent control measures, where attainable until final
   stabilization of the site. Where no such calculation has been performed, the operator would
   be required to provide a temporary (or permanent) sediment basin providing 3,600 cubic feet
June 2002
                                                                                     6-9

-------
Development Document for Construction and Development Proposed Effluent Guidelines	,	
       ^                           .                                               !
     feet of storage per acre drained, or equivalent control measures, where attainable until final
     stabilization of the site. When computing the number of acres draining into a comnjon
     location it is not necessary to include flows from off-site areas and flows from on-site areas
     that are either undisturbed or have undergone final stabilization where such flows are
     diverted around both the disturbed area and the sediment basin.                   :

(2)  In determining whether a sediment basin is attainable, the operator may consider factors
     such as site soils, slope, available area on site, etc. hi any event, the operator wouldibe
     required to consider public safety, especially as it relates to children, as a design factor for
     the sediment basin. Use of alternative sediment controls would be required where site
     limitations preclude a safe basin design.                                        '

(3)  For portions of the site that drain to a common location and have a total contributing
     drainage area of less than 10 acres, the operator would be required to consider installation of
     sediment traps or other sediment control devices.                                •

(4)  Where neither a sediment basin nor equivalent controls  are attainable due to site limitations,
     the operator would be required to install silt fences, vegetative buffer strips or equivalent
     sediment controls for all down slope boundaries of the construction area and for those side
     slope boundaries deemed appropriate for individual site conditions.

Pollution Prevention Measures

The operator would be required to implement the following pollution prevention measures:

(1)  The operator would be required to prevent litter, construction chemicals, and construction
     debris from becoming a pollutant source in storm water discharges; and           ;

(2)  The operator would be required to contain construction  and building materials in   .
     appropriate storage areas and manage the materials to prevent contamination of storm water
     runoff.             .                                                         !

Storm Water Pollution Prevention Plan                                             ;

Permittees would be required to compile Storm Water Pollution Prevention Plans (SWPPPs)
prior to groundbreaking at any construction site. In areas where EPA is not the permit authority,
operators may be required to prepare documents that may serve as the functional equivalent of a
SWPPP. Such alternate documents would satisfy the requirements for a SWPPP so long;as they
contain the necessary elements of a SWPPP.  A SWPPP would be required to incorporate the
following information:
 June 2002
                                                                                     6-10

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

 (1)  A narrative description of the construction activity, including a description of the intended
      sequence of major activities that disturb soils on the site (Major activities include any
      clearing, grubbing, excavating, grading, soil stockpiling, and utilities and infrastructure
      installation, or any other activity that results in significant disturbance of soils.);

 (2)  A general location map (e.g., portion of a city or county map) and a site map. The site map
      shall include descriptions of the following:

        (i)     Drainage patterns and approximate slopes anticipated after major grading
               activities;
        (ii)     The total area of the site and the area of the site that is expected to be disturbed by
               excavation, clearing,, grading and other construction activities during the life of
               the permit;           -    '
        (iii)    Areas that will not be disturbed;
        (iv)    Locations of erosion and sediment controls identified in the SWPPP;
        (v)     Locations where stabilization practices are expected to occur;
        (vi)    Locations of off-site material, waste, borrow or equipment storage areas;
        (vii)    Surface waters (including wetlands); and
        (viii)   Locations where storm water discharges to a surface water;

(3)  A description of available data on soils present at the site;

(4)  A description of BMPs to be used to control pollutants in storm water discharges during
     construction

(5)  A description of the general timing (or  sequence) in relation to the construction schedule
     when each BMP is to be implemented;

(6)  An estimate of the pre-development and post-construction runoff coefficients of the site;

(7)  The name(s) of the receiving water(s);

(8)  Delineation of SWPPP implementation responsibilities for each site owner or operator;

(9)  Any existing data that describe the storm water runoff characteristics at the site (such as data
     that may be collected during a site assessment), and

Updating the SWPPP

The operator would be required to amend the SWPPP and corresponding erosion and sediment
control BMPs whenever:
June 2002
                                                                                       6-11

-------
Development Document for Construction and Development Proposed Effluent Guidelines            '      	

(1)  There is a change in design, construction, or maintenance that is expected to have a
     significant effect on the discharge of pollutants; or

(2)  Inspections or investigations by site operators, local, State, Tribal or Federal officials
     indicate that any BMPs described hi the SWPPP are ineffective in eliminating or  ;
     significantly mhiimizing pollutant discharges.                                  •]

Site Log Book/Certification                                                      ;

The operator would be required to maintain a record of site activities in a site log book, as part of
the SWPPP.  The site log book shall be maintained as follows:                       '.

(1)  A copy of the site log book would be required to be maintained on site and be made
     available to the permitting authority upon request. EPA recommends that the operator make
     a copy of the site log book available to the public upon request within a reasonable period;

(2)  In the site log book, the operator would be required to certify, prior to the commencement of
     construction activities, that the SWPPP meets all Federal, State and local erosion arid
     sediment control requirements and is available to the permitting authority;

(3)  The operator would be required to have a qualified professional conduct an assessment of
     the site prior to groundbreaking and certify that the appropriate BMPs and erosion and
     sediment controls described in the SWPPP have been adequately designed, sized and
     installed to ensure overall preparedness of the site for initiation of groundbreaking activities.
     The operator would be required to record the date of initial groundbreaking in the site log
     book. The operator would be required to certify that the  site inspections, soil stabilization
     activities, and maintenance activities required by the proposed rule have been satisfied
     within 48 hours of actually meeting .such requirements;
                                                                                i
(4)  The operator would be required to post at the site, in a publicly-accessible  location,:a
     summary of the site inspection activities on a monthly basis.  EPA recommends that the
     operator provide contact information for obtaining a copy of the SWPPP and a copy of the
     site inspection log book;                                                     ;

Site Inspections

The operator or designated agent of the operator (such as a consultant, subcontractor, or ^hird-
party inspection firm) would be required to conduct regular inspections of the site and record the
results of such inspection in the site log book. The specific activities that would require j
inspection and certification are:                                            '       ;
 June 2002
                                                                                     6-12

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

 (1)  After initial groundbreaking, operators would be required to conduct site inspections at least
     every 14 calendar days and within 24 hours of the end of a storm event of 0.5 inches or
     greater. These inspections would be required to be conducted by a qualified professional.'
     During each inspection, the operator or designated agent would be required to record the
     following information:

       (i)     On a site map, indicate the extent of all disturbed site areas and drainage
              pathways. Indicate site areas that are expected to undergo initial disturbance or
              significant site work within the next 14-day period;
       (ii)    Indicate on a site map all areas of the site that have undergone temporary or
              permanent stabilization;
       (iii)   Indicate all disturbed site areas that have not undergone active site work during
              the previous 14-day period;                                                  .
       (iv)   Inspect all sediment control practices  and note the approximate degree of
              sediment accumulation as a percentage of the sediment storage volume (for
              example 10 percent, 20 percent, 50 percent, etc.). Record all sediment control
              practices in the site log book that have sediment accumulation of 50 percent or
              more; and
       (v)    Inspect all erosion and sediment control BMPs and record all maintenance
              requirements such as verifying the integrity of barrier or diversion systems
              (earthen berms or silt fencing) and containment systems (sediment basins and
              sediment traps).  Identify any evidence of rill or gully erosion occurring on slopes
              and any loss of stabilizing vegetation  or seedrng/mulching.  Document in the site
              log book any excessive deposition of  sediment or ponding water along barrier or
              diversion systems. Record the depth of sediment within containment structures,
              any erosion near outlet and overflow structures, and verify the ability of rock
              filters around perforated riser pipes to pass water.

(2)  Prior to filing of the Notice of Termination or the end of permit term, a final site erosion and
     sediment, control inspection would be required to be conducted by the operator or designated
     agent. The inspector would be required to certify that the site has undergone final
     stabilization using either vegetative or structural stabilization methods and that  all
     temporary erosion and sediment controls (such as silt fencing) not needed for long-term
     erosion control have been removed.

Stabilization

The operator would be required to initiate stabilization measures as soon as practicable in
portions of the site where construction activities have temporarily or permanently ceased, but in
no case more than 14 days after the construction activity in that portion of the site has
temporarily or permanently ceased. This provision would not apply in the following instances:
 June 2002
                                                                                      6-13

-------
Development Document for Construction and Development Proposed Effluent Guidelines	        •'.	

(1)  Where the initiation of stabilization measures by the 14th day after construction activity
     temporarily or permanently ceased is precluded by snow cover or frozen ground conditions,
     the operator shall initiate stabilization measures as soon as practicable;
(2)  Where construction activity on a portion of the site is temporarily ceased, and earth- ;
     disturbing activities will be resumed within 21 days, temporary stabilization measures need
     not be initiated on that portion of the site.
(3)  In arid areas (areas with an average annual rainfall of 0 to 10 inches), semi-arid areas (areas
     with an average annual rainfall of 10 to 20 inches), and areas experiencing droughts where
     the initiation of stabilization measures by the  14th day after construction activity has
     temporarily or permanently ceased is precluded by seasonably arid conditions, the operator
     shall initiate stabilization measures as soon as practicable.

Maintenance                                                                    |

The operator would be required to remove accumulated sediment from sediment traps and ponds
identified as having sediment accumulation greater than 50 percent to restore the original design
capacity,

6.3.3  Option 3 - No Regulation

EPA also considered an option that would not establish effluent guidelines requirements for this
industry.
6.4    REFERENCES                        ,

USD A. 2000.1997 National Resources Inventory. U.S. Department of Agriculture, Natural
    Resources Conservation Service. Washington, DC.                              (
    http://www.nrcs .usda. gov/technical/NRI/  ,

USEPA. 1983. Final Report of the Nationwide Urban Runoff Program. U.S. Environmental
    Protection Agency. Washington DC.                                           ,

USEPA. 1998.  Reissuance ofNPDES General Permits for Storm Water Discharges from
    Construction Activities. ("Construction General Permit.")  Federal Register, Vo. 63, No. 31,
    p. 7858. February 17, 1998. Washington, DC.
    http://cfpub.epa.gov/npdes/stormwater/cpermit.cftn7program id=6

USEPA. 1999. Economic Analysis of the Final Phase II Storm Water Rule. U.S. Environmental
    Protection Agency. Washington, DC.                                           '   •
 June 2002
                                                                                    6-14.

-------
 Development Document for Construction and Development Proposed Effluent Guidelines          	

 USEPA. 2002. Economic Analysis of Proposed Effluent Limitation Guidelines and New Source
     Performance Standards for the Construction and Development Category; May 2002.  EPA
     821-R-02-008. http://www.epa.gov/waterscience/guide/construction/

 USEPA. 2002a. Environmental Assessment for Proposed Effluent Limitation Guidelines and
     New Source Performance Standards for the Construction and Development Category; May
     2002.  EPA 821-R-02-009.  http://www.epa.gov/waterscience/guide/construction/
June 2002
                                                                                  6-15

-------

-------
  Development Document for Construction and Development Proposed Effluent Guidelines	

  SECTION 7: APPROACH TO ESTIMATING COSTS

  7.1    OVERVIEW

  This section describes EPA's methodology for estimating compliance costs associated with
  implementing the regulatory options proposed for the construction and land development effluent
  limitation guidelines (ELG). EPA estimated three distinct cost categories: (1) erosion and
  sediment control (ESC) costs, including design, installation, operation, and maintenance;
  (2) administrative costs to permittees for activities such as site inspections and certification
  activities; and (3) administrative costs to permit authorities to incorporate the effluent guidelines
 requirements into general permits. Costs contained in categories (1) and (2) are expected to be
 borne directly by the construction and development industry.

 Costs were evaluated individually for 24 site size class and land use types. EPA developed a
 series of model sites for each land use/site size class and estimated costs of proposed options for
 each of these model sites. Using estimates of the population of new construction acreage
 developed using data from the USDA's National Resources Inventory (NRI), the U.S.  Census
 Bureau, EPA's NPDES Storm Water Phase II rulemaking, and other national data sources
 (described in Section 3 of this document), EPA summed the model site costs to the national
 level. A description of this methodology is presented in the Economic Analysis document
 (USEPA, 2002).

 The total costs of'the proposed rule options are presented in Table 7-1.

                     Table 7-1.  Total Costs of Proposed Rule Options
Option
1 - Inspection and Certification sites ^ 1 acre
2 - Codify EPA Construction General Permit (CGP)
with Inspection and Certification sites £5 acres
3 - No Regulation
Annual Cost (millions 2000 dollars)
126
502
0
7.2    METHODS FOR ESTIMATING EROSION AND SEDIMENT CONTROL COSTS

7.2.1         OVERVIEW

EPA used four land use types to account for variations in construction operations and associated
ESCs employed for various development types.  For each land use type, EPA evaluated six site
size classes to account for economies of scale that might occur with certain best management
practice (BMP) design and installation costs (some BMPs are employed only if the site size is
June 2002
                                                                                     7-1

-------
Development Document for Construction and Development Proposed Effluent Guidelines    	.     	

greater than a threshold value). EPA also considered regional cost adjustments due to variations
in labor, supply, and material costs (see Table 7-2). EPA used an industry standard reference to
establish appropriate adjustment factors for regional compliance costs (R.S. Means, 2000).

The costing analysis started by allocating the estimated annual construction acreage and number
of model sites developed in Section 4 (see Table 4-21) for one of 19 EPA-developed ecoregions
shown in Figure 7-1 (see the Environmental Assessment supporting document (EPA, 2002a) for
a complete description of the EPA ecoregions). Matrices of standard BMP quantities for the
technology-based option (Option 2) were developed for the various model site sizes using the
NPDES Phase II economic analysis (USEPA, 1999) and the Agency's engineering judgement.
By multiplying  the two matrices, the total quantity of BMPs for all of the model sites was
determined. EPA estimated the unit costs of each BMP element using R.S. Means (2000), and
data from "The  Economics of Stormwater Treatment: An Update" from the Center for Watershed
Protection's (CWP's) book entitled The Practice of Watershed Protection (Schueler, 2000).
Regional costs were adjusted using cost adjustment factors from R.S. Means (2000), and data
were summed across the different site size categories to determine engineering costs at the
national level. Additional costs for factors such as. design and contingencies (described in the
Economic Analysis) were added to these national costs to arrive at the national cost figures
presented in Table 7-1. All costs presented are incremental over current costs to the industry
from existing Federal and State requirements.

EPA used a similar approach to estimate administrative costs to permittees for conducting the
site inspection and certification provisions contained in Options 1 and 2.  EPA estimated the
number of site inspections needed and the hours required for conducting site inspections and
certifications for each of the model site sizes. By multiplying these hour estimates by a
professional labor rate, EPA was able to estimate the total administrative costs to permittees.
Similarly, EPA estimated the administrative costs to permitting authorities to revise general
permits to incorporate the effluent guideline requirements by multiplying the estimated hours per
entity by the number of entities to arrive at the national costs.
 June 2002
                                                                                     7-2

-------
 Development Document for Construction and Development Proposed Effluent Guidelines
Figure 7-1. EPA Ecoregions
Source: Composited from Omernik, 1987.
June 2002
                                                                                           7-3

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

                 Table 7-2. Regional Compliance Cost Adjustment Factors
EPA
Hydrologic
Region
1
2
3
4
5
6
7
8
9
10 •
11
12
13 .
14
15
16
17
18
19
Regional Compliance Cost
' Adjustment Factor
! 0.855
0.984
0.900
0.782
i 0.857
0.858
0.870
' 1.032
0.877
0.996
0.810
0.854
0.936
0.908
1.094
1.129
1.052
1.046
1.052
                         Source: EPA hydrologic regions are composited
                         from Omernik, 1987. Regional compliance cost
                         adjustment factors are computed based on city
                         data from R.S. Means, 2000.
7.2.2
EROSION AND SEDIMENT CONTROL COSTS
In this analysis, EPA has built upon a number of previous assessments of ESC practices,
including the Economic Analysis of the Final Phase II Storm Water Rule (USEPA, 1999). EPA
estimated types and quantities of ESC BMPs that are commonly employed under baseline
conditions during construction activities to mitigate impacts from construction site runoff for 24
land use/site size class models. In addition, in its analysis EPA estimated that requirements
June 2002
                                                                                     7-4,

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	•

 contained in existing State construction general permit requirements (or, in the non authorized
 states, the region-specific EPA construction general permits (CGPs) would be fully implemented.
 Although Phase II is not fully implemented at this tune, the requirements will be implemented by
 the time final action is taken on these proposed effluent guidelines.  Furthermore, as proposed
 Option 2 (the only option for which EPA is establishing technology-based requirements)
 addressed'only sites with 5 or more acres of disturbed land, the timing of Phase II
 implementation is not an issue.                                              .

 EPA took a model site approach to estimating the baseline ESC usage and quantities of
 materials, as well as design costs and operation and maintenance (O&M) costs that are expected
 to be applicable given a range of physical conditions (1 to 7 percent land slopes and different soil
 types). Table 7-3 lists the construction site BMPs included in the baseline analysis for various
 site sizes. To establish baseline BMP usage, EPA started with the model site estimates generated
 during the Phase II rulemaking, scaling up the BMP quantities to sites larger than 5 acres, and
 adding sediment basins for larger sites. In the final costing analysis of this option, costs  for
 BMPs for sites less than 5 acres were eliminated, consistent with the proposed regulatory
 requirements for Option 2.

          Table 7-3. Construction Site ESC BMP Descriptions and Site Thresholds
ESC BMP Description
Silt Fence, Diversion Dike, Construction Entrances, Stone Check Dams
Mulch
Sediment Traps
Polyacrylamide (PAM)
Sediment Basins
BMP Installation and SWPPP Certifications
Site Inspections - .
Applicable Site Sizes for
ESC BMP Quantity Estimates
> 1 acre
> 1 acre
> 1 acre and < 10 acres
> 1 acres
£ 10 acres
^ 1 acre
£' 1 acres
To determine costs of the regulatory options, EPA first evaluated a variety of State construction
general permits and erosion and sediment control regulations and found that many States have
requirements similar to those contained in the EPA construction general permit, which is the
June 2002
                                                                                      7-5

-------
Development Document for Construction and Development Proposed Effluent Guidelines        	|	

basis for the requirements contained in Option 2 (see Table 7-4)l. In evaluating existing State
programs, EPA specifically examined the major provisions contained in Option 2, namely
sediment basins designed to provide 3,600 cubic feet per acre of storage; requirements fot
stabilization of exposed soil areas within 14 days of reaching final grade; and site inspections at
least every 14 days. In addition, EPA evaluated whether the annual precipitation in each State is
less than 20 niches, since the soil stabilization requirements are linked to this condition.  In the
final analysis of national costs, EPA adjusted the estimates of the national costs for the effluent
guidelines to account for States with programs equivalent to EPA's proposed options.  Table 7-5
summarizes the percentage of national costs eliminated due to equivalent State programs. This is
only applicable to Option 2, as EPA has not determined that a significant number of States have
requirements equivalent to Option 1.

It is expected that on some construction sites there will be some portion of land with steeper
slopes and more erosive soils, which will require more intensive management if built upon than
is assumed by EPA's model. Also, a  State with less than 20 inches of annual rainfall was
considered to be equivalent to a State that has the 14-day cover requirement when assessing
overall equivalence. Local regulations may require use of ESCs that are more stringent than the
Phase I and II requirements.  However, EPA expects that the BMPs selected to develop its model
sites are representative of baseline conditions for the majority of construction activity across the
nation.                                                .            .               ;
       1 Although EPA attempted to obtain comprehensive information, the Agency was not
able to verify the presence of the specific components in Table 7-4 for all States. As a result, the
absence of an entry does not necessarily mean that the State does not currently have an
equivalent requirement.
June 2002
                                                                                      7-6

-------
 Development Document for Construction and Development Proposed Effluent Guidelines
   Table 7-4.  Components of Existing State Erosion and Sediment Control Requirements3
State
Alabama
Alaska
Arizona
Arkansas
California
Colorado
Connecticut
Delaware
District of
Columbia
Florida
Georgia
Hawaii
Idaho
Illinois
Indiana
Iowa
Kansas
Kentucky
Louisiana
Maine
Maryland
Massachusetts
Michigan
Minnesota
Minimum of 3,600
Cubic Feet per Acre
Sediment Basin
Requirement

Yes
Yes

Yes

Yes






Yes

Yes





Yes


, Inspections
Required at Least
Every 14 Days

Yes
Yes












Yes





Yes


14- Day or Less
Stabilization
Requirement

Yes
Yes

Yes

, Yes








Yes





Yes


States with Less than
20 Inches of
Precipitation Per
Year


Yes '

Yes
Yes
Yes





Yes











June 2002
                                                                                     7-7

-------
Development Document for Construction and Development Proposed Effluent Guidelines
State
Mississippi
Missouri
Montana
Nebraska
Nevada
New Hampshire
New Jersey
New Mexico
New York
North Carolina
North Dakota
Ohio
Oklahoma
Oregon
Pennsylvania
Rhode Island
South Carolina
South Dakota
Tennessee
Texas
Utah
Vermont
Virginia
Washington
West Virginia
Wisconsin
Minimum of 3,600
Cubic Feet per Acre
Sediment Basin
Requirement





Yes

Yes

•


Yes

Yes

Yes
Yes
Yes
•Yes
Yes

Yes

Yes

Inspections
Required at Least
Every 14 Days


Yes


Yes

Yes


1
Yes


Yes

Yes
Yes
Yes
Yes
Yes

Yes



14- Day or Less
Stabilization
Requirement





Yes

Yes



Yes


Yes

Yes
Yes
Yes
Yes
Yes

Yes

Yes

States with Less than
20 Inches of
Precipitation Per
Year


Yes

Yes I


Yes


Yes



i


Yes !


Yes ;


i

1
June 2002
                                                                                              7-8

-------
Development Document for Construction and Development Proposed Effluent Guidelines


State
Wyoming

Minimum of 3,600 .
Cubic Feet per Acre
Sediment Basin
Requirement

Inspections
Required at Least
Every 14 Days
Yes
14- Day or Less
Stabilization
Requirement

States with Less than
20 Inches of
Precipitation Per
Year
Yes

     •a. Information is accurate as of May 2002
                 Table 7-5. State Acreage Equivalent to Proposed Option 2

Option 2
Equivalent State
Acreage for
• Sites >5 acres
755,500
Percent of Annual
(>5 acre) Developed
Acreage Equivalent
41
 Once EPA estimated the quantities of ESC BMPs for the model sites, the total baseline cost of
 BMP ^installation was calculated from unit costs provided by R.S. Means (2000) and cost curves
 from "The Economics of Stormwater Treatment: An Update" (Schueler, 2000).  R.S. Means
 provides national average unit costs that include materials, installation, and labor.  Typically,
 users of R.S. Means adjust the national unit costs up or down to obtain their local estimates based
 on city-specific adjustment factors provided by R.S. Means. As described previously, EPA
 developed and used the regional adjustment factors in Table 7-2 to customize unit costs on an
 ecoregion basis, not on a city basis. To  compute region-specific unit costs from the national
 average value, city-specific adjustment factors provided by R.S. Means were converted into
 ecoregion values. First, State-average adjustment factors were estimated based on the values for
 cities they contained. Then ecoregion values were computed based on area-weighting for those
 states that fell within each ecoregion.

 Although R.S. Means is expected to accurately estimate the as-built cost for a particular element,
 in certain cases it might underestimate the cost that a developer or ultimate property owner might
 need to pay a contractor to construct a particular element. This is due to additional site-specific
 cost factors that a contractor may build into a bid package, such as contingencies, allowances for
 change orders, additional time and labor for unseen delays due to weather, unanticipated
 problems with soils^ etc.  However, for the majority of projects, R.S. Means is expected to
 provide accurate cost information. In addition, EPA adjusted for contingency costs in its
 analysis of economic impacts to the industry (see the Economic Analysis document for a
 description of this methodology).

 In calculating the total costs for erosion and sediment control activities, EPA added estimated
 design, operation, and maintenance costs.  Table 7-6 shows design and O&M costs as a fraction
 of the capital cost of ESC BMPs. These cost ratios were obtained from published sources such
June 2002
                                                                                      7-9

-------
Development Document for Construction and Development Proposed Effluent Guidelines             	

as the CWP report and from the Agency's engineering judgement.

In evaluating the proposed rule options, EPA used the model site approach to first determine the
baseline compliance costs, then to modify BMP sizing and BMP quantities to assess the
incremental costs of regulatory options of the proposed rule. The resulting suite of BMPs
evaluated by EPA in establishing costs of the proposed rule are listed in Table 7-7, along with
their unit costs. Appendix B contains additional tables indicating EPA's estimates of the
standard quantity needed for each BMP listed in Table 7-7, for each land use type and site size.
In addition, Appendix B indicates, for key BMPs, the number of equal size BMPs of a single
type that EPA estimates will be needed to serve a single site (i.e., a single 200-acre site will be
served by four equal-size sediment basins, each of which manages 50 acres).

    Table 7-6.  Construction ESC BMP Design and Operation and Maintenance Costs
                            as a Percentage of Capital Costs
, Costed Items
Silt Fence
Diversion Dike
Mulch
Construction Entrance
Stone Check Dam
Sediment Trap
Sediment Basin
Polyacrylamide (PAM)
Effective Life in
Years
1
1
1
1
1
1
1
1
Design Costs as
Percent of
Construction Cost
6%
6%
6%
6%
6%
6%
6%
: 6%
Estimated O&M as Percent of
Original Installation Costs
100% :
10%
2% ;
5%
5% |
20% i
25% '
0
 June 2002
                                                                                     7-10

-------
  Development Document for Construction and Development Proposed Effluent Guidelines	

    Table 7-7. Evaluated Construction Site BMPs that Augment the Suite of Baseline BMPs
         BMP
      Description
                                 Costing Rationale
                                      Erosion and Sediment Controls
   Sediment Basins
   for Sites
   acres
 Standardization to 3,600 cubic feet of storage per watershed acre. Cost based on equation
 for installing permanent dry detention pond, computed from the equation:
 [8.16 x (volume required, cu. ft./number of ponds per site size)0'78] (Schueler, 2000).
   Mulch
                     Mulching of any denuded surface would be required within 14 days of reaching final grade,
                     resulting in more frequent mulching of a portion of the site acreage. Cost of mulching is
                     estimated to be $0.23 per square yard for materials/installation (R.S. Means).  For sites
                     larger than 1 acre, mulching is based on  the total site acreage less the area where structures
                     are being built (estimated as the site impervious coverage). The maximum coverage for
                     single-family and multifamily residential development is 50% of the total site area,
                     assuming the remaining acreage is maintained as open space and/or permanent
                     vegetation/cover is installed.
  Polyacrylamide
  (PAM)
 EPA estimates that a single application of PAM would be used as a temporary stabilization
 method until final cover can be installed. PAM was estimated to be appropriate for only
 20% of construction sites due to physical constraints. PAM is costed at $200 per acre
 treated based on a survey of commercial vendors and the assumption that its application is
 similar to that of herbicide for soil treatment ($0.04 per square yard based on spraying from
 truck) (R.S. Means). The acreage treated is equal to the site size times the ultimate
 impervious area, to a maximum of 50% of.the site size.
                                        Site Administration BMPs
  Site Certifications
For each site, certification activities include certification of storm water pollution
prevention plan (SWPPP) completion, certification of BMP installations, and certification
of final stabilization prior to filing of the notice of termination (NOT). Certification adds an
estimated cost of approximately $ 11 per acre for Options 1 and 2.
  Site Inspections
For each 10-acre unit in the total site, incremental inspection activities over baseline are: (a)
post-BMP installation; (b) once during building; and (c) at end of construction (prior to
filing of the NOT). Inspection adds an estimated cost of approximately $45 per acre for
Options 1 and 2.
Table 7-8 indicates the relative change in quantities of BMPs associated with Options 1 and 2.
Standard quantities of BMPs outlined in Table 7-3 were increased or decreased according to
multiplication factors in Table 7-8 to reflect changes expected due to each option.  For example,
in the case of mulch, EPA estimates that under Option 2, there will be a net increase in the use of
mulch by 20 percent over baseline levels, in part to help meet the requirements for 14-day
coverage of denuded areas. In the case of PAM, EPA anticipates that PAM will be applied to 20
percent of the denuded acreage, as it is an inexpensive and effective means to improve erosion
control.
June 2002
                                                                                                7-11

-------
Development Document for Construction and Development Proposed Effluent Guidelines     	  .  . '	

     Table 7-8. BMP Quantity Adjustment Factors for Baseline and Proposed Options
BMP Type
Silt Fence
Runoff Diversion
Mulch
Construction
Stone Check Dam
Sediment Trap
Sediment Pond
E&S Certification
E&S Inspection
PAM
Baseline
Construction
1.0
1.0
1.0
1.0
1.0
1.0
1.0
0.0
O.Q
0.0
Optionl Inspection/
Certification
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
0.0
Option 2 Inspection/
Certification with
Codification of CGP
1.0 ;
1.0
1.2 ;
1.0 '
1.0 ;
1.0
1.1 '
1.1 ;
1.0
0.2
                                                                                1
Using the information in Table 7-8, EPA estimated baseline costs as well as the costs for Options
1 and 2.  By subtracting the baseline costs from the cost of each option, EPA was able to estimate
the incremental costs of the proposed options.  Table 7-9 indicates the estimated national costs
over baseline of the proposed rule options. Values include design, maintenance, and
opportunity/interest costs. States that are considered to be equivalent to EPA's proposed, options
are removed from the total national cost estimate increases.  The proposed rule is expected to
increase compliance costs for ESCs under Option 1 by $ 126 million, and by $502 million for   .
Option 2 (year 2000 dollars). Option 3 is not expected to have any incremental costs.
 June 2002
                                                                                     7-12

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	

               Table 7-9. National Cost Estimates for Proposed Rule Options
Sector
Single-family Residential
Multifamily Residential
Commercial
Industrial
Total
Option 1 Total Cost
(millions, 1997 dollars)
25.7
12.7
83.8
4.0
126.2
Option 2 Total Cost
(millions, 1997 dollars)
129.7
63.4
296.2
11.8
501.1
 7.3    METHODS FOR ESTIMATING ADMINISTRATIVE COSTS

 7.3.1         OVERVIEW

 The analysis of administrative costs focused on the costs to permit authorities to incorporate the
 effluent guidelines requirements into general permits. Administrative costs are expected to be
 borne by both EPA and States (or surrogate agencies such as conservation districts). EPA's
 assessment is conservative in that it assumes that all States will have to incorporate the effluent
 guideline requirements into their permits. However, EPA estimates that approximately 41
 percent of developed acreage is under state programs that are equivalent to the proposed
 requirements contained in Option 2 and, therefore, will not have to modify their permits to
 incorporate these requirements.
7.3.2
ADMINISTRATIVE COSTS TO PERMITTEES
When considering the administrative costs to permittees for implementation of the proposed
options, EPA estimated the number of CGPs it expects to be issued each year. Table 7-10
indicates the number of construction sites under permit EPA estimates are associated with
current development rates, categorized by Option. In its analysis, EPA estimated that
construction sites not affected by effluent guidelines (those smaller than 5 acres in Option 2, and
those smaller than 1 acre in Option 1) would not incur administrative costs.

Annual administrative costs are expected to be borne by construction firms as a result of site
certification and inspection requirements (See Table 7-7).  Under Options 1 and 2, site operators
will be required to certify that the Storm Water Pollution Prevention Plan (SWPPP) has been
completed, that BMPs are installed according to the SWPPP, that periodic inspections have been
completed, and that the site has been stabilized prior to filing of the notice of termination (NOT).
EPA estimated that it will take 16 hours per 10 acres developed to meet the inspection
requirement. (For construction sites smaller than 10 acres only 16 hours of inspection is
June 2002
                                                                                    7-13

-------
Development Document for Construction and Development Proposed Effluent Guidelines
requked.) EPA used the estimates of construction projects by size presented hi Table 4-21 to
estimate the total hours requked to perform administrative activities.

EPA estimated the total national costs associated with site certifications to be $27,712,000 per
year under Option 1  and $16,727,000 per year under Option 2 (1997 dollars). Based on a review
of States with greater than 50,000 acres per year development, EPA estimates that 34 percent of
acres developed are within States with 14-day inspection requkements that are similar to those
proposed under Options 1  and 2.  As a result, EPA adjusted downward its estimate of national
site inspection costs  to reflect equivalent inspection programs. The total resulting estimates for
site inspection under Options 1 and 2 are $73,161,000 and $44,160,000 per year, respectively.
EPA's estimate of the total annual administrative cost for certification and inspection for Option
1 and 2 are $100,873,000 and $60,887,000, respectively. The total annual administrative costs
are lower for Option 2 because sites of less than 5 acres are not regulated. EPA adjusted these
cost estimates upward to reflect opportunity costs, resulting in overall administrative costs to
permittees of $118,141,000 for Option 1 and $71,290,000 for Option 2. An explanation of this
methodology is presented in the Economic Analysis supporting documentation.         :
7.3.3
ADMINISTRATIVE COSTS FOR GENERAL PERMIT REVISIONS
EPA estimated the total one-time costs for permit authorities to incorporate the erosion and
sediment control effluent guidelines requkements into thek general permits. EPA's estimates of
full tune equivalents (FTEs) and costs for each agency to incorporate effluent guidelines
requirements are indicated in Table 7-10. To determine costs of incorporating the effluent
guidelines requkements into existing State CGPs, EPA estimated that each State will requke 200
hours to evaluate and then modify general permits to incorporate new requirements. All 50
States were estimated to encounter administrative costs, even though many States already have
general permits that meet some of the proposed requkements.  When dividing costs between
Federal and State entities, EPA's estimated costs will be allocated based on the percentage of
States currently authorized to manage the NPDES program (i.e., 44 of 50, or 88 percent).

   Table 7-10. One-Time Hours and Costs to Incorporate Erosion and Sediment Control
          Effluent Guidelines Requirements into General Permits (1997 Dollars)  !
Program Element
Revise General Permits (hours)
Revise General Permits (dollars)
Federal
1,200
$31,000
State
8,800
$229,000
June 2002
                                                                                   7-14

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	'

 7.4    REFERENCES

 Schueler, Thomas R... 2000. "The Economics of Stoimwater Treatment: an Update". Article
    No.. 68 in the Practice of Watershed Protection. Center for Watershed Protection, Ellicott
    City, MD. http://www.stonnwatercenter.net

 Omernik, James M. 1987. Ecoregions of the Conterminous United States. Annal of the
    Association of American Qeorgraphers. 77(1): 118-125.

 R.S. Means.  2000. Site Work & Landscape Cost Data, 19th Edition.  R..S. Means Co., Kingston
    MA.        .

 USEPA. 1999. Economic Analysis of the Final Phase li Storm Water Rule. U.S. Environmental
    Protection Agency. Washington, DC.

 USEPA. 2002. Economic Analysis of Proposed Effluent Limitation Guidelines and New Source
    Performance Standards for the Construction and Development Category; May 2002. EPA
    821-R-02-008. http://www.epa.gov/waterscience/guide/construction/

 USEPA. 2002a. Environmental Assessment for Proposed Effluent Limitation Guidelines and
    New Source Performance Standards for the Construction and Development Category; May
    2002. EPA 821-R-02-009.  http://www.epa.gov/waterscience/guide/construction/
June 2002
                                                                                  7-1.5

-------

-------
 Development Document for Construction and Development Proposed Effluent Guidelines
                             APPENDIX A:
      STATE REGULATIONS ON THE CONTROL OF CONSTRUCTION
                            STORM WATER
June 2002

-------

-------
 a
<§
o:

55












I
1 S1
&i
s \
is
.2

ill*
= g N S
"* s is S
E < 2 |

E .2 w g
K= = &
g g£
Q
- i« I
||I|]

Q 3 3 § *
3 - O* ffl c
E o g •g |
C "g 3 3 n
*u -i
•s
(D *j
•0 •£ c c

W 5 0 g
"i; °~ ° °
E ° a
a
z
" 1 1 I--

3^ O m ^J
m g 0- > .fi
°<-2" ~
1
(Q
Geographic Area N





CO
o
o"
r:
o
re
mpliance date is N
8
o
1

*!
5||
O CO ^
t"

























1
f

__
(O 5 (0
Clean Water Act NPDE
Storm Water program 1
Phase 1 and Phase II M

|
1 o
c £
s 8
E "
m 03
proved erosion and s
ns erosion and sedim
s-f
B 8
re and implement
ilar document that
ra E
Q. .E «
£ » §
f-sf
W C >
3 CO '
5 Q. a
























••



3
L






a
OT
•s
el provided by the dis
c
c
I
lust be qualified pi
E
0
S
Q.
W
C
N.
"* TD >
° S3
n
s









1

(O
g
D
(N

0
5
£
:
o
CM

)
>

o
»
f
M


g-
S
i o o
To (0 £
CD CO 
i sediment control p\i
g a notice of intent. S
sediment basin. Site
ences.
i = = •
S B 1^
must submit erosi
3vention plan befo
temporary or pernr
sediment traps an
§ 3-1?
Q. C CD '
O O C c
> I 2 j
oi.il
1
X
LU










|
F
CM
3
)
?










i

M


:
"5 « co
 §• -a w c
S 2 '? « i ^ i
§ | g .2 |.£ 5
E o 'to •— ns Q.
^ -D p_ CO g c W
ill
2 £ *2 S ° S S.
0 "> g S> g. g 0.
^S fe | g | | Q ;
IHilil!
>OO1O'-J'~O'
g w ~ ir . ^" 'ij ]
|| •§ | J i 1 I
S. 1 S f | = S !
= 1 S. E 8 i.£ j
3
to ^
o .£










i

s
"^
I
)
CM











-"
i


- •


CO
T3
O
ra
^
m
1 be submitted to stati
3.
§ ~
c °
r management pla
/ell as be retained
0) *
§ i
3 '>
en £
o J2 E
C *- CO O)
••§ s § ™
n «> ~ "o
list
< S 2
















g








' O
-" l*-~
M W


~





res, 0.5 inch rainfall.
8
8
:, 1 inch rainfall, <'

:













•
"
s"
o
i-ionaa, south Florida \
Management District
(General, Standard Gen
Noticed General and
Individual Permits)


























>
in
o











i
;
i
o -8
"S i:
hlorlda, southwest Floi
Water Management Pis































=,.

0

j
1

(—

1
:
i
;
'
hlorlda, St. Johns Rivei
Management District


























3
:



)

i
I
i
~

S

1


JU
*" i
hionaa, buwannee Rive
Water Management DIs

g
jj
c '«
m CO
• waters supporting w
Jity units for waters d
•8 €
'£ ti
ometric turbidity u
<10 nephelometrii
^ 0
l| e
/ ia !












j
)

1
I
[ ,
m
CM
w
i

?
*?
i Z
j




-



,

-------











1
o
Z











Visual Inspection
Fraquoncy
I Hi
E < 3 5
g a s rS
iiy
f 1 o|!
1 ? o s |
li!1!
S "• E
^
5 ** T±

3
Z
llilc
ily I
0 < £ 0
e
3
r
!
t
1
I
C
C
fl
•5
| '
o 2
J g

- e
S £
i g
a. S d
sis
g. 1 3
ill
r-o ™
isl
3 S S-
~ OS tZ
2 J5 •••
» re °
§ »- c
5 * :§

5 S g
I S 1
-II
J "O "In
3 jl 0
*? s S
111
After 0.5 inch
rainfall and every 7
days during dry
season, every day
during rainy season













S
CO
CM



S31
X
I























After 0.5 inch
rainfall and every
14 days













1
r*-T



O
1























•a
S
»
UJ



n
c
I
0
2
£
a






a
h-



j
"
i























1 Every 7 calendar
days and within 24
hours of 0.5 Inch of
precipitation













o
§
r-T
CM



re
§
•o
c























At least once per
week








g

i
n
i
-5

X
iS
X) 00_
••- r~-
N CM



V)
ra
( (O
































co S §
Q. '"G
*ill
g S. c i
^1 |1
g S i.0

£ **
1
iC
CM



t
i




























1
CM



CO
40-80% TS
reduction


X3^
S




3
15
:
S
n
i
n
j
_ ^
ra 2
3 W
3 0
(D -C
3) f
3S
D (TJ
s I
0 CM
' 2
S a
s g
(D




!
i
s
to
CM

*
73
% TSS redui
o
CO

„
tn



Maryland










|
5
D
S>
2
0
D

D
n
0 ,
a
CO
Encourage weekly
inspections



per acre
$!
^ o
6 3
D 3

CO
C
:
:
1
n
t
£


i§
N ^



Massachu!
Michigan























llg^ig s




1
0







S
00_
CM



2

-------
3
CO
•8
u
z
1
S- i

— D
•B £
3
>
*Sg
— ON®
^ a E 1
E < 2 2
logo
ill5
* Q

ill. 4
0 € *" § £
£ « >. ? o
io|ii
lllsl
3 2. (0
£ "- E
I- -
I.I II
0) J 0 £
f?l §
E o IE
3
Z
IHI?
3 ra n. > £
« g !_ O S.
D < £ 0
1acre; sediment traps should only be used on drainage
areas > 3 acres; and temporary sediment ponds should only be used on
drainage areas > 10 acres. . .
r-
j: fr
S |
**? "O >,

J? =
< c
2





o.
-C
N
g
CM








1
CM




H
re
c
o

1
•a
i^.

ai
til

§ *
CD o>
3 g1 « '£ S
^o *g ro o> 3
i g •£ f> .-&
1 ° j? ™
S ^
•a








>.




CO
•tr







u
z

f
a
E
CO
s
§



















o






ra
In
I

























q







S
Z

























<="




£
i
E
S.
I

























g





0
to
0
1


^_






















o





8
X
o
£
i













•g
c
IO
o









o







o
1


^_






















o





s
o
^
n
Q
o
z

























D








.2
O
A vegetated buffer zone of at least 100 ft must be retained or successfully
established between the area disturbed during construction and all perennial
or intermittent streams on or adjacent to the construction site. A vegetated
buffer zone at least 50 ft wide must be retained or established between the
area disturbed during construction and all ephemeral streams or drainages.
Treatment volume is the lesser of 3,600 or the runoff volume of a 2 year 24
hour storm.

^_









i
20 acres, erosion and sediment control plan must be prepared by
a Professional Engineer, or Registered Landscape Architect, or Certified
Professional in Erosion and Sediment Control, and plan must be submitted
90 days before construction begins. All permittees must submit an Oregon
Land Use Compatibility Statement if they do not already have one on file
with Oregon DEQ.
jili ||| M>1
-2 21"!? &'c'to'*"'oP:5

>.-D i-a5£-o w § ^ ™
<5->«-2 nic**? E 8
& '§ -5 S ^ g.


















o
NI




?
3

c
Q
ra
S
i
n
c
a
I
j
o
c
(t
a
n
c
a
1
c
o
c
c
I
1
u.
c
V.
re
CO

^










*.
I
to









o
o
CO
CM




"s~
n

S1
I

-------










I

§
1|
tc
-1
I""
,_

9*0.,
f "• 33 c
i 1 8 1
< S 3 |
E < 3 £
Pit
* « OT
°
P *- *"••
Minimum Depth of
Runoff or Storm Retut
Froquency to Treat to
Water Quality
Management (peracn
•s
llUc
If!!
| o "• g
z
•i g o! > j
J2 g i. o —
o < £ 0
r
f
2
^
3
£
£
S



























1
o

2§
81
» ""

S
5
CJ




•t)
1
m
•i
O
f?
1 c » J2 | 1
^|». i-l 1 1 1 1

||| ||§ li||
5 Q o -o S1® 2.>°-c

i If || |1 2 s 1"
|g|^.|gl°-gs;

§-C "o^^^ §"c™
lllllllili
jgisa^m'a0(»<"c;
S g| 2 p-58 S-S-1
QjO^CDCratDJsCOC
il|i|5?ii^
|||liIi|IS
•glm.S^o^giu?^.
o. | •§ $ >- ^ •£ c B 2 §
ra>;cffl"0'05o2i»
«£oo.coEi= — ,^a>
tlllliliill
||§||l|||x8
It II ||!i ill


>•








1


O
CO
1
c
1
i
ss
i
o
CM
§
A



ra
™
I
n
O
j=
1


























1
in





1 217.800 I




c
I
c
C
(
in





d
I
s
ra
s
:ords of chect
5
c
1
1
(0
(D
Q
(D


^

















I 217,800 |





'fl
(I
0
d
C
1

























£
3,600 cubic feet per ac





1 217,800







5
•5.1=5 o«« f
= S 8 S I g f 1 1 1 1 1 1 o

^'»Sn>i»t.»£™S:?'£
g-gS5c3fe.g51ffl = w-|o»
*c ~" Cp {"*"J2 oS3 > o • oj c S ^o e™ >»•—
= ° o *S-o S '^' "o ^- "p ^
2 >»£-tfj ffi!t=3 «"§
OT (D OJ O




1
!



|
1^ '
gl ;
CM





g '
CO






!D
1 -i
































c
(\






c
(
j

Is
w ...
3 -^
5 §
Q- 3
» ^

o '5 %
n*
j_ j2 o
Sll
tesof 10 acre
acres, same
daries of cons
Sediment basins required for si
final stabilization); for sites <10
sideslope and downslope boun


"










£
3,600 cubic feet per ac





o
o
CO_






CO
'E
g
















J3 £ W ^
o ** £* o c*
o J -o ^ S
t« ^~
fe S
ji — o c ^ re
a> o- g d) § w
I**H
•S ro s ro
CM ^

24 hour/ 6 month





1
1
^
^
5

S.
g
|1
3 a.

-------
o
•5
a:

I
55
I
c
z
Visual Inspection
Frequency
Maximum Allowed
Denuded Acreage or
Soil Stablilization
Requirement

°P I

"cL e ~ ®

Q 5 3 e§ ^
3 1- i? 0 S
Illil
= I I i
01 ^ E
I
J5 = E
.= O O
1 2 1 1
| s°=s
Z
^ .ti ^ a>
1 ~a & | "l
5 < £ 0 ^
o
CD
Z
I
.2
£
2
S1
0)
(9





days between October
1 and April 30 (I.e., trie
wet season); 7 days
between May 1 to
September 30 (dry
season)
CM
£
c

E '
CO
S
O
.c







1
/

5
E
(O
5?
£
3)
f I
: £



>





u.
CQ
CM








O
CO
CO
3
T~





C
1
i
3



>
















I
f





C
D
2
3



S ffi "c ^3 ^ "a. S 'S 8 j2
__ « nj s- w .—










2
|
V
i '
1

o
1

-------

-------
 Development Document for Construction and Development Proposed Effluent Guidelines
                               APPENDIX B:
                        SUPPORTING COST DATA
June 2002

-------

-------
 Development Document for Construction and Development Proposed Effluent Guidelines	__^

 APPENDIX B: SUPPORTING COST DATA

 OVERVIEW

 EPA estimated a reference or standard quantity for each costed best management practice (BMP)
 that could be applied to construction erosion and sediment controls (ESCs) (e.g., 621 feet of silt
 fence for a 3-acre single-family residential construction site). These reference quantities were set
 to serve a range of site conditions and slopes consistent with the requirements of the proposed
 rule. Reference quantities were not varied between ecoregions but were varied in response to
 alternative levels of management (i.e., regulatory options explored by EPA) as shown in Table
 B-l. Note that only where values in Table B-l differ between options and baseline values is there
 expected to be a change in the cost for site ESCs.

 Reference quantities of various ESCs (or construction controls) are listed in Tables B-2 through
 B-10, along with unit costing and the assumptions used in EPA's compliance cost assessment.
 Note that for some controls, reference quantities are given in terms of the number of units that
 will be constructed (i.e., the number of construction entrances anticipated for a certain size. site).
 In addition, where unit costs are nonlinear (i.e., the unit cost varies with the size of the unit), both
 a design quantity and a number of units per site size class are required to estimate ESC
 compliance costs. An example of this is sediment basins, where the total volume (the site size in
 acres times 3,600 cubic feet per acre) is apportioned into a number of installations (i.e.,  a 70-acre
 site is estimated to have 2 installations). This process helps ensure that any economies of scale in
 the calculation of compliance costs are reasonable.

 National BMP  costs were determined using the following three equations that relate site size
 class/land use type models to ESC capital, design, and operation and maintenance costs. Note
 that Table B-l 1 contains the regional adjustment factors that customize cost estimates for the
 19 ecoregions defined by EPA to make its analysis more representative  of actual conditions.

 Figure B-l presents a flowchart summarizing the overall costing methodology.
June 2002
                                                                                      B-l

-------
Development Document for Construction and Development Proposed Effluent Guidelines
                                   = RAF*J\\ LM*S*Ni*a\%-
                                              •^"" I    .               I  A/ .•
                                              1=1
                                                                      N,
TRCC - Total Regional Capital (Installation) Cost for a site size class/land use model                   '•
Qi - Quantity of elements required per installation
Nj5* Number of elements required for a single site size class/land use                                  .
S = Estimated number of sites in the site size class/land use
a - Multiplier for converting quantity to national average cost in 2000 dollars
b = Exponent for converting quantity to national average cost in '2000 dollars
RAF = Regional adjustment factor for converting national average costs to region-specific costs
LM - Level of management; values between 0 and 2 that indicate the degree of application of the element. A value
of 1 indicates the full application of an element based on the reference condition

                                           17

                       TRDC = RAF *2>^ *|  LM *s * NI *a\ ^T
TRDC = Total Regional Design Costs
DF{= Design factor, a multiplier which represents the design cost as a percent of capital costs
Qj = Quantity of elements required per installation
Nj - Number of elements required for a single site size class/land use
S- Estimated number of sites in the site size class/land use ;
a - Multiplier for converting quantity to national average cost in 2000 dollars
b = Exponent for converting quantity to national average cost in 2000 dollars
RAF = Regional adjustment factor for converting national average costs to region-specific costs
LM - Level of management; values between 0 and 2 that indicate the degree of application of the element. A value
of 1 indicates the full application of an element based on the reference condition                       ;
                                             17
LM = Level of management; values between 0 and 2 that indicate the degree of application of the element. A value
of 1 indicates the full application of an element based on the reference condition.
TROMC = Total Regional Operation and Maintenance Costs                                         ;
OMj= Operation and Maintenance factor, a multiplier which represents the annual operation and maintenance cost
which ensure proper operation of the element                                                      :
Q. = Quantity of elements required per installation
Nj = Number of elements required for a single site size  class/land use
S m Estimated number of sites in the site size class/land use
a - Multiplier for converting quantity  to national average cost in 2000 dollars                          :
b " Exponent for converting quantity to national average cost in 2000 dollars
RAF = Regional adjustment factor for converting national average costs to region-specific costs
LM = Level of management; values between 0 and 2 that indicate the degree of application of the element. A value
of 1 indicates the full application of an element based on the reference condition.
June 2002
                                                                                                     B-2

-------
  Development Document for Construction and Development Proposed Effluent Guidelines
able B-l. BMP Quantity Adjustment Factors for Baseline and the Pronosed Onlions
Costed Items
Silt Fence
Runoff Diversion
Construction Phasing
Mulch
Seed and Mulch
Construction Entrance
Stone Check Dam
Sediment Trap
Sediment Pond
E&S Certification
E&S Inspection
Polyacrylamide (PAM)
Alum Treatment
Monitoring of Effluent Quality
Baseline
Construction
1.0
1.0
0.0
1.0
0.0
1.0
1.0
1.0
1.0
0.0
0.0
0.0
0.0
0.0
Option 1 — Inspection/
Certification
1.0
1.0
0.0
1.0
0.0
1.0
1.0
1.0
1.0
1.0
1.0
0.0
0.0
0.0
Option 2 — Inspection/
Certification with
Codification of CGP
1.0
1.0
0.0
1.2
0.0
1.0
1.0
1.0
1.1
1.1
1.0
0.2
0.0
0.0

June 2002
                                                                                               B-3

-------
Development Document for Construction and Development Proposed Effluent Guidelines
             Table B-2. Quantities of Erosion and Sediment Control Items For
                  Assessing Compliance Costs for the Construction Industry
Site
size
Acres
1
3
7.5
25
70
200
Feet of Silt Fence
Single-
family
-
621
1,553
5,175
14,490
41,400
Multi-
family
-
722
1,143
3,129
5,238
8,853
Com-
mercial
-
361
600
2,087
3,492
5,902
Indus- '
trial
-
361
600
2,087
3,492
5,902
Feet of Diversion Dike
Single-
family
-
621
1,553
5,175
14,490
41,400
Multi-
family
-
722
1,143
3,129
5,238
8,853
Com-
mercial
-
361
600
2,087
3,492
5,902
Indus-
trial
-
361
600
2,087
3,492
5,902
        Both silt fencing and diversion dike lengths were based on 207 feet per acre on the site.
        Costs for new installation of silt fence are based on $0.92/ft length, excluding profit and overhead (R.S.
        Means, 2000).                                '
        Costs for new installation of diversion ditch are based on $0.55/ft length installation, excluding profit and
        overhead (R.S. Means, 2000).
Tab!
le B-3. Quantities of Erosion and Sediment Control Items
Assessing Compliance Costs for the Construction Industry
Site Size
Acres
1
3
7.5
-25
70
200
Mulched Acreage To Control Erosion
Single-family
0.0
0.8
1.9
6.3
17.5
50.6
Multifamily
6.0
6.8
1.9
6.3
17.5
50-0
Commercial
0.0
0.8
1.9
6.3
17.5
50.0
Industrial
0.0
0.8
1.9
6.3
17.5
50.0
For
        For sites larger man lacre, muicnmg is nmirea to tne sue acreage times nan me peitcmagc ui uiuuuue
        impervious area as a temporary means to stabilize denuded surfaces. The maximum coverage is set to 25%
        of the total site acreage. Cost to mulch is set to $0.20 per square yard for materials/installation without
        overhead and profit (R.S. Means, 2000).
 June 2002
                                                                                                B-4

-------
Development Document for Construction and Development Proposed Effluent Guidelines
             Table B-4. Quantities of Erosion and Sediment Control Items For
                 Assessing Compliance Costs for the Construction Industry
Site Size
Acres
1
3
7.5
-25
70
200
Acres Treated with PAM
Single-family
0.00
0.84
2.10
7.00
•19.60
56.00
Multifamily
0.00
1.32
3.29
10.96
30.70
87.72
Commercial
0.00
1.50
3.75
12.50
35.00
100.00
Industrial
0.00
1.50
3.75
12.50
35.00
100.00
       PAM is costed at $200 per acre per treatment based on a survey of commercial vendors and assuming costs
       are similar to herbicide for soil treatment ($0.04 per square yard without profit and overhead based on
       spraying from truck). The acreage treated is equal to the site size times the ultimate impervious percentage,
       to a maximum of 50% of the site size.
             Table B-5.  Quantities of Erosion and Sediment Control Items For
                 Assessing Compliance Costs for the Construction Industry
The Number of Equal Size Units Installed to Provide Required Protection
Site
Size
Acres
1
3
7.5
-25
70
200
Number of Stone Check Dam
Single-
family
0
0
10
35
50
100
Multi-
family
0
0
10
35
50
100
Com-
mercial
0
0
10
35
50
100
Indus-
trial
0
0
10
35
50
100
Number of Sediment Trap
Single-
family
0
1
1
0
0
0
Multi-
family
0
1
1
0
0
0
Com-
mercial
0
1
1
0
0
0
Indus-
trial
0
1
1
0
0
. 0
       For stone check dam, assume approximately one unit per 5 acres for sites larger than 5 acres at a cost of
       $45.36 per installation, excluding overhead and profit (Phase II Economic Analysis for Phase II Storm
       Water Regulations).,
       Sediment trap of 3,600 cubic feet per acre served at a cost of $0.22 per cubic foot volume (excludes profit
       and overhead).
June 2002
B-5

-------
 Development Document for Construction and Development Proposed Effluent Guidelines
              Table B-6.  Quantities of Erosion and Sediment Control Items For
                  Assessing Compliance Costs for the Construction Industry
Site Size
Acres
1
3
7.5
25
70
200
Number of Sediment Basins
Single-family
0
0
1
2
2
4
Multifaniily
o
0
1
2
2
4
Commercial
0
0
1
2
2.
4
Industrial
0
0
1
2
2
4
        Sediment pond of 3,600 cubic feet per acre served. Cost in dollars is computed from the equation:
        [ 0.76 x 7.47 x (volume required, cubic feet/number of ponds per site size)-78]
        The value of 0.76 removes overhead and profit from cost estimate.                ,          '
             Table B-7. Quantities of Erosion and Sediment Control Items For
                 Assessing Compliance Costs for the Construction Industry
Site Size
Acres
1
3
7.5
25
70
200
Number of Construction Entrances
Single-family
0
1
1
1
2
4
Multifaniily
0
1
1
1
2
4
Commercial
0
1
1
1
2
4
Industrial
0
1
1
1
2
4
       Costs for construction entrance based on $6.92 per square yard (gravel installed) for a footprint covering
       100 square yard, excluding profit and overhead (R.S. Means, 2000).
June 2002
                                                                                         B-6

-------
Development Document for Construction and Development Proposed Effluent Guidelines
             Table B-8. Quantities of Erosion and Sediment Control Items For
                 Assessing Compliance Costs for the Construction Industry
Administrative BMPs for Erosion and Sediment Control
Management
Site Size
Acres
1
3
7.5
25
70
200
E&S Site Inspection
Single-family
0
1
1
2
7
20
Multifamily
0
1
1
2
7
20-
Commercial
0
1
1
2
7
20
Industrial
0
1
1
.2
7^
20
       E&S Site Inspection includes multiple site visits by a certified inspector to verify the proper installation and
       operation of ESC BMPs. Values above are the number of half-day site inspections. Costs are based on 16
       hours of inspection/documentation time per 10-acre unit of a site, at a rate of $28.44 per hour.

             Table B-9. Quantities of Erosion and Sediment Control Items For
                 Assessing Compliance Costs for the Construction Industry
Administrative BMPs for Erosion and Sediment Control
Management
Site Size
Acres
1
3
!
7.5
25
70
200
E&S Site Certification of Sedimentation Basins
Single-family
0
1
1
1
2
4
Multifamily
0
1
1
1
2
4
Commercial
0
1
1
1
2
4
Industrial
0
1
1
1
2
4
       E&S Site Certification includes multiple site visits by a certified inspector to verify the proper installation
       of sedimentation basins. Costs based on 2 hours of inspection/documentation by a licensed engineer per 10-
       acre unit of a site, at a rate of $56.74 per hour.
June 2002
                                                                                            B-7

-------
 Development Document for Construction and Development Proposed Effluent Guidelines
             Table B-10. Quantities of Erosion and Sediment Control Items For
                 Assessing Compliance Costs for the Construction Industry
Site Size
Acres
' 1
3
7.5
25
70
200
Phasing of Construction
Single-family
0
0
0
2
6
19
Multifamily
0
0
o
1 2
6
19
Commercial
0
0
0
2
6
19
Industrial
0
0
0
2
6
19
        For sites larger than 10 acres, the number of remobilizations required is based on a maximum of 10 acres
        denuded at any single time to prevent large unstabilized construction sites. Costs are based on $1,000 per
        remobilization.                                                                  ;
June 2002
B-8

-------
               Table B-ll. Regional Compliance Cost Adjustment Factors
Region or Ecoregion
1
• 2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
Regional Compliance
Cost Adjustment Factors
• 0.85456
0.98351
0.9
0.78103
0.85711
0.85768
0.87
1.03221
0.877
0.99576
6.81034-
0.85357
0.93573
0.9076
1.09438
1.1285
1.05151
1.04609
1.05169
June 2002
                                                                              B-9

-------
Development Document for Construction and Development Proposed Effluent Guidelines
June 2002
                                                                                            B40

-------