Open access
Technical Papers
Jul 11, 2016

Numerical Investigation of Diagonals in Miter Gates: Looking for the Optimum Prestressing

Publication: Journal of Performance of Constructed Facilities
Volume 31, Issue 1

Abstract

This article presents a proposed methodology to acquire optimum prestressing on diagonals in miter gates. The study uses three-dimensional (3D) numerical simulation of miter gates that includes all the geometrical details of the most commonly used miter gates. The simulation is used to attain vertical and lateral displacements at the miter end when the gate is mitered and in the operation for different prestressing forces. The simulation results allow the development of design curves that are used to obtain the optimum prestressing loading. These design curves can also be used to verify that the displacements of a gate are within tolerances for existing levels of prestressing. The analyses are compared with the existing criteria. The comparison shows that the miter gate leaf is not limited to only one prestressing combination, as prescribed by the design criteria, but several prestressing combinations evaluated from the numerical experiments can also provide the adequate torsional capacity. Furthermore, the study shows that the design criteria estimates a higher prestressing that likely causes a reduction in fatigue life in the diagonal connections. The new techniques will allow the prediction of optimum prestressing levels without compromising the torsional capacity.

Introduction

In the last three decades, miter gates (Fig. 1) have experienced a significant amount of distress, the most common being fatigue cracking caused at welded connections. Other problems, such as pintle cracking, buckling plates, out-of-plane distortion, fatigue failure of diagonals, and extensive crack propagation were the result of the deterioration of the design boundary conditions (Riveros and Arredondo 2011; Riveros et al. 2009; Mahmoud et al. 2014; Mahmoud and Riveros 2014). In recent years, there have been a number of failures of diagonals in miter gates (Fig. 2). The authors hypothesize that these diagonal failures are fatigue-induced, driven by the connection details at the diagonal ends, but some evidence also suggests that the current design guidance (Hoffmann 1944; Shermer 1951; Headquarters, U. S. Army Corps of Engineers (USACE) 1994, 2014) results in a much larger prestress in diagonals than may be required. This article presents a three dimensional (3D)-implicit numerical simulation to assess the diagonal prestressing levels, which will provide the beginning for a new design approach. The new techniques will allow the prediction of optimum prestressing levels without compromising the fatigue strength. These results will be compared with the 1944, 1951, 1957, 1994, and 2014 criteria used by agencies responsible for maintenance and operation of miter gates.
Fig. 1. Horizontally framed miter gate (downstream view) (reprinted from Riveros et al. 2014)
Fig. 2. Diagonal failure of end connection bolts (reprinted from Riveros et al. 2009)

Effects of Diagonals on Miter Gates

Lock gates serve different functions, depending on location and condition. The major use of lock gates is to form a damming surface across a lock chamber, but the gates may also be used as guard gates, to fill and empty a lock chamber and to allow ice and debris to pass. A navigation lock requires closure gates at both ends so that the water level in the lock chamber can be varied to coincide with the upper and lower approach channels. Many locks in the United States are equipped with double-leaf miter gates that are used for moderate- and high-lift locks, with a height of 6–24 m and a chamber width of 17–33.5 m (Fig. 3).
Fig. 3. Lock and dam with double-leaf miter gates (image courtesy of the U.S. Army Corps of Engineers)
A miter gate is a very deep cantilever girder with a relatively short span. The skin plate is the web of this girder. If the ordinary formulas for the deflection of a cantilever under shearing and bending stresses are applied, the vertical deflection of the miter gate leaf will be found as only a few tenths of a millimeter. Because the skin plate imparts such an abundant vertical stiffness to the leaf, the stresses in the diagonals are a function of only the torsional forces acting upon the leaf. These forces produce a considerable torsional deflection when the gate is being operated (USAED 1960).
The diagonals must be prestressed to maintain adequate torsional rigidity, typically during operation. The main functions of the diagonals prestressing is to reduce the torsional deflection of the leaf, to minimize the shear stress on the pintle region, to maintain the gate within tolerances regarding the vertical displacement, and to reduce the geometrical torsion (USACE 1994).
The Hoffman (1944), Shermer (1951, 1957), and USACE (1994, 2014) criteria define the shape of the twisted leaf geometrically. Then, the work done by the loads is equated to the internal work realized by the structure. From this, the resistance offered by each diagonal to twisting of the leaf is calculated as a function of the torsional deflection of the structure. The deflection is positive when the top of the leaf moves upstream in relation to the bottom. With the positive deflection, those diagonals that decrease in length are considered positive diagonals. The deflection is negative when the top of the gate moves downstream in relation to the bottom. With the negative deflection, those diagonals that decrease in length are considered negative diagonals (Fig. 4).
Fig. 4. Typical miter gate leaf (downstream elevation)

Diagonal Design Criteria of Miter Gates

Hoffman (1944), Shermer (1951, 1957), and USACE (1994, 2014) assumed that the diagonals carry the vertical shear in a panel similar to the lateral bracing on buildings. The diagonals on a miter gate transfer vertical shear to the supports at the quoin and miter ends (Riveros 1995) and can be seen in Fig. 4. The stresses due to the dead load are assumed to be resisted by the truss formed by the top and bottom horizontal girders, the vertical quoin and miter post, and the skin plate and diagonal system. They also assumed that diagonal tensile stresses are divided equally between the skin plate and the diagonals on the opposite side of the gate. Fig. 4 shows a typical miter gate leaf reflecting the 1944, 1951, 1957, 1994, and 2014 design provisions.
The process described by the 1944, 1951, 1957, 1994, and 2014 criteria to design the diagonals first defines the stiffness of the leaf in deforming the diagonal (A). This value is conservative because it is taken as the sum of the average cross-sectional areas of the two-end diaphragms and the top and bottom girders, which bound a panel; this value must be increased by 12.5% for the heavier, welded, and horizontally framed leaves. For this example, A is 287cm2. Secondly, the elasticity constant Qo, which is a measure of the torsional stiffness of a leaf without diagonals, is determined with a constant obtained from a limited amount of experiments conducted in 1944. Qo is 5,609kN·m for the present example. Thirdly, the load torque area (Tz), which is the product of an applied load (F); the distance (z) from the pintle to the applied load; and the moment arm (r) of the applied load with respect to the center of moments (located at the operating strut elevation) (Figs. 4 and 5) are determined. Table 1 shows the load torque areas with the corresponding distances for this example. The loads considered are the dead load, operating load, hydrodynamic load, and the temporal load that represent differential heads, wind waves, ship waves, and propeller wash. Fourthly, the ratio of change in length of diagonal to the deflection of the leaf when diagonal offers no resistance (Ro) is obtained. Ro is positive for positive diagonals and negative for negative diagonals. For the present example, Ro=±0.1238. The summation of the maximum positive and negative load torque areas (ΣTz) will provide the total area required in the set of negative and positive diagonals (Ap and An) for the predefined load combinations. For this example, Ap=483cm2 and An=419cm2. The ratios of the actual change in length of the diagonal to the prestress deflection of the leaf (R) are Rp=0.00461 and Rn=0.0503.
Fig. 5. Plan and elevation views of the gate leaf
Table 1. Load Torques Areas
Type of loadingForce (kN)Moment arm r (m)z (m)Tz(kNm2)
Dead load, D2,415.831.198.8725,499.81
Operating load, Q1,112.0619.4913.72±297,367.95
Hydrodynamic, Hd369.3417.458.87±57,166.99
Temporal load, Ht1,355.379.208.87±110,603.61
The hydrodynamic load (Hd) is the inertial resistance to water while a leaf is operated. From tests performed to determine operating machinery design loads, the maximum value of the hydrodynamic load was found to be equivalent to a resistance of 1.44 kPa on the submerged portion of the gate leaf.
Once A, Q0, Tz, R0, Ap, An, Rp, Rn, Qn, and Qp are calculated, the elasticity constant of the diagonal Q is defined. The elasticity constant of the diagonal (Q) is a measure of the resistance in which a diagonal member offers to the torsional deflection of a leaf. Q for this example is Qp=8.30×103kN·m and Qn=7.85×103kN·m. Then, the maximum positive and maximum negative deflections of the leaf are calculated as Δopening=183mm and Δclosing=183mm. The dead load will cause no torsional deflection because the gate leaf will be adjusted so that it has no vertical deformation (D=0) under dead load. If the gate leaf is to hang plumb under torsional action, the following equation is to be satisfied:
Σ(Tz)D+Σ(QD)=0
(1)
where (Tz)D = dead load torque area; Q = elasticity constant of a diagonal; and D = prestress deflection for a diagonal.
The values of D must be selected so that they follow the statement specified in Eq. (1). For a negative diagonal, let D=201mm, then, QD=1.57×103kN·m2. The dead load torque area is 2.54×103kN·m2. In order to satisfy Eq. (1), the quantity QD for the positive diagonal D must equal 220 mm to guarantee that the leaf will not have vertical deformation under dead load. Prestressed deflections and stress in diagonals follow Eq. (2). The required stresses in the diagonals during operation are described in Tables 2 and 3
s=RE(DΔ)L
(2)
where s = unit stress in diagonal; R = ratio of actual change length of diagonal to deflection of leaf when diagonal offers no resistance; E = modulus of elasticity; D = prestress deflection for a diagonal; Δ = total torsional deflection of the leaf measured at the miter end, by the movement of the top girder relative to the bottom girder; and L = length of diagonal.
Table 2. Diagonal Design Parameters
ParameterPositive diagonalNegative diagonal
R0.04610.0503
Q(m·kN)830,1786,785,193
Minimum numerical value of D (mm)183183
Maximum numerical value of ΔD (mm)434395
Maximum numerical value of D (mm)250214
D (selected value) (mm)221201
QD (kN·m^2)183,452157,950
Table 3. Stresses in Diagonals during Operation
OperationStress (MPa)Stress (MPa)
Stationary Δ=0mm (0 in.)8787
Opening Δ=183mm (7.22 in.)15166
Closing Δ=183mm (7.22in.)1608
The design criteria only takes into account the top horizontal girder, bottom horizontal girder, the vertical quoin, and miter posts, the skin plate, and diagonal system contributing to the rotational stiffness of the gate require a higher prestressing. The proposed numerical approach will consider all of the miter gate elements contributing to the torsional stiffness of the gate.

Numerical Model Validation

Verification efforts were focused on the overall model methodology. The analyses were carried out as a quasistatic analysis using plate and shell finite elements. The numerical methodology used in this study was verified through a benchmark problem, such as the miter gates at Lock and Dam 27, which were modeled using the same techniques and have been validated with experimental data.

Lock and Dam 27 Miter Gate

Lock and Dam 27 (LD27) is located on the Mississippi River Chain of Rocks Canal in Granite City, IL, approximately 16 km from St. Louis, MO. Lock and Dam 27 was originally constructed between 1947 and 1953. It consists of two locks at the site: the main 366 m chamber and an auxiliary 183 m chamber (Fig. 6).
Fig. 6. Lock and Dam 27. (image courtesy of the U.S. Army Corps of Engineers)
The miter gate at the downstream end of the main chamber has been in service since 1953 and was approaching the end of its design life. In August 1996, several alterations were designed for the gate and were constructed the following year. From 1996 to 2012, lock closures were frequently occurring for emergency repairs to the existing miter gate. The principal repairs were related to girder flange fatigue cracking and failure of the new diagonals (1996 repairs). Because of the severity of the issues, it was decided that replacement was the more economical and advantageous option than extensive repairs to this aging gate. The replacement was completed in 2013. The LD27 miter gate is a typical horizontally framed miter gate with a flat skin plate, flexible diagonal braces, and 13 horizontal girders [Fig. 7(a)].
Fig. 7. (a) Lock and Dam 27 geometrical model; (b) Lock and Dam 27 finite-element mesh
The numerical model recreates the true in-service condition of the gate. The finite-element mesh shown in Fig. 7(b) uses a shell element with a quadratic formulation and six degrees of freedom per node. The total number of degrees of freedom for this model is 1.8 million. The displacement boundary conditions restrained the miter block, quoin block, and gudgeon pin in the 1-1 and 2-2 direction, whereas the pintle was restrained in the three principal axes. The loads boundary conditions were gravity, bolt load, hydrostatic pressure, and the diagonal prestressing.
One hundred and five channels of instruments were used to study the long-term performance of the gate and to understand the different modes of deterioration that have been observed in the existing gates. A diagram showing the locations of the instrumentations can be seen in Figs. 8(a–c). Comparison of the results for the finite-element analyses and the field experiments show excellent correlation with the experiments for the horizontal gauges at the tapered end section on Girders 9 [Fig. 9(a)] and 11 [Fig. 9(b)].
Fig. 8. (a) Location of strain gauges on downstream side of miter gate; (b) location of strain gauges on horizontal girders; (c) location of strain gauges on quoin and miter tapered end sections
Fig. 9. (a) Model calibration with horizontal strain gauges at the tapered end section for Girder 9; (b) model calibration with horizontal strain gauges at the tapered end section for Girder 11
The instrumentation is based on the ERDC Smart Gate data acquisition system (Version 1.0). The system uses the Campbell Scientific, Inc. (CSI), Model CR1000 data logger. Each 8-channel CR1000 can accommodate up to 8 CSI AM16/32B multiplexers, thus allowing up to 128 total channels per logger. In this case, each gate leaf has two data loggers on each leaf, connected to 14 multiplexers that operate continually at a rate of 10 seconds per sample (6 samples per minute) per second. In addition to the gate sensors, other data were also collected from water level sensors, lock valves, and gate controls to provide a full picture of the lock operations being conducted, such as if the lock is filling or emptying or when the gate is in recess, miter, or moving.

Numerical Investigation of Diagonals Prestress

A series of analyses were carried out using the 3D-finite-element model of a miter gate shown in Fig. 10. The model consisted of 3D shell elements with six degrees of freedom per node. There are two main numerical experiments that will be discussed. The first is gravity with positive and negative diagonals (Case 1). The second is gravity with positive and negative diagonals, hydrostatic resistance pressure, and operation of the gate (Case 2). For both experiments, different prestressing diagonals were evaluated.
Fig. 10. Finite-element model
Displacement and loads boundary conditions: During the filling and lowering of the chamber pool, the gate is restrained in the 1, 2, and 3 directions in the pintle and in the 1 and 2 directions in the gudgeon pin, quoin block, and miter block (Fig. 10). The dead load, diagonals prestressing, and hydrostatic pressure induced by the upper and lower pool define the load boundary conditions. For this analysis, the upper pool is steadily lowered until it reaches hydraulic equilibrium. During gate operation, the displacement boundary conditions are located in the pintle and gudgeon pin. The load boundary conditions for the gate operation are the self-weight of the gate, prestressing diagonals, and water resistance pressure of 1.4 kPa (USACE 1994).

Results and Discussion

Gravity with Positive and Negative Diagonals

Case 1 includes the gravity loads and the positive and negative diagonals. The selected prestressing values were 0, 34, 69, 103, 138, 172, and 207 MPa. Fig. 11 shows a schematic view of the structure for this case.
Fig. 11. Schematic drawing of Case 1

Vertical Displacement

The analysis confirmed that the structure had a negative vertical displacement caused by the self-weight of the miter gate. Diagonals assist in the reduction of deflection and increase of the torsional stiffness; the greater the prestressing of positive diagonals, the smaller the vertical deflection. However, the negative diagonals had the inverse effect; the greater the prestressing of the negative diagonals, the greater the vertical displacement. Fig. 12 presents the variation in vertical displacement of the miter gate end for the different combinations of diagonal prestresses. Each series corresponds to a positive diagonal prestress, and the vertical axis represents the negative diagonal prestress. The horizontal axis represents the vertical displacement of the miter end. In Fig. 12, the negative diagonal prestress required to obtain zero vertical displacement is the intersection between the lines with the vertical axis. For example, for 138 MPa of prestressing in the positive diagonal, 76 MPa of prestressing in the negative diagonal is required to achieve zero vertical displacement in the miter end.
Fig. 12. Vertical displacement variation for the different diagonal prestressing

Lateral Displacement

The self-weight of the miter gate produced a downstream movement of the top portion of the miter end. The positive diagonal generated an upstream movement of the top portion of the miter end as the prestressing increased, whereas the bottom portion of the miter gate moved downstream.
Fig. 13 presents top (TD) and bottom (BD) lateral displacements for all analyses conducted. It shows the required positive and negative diagonal prestressing necessary to obtain zero relative displacement between the top and bottom ends and was given by the point of intersection of the series representing top and bottom displacements for the same positive diagonal prestressing. This point of intersection was associated with a downstream displacement. This displacement occurred because the diagonals were located in the downstream side of the gate, and an increase in the prestressing of the diagonals produced a small rotation of the gate in the downstream direction. This rotation was less than 25 mm for all the prestressing combinations. An approximation of the top and bottom displacements can also be obtained for different combinations of diagonal prestressing.
Fig. 13. Lateral displacement variation for the different diagonal prestressing

Gravity with Positive and Negative Diagonals, Operation, and Water Resistance

Case 2 was divided into two steps, opening the gate and closing the gate. EM 1110-2-2703 (USACE 1994), EM 1110-2-2105 (USACE 1993), and ETL 1110-2-584, (USACE 2014) suggest water-resistance pressures should be either 1.4 or 2.2 kPa, depending on gate geometrical properties, such as height and length. For gates similar to the one analyzed in this study, 1.4 kPa of resistance pressure was used. The water resistance pressure produced downstream movement of the bottom portion of the gate miter end, whereas the upper portion moved in the upstream direction because of the operational load. In this study, the operating load magnitude was 1.11 MN. Fig. 14 shows a schematic of the gate for both steps, opening and closing.
Fig. 14. Schematic drawing of Case 2

Vertical Displacement

The results revealed an upward movement of the miter end during opening and a downward movement of the miter end during closing. These values were less than 15 mm, negligible in comparison with the gate dimensions. Table 4 and Fig. 15 present the vertical displacement results for the different diagonal prestressing combinations when the gate was opening and closing. The best combination when the gate was opening (.09mm) was with 69 and 138 MPa of positive and negative diagonals prestressing, respectively. However, 207 and 69 MPa of positive and negative prestressing, respectively, achieved a vertical displacement of 2.35mm when the gate was closing.
Table 4. Vertical Displacement Results (mm)
Gate positionNegative diagonal (MPa)Positive diagonal
0 (MPa)69 (MPa)138 (MPa)207 (MPa)
Opening06.256.7410.3914.36
690.093.207.1311.13
1384.020.093.907.89
2077.323.350.644.66
Closing018.1710.973.780.85
6917.8310.616.372.35
13817.6213.599.605.61
20720.8216.8612.868.84
Fig. 15. Vertical displacements when gate is opening and closing

Lateral Displacement

The critical lateral displacements are summarized in Table 5, when the gate is opening, and in Table 5, when the gate is closing. The maximum values occur immediately after the operation load is applied.
Table 5. Lateral Displacement Results for the Opening Cases (mm)
RegionNegative diagonalPositive diagonal
069138207
Top end032.9435.4852.3871.01
693.6019.0537.6156.36
13814.394.2222.8741.69
20729.4310.778.1426.98
Bottom end020.1921.7532.1043.52
692.2011.6723.0534.54
1388.822.5814.0225.55
20718.046.604.9916.54
Fig. 16 and Tables 5 and 6 summarize all the lateral displacements for the positive and negative diagonal combinations. In order to obtain the optimal diagonal prestressing combination when the gate opens and closes, a maximum lateral displacement limit needs to be defined (tolerance). For example, if the tolerance is 25 mm, then based on Fig. 16, 69 and 152 MPa of positive and negative diagonal prestressing is needed when the gate is opening to reach zero relative displacement. Similarly, when the gate is closing, 207 and 83 MPa of positive and negative diagonal prestressing, respectively, will produce no relative displacement. Tables 5 and 6 provide the opportunity to define different levels of prestressing when a relative displacement is allowed. If 69 and 138 MPa of positive and negative diagonal prestressing are applied when the gate is opening, 6.86 mm of relative displacement will be developed. Similarly, 138 and 0 MPa of positive and negative diagonal prestressing will develop a relative displacement of 1.62 mm.
Fig. 16. Lateral displacements with different diagonal prestressing combination when the gate is opening and closing
Table 6. Lateral Displacement Results for the Closing Cases (mm)
RegionNegative diagonal (MPa)Positive diagonal
0 (MPa)69 (MPa)138 (MPa)207 (MPa)
Top end069.1737.114.7218.17
6968.8235.6515.824.27
13868.1649.1830.3811.69
20782.6763.8945.1126.29
Bottom end042.3922.742.8911.14
6942.1821.859.702.62
13841.7730.1518.627.17
20750.6739.1627.6516.11

Conclusions

The comparison of the existing criteria developed in 1944, 1951, 1957, 1994, and 2014 to the numerical experiments shows that the numerical investigation can be used to obtain the optimum prestressing for miter gate diagonals. The comprehensive miter gate diagonals prestress analysis demonstrated the importance of maintaining the adequate levels of prestress. Opening and closing the gate produced an additional lateral displacement. Relaxation will develop over time, and vertical and lateral displacement on the miter end will increase, causing misalignment and additional torsional stresses on the pintle region. The magnitude of the misalignments can be obtained graphically from the developed curves.
Optimum values of prestressing can easily be calculated using numerical simulations. Once the different diagonal prestressing models are obtained, a series of curves relating prestressing levels with relative displacements can be accomplished. Then, for a predefined tolerance, the optimum prestressing levels can be easily extracted.
When the gate was closing, it required 161 and 8.0 MPa of positive and negative prestressing. This displacement was relatively large compared to the displacements from the numerical investigation.
The 1944, 1951, 1957, 1994, and 2014 criteria limited its design procedure of the amount of elements that can contribute to stiffening of the gate. In comparing the 1944, 1951, 1957, 1994, and 2014 criteria to the numerical simulation, the authors found that there was no one set of prestressing combinations for the positive and negative diagonals.
Finally, the numerical solution provides a reduction of 14 and 17% for the positive and negative diagonal prestressing. This indicates that the current design guidance results in larger prestress in miter gate diagonals than is required. Therefore, if this new approach to obtain the optimum diagonal prestressing is adopted, miter gates’ diagonals and their connections should have a better performance.

References

Hoffman, E. G. (1944). “Torsion deflection of miter-type lock gates and diagonal and design of the diagonals.” M.S. thesis, Dept. of Applied Mathematics, Washington Univ., Saint Louis, MO.
HQUSACE (Headquarters, U.S. Army Corps of Engineers). (1993). “Design of hydraulic steel structures.” EM 1110-2-2105, Washington, DC.
HQUSACE (Headquarters, U.S. Army Corps of Engineers). (1994). “Lock gates and operating equipment.” EM 1110-2-2703, Washington, DC.
HQUSACE (Headquarters, U.S. Army Corps of Engineers). (2014). “Design of hydraulic steel structures.” ETL 1110-2-584, Washington, DC.
Mahmoud, H., Como, A., and Riveros, G. A. (2014). “Fatigue assessment of underwater CFRP-repaired steel panels using finite element analysis.”, U.S. Army Engineer Research and Development Center (ERDC), Vicksburg, MS.
Mahmoud, H., and Riveros, G. A. (2014). “Fatigue reliability of a single stiffened ship hull panel.” Eng. Struct., 66, 89–99.
Riveros, G. A. (1995). “User guide: Computer program for the design and investigation of horizontally framed miter gates using the load and resistance factor design criteria.”, U.S. Army Engineer Research and Development Center (ERDC), Vicksburg, MS.
Riveros, G. A., et al. (2014). “A procedure for predicting the deterioration of steel hydraulic structures to enhance their maintenance, management, and rehabilitation.”, U.S. Army Engineer Research and Development Center (ERDC), Vicksburg, MS.
Riveros, G. A., and Arredondo, E. (2011). “Predicting deterioration of navigation steel structures with Markov Chain and lation hybercube simulation.” Rev. Int Nat. Disasters Accidents Civ. Infrastruct., 11(1), 3–15.
Riveros, G. A., Ayala-Burgos, J. L., and Perez, J. (2009). “Numerical investigation of miter gates.”, U.S. Army Engineer Research and Development Center (ERDC), Vicksburg, MS.
Shermer, C. L. (1951). “Torsion in lock gates and prestressing of diagonals.” Ph.D. dissertation, Univ. of Michigan, Ann Arbor, MI.
Shermer, C. L. (1957). “Stiffening lock gates by prestressing diagonals.” Transactions, 122(1), 106–114.
U.S. AED (U.S. Army Engineer District). (1960). “Torsional deflection of miter type lock gates and design of the diagonals.” U.S. Army Engineer District, Chicago.

Information & Authors

Information

Published In

Go to Journal of Performance of Constructed Facilities
Journal of Performance of Constructed Facilities
Volume 31Issue 1February 2017

History

Received: May 27, 2015
Accepted: Feb 1, 2016
Published online: Jul 11, 2016
Discussion open until: Dec 11, 2016
Published in print: Feb 1, 2017

Authors

Affiliations

Guillermo A. Riveros, Ph.D., M.ASCE [email protected]
P.E.
U.S. Army Engineer Research and Development Center, Vicksburg, MS 39180 (corresponding author). E-mail: [email protected]
Jorge L. Ayala-Burgos [email protected]
U.S. Army Corps of Engineers, Jacksonville District, Ft. Myer, FL 33919. E-mail: [email protected]
DeAnna Dixon [email protected]
U.S. Army Engineer Research and Development Center, Vicksburg, MS 39180. E-mail: [email protected]

Metrics & Citations

Metrics

Citations

Download citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click Download.

Cited by

View Options

Media

Figures

Other

Tables

Share

Share

Copy the content Link

Share with email

Email a colleague

Share