Skip to main content

Prograding and Retrograding Hypo- and Hyper-Pycnal Deltaic Formations into Quiescent Ambients

  • Chapter
  • First Online:
Physics of Lakes

Abstract

Sediment transport from mountainous rivers into a quiescent ambient with simultaneous formation of deltas is reviewed. The focus is restricted to flow in vertical cross-sections with no changes perpendicular to the plane of the flow. The bed load transport in the river is derived for quasi-steady situations using sediment mass balance and the Mohr-Terzaghi shear stress-pressure relation with the angle of internal friction, \(\varphi \), as the essential frictional parameter. The emerging model is a diffusion equation for the upper surface of the moving sediment layer and corresponding boundary conditions. Its diffusivity is expressible in terms of the hydraulic discharge, the densities of the sediment and the turbid water, the angle of internal friction and a parameter characterizing the bed-parallel sediment velocity in terms of the average velocity in the turbid layer. When this river flow enters quiescent water, two different classes of deltas can be formed. When the entering water is either neutrally buoyant or lighter than the ambient water, the sudden reduction in tractive force along the bed generates a conspicuous avalanching flow to depth. This leads to steep-sloped foreset deposits with delta fronts inclined by the angle of internal friction. Such so-called Gilbert-type deltas are governed by a jump requirement of the sediment flux across the shore line and the geometry of the receiving basin. When the inflowing discharge is denser than the receiving ambient water, it will dive down as a turbulent under-current. The basal sediment transport in this subaqueous density current is analogous to the subaerial case and again described by a diffusion equation with similarly determined diffusivity. The combined dual subaerial-subaqueous sedimenting process is mathematically very similar to a (generalized) Stefan problem, e.g. the freezing of an ice cover on a lake. We present (mostly analytical) solutions for (1) bedrock-alluvial transitions, (2) overtopping failure of a dam, (3) topset-foreset diffusion processes for hypo- and hyper-pycnal deltas. Laboratory experiments demonstrate the adequacy of the models.

This chapter is a near-reprint of ‘A tutorial on prograding and retrograding hypo- and hyper-pycnal deltaic formations into quiescent ambients’ [25]. Hutter thanks Prof. P. Rutschmann for granting permission of reproduction.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 129.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Hardcover Book
USD 169.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    Paola et al. 1992 [53] cite a large number of references, of which we mention here Soni 1981 [63], Gill 1983 [20] and Zhang and Kahawita 1987 [71] but Culling 1960 [10] remains unmentioned as an early example of derivation of the diffusion equation for erosion problems.

  2. 2.

    Equation (31.1) is used and \([(1-n_0)\rho _s+n_0\rho _w]\simeq 0 \) is assumed.

  3. 3.

    Other books are e.g. Sokolnikoff–Redheffer (1966), [62], Abramowitz and Stegun (1964), [1], Kreyszig (2006), [32].

  4. 4.

    We compute this indefinite integral as follows:

    $$\begin{aligned} \begin{array}{l} \displaystyle {\int _{\varXi }^{A}}\left( {\frac{1}{x^2}}\right) \exp (-x^2)\text {d}{x} = {\int _{\varXi }^{A}}{\frac{\text {d}}{\text {d}x}}\left( -{\frac{1}{x}}\right) \exp (-x^2)\text {d}x\\ = \displaystyle {\int _{\varXi }^{A}}{\frac{\text {d}}{\text {d}x}}\left[ \left( -{\frac{1}{x}} \right) \exp (-x^2)\right] \text {d}x - {\int _{\varXi }^{A}}\left( -{\frac{1}{x}} \right) (-2x)\exp (-x^2)\text {d}x\\ = -{\frac{1}{x}}\exp (-x^2){|_{\varXi }^{A}} - 2 \displaystyle {\int _{\varXi }^{A}}\exp (-x^2)\text {d}x\\ = -{\frac{1}{A}}\exp (-A^2) + {\frac{1}{\varXi }}\exp (-\varXi ^2) + 2\underbrace{{\int _{A}^{\varXi }}\exp (-x^2)\text {d}x}_{{\int _{A}^{0}}+ {\int _{0}^{\varXi }}}\\ = {\frac{\exp (-\varXi ^2}{\varXi })} + \underbrace{2{\int _{0}^{\varXi }}\exp (-x^2)\text {d}x}_{2\sqrt{\pi }\text {erf}(\varXi )} - \underbrace{{\frac{\exp (-A^2)}{A}} - 2{\int _{0}^{A}}\exp (-x^2) \text {d}x}_{C}\\ = {\frac{\exp (-\varXi ^2)}{\varXi }} + 2\sqrt{\pi } \text {erf}(\varXi ) + C , \nonumber \end{array} \end{aligned}$$

    which agrees with (31.25).

  5. 5.

    Had we chosen \(C\) differently from \(-1\), then (31.36)\(_{1}\) would read \(A = S_{2} + \sqrt{\pi }(1 + C)\) and the initial value for \(s\) at \(t =0\) would no longer be zero. Requesting that \(s(0) = 0\) would in this case fix \(C\) to be again \(-1\).

  6. 6.

    \(s^{\pm }(t) = s(t)\pm \varepsilon ,\,\varepsilon >0,\,\varepsilon \rightarrow 0\).

  7. 7.

    An alternative derivation of formula (31.42) when \(\sigma = 0\) is given in Appendix A.

  8. 8.

    From now on \(\phi \) is counted as positive.

  9. 9.

    We regard the denotation ‘equilibrium’ as introduced by geologists as a misnomer, since the graded state is not a thermodynamic equilibrium.

  10. 10.

    With \(\alpha _{1} = \alpha _2\) and \(\rho _1 = 2100\) kg m\(^{-3}\) and \(\rho _{\infty } = 1100\) kg m\(^{-3}\) one obtains \(D_{2}/D_{1} = 0.48\) (foreset conditions). Alternatively, with \(\rho _{1} = 1200\) kg m\(^{-3}\), \(\rho _{\infty } = 1100\) kg m\(^{-3}\), we get \(D_2/D_1 = 0.083\).

  11. 11.

    Note that the primes in these functions designate differentiations with respect to different variables.

  12. 12.

    Lai and Capart [34] choose \(d = [\![D]\!]S^\mathrm{min}\), assuming that \(S_{1}^\mathrm{min}=S_{2}^\mathrm{min}\). They say that this choice may be too restrictive, but it is clear from above that they did not restrict the flexibility of the model by this choice.

  13. 13.

    Similar, somewhat reminiscent alterations of sediment regimes have frequently occurred in the Alps during the Middle Ages when ice avalanches, formed from hanging glaciers, dammed riverine valleys, Roethliberger (1978), [58]. As long as the ice deposit existed, an ice-dammed lake formed and changed the upslope and downslope sediment flows. More significantly, floods due to sudden dam break caused devastating debris flows. Other, related processes are artificially formed when barrages are built for valley reservoirs. They change the upstream sediment regimes and slowly fill the reservoir, thereby reducing the power-generating capacity. Through a base opening in the barrage or side channel and judicious flushing operations, in which the discharge and the lake level are monitored, the sediment deposit is partly removed, a process which affects the sediment regimes in the lake and its topset as well as the sediment flow in the river stretch below the barrage.

  14. 14.

    Unsteady situations can also be analysed, but may need pure numerical solution techniques.

  15. 15.

    \(h_{0}\) is the water depth on the assumption that the water flux leaving the lake is the same as the in-flux.

References

  1. Abramowitz, M. and Stegun, I. A.: Handbook of Mathematical Functions, Dover Publ. Inc, New York, (1964)

    Google Scholar 

  2. Adams, E. W., Schlager, W. and Anselmetti, F. S.: Morphology and curvature of delta slopes in Swiss lakes: Lessons for the interpretation of clinoforms in seismic data. Sedientology, 48, 661–679 (2001)

    Google Scholar 

  3. Begin, Z. B., Meyer, D. F. and Schumm, S. A.: Development of longitudinal profiles of alluvial channels in reponse to base level lowering. Earth Surface Processes & Landforms, 6, 49–68 (1981)

    Google Scholar 

  4. Bell, H. S.: Density currents as agents for transporting sediments. J. Geology, 50, 512–547 (1942)

    Google Scholar 

  5. Bornhold, B. D. and Prior, D. B.: Morphology and sedimentary processes on the subaqeous Noeick River delta, British Columbia, Canada. In: Coarse-Grained Deltas, Spec. Publs. Int. Ass. Sediment., (Collela, A. and Prior, D. B., Eds.) 10, 169–181 (1960)

    Google Scholar 

  6. Capart, H., Bellal, M. and Young, D. L.: Self-similar evolution of semi-infinite alluvial channels with moving boundaries. J. Sedimentary Research, 77, 13–22, (2007)

    Google Scholar 

  7. Capart, H., Hsu, J. P. C., Lai, S. Y. J. and Hsieh, M.-L.: Formation and decay of a tributary-dammed lake, Laonong River, Taiwan, Water Resources Res., doi:10.1029/2010WR009159 (2010)

  8. Carslaw, H. S. and Jaeger, J. C.: Conduction of Heat in Solids. Oxford University Press, Oxford (1959)

    Google Scholar 

  9. Crank, J.: Free and Moving Boundary Problems, Cambridge University Press, Cambridge (1984)

    Google Scholar 

  10. Culling, W. E. H.: Analytical theory of erosion. J. Geology, 68, 336–344

    Google Scholar 

  11. Davis, W.M.: The lakes of California. Calif. J. Mines, 29, 175–236 (1933)

    Google Scholar 

  12. Ellison, T. H. and Turner, J. S.: Turbulent entrainment in stratified flows J. Fluid Mech. 6, 423–448 (1959)

    Google Scholar 

  13. Fan, J. and Morris G.L.: Reservoir sedimentation VI: Delta and density current deposits. J. Hydraul. Eng bf 118(3), 354–369 (1992)

    Google Scholar 

  14. Fleming, P. B. and Jordan, T. E.: A synthetic stratigraphic model of foreland basin development. Geophys, Res., 94, 3851–3866 (1989)

    Google Scholar 

  15. Fracarollo, L. and Capart, H.: Riemann wave description of erosional dam-break flows. J. Fluid Mech., 461, 183–228 (2002)

    Google Scholar 

  16. Galay, V. J., Tutt, D. B. and Kellerhals, R.: The meandering distributary channels of the upper Columbia river. In: River Meandering, edited by Eliott, C. M., pp 113–125, Am. Soc. Civil. Engr., New York (1983)

    Google Scholar 

  17. Garcia, M. H.: Hydraulic jumps in sediment-driven bottom currents. J. Hydraul. Eng., 119, (10), 1094–1117 (1993)

    Google Scholar 

  18. Gilbert, G. K.: Geology of the Henri Mountains 170 pp. U.S. Government Print. Office, Washington, D. C. (1877)

    Google Scholar 

  19. Gilbert, G. K.: Lake Bonneville U.S. Geological Survey Monograph, no. 1, 438 p. (1890)

    Google Scholar 

  20. Gill, M. A.: Diffusion model for degrading channels J. Hydraul. Res., 21, 369–378 (1983)

    Google Scholar 

  21. Graf, W. H.: Hydraulics of Sediment transport, McGraw-GHill, New York (1971)

    Google Scholar 

  22. Grover, N. C. and Howard, C. S.: The passage of turbid water through Lake Mead. Transactions ASCE, Paper No. 1994, 720–732, (1937)

    Google Scholar 

  23. Hinderer, M.: Late quarternary denudation of the Alps, valley and lake fillings and modern river loads. Geodinamica Acta, 14, 231–263 (2001)

    Google Scholar 

  24. Hsu, J. P. C. and Capart, H.: Onset and growth of tributary dammed lakes. Water Resources Research, 44, W11201 (2008)

    Google Scholar 

  25. Hutter, K.: A tutorial on prograding and retrograding hypo- and hyper-pycnal deltaic formations into quiescent ambients. In: Contributions on Sediment Transport. Bericht des Lehrstuhls und der Versuchsanstalt für Wasserbau und Wasserwirtschaft der TU München Nr. 127, Herausgeber: Prof. P. Rutschmann (2013)

    Google Scholar 

  26. Hydon, P. E.: Symmetry methods for differential equations, Cambridge University press, Cambridge

    Google Scholar 

  27. Jordan, T. E. and Fleming, P. B.: From geodynamic models to basin fill, a stratigraphic perspective. In: Quantitative Dynamic Stratigraphy (Ed.: T.A. Cross), 149–163 (1990)

    Google Scholar 

  28. Kenyon, P. M. and Turcotte, D. L.: Morphology of a delta prograding by bulk sediment transport. Geol. Soc. Amer. Bull., 96, 1457–1465 (1985)

    Google Scholar 

  29. Kostic, S. Parker, G. and Marr, J. G.: Role of turbidity currents in setting the foreset slope of clinoforms prograding into standing fresh water. J. Sedimentary Res., 72 (3), 353–362 (2002)

    Google Scholar 

  30. Kostic, S. and Parker, G.: Progradational sand mud deltas in lakes and reservoirs. Part I. Theory and numerical modeling. J. Hydr. Res. 41(2), 127–140 (2003)

    Google Scholar 

  31. Kostic, S. and Parker, G.: Progradational sand-mud deltas in lakes and reservoirs. Part II. Experiment and numerical simulation, J. Hydr. Res. 41(2), 141–152 (2003)

    Google Scholar 

  32. Kreyszig, E.: Advanced Engineering Mathematics, Wiley (2006)

    Google Scholar 

  33. Lai, S. Y. J.: Self-similar delta formation by hyperpycnal flows: theory and experiments. M. S. Thesis, Graduate Institute of Civil Engineering, Natl. Taiwan University, Taiwan, June 2006

    Google Scholar 

  34. Lai, S. Y. J. and Capart, H.: Two-diffusion description of hypopycnal deltas. J. Geophys Res. Earth Surface, 112, art F05005, 1–20 (2007) doi: 10.1029/2006JF00617

  35. Lai, S. Y. J. and Capart, H.: Reservoir infill by hyperpycnal deltas over bedrock. Geophys. Res. Lett., 36 (2009) L08402, doi:10.1029/2008GL037139(2009)

  36. Lai, S. Y. J. and Capart, H.: Response of hyperpycnal deltas to steady rise in base level. 5\(^{th}\) IAHR Symposium on River, Coastal and Estuarine Morphodynamics in press (2011)

    Google Scholar 

  37. Lambert. A.: Turbidity currents from the Rhine River on the bottom of Lake Constance. Wasserwirtschaft, 72(4), 14 (in German), (1982)

    Google Scholar 

  38. Lane, E. W.: The importance of fluvial morphology in hydraulic engineering. Proceedings Am. Soc. Civ. Eng. 81, 745-1-745-17 (1955)

    Google Scholar 

  39. Lee, H.-Y., Lin, Y.-T., Chiu, Y. J. Quantitative estimation of reservoir sedimentation from three typhoon events. J. Hydraul. Eng. 11, 362–370 (2006)

    Google Scholar 

  40. Lorenzo-Trueba, J., Voller, V. R., Muto, T., Kim, W., Paola, C. and Swenson, J. B.: A similarity solution for a dual moving boundary problem associated with a coastal-plain depositional system. J. Fluid Mech., 628, 427–443 (2009)

    Google Scholar 

  41. Mackin, J. H.: Concept of the graded river. Bulletin Geol. Soc. America  59, 463–512 (1948)

    Google Scholar 

  42. Meyer-Peter, E. and Müller, R.: Formulas for bedload transport. 2nd Meeting, Int. Hydraul. Structures Research, Stockholm, 39–64 (1948)

    Google Scholar 

  43. Mitchell, N. C.: Morphologies of kinkpoints in submarine canyons. Geol. Soc. Amer. Bull., 118, 589–605 (2006)

    Google Scholar 

  44. Müller, G.: The new Rhine delta in Lake Constance. In: Deltas in their geologic framework, Shirley, M.L. and Ragsdale, J. E., eds. Houston Geological Society, 107–124 (1966)

    Google Scholar 

  45. Müller, G. and Förstner, U.: General relationship between suspended sediment concentration and water discharge in the Alpenrhein and some other rivers. Nature 217, 244–245 (1966)

    Google Scholar 

  46. Muto, T. Shoreline autoretreat substantiated in flume experiments. J. Sedimentary Res., 71(2), 246–254 (2001)

    Google Scholar 

  47. Muto, T., and Steel, R. J.: Retreat of the front in a prograding delta. Geology, 20, 967–970 (1992)

    Google Scholar 

  48. Muto, T. and Steel, R. J.: Autogenic attainment of large-scale alluvial grad with steady sea-level fall. Implications from flume-tank experiments. Geology, 32(5), 401–404; doi:10.1130/G20269 (2004)

    Google Scholar 

  49. Muto, T. and Swenson, J. B.: Large-scale fluvial grade as a non-equilibrium state in linked depositional systems: Theory and experiment. J. Geophys. Res., 110(F3), art.no F03002 (2005)

    Google Scholar 

  50. Muto, T. and Swenson, J. B.: Autogenic attainment of large-scale alluvial grad with steady sea-level fall. An analog tank-flume experiment, Geology, 34(3), 161–164; doi:10.1130/G21923.1 (2006)

    Google Scholar 

  51. Muto, T., Steel, R. J. and Swenson, J. B.: Autostratigraphy: A framework norm for genetic stratigraphy J. Sedimentary Res., 77, 2–12, doi:10.2110/JSR2007.005 (2007)

    Google Scholar 

  52. Paola, C.: Subsidence and gravel transport in alluvial basins. In: New Perspectives in Basin Analysis (K. L. Kleinsphen and C. Paola, eds), 231–243 (1988)

    Google Scholar 

  53. Paola, G., Heller, P. L. and Angevine, C. L.: The large scale dynamics of grain-size variations in alluvial basins. 1: Theory. Basin Res. 4, 73–90 (1992)

    Google Scholar 

  54. Parker, G.: Basic principles of river hydraulics. Hydraul. Div. ASCE 103, 1077–1087 (1977)

    Google Scholar 

  55. Parker, G., Muto, T.: One-dimensional numerical model of delta response to rising sea-level. In: Proceedings third IAHR Symposium on River, Coastal and Estuarine Morphodynamics Editors: A Sánches-Arcilla and A. Bateman, pp. 558–570. Int Assoc. for Hydraulic Research, Madrid (2003)

    Google Scholar 

  56. Petter, A. L. and Muto, T.: Sustained alluvial aggradation and autogenic detachment of the alluvial river from the shoreline in response to steady fall of relative sea level. J. Sedimentary Res., 78, 98–111; doi: 10.2110/JRS.2008.012 (2008)

    Google Scholar 

  57. Pitman, W. C.: Relationship between eustacy and stratigraphic sequences on passive margins. Geol. Soc. Am. Bull. 89, 1389–1403 (1978)

    Google Scholar 

  58. Roethlisberger, H.: Eislawinen und Ausbrüche von Gletscherseen. Jahrbuch der Schwei- zerischen Naturforschenden Gesellschaft, wissenschaftlicher Teil, 170–211 (1978)

    Google Scholar 

  59. Roth, M. M., Weber, M. and Bezzola, G. R.: Physical modeling of sediment deposits in a river delta: The Alpenrhein Delta in Lake Constance. Proc. XXIX IAHR Congress, Beijing, China, Theme E., 187–194 (2001)

    Google Scholar 

  60. Schumm, S. A.: The Fluvial System, 338 pp. John Wiley, Hoboken, N. J. (1977)

    Google Scholar 

  61. Smith, W. O., Vettere, C. P. and Cummings, G. B.: Comprehensive survey of sedimentation in Lake Mead, 1948–1949. U.S. Geol. Survey Prof. Papier 295, 253 p. (1960)

    Google Scholar 

  62. Sokolnikoff, I. S. and Redheffer, R. M.: Mathematics of Physics and Modern Engineering, McGraw Hill Book Company, New York, etc. (1966)

    Google Scholar 

  63. Soni, J. P.: Unsteady sediment teransport law and prediction of aggradation parameters Water Resour. Res., 17, 33–40 (1981)

    Google Scholar 

  64. Swenson, J. B., Voller, V. R., Paola, C., Parker, G. and Marr, J. G.: Fluvio-deltaic sedimentation: A generalized Stefan problem. Euro J. of Applied Mathematics, 11, 433–452 (2000)

    Google Scholar 

  65. Tonioli, H. and Schultz, J.: Experiments on sediment trap efficiency in reservoirs. Lakes and Reservoirs  10, 13–24 (2005)

    Google Scholar 

  66. Turner, T. H.: Buoyancy effects in fluids. Cambridge University Press, Cambridge UK (1973)

    Google Scholar 

  67. Voller, V. R., Swenson, J. B., Kim, W. and Paola, C.: An enthalpy method for moving boundary problems on the Earth’s surface. Int. J. Numerical Methods for Heat & Fluid Flow, 16(5), 641–654, (2006)

    Google Scholar 

  68. Wright, H. E., Jr., Lease, K. and Johnson, S. Glacier River Warren, Lake Pepin, and the environmental history of southeastern Minnesota. In: Contributions to Quarternary Studies in Minnesota, Minn. Geol. Surv. Rep. Invest. Vol 49, (C.J. Patterson and H.E. Wright, Jr., eds., pp. 131–140, Univ. of Minn. Press, Minneapolis, Minn (1998)

    Google Scholar 

  69. Write, L. D., Wiseman, W. J., Bornhold, B. D., Prior, D. B., Suhayda, J. N., Keller, G. H. Yang, Z.-S. and Fan, J. B.: Marine dispersal and deposition of Yellow River silts by gravity driven underflows. Nature, 332, 629–632 (1988)

    Google Scholar 

  70. Yu, W.-S., Lee, H.-Y. and Hsu, S. M.: Experimental study on delta formation in a reservoir. J. Chin. Inst. Civ. Hydraul. Eng., 12(1), 171–177 (2000) [in Chinese]

    Google Scholar 

  71. Zumberge, J. H.: The lakes of Minnesota, their Origin and Classification Univ. of Minn. Press, Minneapolis, Minn. (1952)

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Kolumban Hutter .

Appendices

Appendix A: Derivation of the Sediment Flux Boundary Condition at the Plunge Point of a Gilbert-Type Delta

In this appendix an explicit derivation of the flux boundary condition (31.42) at the plunge point of a hypopycnal delta will be given. The derivation follows Kostic and Parker [30] but in the notation of this chapter.

Consider Fig. 31.36. With reference to this figure the front surface of the foreset delta can be described as

$$\begin{aligned} \hat{\zeta }(X, t) = \hat{\zeta }_{s} - \tan \phi \left( X - s(t)\right) ,\quad \text {where}\quad \hat{\zeta }_{s} = \zeta (s(t), t). \end{aligned}$$
(31.167)
Fig. 31.36
figure 36

Definition sketch for a Gilbert-type deltaic deposition on a non-erodible basement of slope angle\(\alpha _{1}\). The origin of the \((X, Z)\)-coordinates is at the intersection of the basement, \(Z = b(X)\) and the lake surface at time \(t = 0, Z = Z_{\ell }(0)\), the plunge point is at \([X = s(t), Z = \hat{\zeta }_{s}(t)]\) and the front of the wedge is at \([X = u(t), Z = \hat{\zeta }_{toe}(t)]\). The sediment flow through the plunge point from the topset is \(q_{s}(s(t), t)\) and the conservation of mass of the sediment, expressed in formula (31.169) is explained in the inset. The river water depth at the plunge point is \(h_{s}(t)\) and the level of the lake may vary with time, \(Z = Z_{\ell }(t)\)

If this is evaluated at the toe of the alluvial deposit,

$$\begin{aligned} \hat{\zeta }_{b} = \hat{\zeta }_{s} - \tan \phi \left( u(t) - s(t)\right) \end{aligned}$$
(31.168)

is obtained.

If conservation of mass is formulated for a sediment element as shown in the inset of Fig. 31.36, then one may deduce

$$\begin{aligned} \begin{array}{ll} n_{s}\displaystyle {\frac{\partial \, \hat{\zeta }}{\partial t}}\text {d}X &{}= \bar{q}_{s}(X) - \bar{q}_{s}(X + \text {d}X)\\ &{}\simeq \bar{q}_{s}(X) - \bar{q}_{s}(X) - \displaystyle {\frac{\partial \bar{q}_{s}}{\partial X}}\text {d}X = - \frac{\partial \bar{q}_{s}}{\partial X}\text {d}X , \end{array} \end{aligned}$$

or, since the solid volume fraction is assumed to be constant,

$$\begin{aligned} \frac{\partial \, \hat{\zeta }}{\partial t} = - \frac{\partial (\bar{q}_{s}/n_{s})}{\partial X} = - \frac{\partial q_{s}}{\partial X}. \end{aligned}$$
(31.169)

In the above, \(\bar{q}_{s}\) is the sediment flux at a certain volume fraction, whereas \(q_{s}\) is the corresponding effective flux.

An explicit expression for \(q_{s}\) is obtained, if Eq. (31.169) is integrated from \(X = s^{-}\) to \(X = u(t)\). This integration is composed of an ‘integration’ from \(X = s^{-}(t)\) to \(X = s^{+}(t)\) plus the integration from \(X = s^{+}(t)\) to \(X = u(t)\). Thus,

$$\begin{aligned} q_{s} |^{s^{+}(t)}_{s^{-}(t)} = q_{s}(s^{+}(t)) - q_{s}(s^{-}(t)) = - \int _{s^{-}(t)}^{s^{+}(t)} \frac{\partial \hat{\zeta }}{\partial t}\,\text {d}\,X = 0. \end{aligned}$$

Here, the integral on the far right vanishes because of continuity requirements for \(\hat{\zeta }(\cdot )\). It follows that the flux \(q_{s}\) is continuous across the plunge point. Therefore, we may write

$$\begin{aligned} \begin{array}{l} \displaystyle {\int _{s^{-}(t)}^{u(t)} \frac{\partial q_{s}}{\partial X}}\text {d}X = \underbrace{q_{{s}_{\mid u(t)}}}_{=0} - q_{s^{-}\mid s(t)} = \int _{s^{-}(t)}^{u(t)}\,{\frac{\partial \, \hat{\zeta }}{\partial \,t}}\text {d}\,X \\ \qquad \qquad \qquad = -q_{s\mid s^{-}(t)} = -\displaystyle {\int _{s^{-}(t)}^{u(t)}} {\frac{\partial \hat{\zeta }}{\partial t}}\text {d}X, \end{array} \end{aligned}$$
(31.170)

in which integration can now be restricted to \(X > s(t)\). Moreover, it was assumed that the sediment flux at the toe of the frontal surface of the delta vanishes, which is realistic. With \(\hat{\zeta }\) as given in (31.167) one may write

$$\begin{aligned} \begin{array}{ll} \displaystyle \frac{\partial \,\hat{\zeta }}{\partial \,t} &{}= {\dfrac{\text {d}}{\text {d}\,t}}\hat{\zeta }\left( s^{-}(t), t\right) - {\dfrac{\partial }{\partial \,t}}\left( \tan \phi (X - s(t))\right) \\ &{}= \displaystyle {\frac{\text {d}}{\text {d}\,t}} \hat{\zeta }\left( s^{-}(t), t\right) + \tan \phi {\frac{\text {d}s}{\text {d}\,t}} - \frac{\text {d}(\tan \phi )}{\text {d}\,t}(X - s(t)) \end{array} \end{aligned}$$
(31.171)

in which

$$\begin{aligned} \begin{array}{ll} \displaystyle {\frac{\text {d}}{\text {d}\,t}}\left( \hat{\zeta }(s^{-}(t), t)\right) &{}= {\dfrac{\partial \,\hat{\zeta }}{\partial t}}_{{{\mid _{s(t)}}}} + {\dfrac{\partial \,\hat{\zeta }}{\partial X}}\dot{s}(t) \\ &{}= \displaystyle {\frac{\partial \,\hat{\zeta }}{\partial t}}_{{{\mid _{s(t)}}}} - \tan {\alpha _{1}}_{{{\mid _{s^{-}(t)}}}}\dot{s}(t). \end{array} \end{aligned}$$
(31.172)

Here, \(\tan \alpha _{1}\) is the slope of the sediment bed in the topset of the plunge point. Substituting (31.172) into (31.171) and the resulting expression for \(\partial \,\hat{\zeta }/\partial \,t\) into (31.170) yields

$$\begin{aligned} \begin{array}{l} {q_{s}}_{{{\mid _{s(t)}}}} = \displaystyle {\int _{s(t)}^{u(t)}}\left\{ {\dfrac{\partial \,\hat{\zeta }}{\partial \,t}}_{{ {\mid _{s(t)}}}} - (\tan \alpha _{1} - \tan \phi )\dot{s}(t) -(\tan \phi )^{\cdot }(X - s(t))\right\} \text {d}X \\ = \left\{ {{\dfrac{\partial \,\hat{\zeta }}{\partial \,t}}}_{{{\mid _{s(t)}}}} + (\tan \phi - \tan \alpha _{1}){\dfrac{\text {d} s}{\text {d}\,t}}\right\} \left( u(t) - s(t)\right) - \dfrac{1}{2}(\tan \phi )^{\cdot }(u(t) - s(t))^{2}. \end{array} \end{aligned}$$
(31.173)

The surface point \(\hat{\zeta }_{\mid s(t)}\) is given by the level of the lake surface and the water depth above the sediment as follows: \(\hat{\zeta }_{\mid s(t)} = Z_{\ell }(t) - h_{\mid s(t)}\). Consequently,

$$\begin{aligned} \frac{\partial \, \hat{\zeta }}{\partial t}_{{{\mid _{s(t)}}}} = \frac{\partial }{\partial t}\left( Z_{\ell } - h\right) _{{{\mid _{s(t)}}}} = \dot{Z}_{\ell }(t) - \frac{\partial h}{\partial t}_{{{\mid _{s(t)}}}}. \end{aligned}$$
(31.174)

Substituting this into (31.173) yields a first variant of the final formulae for \(q_{s}\):

$$\begin{aligned} \begin{array}{ll} {q_{s}}_{{{\mid _{s(t)}}}} = \left\{ \left[ \left( \dot{Z}_{\ell } - {\dfrac{\partial \,h}{\partial \,t}}_{{{\mid _{s(t)}}}}\right) \right. \right. &{}\!\!\!\Biggl . + \left( \tan \phi - \tan \alpha _{1}\right) {\dfrac{\text {d} s}{\text {d}\,t}}\Biggr ]\left( u(t) - s(t)\right) \\ &{}\Biggl .-\dfrac{1}{2}\big (\tan \phi \big )^{\cdot }\big (u(t) - s(t)\big )^{2}\Biggr \}. \end{array} \end{aligned}$$
(31.175)

Sometimes it is more convenient to additionally use the trigonometric relation

$$\begin{aligned} \left( u(t) - s(t)\right) = \dfrac{\hat{\zeta } - \hat{\zeta }_{toe}}{\tan \phi }. \end{aligned}$$
(31.176)

We then obtain

$$\begin{aligned} \begin{array}{ll} {q_{s}}_{\mid \,s(t)} = \left\{ \left[ \left( \displaystyle {\frac{\dot{Z}_{\ell }(t) - {\dfrac{\partial \,h}{\partial \,t}}_{{{\mid _{s(t)}}}}}{\tan \phi }}\right) \right. \right. &{} + \Biggl .\left( 1 - {\dfrac{\tan \alpha _{1}}{\tan \phi }}\right) {\dfrac{\text {d} s}{\text {d}\,t}}\Biggr ]\left( \hat{\zeta }_{s} - \hat{\zeta }_{toe}\right) \\ &{} \left. -\displaystyle \frac{1}{2}(\tan \alpha _{1})^{\cdot }{\frac{\left( \hat{\zeta }_{s} -\hat{\zeta }_{toe}\right) ^{2}}{(\tan \phi )^{2}}}\right\} . \end{array} \end{aligned}$$
(31.177)

Even though \(\tan \alpha _1 <(\ll ) \tan \phi \), it is not justified in general, to ignore the expression \(\tan \alpha _{1}/\tan \phi \) in the above formulae. However, it is justified to ignore the term associated with \((\tan \phi )^{\cdot }\). Ignoring also \((\partial \,h/\partial \,t)_{\mid s(t)}\) and \(\dot{Z}_{\ell }(t)\) yields

$$\begin{aligned} q_{s\mid s(t)} = \left( \hat{\zeta }_{s} - \hat{\zeta }_{toe}\right) \frac{\text {d}\,s(t)}{\text {d}\,t}. \end{aligned}$$
(31.178)

Formula (31.42) with \(\sigma = 0\) is obtained from (31.175), if the river water depth is ignored, \((\partial \,h/\partial \,t)_{\mid s(t)}\simeq 0\) and \(\tan \alpha _{1}\) is ignored in comparison to \(\tan \phi \).

Appendix B: Characteristics of Error Functions

In this Appendix we collect a number of properties of mathematical expressions which are connected to the error function. These have been collected and/or derived in Carslaw and Jaeger (1959) [8]

  • Definition of the error function and complementary error function

    $$\begin{aligned} \mathrm{{erf}}(x)&= {\frac{2}{\sqrt{\pi }}}\int _{0}^{x} \exp (-\xi ^2)\mathrm{{d}}\xi , \end{aligned}$$
    (31.179)
    $$\begin{aligned} \mathrm{{erfc}}(x)&= 1 - \mathrm{{erf}}(x) = {\frac{2}{\sqrt{\pi }}}\int _{x}^{\infty } \exp (-\xi ^2)\mathrm{{d}}\xi , \end{aligned}$$
    (31.180)
  • The above definitions imply

    $$\begin{aligned} \begin{array}{lll} \mathrm{{erf}}(0) = 0, \quad &{} \mathrm{{erf}}(\infty ) = 1, \quad &{} \mathrm{{erf}}(-x) = -\mathrm{{erf}}(x),\\ \mathrm{{erfc}}(0) = 1, \quad &{} \mathrm{{erfc}}(\infty ) = 0, \quad &{} \mathrm{{erfc}}(-x) = 2 - \mathrm{{erfc}}(x). \end{array} \end{aligned}$$
    (31.181)
  • Both,

    $$\begin{aligned} \mathrm{{erf}}\left( \frac{x}{2\sqrt{Dt}}\right) \quad \mathrm{{and}}\quad \mathrm{{erfc}}\left( \frac{x}{2\sqrt{Dt}}\right) \end{aligned}$$

    satisfy the diffusion equation

    $$\begin{aligned} \frac{\partial \,f}{\partial \,t} - D \frac{\partial ^{2}f}{\partial \,x^{2}} = 0. \end{aligned}$$
  • The \(n\)th integral complementary error functions are defined as

    $$\begin{aligned} \mathrm{i}^{n}\mathrm{{erfc}}(x)&\qquad = \qquad \int _{x}^{\infty }\mathrm{i}^{n-1}\mathrm{{erfc}}\,\xi \mathrm{{d}}\xi , \quad n = 2, 3, 4, \dots \nonumber \\ \mathrm{i}^{0}\mathrm{{erfc}}(x)&\qquad = \qquad \mathrm{{erfc}}(x), \end{aligned}$$
    (31.182)
    $$\begin{aligned} \mathrm{i}\mathrm{{erfc}}(x)&\qquad = \qquad \int _{x}^{\infty }\mathrm{{erf}}\,\xi \,\mathrm{{d}}\xi \nonumber \\&\mathop {=}\limits ^{\text {integr. by parts}} \xi \mathrm{{erfc}}(\xi )|_{x}^{\infty } - \frac{1}{\sqrt{\pi }}\int _{x}^{\infty }\underbrace{(-2\xi )\exp (-\xi ^2)}_{\frac{\mathrm{{d}}}{\mathrm{{d}}\xi } \left( \exp (-\xi ^2)\right) }\mathrm{{d}}\xi \end{aligned}$$
    (31.183)
    $$\begin{aligned}&\qquad = \qquad - x\,\mathrm{{erfc}}(x) + \frac{\exp \,(-x^2)}{\sqrt{\pi }} . \end{aligned}$$
    (31.184)
  • The function \(\mathrm{{erfc}}(x)\) exhibits the following properties:

    1. (1)
      $$\begin{aligned}&\hat{\zeta }_{s}^{(2)}(x, t) = \frac{A}{D}\sqrt{Dt}\,\mathrm{{ierfc}}\left( \frac{x}{2\sqrt{Dt}}\right) , \quad x \geqslant 0,\nonumber \\ - \qquad&\lim _{x \rightarrow \infty }\hat{\zeta }_{s}^{(2)}(x, t) \rightarrow 0, \nonumber \\ - \qquad&\frac{\partial \,\hat{\zeta }_{s}^{(2)}}{\partial \,x} = \frac{A}{2D}\left( \mathrm{{erf}}\left( \frac{x}{2\sqrt{Dt}}\right) - 1\right) \\ - \qquad&\zeta _{s}^{2}(x, t) \quad {\text {satisfies the diffusion equation}} \nonumber \\&\frac{\partial \,\hat{\zeta }_{s}^{(2)}(x, t)}{\partial \,t} - D \frac{\partial ^{2}\,\hat{\zeta }_{s}^{(2)}(x, t)}{\partial \,x^{2}} = 0,\nonumber \end{aligned}$$
      (31.185)
    2. (2)
      $$\begin{aligned}&\hat{\zeta }_{s}^{(3)}(x, t) = \frac{A}{D}\sqrt{Dt}\,\mathrm{{ierfc}}\left( \frac{|x|}{2\sqrt{Dt}}\right) , \quad x \leqslant 0,\nonumber \\ - \qquad&\lim _{x \rightarrow -\infty } \hat{\zeta }_{s}^{(3)}(x, t) \rightarrow 0 ,\nonumber \\ - \qquad&\frac{\partial \,\hat{\zeta }_{s}^{(3)}(x, t)}{\partial \,x} = \frac{A}{2D}\left( -\mathrm{{erf}}\left( \frac{|x|}{2\sqrt{Dt}}\right) + 1\right) \\ - \qquad&\hat{\zeta }_{s}^{(3)}(x, t) \quad {\text {satisfies the diffusion equation}} \nonumber \\&\frac{\partial \,\hat{\zeta }_{s}^{(3)}(x, t)}{\partial \,t} - D \frac{\partial ^{2}\,\hat{\zeta }_{s}^{(3)}(x, t)}{\partial \,x^{2}} = 0. \nonumber \end{aligned}$$
      (31.186)

Rights and permissions

Reprints and permissions

Copyright information

© 2014 Springer International Publishing Switzerland

About this chapter

Cite this chapter

Hutter, K., Wang, Y., Chubarenko, I.P. (2014). Prograding and Retrograding Hypo- and Hyper-Pycnal Deltaic Formations into Quiescent Ambients. In: Physics of Lakes. Advances in Geophysical and Environmental Mechanics and Mathematics. Springer, Cham. https://doi.org/10.1007/978-3-319-00473-0_31

Download citation

Publish with us

Policies and ethics