9
RESEARCH ARTICLE Testing a theoretical resistance law for overland flow on a stony hillslope Alessio Nicosia 1 | Costanza Di Stefano 1 | Vincenzo Pampalone 1 | Vincenzo Palmeri 2 | Vito Ferro 2 | Mark A. Nearing 3 1 Department of Agricultural, Food and Forest Sciences, University of Palermo, Palermo, Italy 2 Department of Earth and Marine Science, University of Palermo, Palermo, Italy 3 Southwest Watershed Research Center, USDA-Agricultural Research Service, Tucson, AZ, USA Correspondence Ferro Vito, Department of Earth and Marine Science, University of Palermo, Via Archirafi 20, 90123 Palermo, Italy. Email: [email protected] Abstract Overland flow, sediments, and nutrients transported in runoff are important pro- cesses involved in soil erosion and water pollution. Modelling transport of sediments and chemicals requires accurate estimates of hydraulic resistance, which is one of the key variables characterizing runoff water depth and velocity. In this paper, a new the- oretical powervelocity profile, originally deduced neglecting the impact effect of rainfall, was initially modified for taking into account the effect of rainfall intensity. Then a theoretical flow resistance law was obtained by integration of the new flow velocity distribution. This flow resistance law was tested using field measurements by Nearing for the condition of overland flow under simulated rainfall. Measurements of the DarcyWeisbach friction factor, corresponding to flow Reynolds number rang- ing from 48 to 194, were obtained for simulated rainfall with two different rainfall intensity values (59 and 178 mm hr -1 ). The database, including measurements of flow velocity, water depth, cross-sectional area, wetted perimeter, and bed slope, allowed for calibration of the relationship between the velocity profile parameter Γ, the slope steepness s, and the flow Froude number F, taking also into account the influence of rainfall intensity i. Results yielded the following conclusions: (a) The pro- posed theoretical flow resistance equation accurately estimated the DarcyWeisbach friction factor for overland flow under simulated rainfall, (b) the flow resistance increased with rainfall intensity for laminar overland flow, and (c) the mean flow velocity was quasi-independent of the slope gradient. KEYWORDS dimensional analysis, flow resistance, overland flow, rainfall, self-similarity, stony hillslopes, velocity profile 1 | INTRODUCTION Erosion by rainfall and runoff causes detachment, transport, and deposition of soil on hillslopes and is a major driver in landscape evolution. Water erosion processes are generally characterized as either interrill or rill erosion, which behave differently. On interrill areas, the main erosion process is due to raindrop impact, and the sediment is transported by flow from these areas to rills. Soil particle detachment in interrill areas is due to rainsplash, whereas the transport of detached sediments to rills is due to overland flow, characterized by shallow water depth and bed shear stress (Mutchler & Young, 1975; Toy, Foster, & Renard, 2002). Research on this topic aims to improve the understanding of overland flow and sediment transport under different conditions of morphology, soil type, vegetation cover, and rainfall regime (Dunkerley, 2018; Smith, Cox, & Bracken, 2007). Received: 26 October 2019 Accepted: 15 January 2020 DOI: 10.1002/hyp.13709 Hydrological Processes. 2020;19. wileyonlinelibrary.com/journal/hyp © 2020 John Wiley & Sons Ltd 1

Testing a theoretical resistance law for overland flow on a ......Ferro Vito, Department of Earth and Marine Science, University of Palermo, Via Archirafi 20, 90123 Palermo, Italy

  • Upload
    others

  • View
    5

  • Download
    0

Embed Size (px)

Citation preview

  • R E S E A R CH A R T I C L E

    Testing a theoretical resistance law for overland flowon a stony hillslope

    Alessio Nicosia1 | Costanza Di Stefano1 | Vincenzo Pampalone1 |

    Vincenzo Palmeri2 | Vito Ferro2 | Mark A. Nearing3

    1Department of Agricultural, Food and Forest

    Sciences, University of Palermo, Palermo, Italy

    2Department of Earth and Marine Science,

    University of Palermo, Palermo, Italy

    3Southwest Watershed Research Center,

    USDA-Agricultural Research Service, Tucson,

    AZ, USA

    Correspondence

    Ferro Vito, Department of Earth and Marine

    Science, University of Palermo, Via Archirafi

    20, 90123 Palermo, Italy.

    Email: [email protected]

    Abstract

    Overland flow, sediments, and nutrients transported in runoff are important pro-

    cesses involved in soil erosion and water pollution. Modelling transport of sediments

    and chemicals requires accurate estimates of hydraulic resistance, which is one of the

    key variables characterizing runoff water depth and velocity. In this paper, a new the-

    oretical power–velocity profile, originally deduced neglecting the impact effect of

    rainfall, was initially modified for taking into account the effect of rainfall intensity.

    Then a theoretical flow resistance law was obtained by integration of the new flow

    velocity distribution. This flow resistance law was tested using field measurements

    by Nearing for the condition of overland flow under simulated rainfall. Measurements

    of the Darcy–Weisbach friction factor, corresponding to flow Reynolds number rang-

    ing from 48 to 194, were obtained for simulated rainfall with two different rainfall

    intensity values (59 and 178 mm hr−1). The database, including measurements of

    flow velocity, water depth, cross-sectional area, wetted perimeter, and bed slope,

    allowed for calibration of the relationship between the velocity profile parameter Γ,

    the slope steepness s, and the flow Froude number F, taking also into account the

    influence of rainfall intensity i. Results yielded the following conclusions: (a) The pro-

    posed theoretical flow resistance equation accurately estimated the Darcy–Weisbach

    friction factor for overland flow under simulated rainfall, (b) the flow resistance

    increased with rainfall intensity for laminar overland flow, and (c) the mean flow

    velocity was quasi-independent of the slope gradient.

    K E YWORD S

    dimensional analysis, flow resistance, overland flow, rainfall, self-similarity, stony hillslopes,

    velocity profile

    1 | INTRODUCTION

    Erosion by rainfall and runoff causes detachment, transport, and

    deposition of soil on hillslopes and is a major driver in landscape

    evolution. Water erosion processes are generally characterized as

    either interrill or rill erosion, which behave differently. On interrill

    areas, the main erosion process is due to raindrop impact, and the

    sediment is transported by flow from these areas to rills. Soil

    particle detachment in interrill areas is due to rainsplash, whereas

    the transport of detached sediments to rills is due to overland flow,

    characterized by shallow water depth and bed shear stress

    (Mutchler & Young, 1975; Toy, Foster, & Renard, 2002). Research

    on this topic aims to improve the understanding of overland flow

    and sediment transport under different conditions of morphology,

    soil type, vegetation cover, and rainfall regime (Dunkerley, 2018;

    Smith, Cox, & Bracken, 2007).

    Received: 26 October 2019 Accepted: 15 January 2020

    DOI: 10.1002/hyp.13709

    Hydrological Processes. 2020;1–9. wileyonlinelibrary.com/journal/hyp © 2020 John Wiley & Sons Ltd 1

    https://orcid.org/0000-0003-0540-8788https://orcid.org/0000-0003-2941-4664https://orcid.org/0000-0002-5195-9209https://orcid.org/0000-0001-6594-9530https://orcid.org/0000-0003-3020-3119mailto:[email protected]://wileyonlinelibrary.com/journal/hyp

  • Flow resistance is fundamental to evaluate overland flow velocity,

    which can be also be used to estimate flow transport capacity (Ferro,

    1998). Overland flow resistance is characterized by two main vari-

    ables, which are the soil roughness and the rainfall intensity.

    Katz, Watts, and Burroughs (1995) suggested that flow resistance

    over a smooth and sloping bed is the result of both the internal fluid

    resistance and the frictional resistance due to the channel boundaries.

    In the case of uniform laminar flow, if the flow shear stress varies line-

    arly with the distance from the bed and the flow velocity is zero at

    the boundary, the local velocity increases parabolically upwards in the

    flow, reaching a maximum value of velocity at the free surface (Yalin,

    1977). For laminar flow conditions, the Darcy–Weisbach friction fac-

    tor f assumes the following expression:

    f =KRe

    , ð1Þ

    in which K is a constant equal to 24 for uniform flow on a smooth

    bed. The overland flow under rainfall may be considered to be as a

    “disturbed laminar flow,” because the impact of raindrops on a flow

    causes both an increase of flow turbulence and a retarding effect on

    mean flow velocity (Emmett, 1970).

    Yoon andWenzel (1971) suggested that, for Re > 2,000, the rainfall

    intensity effects on flow resistance are negligible, whereas for

    Re < 2,000 flow resistance always increases with rainfall intensity. The

    local flow velocity is influenced by rainfall with a retarding effect, which

    is higher near the free surface. This retarding effect is explainable by

    the concept of momentum transfer. The component of momentum car-

    ried by raindrops in the mean flow direction is negligible, if the direction

    of fall of raindrops is more or less normal to the water surface. For this

    reason, a part of mean flow momentum must be transferred to the

    added water in order to accelerate the raindrop mass. The transfer of

    mean flow momentum is not equally distributed, but it is greater near

    the surface, and so, the velocity retardation decreases from the surface

    downward. Taking into account that for given rainfall intensity and drop

    size, the rate of raindrop impact is constant, whereas the mean flow

    momentum increases with the flow Reynolds number; the ratio of

    momentum transfer diminishes with increasing Re values.

    Unfortunately, after the experiments by Yoon and Wenzel (1971),

    few authors have carried out studies on flow resistance regarding

    overland flow with rainfall simulation.

    Shen and Li (1973) investigated an overland flow in a 0.6-m-wide

    and 18.3-m-long flume, with plexiglass walls and a stainless steel bot-

    tom, using a rainfall simulator (with rainfall intensities ranging

    between 190 and 444 mm hr−1). These authors confirmed that the

    Darcy–Weisbach friction factor depends on both the flow Reynolds

    number and the rainfall intensity for Re values less than 2,000.

    Savat (1977) carried out some experimental runs, characterized

    by a single rainfall intensity (60 mm hr−1), for the determination of the

    average depth of a sheet flow using a simple weighing of the entire

    run and the water flowing in it. The experiments demonstrated that

    the effect of rainfall diminishes when the degree of flow turbulence

    or the slope angle increases.

    Laboratory investigations were carried out by Katz et al. (1995) in

    a flume, 4.87 m long and 1 m wide, with a rough bed obtained by

    glued sand grains. Three slopes (4%, 6%, and 8%) and two simulated

    rainfall intensities (40.9 and 115.1 mm hr−1) were used. These authors

    demonstrated that the roughness due to raindrop impact and the

    channel bed friction affect the K coefficient of Equation (1).

    Mügler et al. (2011) carried out an experiment, on a 10 m × 4 m

    rainfall simulation plot, to test four different roughness models with

    high-resolution velocity data. The analysis showed that the best

    results were obtained using a calibrated Manning coefficient n model

    with a depth-dependent roughness formulation (Jain, Kothyari, &

    Ranga Raju, 2004).

    Nearing et al. (2017) carried out experiments applying artificial rain-

    fall (with rainfall intensities of 59 and 178 mm h−1) to 2 m × 6 m plots

    at 5%, 12%, and 20% slope gradients. The experiments aimed to study

    the evolution of the surface roughness on a stony hillslope and to test

    the slope–velocity equilibrium hypothesis. Nearing, Kimoto, Nichols,

    and Ritchie (2005) hypothesized that stony hillslopes in the semi-arid

    environment evolve to a state of “slope–velocity equilibrium.” The

    slope–velocity equilibrium is a state that evolves naturally over time

    due to the interactions between overland flow, erosion processes, and

    bed surface morphology, wherein steeper areas develop a relative

    increase in physical and hydraulic roughness such that flow velocity is a

    unique function of overland flow discharge independent of slope

    gradient.

    Nearing et al. (2017) noticed that there was no relationship between

    the final rock cover of the investigated plots and either the slope gradi-

    ent or the initial rock cover. The final random roughness (at the end of

    the experiment) varied with the slope gradient, with a rougher final sur-

    face resulting on steeper slopes. Measured runoff velocities supported

    the hypothesis that flow velocity is only dependent on unit flow dis-

    charge and, as a consequence, is independent of slope gradient. In other

    words, according to the experimental results by Nearing et al. (2017), the

    steeper slopes evolve to rougher surfaces compared with shallower

    slopes, and this increase of roughness with slope balanced the increase

    of flow velocity with slope resulting in a slope-independent condition.

    The hypothesis of slope–velocity equilibrium implies that the use

    of hydraulic equations, such as Manning and Darcy–Weisbach, in

    hillslope-scale runoff models can be problematic because the friction

    factors vary with both slope and rainfall intensity.

    The slope–velocity equilibrium condition is a result similar to the

    feedback mechanism detected by Govers (1992); Di Stefano, Ferro,

    Palmeri, & Pampalone, 2018; Palmeri, Pampalone, Di Stefano, Nico-

    sia, & Ferro, 2018) for rill flows and highlights the conclusion that if a

    constant roughness coefficient is assumed, then “velocity will be

    over-predicted on steep slopes and under-predicted on shallower

    slopes” (Govers, Takken, & Helming, 2000).

    Despite the fact that the retarding effect of rainfall impact and

    the increase of flow turbulence are recognized effects, few studies

    have been undertaken on the effects of rainfall on hydraulic resis-

    tance (Ferro & Nicosia, 2020).

    The aim of this paper was to modify a theoretical power–velocity

    profile, originally deduced neglecting the effect of rainfall impact,

    2 NICOSIA ET AL.

  • (Di Stefano, Ferro, Palmeri, & Pampalone, 2017; Di Stefano, Ferro,

    Palmeri, Pampalone, & Agnello, 2017; Ferro, 2017, 2018; Ferro &

    Porto, 2018; Palmeri, Pampalone, Di Stefano, Nicosia, & Ferro, 2018),

    taking into account the rainfall intensity effect. Then, a theoretical

    flow resistance law was obtained by integration of the new flow

    velocity distribution. The flow resistance law was tested using field

    measurements by Nearing et al. (2017) for an overland flow with sim-

    ulated rainfall on stony hillslopes. The available measurements

    allowed (a) calibration of the relationship between the velocity profile

    parameter Γ (from the theoretical flow resistance equations discussed

    in detail in the next section), the channel slope, the flow Froude num-

    ber, and a dimensionless group representative of rainfall intensity

    effects (rain Reynolds number); (b) assessment of the influence of

    rainfall on flow resistance; (c) assessment of the applicability of the

    theoretical flow resistance equation for an overland flow disturbed by

    rainfall impact; and (d) verification of the hypothesis of quasi-

    independence between the flow velocity and the slope gradient.

    2 | THE THEORETICAL FLOW RESISTANCEEQUATION

    One of the main tasks of open-channel flow hydraulics is deducing a

    theoretical flow resistance law by integration of a known flow velocity

    distribution in the cross section (Ferro, 1997; Powell, 2014).

    For overland flow, subjected to rainfall, on a stony hillslope, the

    local flow velocity profile v(y) along a given vertical can be expressed

    by the following functional relationship (Barenblatt, 1987, 1993;

    Ferro, 1997, 2017, 2018):

    ϕdvdy

    ,y,h,d,u*,s, i,ρ,μ,g

    � �=0, ð2Þ

    in which ϕ is a functional symbol, y is the distance from the bottom,

    h is water depth, d is a characteristic diameter of the soil particles rep-

    resentative of the grain roughness,u* =ffiffiffiffiffiffiffiffiffig R s

    pis the shear velocity,

    R is the hydraulic radius, s is the slope, i is the rainfall intensity, ρ is the

    water density, μ is the water viscosity, and g is acceleration due to

    gravity.

    According to the Π-theorem (Barenblatt, 1987), Equation (2) can

    be expressed in a dimensionless form as follows:

    Π1 =ϕ1 Π2,Π3,Π4,Π5,Π6,Π7ð Þ, ð3Þ

    where Π1, Π2, Π3, Π4, Π5, Π6, and Π7 are dimensionless groups and ϕ1is a functional symbol.

    Using as dimensionally independent variables y, u*, and μ, the fol-

    lowing expression of the dimensionless groups Π1, Π2, Π3, Π4, Π5, Π6,

    and Π7 can be obtained:

    Π1 =yu*

    dvdy

    , ð4Þ

    Π2 =hy, ð5Þ

    Π3 =dy, ð6Þ

    Π4 = s, ð7Þ

    Π5 =iu*

    , ð8Þ

    Π6 =u* yυk

    , ð9Þ

    Π7 =g y

    u2*, ð10Þ

    in which υk = μ/ρ is the kinematic viscosity.

    According to Barenblatt (1987), “In some cases, it turns out

    to be convenient to choose new similarity parameters—products

    of powers of the similarity parameters obtained in the previous

    step.” In other words, Barenblatt (1987) suggested combining

    the original dimensionless groups to obtain new similarity

    parameters.

    From Equation (5) to (6) it follows:

    Π2,3 =Π2Π3

    =hyyd=hd: ð11Þ

    Coupling Equations (8), (5), and (9), the following dimensionless

    group is obtained:

    Π5,2,6 =Π5Π2Π6 =iu*

    hyu* yυk

    =i hυk

    =Rei, ð12Þ

    which can be referred to as rain Reynolds number.

    Coupling Equations (9), (6), and (11), the following equation is

    obtained:

    Π6,3,2 =Π6Π3Π2,3 =u* yυk

    dyhd=u* hυk

    , ð13Þ

    whereas from Equations (10) and (5), the following dimensionless

    group is obtained:

    Π7,2 =ffiffiffi8f

    r1

    Π1=27 Π1=22

    =

    ffiffiffi8f

    ru*

    g1=2 y1=2y1=2

    h1=2=

    ffiffiffiffiffiffiffi8 u2*f

    qg1=2 h1=2

    =Vffiffiffiffiffiffig h

    p = F, ð14Þ

    in which V is the mean flow velocity and F is the flow Froude number.

    The functional relationship (3) can be rewritten in the follow-

    ing form:

    NICOSIA ET AL. 3

  • Π1 =ϕ2 Π2,3,Π4,Π5,2,6,Π6,3,2,Π7,2ð Þ, ð15Þ

    where ϕ2 is a functional symbol.

    Introducing into Equation (15) the expression of each dimension-

    less group, this functional relationship can be rewritten as

    yu*

    dvdy

    =ϕ2hd,s,

    i hυk,u* yυk

    ,u* hυk

    ,F

    � �: ð16Þ

    Assuming Incomplete Self-Similarity in u*y/νk (Barenblatt &

    Monin, 1979; Barenblatt & Prostokishin, 1993; Butera, Ridolfi, &

    Sordo, 1993; Ferro, 2017; Ferro & Pecoraro, 2000), Equation (16)

    yields

    1u*

    dvdy

    =1yu* yυk

    υku* y

    � �u* yυk

    � �δϕ3

    u* hυk

    ,hd,s,Rei,F

    � �, ð17Þ

    in which ϕ3 is a functional symbol.

    Rearranging Equation (17) results in

    1u*

    dvdy

    =u*υk

    � �u* yυk

    � �δ−1ϕ3

    u* hυk

    ,hd,s,Re,i,F

    � �: ð18Þ

    Integrating Equation (18), and assuming equal to zero the integra-

    tion constant (Barenblatt & Prostokishin, 1993; Butera et al., 1993;

    Ferro & Pecoraro, 2000), the following power velocity distribution is

    obtained:

    vu*

    =1δϕ3

    u* hυk

    ,hd,s,Rei ,F

    � �� �u* yυk

    � �δ, ð19Þ

    in which δ is an exponent which can be calculated by the following

    theoretical equation (Barenblatt, 1991; Castaing, Gagne, &

    Hopfinger, 1990):

    δ=1:5lnRe

    , ð20Þ

    in which Re = V h/νk is the flow Reynolds number.

    Equation (19) can be rewritten as follows:

    vu*

    =Γu* hυk

    ,hd,s,Rei,F

    � �u* yυk

    � �δ: ð21Þ

    Taking into account that the shear Reynolds number Re* can be

    rewritten as

    Re* =u* hυk

    dh: ð22Þ

    Equation (22) demonstrates that Re* implicitly considers the vari-

    ability of the depth to sediment ratio h/d.

    Ferro (2018) also deduced the following relationship, which states

    that F is related to Shields parameter, which is dependent on Re*

    (Buffington & Montgomery, 1997; Di Stefano & Ferro, 2005;

    Shields, 1936):

    F2hd

    γ

    γs−γð Þ=V2

    g dγ

    γs−γð Þ=g R sg d

    8f

    γ

    γs−γð Þ=

    γ R sγs−γð Þd

    8f=8fθ =

    8fϕ4 Re*ð Þ,

    ð23Þ

    in which ϕ4 is a functional symbol, γs is the soil particle specific

    weight, and γ is the water specific weight. As according to Equa-

    tion (23) F takes into account both the depth sediment ratio h/d and

    the shear Reynolds number Re*. The velocity profile can be rewritten

    as follows:

    vu*

    =Γ s,Rei,Fð Þ u* yυk

    � �δ: ð24Þ

    Integrating the power velocity distribution (Equation 24), the fol-

    lowing expression of the Darcy–Weisbach friction factor f is obtained

    (Barenblatt, 1993;Ferro, 2017 ; Ferro & Porto, 2018):

    f =821−δΓReδ

    δ+1ð Þ δ+2ð Þ

    " #− 21+ δ: ð25Þ

    Starting from Equation (24) and establishing that y = α h, the dis-

    tance from the bottom at which the local velocity is equal to the cross

    section average velocity V, the following value Γv of Γ can be obtained

    (Ferro, 2017; Ferro & Porto, 2018):

    Γv =V

    u*u*αhνk

    � �δ , ð26Þ

    in which α is a coefficient, smaller than unity, that allows consider-

    ation both that the average velocity V is located below the water sur-

    face and that the mean velocity profile, whose integration gives V, is

    calculated by averaging, for each distance y, the velocity values

    v measured in different verticals.

    Ferro (2017) deduced the following theoretical equation for cal-

    culating α:

    α=21−δ

    δ+1ð Þ δ+2ð Þ

    " #1=δ: ð27Þ

    For flow Reynolds number less than 2,000, the calculated α values

    vary in a very narrow range, and the mean value equal to 0.132 is

    applicable.

    The applicability of the flow resistance Equation (25) was tested

    in previous studies for different hydraulic conditions without the rain-

    fall effect (Di Stefano, Ferro, Palmeri, & Pampalone, 2017; Di Stefano,

    4 NICOSIA ET AL.

  • Ferro, Palmeri, Pampalone, & Agnello, 2017; Ferro, 2017, 2018;

    Ferro & Porto, 2018; Palmeri et al., 2018), and the following theoreti-

    cal expression of Γv function (Ferro, 2018) was proposed:

    Γv =aFb

    sc, ð28Þ

    in which a, b, and c are coefficients to be estimated by the

    measurements.

    When a flow is subjected to the rainfall impact, the Γ parameter

    of the velocity profile also depends, according to Equation (24), on Rei

    and the following expression of Γv can be used:

    Γv =aFb

    scReie , ð29Þ

    in which a, b, c, and e are coefficients to be estimated by the measure-

    ments. When the rainfall effect is neglected, Equation (29) reduces to

    Equation (28) for e = 0.

    3 | MEASUREMENTS OF OVERLAND FLOWRESISTANCE UNDER RAINFALL USED INTHIS INVESTIGATION

    Nearing et al. (2017) carried out experiments applying simulated rain-

    fall to plots at three different values of slope gradient (5%, 12%, and

    20%). The experiments were conducted on 6-m-long, 2-m-wide, and

    0.3-m-deep pivoted steel boxes with an adjustable slope (Figure 1).

    The soil used in the experiment (Luckyhills–McNeal) was gravely

    sandy loam, containing approximately 52% sand, 26% silt, and

    22% clay.

    The Walnut Gulch Rainfall Simulator was equipped with a single

    oscillating boom with four V-jet nozzles that could produce rainfall

    rates ranging between 13 and 190 mm hr−1. Wind shields were used

    to avoid wind effects on raindrops. Before the beginning of each

    experiment, the simulator was positioned over the soil box and cali-

    brated, and then the soil layer, approximately 20 cm deep, was wetted

    with a rainfall having a constant intensity (35 mm hr−1) to obtain equal

    moisture conditions.

    The experiments were carried out using two different values of

    rainfall intensity, namely, 59 and 178 mm hr−1, and runoff rate was

    measured using a V-shaped flume equipped with an electronic depth

    gauge, which was calibrated periodically during the experiment with

    timed runoff weight samples. Overland flow velocity was measured

    using a salt solution (three different point of application located at

    1.65, 3.5, and 5.8 m from the outlet) and by monitoring the electrical

    resistivity by electrical sensors embedded at the end of the plot. Two

    litres of the solution were uniformly applied on the surface using a

    perforated PVC pipe placed across the plot. The data were collected

    at 0.37 s intervals with real-time graphical output using LoggerNet3

    software and a CR10X3 data-logger by Campbell Scientific. The salt

    curve started at salt application time, and the measurement was

    stopped when the residual resistivity corresponding to the base level

    was reached. Peak values from the salt curve were used to measure

    flow velocity. The average flow depth was calculated by the runoff

    rate and the mean flow velocity assuming a rectangular cross

    section 2 m wide.

    Surface rock cover was measured at 300 points on a 20 cm ×

    20 cm grid using a hand-held, transparent, size guide arranged over

    the surface. A single-point laser sliding along a notched rail placed

    across the plot and pointed down on the plot was used to objectively

    identify the sample point locations. The technique ensured that sur-

    face rock was measured at the same points every time during the

    experiment. Rock cover percentage was considered to be the percent-

    age of points with rocks greater than 0.5 cm.

    Surface elevations were measured along three 2-m-long transects

    oriented across the plot located at 0.9, 2.9, and 4.9 m from the lower

    edge of the plot. Elevation points along these transects were mea-

    sured at 5-mm intervals with a 0.2-mm vertical resolution using a

    Leica3 E7500i laser distance metre mounted on an automated linear

    motion system.

    Two replications for each treatment were made, with measure-

    ments of runoff rate, flow velocity, rock cover, and surface roughness.

    These experiments were characterized by Reynolds numbers varying

    between 48 and 194 (which suggest laminar flow in the absence of

    raindrop impacts) and average Froude numbers between 0.08 and

    1.05 (99.4% were subcritical).

    The experimental data used in this paper are available online at

    https://doi.org/10.5194/hess-21-3221-2017-supplement.

    4 | RESULTS

    The variable Rei, accounting for rainfall intensity, describes this rainfall

    effect and is not correlated with Re (Figure 2). Therefore, the Reynolds

    number Re is not able to explain the rainfall effect, and both Rei andF IGURE 1 View of the soil box used by Nearing et al. (2017) forthe experiments

    NICOSIA ET AL. 5

    https://doi.org/10.5194/hess-21-3221-2017-supplement

  • Re are considered to be predictive variables of the Darcy–Weisbach

    friction factor f.

    The effect of rainfall on the flow resistance law was tested cali-

    brating the theoretical relationship (Equation 29) among the velocity

    profile parameter Γ, the channel slope, the flow Froude number, and

    the rain Reynolds number Rei.

    Using the whole database, constituted by 168 data points, the

    following equation to calculate the parameter Γv was found:

    Γv = 0:323F1:2833

    s0:6698 Re0:0787i: ð30Þ

    Equation (30) is characterized by a coefficient of determination

    equal to 0.999, and in Figure 3, the comparison between the Γv values

    calculated by Equation (26), with α = 0.132, and those calculated by

    Equation (30) is shown.

    Introducing Equation (30) into Equation (25), taking into account

    that Reδ is always equal to 4.4817 and using the mean value of δ

    (0.327), being as its range is narrow, the following equation to esti-

    mate the Darcy–Weisbach friction factor was obtained:

    f =12:408s1:0095Rei

    0:1187

    F1:9342: ð31Þ

    Figure 4 shows the comparison between the measured values of

    the Darcy–Weisbach friction factor and those calculated by Equa-

    tion (31); this agreement is characterized by a root mean square error

    (RMSE) equal to 1.4474. The friction factor values calculated by Equa-

    tion (31) were characterized by errors that were less than or equal to

    ±10% for 99.4% of cases and less than or equal to ±5% for 88.1% of

    cases.

    The analysis was also developed neglecting the influence of the

    rain Reynolds number and calibrating Equation (28) using the entire

    database, the following equation to calculate the parameter Γv was

    obtained:

    Γv =0:4054F1:2794

    s0:6633: ð32Þ

    Equation (32) is characterized by a coefficient of determination

    equal to 0.990. Figure 5 shows the comparison between Γv values cal-

    culated by Equation (26), with α = 0.132, and those calculated by

    Equation (32). Figure 5 clearly shows that Equation (32) under-

    estimated Γv values corresponding to a rainfall intensity i of

    59 mm hr−1, whereas it overestimated those corresponding to

    i = 178 mm hr−1.

    Introducing Equation (32) into Equation (25) and substituting

    Reδ = 4.4817 and δ = 0.327, the following flow resistance equation

    was obtained:F IGURE 2 Comparison between the experimental values of Reand Rei

    F IGURE 3 Comparison between Γv values calculated byEquation (26), with α = 0.132, and those calculated by Equation (30)

    F IGURE 4 Comparison between measured Darcy–Weisbachfriction factor values and those calculated by Equation (31)

    6 NICOSIA ET AL.

  • f =8:8088s1:008

    F1:9283: ð33Þ

    In Figure 6, the agreement between the measured friction factor

    values and those calculated by Equation (33) are shown. This agree-

    ment is characterized by RMSE equal to 1.4926, and the friction fac-

    tor values calculated by Equation (33) are characterized by errors that

    are less than or equal to ±10% for 71.4% of cases and less than or

    equal to ±5% for 22.6% of cases.

    Notwithstanding Equations (31) and (33) are characterized by

    similar values of RMSE (1.4474 and 1.4926, respectively), the differ-

    ence in terms of errors in calculating the Darcy–Weisbach friction fac-

    tor is substantial. Figure 7, which shows the frequency distribution of

    the errors corresponding to the two examined cases, demonstrates

    that errors in the estimate of the Darcy–Weisbach friction factor by

    Equation (31) were appreciably lower than those obtained by

    Equation (33).

    Finally, for both Equations (31) and (33), the Darcy–Weisbach

    friction factor f increased with a power of slope gradient having an

    exponent approximately equal to 1.

    Taking into account that V =ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi8 g R s=f

    p, this result confirms that

    flow velocity is approximately independent of slope.

    5 | DISCUSSION

    Figure 2 confirms that raindrop impact affects the f–Re relationship,

    and the K coefficient of Equation (1) is dependent on both channel

    grain resistance and rainfall intensity. Taking into account that the

    measurements used in this investigation are characterized by Reyn-

    olds numbers that suggest laminar flow in the absence of raindrop

    impacts, rain resistance can appreciably contribute to total resistance

    (Smith et al., 2007).

    Equation (31) takes into account two different effects on flow

    resistance: the friction factor due to bottom roughness, represented

    by F, and to rainfall impact represented by Rei. Equation (31) also

    shows that the flow resistance increased for increasing rain Reynolds

    number values. This result can be explained because the Rei increases

    with rainfall intensity, and for Re values less than 2,000, in agreement

    with previous studies (Yoon & Wenzel, 1971), flow resistance

    increases with rainfall intensity.

    The underestimation and overestimation of Equation (32) for

    i = 59 and 178 mm hr−1, respectively, shown in Figure 5, produces an

    opposite bias on the calculation of the Darcy–Weisbach friction factor

    (Figure 6), which establishes that the effect of rainfall impact cannot

    be neglected for the laminar flow regime.

    These influences of the rainfall intensity on flow resistance are

    also highlighted by the frequency distribution of the errors shown in

    Figure 7. The errors associated to Equation (31) vary in a very narrow

    range (−6%, +10%). In other words, the frequency distribution of the

    errors associated with Equation (33), which denotes a bias in the fric-

    tion factor estimate (the errors vary from −12% to +30%),

    F IGURE 5 Comparison between Γv values calculated byEquation (26), with α = 0.132, and those calculated by Equation (32)

    F IGURE 6 Comparison between measured Darcy–Weisbachfriction factor values and those calculated by Equation (33)

    F IGURE 7 Frequency distributions of errors in Darcy–Weisbachfriction factor calculation by Equations (31) and (33)

    NICOSIA ET AL. 7

  • demonstrates that the roughness model is incomplete. Other variables

    must be used to explain the friction factor variability, including the

    effect of rainfall impact on flow resistance.

    Finally, the exponent of slope quasi-equal to 1 for Equa-

    tions (31) and (33) demonstrates that, for a laminar overland flow, the

    increase of velocity due to slope can be balanced by the retarding

    effect due to the rainfall impact.

    Further experiments should be carried out for testing the appli-

    cability of the theoretical flow resistance equation for turbulent

    flow subjected to the impact of rainfall over a wider range of

    intensities.

    6 | CONCLUSIONS

    The approach using the theoretical flow resistance law, deduced by

    the Π-Theorem of dimensional analysis and self-similarity theory, was

    modified for taking into account the effect of rainfall intensity and

    was applied to the overland flow data by Nearing et al. (2017)

    obtained by rainfall simulator experiments.

    The available measurements, corresponding to two different rain-

    fall intensities and three different values of slope, were used to cali-

    brate the theoretical relationship between the velocity profile

    parameter Γv, the channel slope, the flow Froude number and the rain

    Reynolds number, which is a dimensionless group representing the

    effect of rainfall intensity on flow resistance.

    The results confirmed that, for flow Reynolds number less than

    2,000, flow resistance always increases for greater rainfall intensity

    values.

    This investigation also showed that the Darcy–Weisbach friction

    factor f increased with a power of slope gradient, having an exponent

    approximately equal to 1 and, as a consequence, showed that the

    overland flow mean velocity was quasi-independent of slope.

    In conclusion, the lack of knowledge in the field of overland flow

    subjected to rainfall should stimulate experimental activity for study-

    ing overland flow in different rainfall and surface roughness

    conditions.

    ACKNOWLEDGMENT

    All Authors set up the research and contributed to both the analysis

    of the data and writing of the manuscript. Dr. A. Nicosia developed

    this research activity, under the supervision of Dr. Nearing, during his

    stay at USDA-Agricultural Research Service, Southwest Watershed

    Research Center, Tucson, Arizona.

    DATA AVAILABILITY STATEMENT

    The data that support the findings of this study are openly available in

    https://doi.org/10.5194/hess-21-3221-2017-supplement.

    ORCID

    Alessio Nicosia https://orcid.org/0000-0003-0540-8788

    Costanza Di Stefano https://orcid.org/0000-0003-2941-4664

    Vincenzo Pampalone https://orcid.org/0000-0002-5195-9209

    Vincenzo Palmeri https://orcid.org/0000-0001-6594-9530

    Vito Ferro https://orcid.org/0000-0003-3020-3119

    REFERENCES

    Barenblatt, G. I. (1987). Dimensional analysis. In Gordon &breach. Amster-

    dam: Science Publishers Inc..

    Barenblatt, G. I. (1991). On the scaling laws (incomplete self-similarity with

    respect to Reynolds numbers) for the developed turbulent flows in

    tubes. C. R. Acad. Sci., Ser. II, 313, 307–312.Barenblatt, G. I. (1993). Scaling laws for fully developed turbulent shear

    flows, part 1, basic hypothesis and analysis. Journal of Fluid Mechanics,

    248, 513–520.Barenblatt, G. I., & Monin, A. S. (1979). Similarity laws for turbulent strati-

    fied flows. Archive for Rational Mechanics and Analysis, 70, 307–317.Barenblatt, G. I., & Prostokishin, V. M. (1993). Scaling laws for fully devel-

    oped turbulent shear flows, part 2. Processing of experimental data.

    Journal of Fluid Mechanics, 248, 521–529.Buffington, J. M., & Montgomery, D. R. (1997). A systematic analysis of

    eight decades of incipient motion studies with special reference to

    gravel-bedded rivers. Water Resources Research, 33, 1993–2029.Butera, L., Ridolfi, L., & Sordo, S. (1993). On the hypothesis of self-

    similarity for the velocity distribution in turbulent flows. Excerpta, 8,

    63–94.Castaing, B., Gagne, Y., & Hopfinger, E. J. (1990). Velocity probability density

    functions of high Reynolds number turbulence. Phisica D, 46, 177–200.Di Stefano, C., Ferro, V., 2005. Storia ed attualità dell'Abaco di Shields.

    Rivista di Ingegneria Agraria, 1, 65–78 (in Italian).Di Stefano, C., Ferro, V., Palmeri, V., & Pampalone, V. (2017). Flow resis-

    tance equation for rills. Hydrological Processes, 31, 2793–2801.Di Stefano, C., Ferro, V., Palmeri, V., & Pampalone, V. (2018). Testing slope

    effect on flow resistance equation for mobile bed rills. Hydrological

    Processes, 32, 664–671. https://doi.org/10.1002/hyp.11448Di Stefano, C., Ferro, V., Palmeri, V., Pampalone, V., & Agnello, F. (2017). Test-

    ing the use of an image-based technique to measure gully erosion at

    Sparacia experimental area. Hydrological Processes, 31, 573–585.Dunkerley, D. (2018). How is overland flow produced under intermittent

    rainfall? An analysis using plot-scale rainfall simulation on dryland soils.

    Journal of Hydrology, 556, 119–130.Emmett, W. W. (1970). The hydraulics of overland flow on hillslopes:

    Dynamic and descriptive studies of hillslopes. Geological Survey Profes-

    sional Paper, A1–A67.Ferro, V. (1997). Applying hypothesis of self-similarity for flow-resistance

    law of small-diameter plastic pipes. Journal of Irrigation and Drainage

    Engineering, ASCE, 123, 175–179.Ferro, V. (1998). Evaluating overland flow sediment transport capacity.

    Hydrological Processes, 12, 1895–1910.Ferro, V. (2017). New flow resistance law for steep mountain streams

    based on velocity profile. Journal of Irrigation and Drainage Engineering

    ASCE, 143(04017024), 1–6. https://doi.org/10.1061/(ASCE)IR.1943-4774.0001208

    Ferro, V. (2018). Assessing flow resistance in gravel bed channels by

    dimensional analysis and self similarity. Catena, 169, 119–127. https://doi.org/10.1016/j.catena.2018.05.034

    Ferro, V., Nicosia, A., (2020). Comment on “Rill erosion processes on steepcolluvial deposit slope under heavy rainfall in flume experiments with

    artificial rain by F. Jiang et al.” Catena, 185, 103793. https://doi.org/10.1016/j.catena.2018.10.022https://doi.org/10.1016/j.catena.

    2018.10.022

    Ferro, V., & Pecoraro, R. (2000). Incomplete self-similarity and flow veloc-

    ity in gravel bed channels. Water Resources Research, 36, 2761–2770.Ferro, V., & Porto, P. (2018). Applying hypothesis of self-similarity for flow

    resistance law in Calabrian gravel bed rivers (Fiumare). Journal of

    Hydraulic Engineering ASCE, 144, 1–11. https://doi.org/10.1061/(ASCE)HY.1943-7900.0001385

    8 NICOSIA ET AL.

    https://orcid.org/0000-0003-0540-8788https://orcid.org/0000-0003-0540-8788https://orcid.org/0000-0003-2941-4664https://orcid.org/0000-0003-2941-4664https://orcid.org/0000-0002-5195-9209https://orcid.org/0000-0002-5195-9209https://orcid.org/0000-0001-6594-9530https://orcid.org/0000-0001-6594-9530https://orcid.org/0000-0003-3020-3119https://orcid.org/0000-0003-3020-3119https://doi.org/10.1002/hyp.11448https://doi.org/10.1061/(ASCE)IR.1943-4774.0001208https://doi.org/10.1061/(ASCE)IR.1943-4774.0001208https://doi.org/10.1016/j.catena.2018.05.034https://doi.org/10.1016/j.catena.2018.05.034https://doi.org/10.1016/j.catena.2018.10.022https://doi.org/10.1016/j.catena.2018.10.022https://doi.org/10.1061/(ASCE)HY.1943-7900.0001385https://doi.org/10.1061/(ASCE)HY.1943-7900.0001385

  • Govers, G. (1992). Relationship between discharge, velocity and flow area

    for rills eroding loose, non-layered materials. Earth Surface Processes

    and Landforms, 17, 515–528.Govers, G., Takken, I., & Helming, K. (2000). Soil roughness and overland

    flow. Agronomie, 20, 131–146.Jain, M. K., Kothyari, U. C., & Ranga Raju, K. G. (2004). A GIS distributed

    rainfall-runoff model. Journal of Hydrology, 299, 107–135.Katz, D. M., Watts, F. J., & Burroughs, E. R. (1995). Effects of surface

    roughness and rainfall impact on overland flow. Journal of Hydraulic

    Engineering ASCE, 121, 546–553.Mügler, C., Planchon, O., Patin, J., Weill, S., Silvera, N., Richard, P., &

    Mouche, E. (2011). Comparison of roughness models to simulate over-

    land flow and tracer transport experiments under simulated rainfall at

    plot scale. Journal of Hydrology, 402, 25–40.Mutchler, C. K., & Young, R. A. (1975). Soil detachment by raindrops. Present

    and prospective technology for predicting sediment yield and sources,

    ARS-S-40 (pp. 113–117). Washington, DC: USDA Science and Educa-tion Administration.

    Nearing, M.A., Kimoto, A., Nichols, M. H., and Ritchie, J. C., 2005. Spatial

    patterns of soil erosion and deposition in two small, semiarid water-

    sheds. Journal of Geophysical Research, 110, 1-11. https://doi.

    org/10.1029/2005JF000290.

    Nearing, M. A., Polyakov, V. O., Nichols, M. H., Hernandez, M., Li, L.,

    Zhao, Y., & Armendariz, G. (2017). Slope–velocity equilibrium and evo-lution of surface roughness on a stony hillslope. Hydrology and Earth

    System Sciences, 21, 3221–3229.Palmeri, V., Pampalone, V., Di Stefano, C., Nicosia, A., & Ferro, V. (2018).

    Experiments for testing soil texture effects on flow resistance in

    mobile bed rills. Catena, 171, 176–184.

    Powell, D. M. (2014). Flow resistance in gravel bed rivers: Progress in

    research. Earth Science Reviews, 136, 301–338.Savat, J. (1977). The hydraulics of sheet flow on a smooth surface and the

    effect of simulated rainfall. Earth Surface Processes, 2, 125–140.Shen, H. W., & Li, R. (1973). Rainfall effect on sheet flow over smooth sur-

    face. Journal of Hydraulics Division ASCE, 99, 771–792.Shields, A., 1936. Anwendung der Aenlichkeitsmechanik und der

    Turbolenzforschung auf die Geschiebebewegung. Mitteil, PVWES,

    Heft 26, Berlin.

    Smith, M. W., Cox, N. J., & Bracken, L. J. (2007). Applying flow resistance

    equations to overland flows. Progress of Physical Geography, 31,

    363–387.Toy, T. J., Foster, G. R., & Renard, K. G. (2002). Soil erosion: Processes, pre-

    diction, measurement, and control. Inc, New York, USA: John Wiley &

    Sons.

    Yalin, M. S. (1977). Mechanics of sediment transport. Oxford, U.K.: Per-

    gamon Press.

    Yoon, Y. N., & Wenzel, H. G. (1971). Mechanics of sheet flow under simu-

    lated rainfall. Journal of the Hydraulics Division ASCE, 97, 1367–1385.

    How to cite this article: Nicosia A, Di Stefano C, Pampalone V,

    Palmeri V, Ferro V, Nearing MA. Testing a theoretical

    resistance law for overland flow on a stony hillslope.

    Hydrological Processes. 2020;1–9. https://doi.org/10.1002/

    hyp.13709

    NICOSIA ET AL. 9

    https://doi.org/10.1002/hyp.13709https://doi.org/10.1002/hyp.13709

    Testing a theoretical resistance law for overland flow on a stony hillslope1 INTRODUCTION2 THE THEORETICAL FLOW RESISTANCE EQUATION3 MEASUREMENTS OF OVERLAND FLOW RESISTANCE UNDER RAINFALL USED IN THIS INVESTIGATION4 RESULTS5 DISCUSSION6 CONCLUSIONSACKNOWLEDGMENT DATA AVAILABILITY STATEMENT

    REFERENCES