ArticlePDF Available

Modeling Cosmogenic Nuclides in Transiently Evolving Topography and Chemically Weathering Soils

Wiley
Journal of Geophysical Research: Earth Surface
Authors:

Abstract and Figures

Terrestrial cosmogenic nuclides (TCN) are widely employed to infer denudation rates in mountainous landscapes. The calculation of an inferred denudation rate (Dinf) from TCN concentrations is typically performed under the assumptions that denudation rates were steady during TCN accumulation and that soil chemical weathering negligibly impacted soil mineral abundances. In many landscapes, however, denudation rates were not steady and soil composition was significantly impacted by chemical weathering, which complicates interpretation of TCN concentrations. We present a landscape evolution model that computes transient changes in topography, soil thickness, soil mineralogy, and soil TCN concentrations. We used this model to investigate TCN responses in transient landscapes by imposing idealized perturbations in tectonically (rock uplift rate) and climatically sensitive parameters (soil production efficiency, hillslope transport efficiency, and mineral dissolution rate) on initially steady‐state landscapes. These experiments revealed key insights about TCN responses in transient landscapes. (a) Accounting for soil chemical erosion is necessary to accurately calculate Dinf. (b) Responses of Dinf to tectonic perturbations differ from those to climatic perturbations, suggesting that spatial and temporal patterns in Dinf are signatures of perturbation type and magnitude. (c) If soil chemical erosion is accounted for, basin‐averaged Dinf inferred from TCN in stream sediment closely tracks actual basin‐averaged denudation rate, showing that Dinf is a reasonable proxy for actual denudation rate, even in many transient landscapes. (d) Response times of Dinf to perturbations increase with hillslope length, implying that response times should be sensitive to the climatic, biological, and lithologic processes that control hillslope length.
This content is subject to copyright. Terms and conditions apply.
1. Introduction
The denudation of mountainous landscapes is central to many Earth processes. It modifies topography (e.g.,
Gilbert, 1877), accelerates soil production (e.g., Heimsath et al., 1997), affects soil nutrient supply (e.g., Y.
Lucas,2001), focuses rock exhumation (e.g., Willett,1999), perturbs sea level by unloading continental surfaces
(e.g., Dalca etal.,2013; Ferrier etal.,2017), and modulates Earth's climate by affecting weathering rates of sili-
cates and sulfides (e.g., Torres etal.,2014; Walker etal.,1981). Deciphering the controls on denudation rate is
therefore of wide interest across the geosciences.
Abstract Terrestrial cosmogenic nuclides (TCN) are widely employed to infer denudation rates in
mountainous landscapes. The calculation of an inferred denudation rate (Dinf) from TCN concentrations is
typically performed under the assumptions that denudation rates were steady during TCN accumulation and
that soil chemical weathering negligibly impacted soil mineral abundances. In many landscapes, however,
denudation rates were not steady and soil composition was significantly impacted by chemical weathering,
which complicates interpretation of TCN concentrations. We present a landscape evolution model that
computes transient changes in topography, soil thickness, soil mineralogy, and soil TCN concentrations. We
used this model to investigate TCN responses in transient landscapes by imposing idealized perturbations
in tectonically (rock uplift rate) and climatically sensitive parameters (soil production efficiency, hillslope
transport efficiency, and mineral dissolution rate) on initially steady-state landscapes. These experiments
revealed key insights about TCN responses in transient landscapes. (a) Accounting for soil chemical erosion
is necessary to accurately calculate Dinf. (b) Responses of Dinf to tectonic perturbations differ from those to
climatic perturbations, suggesting that spatial and temporal patterns in Dinf are signatures of perturbation type
and magnitude. (c) If soil chemical erosion is accounted for, basin-averaged Dinf inferred from TCN in stream
sediment closely tracks actual basin-averaged denudation rate, showing that Dinf is a reasonable proxy for actual
denudation rate, even in many transient landscapes. (d) Response times of Dinf to perturbations increase with
hillslope length, implying that response times should be sensitive to the climatic, biological, and lithologic
processes that control hillslope length.
Plain Language Summary In geomorphology, the rate at which mountains wear down is
commonly inferred from concentrations of cosmogenic nuclides in minerals in soil or river sand. Cosmogenic
nuclides are isotopes that build up in minerals during exposure to high energy particles from space,
accumulating at a rate that depends on the degree to which the soil is chemically altered and the mountain's
erosion rate itself. To explore how cosmogenic nuclide concentrations change in soil over time, we developed
a computer model that tracks cosmogenic nuclide concentrations in soil across a landscape. We used this
model to perform two model experiments: one driven by a change in the rate at which the bedrock is rising
(the so-called tectonic experiment), and the other driven by a change in rainfall (the climatic experiment).
Our experiments confirm that accounting for soil composition is necessary to accurately infer erosion rate.
They also show that cosmogenic nuclide concentrations respond differently to tectonic changes than climatic
changes, implying that patterns of cosmogenic nuclide concentrations may reflect perturbation type. The
time it takes cosmogenic nuclide concentrations to respond to a perturbation increases with the length of the
hills from the ridge to the valley, implying that response times should be influenced by climate, life, and rock
type. These results show how this model can be used to explore how mountainous topography, soils, and
cosmogenic nuclides change simultaneously. Our model will be a useful tool for improving field measurements
of cosmogenic nuclide concentrations in soil and in stream sediment.
REED ETAL.
© 2023. The Authors.
This is an open access article under
the terms of the Creative Commons
Attribution License, which permits use,
distribution and reproduction in any
medium, provided the original work is
properly cited.
Modeling Cosmogenic Nuclides in Transiently Evolving
Topography and Chemically Weathering Soils
Miles M. Reed1 , Ken L. Ferrier1 , and J. Taylor Perron2
1Department of Geoscience, University of Wisconsin-Madison, Madison, WI, USA, 2Department of Earth, Atmospheric, and
Planetary Sciences, Massachusetts Institute of Technology, Cambridge, MA, USA
Key Points:
We developed a model to compute
terrestrial cosmogenic nuclide
(TCN) concentrations in transiently
evolving topography and chemically
weathering soils
TCN-based denudation rates track
actual denudation rates more closely
during responses to changes in uplift
rate than to changes in climate
Soil chemical weathering inf luences
modeled TCN concentrations,
confirming that this should be
accounted for in TCN-based
denudation rates
Supporting Information:
Supporting Information may be found in
the online version of this article.
Correspondence to:
M. M. Reed,
miles.reed@wisc.edu
Citation:
Reed, M. M., Ferrier, K. L., & Perron, J.
T. (2023). Modeling cosmogenic nuclides
in transiently evolving topography and
chemically weathering soils. Journal of
Geophysical Research: Earth Surface,
128, e2023JF007201. https://doi.
org/10.1029/2023JF007201
Received 12 APR 2023
Accepted 25 SEP 2023
10.1029/2023JF007201
RESEARCH ARTICLE
1 of 20
Journal of Geophysical Research: Earth Surface
REED ETAL.
10.1029/2023JF007201
2 of 20
At present, the most common tool for measuring millennial-scale denudation rates is terrestrial cosmogenic
nuclides (TCN). Concentrations of TCN in host minerals (e.g.,
10Be in quartz) can be used to infer denudation
rates averaged over a catchment drainage area (Bierman & Steig,1996; Brown etal.,1995; Granger etal.,1996)
or at a point on a hillslope (Dixon & von Blanckenburg,2012; Lal,1991; Riebe etal.,2003). Over the past few
decades, TCN-inferred estimates of denudation rate (which we refer to as Dinf) have been measured in thousands
of places (Codilean etal.,2022) and have been harnessed to investigate many things, including the influence of
climate and tectonics on denudation rate (e.g., Ferrier etal.,2012; Godard etal.,2020; Riebe etal.,2004; West
etal.,2005). The standard equations used to invert TCN concentrations for Dinf assume steady state—that is,
TCN-bearing minerals have been exhumed from depth to the surface at a steady rate and that the subsurface depth
profile of TCN concentrations is invariant in time (Lal,1991).
In landscapes that are out of steady state, TCN concentrations differ from what they would be in steady state.
If conventional expressions for steady-state Dinf were applied to TCN concentrations in transient landscapes,
then the resulting estimates of Dinf would differ from the actual denudation rates by an amount that depends on
the magnitude of the landscape's deviation from steady state (e.g., Bierman & Steig,1996; Mudd,2017; von
Blanckenburg,2006). Ferrier and Kirchner(2008) observed that Dinf is robust to minor deviations from steady
state, which suggests that Dinf is likely to be an accurate approximation of the true denudation rate in many land-
scapes. More practically, the Ferrier and Kirchner(2008) simulations imply that errors in Dinf associated with
transience are smaller than our ignorance of what denudation rates actually were over the past few thousand years
(an ignorance that is large in most landscapes, and in some landscapes is unbounded), given the absence of alter-
native methods for measuring millennial-scale denudation rates. This underscores the usefulness of measuring
Dinf with TCN, despite uncertainties associated with transient conditions. Nonetheless, if deviations from steady
state are not accounted for, they could lead to errors in Dinf of unknown size. It would be useful to be able to
interpret those sources of error as accurately as possible.
Concerns about transients are common because transient landscapes are common. Stream capture (Beeson
etal.,2017), exhumation of rocks of different strengths (Forte etal.,2016; Zondervan et al.,2020), climatic
shifts (Marshall etal.,2017), and changes in tectonic movements (Hurst etal.,2019) can all give rise to transients
in topography and soil composition (Ferrier & West,2017). Moreover, detecting landscape transience can be
difficult at the soil profile or hillslope scale (Heimsath etal.,2002; Hippe etal.,2021). Recent work with paired
nuclides has presented a method for detecting the magnitude but not the timing of transient conditions (Skov
etal.,2019), but quantifying the history of deviations from steady state remains challenging.
The recognition that landscapes are rarely in perfect steady state has inspired efforts to investigate transient
effects on TCN through numerical models. Small etal.(1999) developed the first hillslope-profile model of
TCN conservation that included process representations of hillslope sediment transport and soil production.
Anderson(2015) extended the hillslope-profile framework to investigate the influence of soil mixing on particle
transit times and TCN concentrations. Heimsath(2006) modeled the response of TCN depth profiles to changes in
physical erosion rate and showed that TCN-inferred erosion rates lag behind actual rates. Mudd(2017) modeled
TCN responses in a 2D landscape to changes in uplift rate, hillslope transport, and fluvial incision efficiency and
showed that Dinf responses to climate-related parameters closely matched the forcing while Dinf responses to uplift
rate were damped and lagged. These studies and many others have advanced our understanding of the effects of
transients on TCN concentrations.
To date, no study has investigated the sensitivity of TCN to transient conditions in a full 2D landscape evolution
model that includes explicit treatment of eroding and chemically weathering soil. That is our goal here. In this
study, we present a model to compute TCN concentrations in soil and the underlying rock, and we show how to
apply this model across landscapes undergoing simultaneous changes in topography, soil production rate, physi-
cal erosion rate, chemical erosion rate, soil thickness, and soil composition. This extends the model of Ferrier and
Perron(2020), which accounted for production, transport, and chemical erosion of soils in transiently evolving
topography but did not include TCN.
A key aspect of this model is that it is able to account for the effects of soil chemical erosion on TCN. These
effects can be significant. Small etal.(1999) and Riebe etal.(2001) recognized that chemical weathering of solu-
ble minerals increases the exposure times of dissolution-resistant host minerals such as quartz to cosmic rays, and
hence increases TCN concentrations in quartz beyond what they would be in the absence of chemical weathering.
If this additional exposure were neglected, this would yield erroneously low estimates of TCN-based denudation
Journal of Geophysical Research: Earth Surface
REED ETAL.
10.1029/2023JF007201
3 of 20
rate. Riebe and Granger(2013) extended this work to consider chemical erosion in deep saprolite, showing that
Dinf can be in error by tens of percent if chemical erosion is neglected. Because our simulations calculate soil
chemical depletion directly, they are able to confirm the importance of accounting for soil chemical erosion in
TCN-based estimates of denudation rate. More generally, our simulations show how this model can be used to
investigate the influence of tectonic and climatic perturbations on transiently evolving hillslopes and illustrate
how to identify transient conditions from multiple TCN measurements across a landscape.
2. Methods
2.1. Model Description
We designed our model to capture simultaneous responses of TCN, topography, soil thickness, and soil compo-
sition to changes in boundary conditions (e.g., bedrock uplift rate) and rates of soil production, transport, and
chemical weathering. The model does not treat bare-bedrock hillslopes, so application of this model is restricted
to soil-mantled landscapes. We incorporated nonlinear hillslope soil transport and TCN conservation into the
coupled landscape evolution-soil mineralogy model of Ferrier and Perron(2020). The model tracks and conserves
TCN concentrations in the soil, at the soil-bedrock interface, and in deeper bedrock (Figure1). Our aim was to
construct and test a model that simulates cosmogenic nuclide production and transport in soil-mantled uplands,
a common type of landscape.
In our model, stream channels are restricted to vertical movement relative to the hillslopes and set the local
base-level of the hillslopes (Mudd & Furbish,2007). We do have the capacity to specify a rate law for stream
channel incision (Text S4 in Supporting Information S1), but, in this work, we wish to focus solely on the
distributed hillslope response for simplicity. During transient conditions, rates of hillslope soil transport, soil
production, and mineral dissolution vary across the landscape. This results in temporal and spatial variations in
soil thickness, soil mineral concentrations, and TCN concentrations. In this section, we detail the governing equa-
tions, numerical implementation, and several methods for solving for inferred denudation rate using the modeled
TCN concentrations.
2.2. Governing Equations for Topography, Soil Thickness, and Soil Composition
The elevation of the bedrock-soil boundary, zb, increases at a bedrock uplift rate, U (L T
−1), and decreases at a soil
production rate, ϵ (L T
−1). The change in zb is expressed as
Figure 1. Schematic of model framework (Section2.1). At each grid point within a synthetic landscape (panel a), soil mass (panel b) and cosmogenic nuclides (panel
c) are conserved. (a) Stream channels indicated in blue, soil-mantled hillslopes in shaded relief with elevation contours. (b) Soil thickness (H) is governed by physical
erosion rate (E), chemical erosion rate (W), and soil production rate (ϵ). Elevation of the bedrock-soil boundary (zb) is governed by bedrock uplift rate (U) and ϵ.
Soil mineral concentrations (CXs) are determined by mineral supply from ϵ and losses through E and W. Soil and bedrock densities (ρs and ρr) are held constant. (c)
Soil terrestrial cosmogenic nuclides (TCN) concentrations (Ns) are determined by supply from bedrock, hillslope soil transport, and cosmogenic nuclide production
and decay (star symbol). In the soil, Ns (
10Be in this case) is uniform with depth, consistent with a vertically well-mixed soil. The TCN concentration at soil-bedrock
boundary (Nr,zb) arises through the calculation of a bedrock TCN profile Nr at a depth ξ below the soil-bedrock interface through time.
Journal of Geophysical Research: Earth Surface
REED ETAL.
10.1029/2023JF007201
4 of 20


=00
.
(1)
In Equation1, the second term is the exponential soil production function (Heimsath etal.,1997,1999), where
H (L) is soil thickness, H0 (L) is the characteristic length that regulates soil production rate with regard to H, and
ϵ0 (L T
−1) is the maximum soil production rate at zero H. U is imposed upon the landscape, and a static channel
network incises at a rate equal to U. H changes via soil production rate, the divergence of downslope soil trans-
port, and chemical erosion rate:


=
00−∇
𝑊
(2)
where ρr and ρs are bedrock and soil density (M L
−3), respectively. In our model, ρr and ρs are time-invariant.
The downslope volumetric flux of soil per unit contour width, qs (L
2 T
−1), increases nonlinearly with respect to
hillslope gradient of the model surface elevation, (zs=zb+H), and is defined as
𝑞
𝑠=
𝐾nl𝑧
𝑠
1−(|𝑧𝑠|𝑆𝑐)
2
.
(3)
Here, Knl is the soil transport coefficient (L
2 T
−1), ∇zs is the local hillslope gradient, and Sc is the critical gradi-
ent (Roering etal., 1999,2001). Our framework is flexible enough to adopt other formulations for qs, such as
those that depend on soil thickness (Braun etal., 2001). For simplicity, the simulations we show in this study
are governed by Equation3. This translates into a physical erosion rate E (L T
−1) by taking the divergence of qs
(Roering etal.,1999):
𝐸
=−𝐾nl
2𝑧𝑠
1−(
𝑧𝑠
𝑆𝑐)2+2
𝜕𝑧𝑠
𝜕𝑥
2
𝜕2𝑧𝑠
𝜕𝑥2
+
𝜕𝑧𝑠
𝜕𝑦
2
𝜕2𝑧𝑠
𝜕𝑦2
+2
𝜕𝑧𝑠
𝜕𝑥
𝜕𝑧𝑠
𝜕𝑦
𝜕2𝑧𝑠
𝜕𝑥𝜕𝑦
𝑆2
𝑐
1−(
𝑧𝑠
𝑆𝑐)2
2
,
(4)
where x and y are the two axes of a regular grid. As implemented, E is limited by the thickness of the explicitly
modeled soil layer, H, such that, for a given timestep, erosion can never be greater than H (Text S2 in Supporting
InformationS1). This parameterization for soil transport is most applicable to nonlinear creep and is not directly
applicable to large discrete events such as deep-seated landslides. Chemical erosion rate, W (L T
−1), is calculated
following Ferrier and Kirchner (2008) as the difference between mineral dissolution and secondary mineral
formation rates summed over n mineral species throughout the soil:
=
=1 (
)
(5)
where kj (mol L
−2T
−1), Aj (L
2 mol
−1), Cjs (M M
−1), sj (mol L
−3T
−1), and wj (M mol
−1) are the mineral dissolution
constant, specific surface area, soil mineral concentration, secondary mineral production rate, and molar mass of
the jth mineral from the set of n soil minerals, respectively.
The concentration of an individual soil mineral X, CXs (M M
−1), evolves through the competition between disso-
lution and mineral supply. We assume a well-mixed soil such that mineral concentration is independent of soil
depth (Ferrier & West,2017). Soil production, hillslope soil transport, mineral dissolution, secondary mineral
formation, and the mass losses of the other mineral phases combine to modify a mineral's concentration:

Xs
 =
00(Xr Xs)
Xs Xs +
+Xs
=1
(

).
(6)
Here, CXr (M M
−1) is the concentration of mineral X in the bedrock. This formulation differs from that in Ferrier
and Perron(2020) only in the transport term (second term), where qs is represented as a nonlinear function of
hillslope gradient (Equation3) rather than a linear function of it (Yoo etal.,2007). Combining Equations1 and2,
the rate of change of zs is expressed as
Journal of Geophysical Research: Earth Surface
REED ETAL.
10.1029/2023JF007201
5 of 20

 =+(
−1
)00
=1
(
)
.
(7)
Model outputs can exhibit nonlinear behavior during transient conditions through the interacting changes in
topography, mineral supply from soil production, soil thickness, and chemical erosion (Ferrier & Perron,2020).
2.3. Governing Equations for Cosmogenic Nuclides
We extended the model of Ferrier and Perron(2020) to track concentrations of cosmogenic nuclides in soil and
the underlying rock (Riebe & Granger, 2013). First, we implemented TCN production in the soil and bedrock
such that production rates evolve with topography. Here, we outline how the processes described above modify
TCN concentrations in soil and bedrock.
To compute changes in TCN concentrations in the soil, we account for TCN production in the soil, addition
of TCN from soil production, changes in TCN through the divergence of soil transport, and losses of TCN to
radioactive decay and host mineral dissolution. The full expression for the change in soil TCN concentration, Ns
(atoms
M−1
host
), is

 =1
=1(0)Λ(1−⁄Λ)
host,
host,
00(,)
∇(host,)−host host
.
(8)
Here, i is the ith of n cosmogenic production pathways, Pi (atoms
M−1
host
T
−1) is the cosmogenic production rate
of the ith production pathway (at zero depth), and Λi (M L
−2) is the attenuation length associated with the ith
production pathway. Chost,r and Chost,s (M M
−1) are the concentrations of the TCN host mineral in the bedrock and
soil, respectively. Nr,zb is the concentration of TCN at the soil-bedrock interface, and λ (T
−1) is the decay constant
of the TCN. The last term in Equation8 represents the loss of TCN concentration by chemical erosion, which
can be ignored if the host mineral is inert or nearly so. A more complete derivation of Equation8 can be found in
Text S2 of Supporting InformationS1.
In this study, we used n=3 production pathways representing spallation, negative muon capture, and fast muogenic
production, which is appropriate for several TCNs, including
10Be,
26Al, and
14C in quartz (Dunai,2010). For all
production pathways, we employed the Lifton-Sato-Dunai scaling scheme (Lifton etal.,2014) within CRONUS-
calc (Marrero etal.,2016) using a sea-level, high-latitude
10Be production rate of 3.92 atoms
g−1
qtz
yr
−1 (Borchers
et al., 2016). We used Madison, Wisconsin, as a reference location to scale production rates. We obtained a
uniform spallation pathway attenuation length (Λs) from CRONUScalc using the mean elevation of the reference
model topography. To parameterize muogenic production rates and attenuation lengths, we used the codes of
Balco(2017) to create 1,000 synthetic muogenic production profiles for both negative muon capture and fast
production pathways across the entire range of elevations comprising our modeling runs. In keeping with previ-
ous work (Braucher etal.,2013), profiles were fitted with a single-term exponential function to arrive at a surface
production rate and an attenuation length for each pathway. Similar to spallation, we used a mean attenuation
length for each muogenic pathway as the range was small. We then created piecewise cubic Hermite interpolants
(Fritsch & Carlson,1980) for each production pathway using the corresponding surface production rates and
elevations, allowing the model to update surface production rates as elevations change.
Calculating Nr,zb in Equation8 requires determining the rate of change in bedrock TCN concentration, Nr, at an
arbitrary depth below the soil-bedrock interface, ξ (L), as
𝜕𝑁
𝑟(𝜉)
𝜕𝑡
=
3
𝑖
𝑃𝑖(0)𝑒𝜌𝑠𝐻∕Λ𝑖𝑒𝜌𝑟𝜉∕Λ𝑖𝜆𝑁𝑟(𝜉)−𝜖0𝑒𝐻𝐻0𝜕𝑁𝑟
𝜕𝜉 ||||𝜉
.
(9)
This equation represents the sum of nuclide production in bedrock and decay while recognizing that the soil
production rate history influences the bedrock nuclide concentration profile. By including deep production,
periods of transient denudation rate can induce changes in the bedrock TCN concentration profile (Knudsen
etal.,2019), which can alter soil TCN (Skov etal.,2019). The change in Nr,zb is then evaluated at the soil-bedrock
interface (i.e., ξ=0) as
Journal of Geophysical Research: Earth Surface
REED ETAL.
10.1029/2023JF007201
6 of 20
𝜕𝑁
𝑟,𝑧𝑏
𝜕𝑡
=
3
𝑖
𝑃𝑖(0)𝑒𝜌𝑠𝐻∕Λ𝑖𝜆𝑁𝑟(𝜉)−𝜖0𝑒𝐻𝐻0𝜕𝑁𝑟
𝜕𝜉 ||||𝜉=0
.
(10)
We solved an approximation of these two equations with a semi-Lagrangian advection scheme (Blackburn
etal.,2018; Spiegelman & Katz,2006) described in Section2.4.
2.4. Inferring Denudation Rates From Cosmogenic Nuclides
To fully exploit a landscape evolution model that conserves TCN, we must be able to efficiently infer denudation rates
from TCN concentrations. Throughout this study, we applied Equation11 to infer denudation rates (Dinf) from Ns.
=
3
(0)
+inf ∕Λ
[
host,
host,
(
1−∕Λ
)
+∕Λ
]
(11)
Equation 11 is nearly identical to an analogous expression from Riebe and Granger (2013) for a two-layer
bedrock-soil system similar to our model; the only difference is that it includes radioactive decay. It was derived
under assumptions of steady state, with constant Dinf and H in a well-mixed, chemically weathered soil produced
from an underlying parent rock in which negligible chemical erosion occurred. Equation 11 does not have an
analytical solution for Dinf, so in practice it must be solved numerically. Our goal is to investigate the extent to
which denudation rate estimates inferred from Equation11 would deviate from actual denudation rates during
landscape transience. At all soil-mantled cells, we solved for Dinf in Equation11 and others presented in Section3
using local values of Pi, Ns, H, and Chost,s/Chost,r, all of which are measurable in the field or in the lab.
Calculators such as CRONUScalc or CRONUS solve Dinf from Ns numerically (Balco et al., 2008; Marrero
etal.,2016). Similarly, we used model outputs of Ns, Chost,s, and H to solve Dinf in Equation11 as a post-processing
step as it is too computationally expensive to solve at each timestep.Using a Newton-Raphson method implemented
within MATLAB's variable-precision solver for Equation11 under steady-state conditions at ridge locations not
subject to downslope soil transport, Dinf deviates by only ∼0.18% from the modeled, actual denudation rate, Dact.
For basin-averaged inferred denudation rate (Dinf,basin), we solved Equation11 using a mean H and Chost,s/Chost,r,
and hypsometrically averaged Pi(0) for the entire basin. We did this as a post-processing step on a delineated basin
through time using outputs of Ns, Chost, and E. We constructed a virtual stream sediment TCN concentration using
these variables for all stream-side grid points. This concentration is calculated as
𝑁
𝑠,stream =
𝑁
𝑠,
stream-side𝐸stream-side 𝜌
𝑠
𝐶
ℎ𝑜𝑠𝑡,𝑠,
stream-sideΔ𝑥Δ𝑦Δ𝑡
𝐸stream-side𝜌𝑠𝐶ℎ𝑜𝑠𝑡,𝑠,stream-side Δ𝑥Δ𝑦Δ𝑡.
(12)
Here, ∆x and ∆y are the grid point lengths in the x and y directions, and ∆t is the duration of the timestep.Thus
far, few studies have accounted for enrichment or depletion of the host mineral in cosmogenic nuclide-based
estimates of basin-averaged denudation rate, but recent work in carbonate landscapes underscores the importance
of doing so (Ott etal., 2022,2023). In Section 3, we show the implications of neglecting chemical erosion in
point-based and basin-averaged estimates of denudation rate.
2.5. Numerical Implementation
Here, we describe some details associated with the model domain, boundary conditions, and solutions to the
governing equations. In all modeling scenarios, we used a 150× 150 regular grid with a resolution of 10 m.
The boundaries of the model domain were pinned to zero elevation, allowing the boundaries to act as a regional
base-level. Mass transfers to the channel network do not change the channel elevation (Ferrier & Perron,2020).
We used a timestep of 10years to maintain an acceptable agreement between Dact and Dinf at steady state. The
governing equations are solved using explicit finite difference methods. The supplement contains a detailed
description of the implementation of nonlinear hillslope soil transport based on Perron(2011).
In order to solve Equations9 and10, we used a semi-Lagrangian advection scheme to account for deep cosmo-
genic production under variable denudation rates (Figure S1 in Supporting InformationS1). The scheme can
solve for Nr(ξ) at regularly spaced depths of an evolving bedrock concentration profile (Knudsen etal.,2019).
The soil production rate at a grid cell provides a vertical velocity from which we find a departure depth via cubic
interpolation (Fletcher,2019). During each timestep, TCN production rate is calculated using a depth obtained by
the averaging the depth of evaluation (i.e., ξ) and the departure depth. Nr,zb is then filled with the value at ξ=0,
Journal of Geophysical Research: Earth Surface
REED ETAL.
10.1029/2023JF007201
7 of 20
which is the bedrock-soil interface. The lowermost value of the profile is filled with a steady-state TCN concen-
tration corresponding to a running average of bedrock exhumation velocities. This mimics a pre-perturbation
history of TCN accumulation below the profile, which cannot be accounted for analytically. The number and
spacing of evaluation points below the surface affect the accuracy of the TCN concentration relative to an analyt-
ical steady-state value due to interpolation errors (T. R. Lucas,1974). We observed a depth of 10m and point
spacing of 10mm to be a good compromise between accuracy and computational speed. In the supplement, we
detail an alternative, speedier method for calculating Nr,zb that does not directly track TCN profiles (Figure S2 in
Supporting InformationS1).
2.6. Attaining Steady-State Topography
To prepare for simulations under transient conditions (Section4), we ran two successive simulations to obtain
steady-state initial conditions. First, we generated a steady-state topography with the Tadpole landscape evolution
model (Perron etal.,2008,2009,2012; Richardson etal.,2020). This model differs from that in Section2 in that
it includes not only soil transport but also stream incision, and in that it does not track changes in soil thickness,
soil composition, or chemical erosion rate. In this model, under a threshold stream-power incision law, the change
in elevation with time due to fluvial erosion is expressed as

 =
0
()>
,
(13)
where Kf (L
1–2m T
−1) is an effective erodibility, A (L
2) is drainage area, S is slope (L L
−1), m and n are unit-
less constants, and θc (L
2m) is an incision threshold (e.g., Pelletier, 2012; Theodoratos & Kirchner, 2020).
Starting from a surface generated by red noise, we ran the model until the topography reached steady state
using the same boundary conditions and model parameters used in the transient simulations (U = 50 mm
kyr
−1, Knl=0.0032m
2yr
−1 and Sc=1.2; Section4.1) and the following values for stream incision parameters:
Kf=0.0001year
−1, m=0.5, n=1, and θc=8m. We extracted the channel network from the resulting steady-state
topography using a threshold drainage area of 5,000m
2.
We then used the topography output from this first simulation as the initial condition for a second simulation with
the model in Sections2.1 and2.2. We ran this simulation to steady state to obtain steady state values of zs, H, CXs,
and TCN concentrations. In this simulation, the channel network was fixed in the geometry it was in at the end
of the Tadpole simulation. The median difference in elevation between the initial topography and the steady-state
topography is ∼3.5m and is due to the inclusion of soil chemical erosion, soil production, bedrock lithology, and
the fixed stream channel network.
3. Results: Steady-State TCN-Inferred Denudation Rates Are Sensitive to Soil
Chemical Erosion
How do inferred denudation rates computed with Equation11 compare to denudation rate estimates inferred from
Ns via some other common formulations? Since we solved Equation11 numerically as a post-processing step,
we tracked and outputed Dinf by ignoring radioactive decay. If decay is neglected, Equation11 would simplify to
inf,CEF =
3
(0)Λ
[host,
host,
(
1−∕Λ
)
+∕Λ]
,
(14)
which is identical to that in Riebe and Granger(2013). Most inferred denudation rates in the literature do not correct
for chemical erosion (Riebe & Granger,2013). If chemical erosion is neglected, Equation11 would simplify to
=
3
(0)
+inf,Lal ∕Λ
,
(15)
which is similar to that originally derived in Lal(1991) and identical to those used in denudation rate calculators
when evaluating surface samples (Marrero etal., 2016) or stream sediment samples (Mudd etal.,2016). We
solved Equation15 via the same method as Equation11.
Journal of Geophysical Research: Earth Surface
REED ETAL.
10.1029/2023JF007201
8 of 20
We computed Dinf with Equations11, 14, and 15 in five steady-state landscapes that differ in their bedrock
concentration of the soluble phase plagioclase (5%, 20%, 35%, 50%, and 65%) and the complementary concen-
tration of the bedrock's only other mineral phases, quartz, which is treated as insoluble (95%, 80%, 65%, 50%, and
35%, respectively). All the landscapes had a U of 50mm kyr
−1. Figure2 shows Dinf/Dact for these five landscapes
using Equations11, 14 and15 at a ridge location not subject to downslope soil transport (left panel) and a small
basin (right panel). Comparing Dinf/Dact for Equations11 and14 in Figure2 shows that ignoring decay results in
a small error in Dinf (∼2.5%) in these simulations. By contrast, comparing Dinf/Dact for Equations11 and15 shows
that neglecting the enrichment of Chost,s by chemical erosion leads to larger errors in these simulations, up to 29%
in the simulation with the highest W/D (and plagioclase concentration). These errors increase with increasing
W/D, consistent with the observations in Small etal.(1999), Riebe etal.(2001), and Riebe and Granger(2013).
Figure2 also shows that the ratio diverges from 1 with increasing W/D for basin-averaged Dinf. In the model, the
effect of W on Dinf is indirectly affected by U, to the extent that U affects properties like soil thickness and soluble
mineral concentrations (Text S5 in Supporting InformationS1).
To summarize, Figure2 implies that calculating Dinf requires accounting for soil chemical depletion if the cosmogenic
host mineral has a different soil residence time than the average soil material (e.g., quartz, which stays in the soil while
other minerals are lost to dissolution). Our study is not the first to make this point—Small etal.(1999) and Riebe
etal.(2001) did this more than 20years ago—but our model is well-equipped to show the importance of this effect.
4. Results: Transient Landscapes
4.1. Responses to Tectonic and Climatic Perturbations
To investigate the transient response of TCN to perturbations, we performed two numerical experiments. In the
first experiment, we simulated a simple tectonic perturbation in which initially steady-state landscape experi-
ences an instantaneous doubling of rock uplift rate (or equivalently a doubling of the lowering rate of the domain
boundaries and the channels) from 50mm kyr
−1 to 100mm kyr
−1. In the second experiment, we simulated a
Figure 2. Ratio of inferred denudation rate (Dinf, computed with Equations11, 14 and15) to actual denudation rate,
Dact, in steady-state landscapes at a ridge (left panel) and averaged over a small basin (right panel). Markers represent
simulations with a range of ratios of chemical erosion rate to denudation rate (W/D), resulting from a range of bedrock
plagioclase concentrations (5%–65%). Values of Dinf/Dact close to 1 for Equations11 and14 show that accounting for soil
chemical erosion yields denudation rate estimates close to actual denudation rates. Values of Dinf/Dact far from 1 computed
with Equation15, which neglects chemical erosion, show that Dinf grows increasingly farther from Dact with increasing
W/D. This underscores the importance of accounting for soil chemical erosion in TCN-based estimates of denudation rate.
Basin-averaged rates are nearly identical to those on ridges because these simulations are steady-state landscapes.
Journal of Geophysical Research: Earth Surface
REED ETAL.
10.1029/2023JF007201
9 of 20
simple climatic perturbation in which the maximum soil production rate, hillslope transport efficiency, and the
dissolution rate constant of plagioclase were simultaneously increased by 50%. This tests the response of TCN to
a “top-down” perturbation where the entirety of the hillslope area is affected at the same time (Mudd,2017). In
both experiments, we restricted our attention to the responses of the hillslopes by fixing the lateral position of the
stream network in place during the simulation (Hurst etal.,2012; Mudd & Furbish,2007). In each simulation,
we ran the model until CXs, H, and Ns reached new steady-state values, which occurred at 3 and 3.1 Myr of model
run time in the tectonic and climatic experiments, respectively.
To quantify the difference between the actual instantaneous denudation rate Dact and the TCN-based estimate
of denudation rate Dinf (Equation11), we introduce the metric Ddiff=Dinf−Dact. This reveals the magnitude of
the error in the TCN-inferred denudation rate and is a measure of the deviation from steady-state conditions in
a non-steady landscape. This can be calculated at any point on the landscape, and therefore can reveal spatial
patterns in deviations from steady state. It can also be calculated for a drainage basin as the difference between
the basin-averaged inferred denudation rate, Dinf, basin, and the actual basin-averaged denudation rate, Dact, basin.
In both experiments, the initial U is 50mm kyr
−1. All model parameters used in the perturbation experiments
are listed in Table1. We use values of Knl, Sc, and ϵ0 from the Oregon Coast Range as it is one of the few land-
scapes where these parameters have been calibrated from field measurements (Heimsath etal.,2001; Roering
etal.,1999). The kplag parameter comes from the field measurements of Clow and Drever(1996). Both scenarios
are assigned the same bedrock mineralogy (50% plagioclase and 50% quartz) to simplify the interpretation of soil
composition in terms of one easily weathered phase and one insoluble phase.
The responses of landscape-wide Dinf and Dact differed spatially and temporally between the two experiments. In the
tectonic scenario, the areas around channels responded first to the increase in uplift rate, steepening and increasing
Dact. This was followed by a wave of increased physical erosion rate that gradually propagated up the hillslopes to
the ridges (Figure3). Although Dact exhibited notable change around channels by 100 kyr, Dinf showed little change
(Figure3). This is due to downslope soil transport from upslope areas that have not yet reached the new steady-state
conditions, which buffers the response in convergent hollows and footslopes. Both Dact and Dinf along ridges equili-
brated slowly to the perturbation. Dinf stayed close to local steady state (Dinf≈Dact) on ridges throughout the experi-
ment, such that the minimum Ddiff stayed above −10mm kyr
−1. By comparison, Ddiff on footslopes and hollows was
larger and persisted due to the buffering effect of hillslope transport (Figure3). Across the entire experiment, Dinf
deviated from Dact by at most 64% during the experiment (Ddiff=−46mm kyr
−1), which occurred at 71 kyr in hollows
directly above channel heads and along channels below long hillslopes. Positive Ddiff values were low (Ddiff=3mm
kyr
−1) during the entirety of the experiment, implying that Dinf did not exceed Dact by much at any time or place.
Model parameter Tectonic Climatic
t—model timestep 10yr 10yr
x, ∆y—grid resolution 10m 10m
U—bedrock uplift rate 0.05 to 0.1mm yr
−1 0.05mm yr
−1
Knl—hillslope transport efficiency (nonlinear) 0.0032m
2 yr
−1 a 0.0032–0.0048m
2 yr
−1
SC—critical hillslope gradient 1.2m m
−1 b 1.2m m
−1
ε0—maximum soil production rate 0.268mm yr
−1 c 0.268–0.402mm yr
−1
H0—soil production rate scaling length 1/3m c 1/3m
ρs—soil density 1,325kg m
−3 1,325kg m
−3
ρr—bedrock density 2,650kg m
−3 2,650kg m
−3
kplag—plagioclase dissolution rate constant 0.000005mol m
−2yr
−1 d 0.000005–0.0000075mol m
−2yr
−1
Aplag—plagioclase mineral specific surface area 117m
2 mol
−1 d 117m
2 mol
−1
ξmax—max depth of bedrock concentration profile 10m 10m
ξ spacing—distance between profile evaluation points 0.01m 0.01m
# averaging timesteps for cosmogenic profile infilling 20,000 20,000
aRoering etal.(1999).
bRoering etal.(2001).
cHeimsath etal.(2001).
dFerrier and Kirchner(2008) and Clow and Drever(1996).
Table 1
Model Parameters for Transient Perturbation Experiments
Journal of Geophysical Research: Earth Surface
REED ETAL.
10.1029/2023JF007201
10 of 20
Closer investigation of changes in H, Ns, and CXs at different hillslope positions through time showed how large
Ddiff developed in hollows and footslopes through the transport of Ns and CXs (Figure 5b–5e). In the hollows,
the decrease in H far outpaced the slower response of Ns and CXs. This suggests that estimates of Dinf inferred
from samples on ridges (Ferrier etal.,2012; Larsen etal.,2014) should be good indicators of current Dact during
tectonic-style perturbations like the one imposed here. At the ridge, larger magnitude changes in U induced larger
Ddiff that equilibrated faster due to hillslope steepening and soil thinning as the landscape approached SC (Figure
S3 in Supporting InformationS1).
Across the landscape in the climatic experiment, Dact increased rapidly after the initial perturbation, and Dinf
increased nearly as rapidly. This resulted in negative Ddiff values at 10 kyr (Figure4). Dinf and Dact attained a mean
maximum difference of around 44% (Ddiff=−21mm kyr
−1) across all hillslope positions at 2 kyr. By 100 kyr,
stream channel incision rate started to dominate the response, and a wave of decreased Dact migrated upslope,
which generated positive Ddiff (Figure4). This was again due to the buffering capacity of soil transported from
hillslope positions at different stages of transient evolution. When Ns, CXs, and H from different hillslope posi-
tions are considered, the perturbation to ϵ0, soil production efficiency, rapidly increased H at all positions, which
resulted in a decrease in Ns. The relatively fast change in these variables explains the quick recovery of Ddiff to
near zero (Figures5g and 5h). As the channels began to dominate, the hillslope positions displayed differing
responses. Soil transport delayed the recovery to steady-state conditions by supplying Ns and CXs from ridges to
footslopes and hollows, leading to prolonged positive Ddiff in these locations (Figures5g and 5h). Given these
results, perturbations that synchronously disrupt the entire landscape may cause inaccuracies in real-world esti-
mates of Dinf on ridges directly after the perturbation. These inaccuracies would grow and be prolonged if the
magnitude of the perturbation change were increased (Figure S4 in Supporting InformationS1).
Figure 3. Landscape distributions of actual and inferred denudation rates Dact and Dinf (Equation11), and the difference Ddiff=Dinf−Dact for the tectonic experiment
at 10, 100, and 1,000 kyr after the perturbation. Topography shown at 2x vertical exaggeration. The model domain is 1.5 km by 1.5 km. Topographic contours are every
15m. Maximum elevation is 180m at 1,000 kyr. A movie of the experiment is available as a supplement (MovieS1). Ridges not subject to hillslope soil transport
remain near local steady state (Dinf≈Dact) throughout the simulation.
Journal of Geophysical Research: Earth Surface
REED ETAL.
10.1029/2023JF007201
11 of 20
Figure6 shows Ddiff, basin, the difference between the basin-averaged TCN-inferred denudation rate, Dinf, basin, and
the basin-averaged instantaneous denudation rate, Dact, basin, for the drainage basin outlined in Figure6. Ddiff, basin
was relatively small throughout most of each simulation, yet the time series of Ddiff, basin differed in style between
the tectonic and climatic experiments. In the tectonic experiment, Ddiff, basin progressed to −7mm kyr
−1 (10%
difference) at ∼158 kyr and then declined gradually to nearly zero over ∼1 Myr (Figure6b). In contrast, in the
climatic experiment, Ddiff, basin evolved as a sharp initial pulse to a maximum of −23mm kyr
−1 (37% difference)
at 2 kyr and then returned to near zero within 83 kyr (Figure6c), a much faster response than that in the tectonic
experiment. This was followed by a longer, lower-amplitude response in which Ddiff, basin grew as large as 2mm
kyr
−1 and then decayed to nearly zero over ∼1 Myr, similar to the duration of the response in the tectonic exper-
iment (Figure6b). Larger magnitude changes in the perturbed parameters for each scenario generated greater
Ddiff, basin which equilibrated either more rapidly (tectonic) or slowly (climatic) due to steeper (tectonic) or gentler
(climatic) gradients (Figures S5 in Supporting InformationS1).
The results show that these transient changes in bedrock uplift rate, soil production rate, dissolution rate, and soil trans-
port rate do not generate large errors in estimates of Dinf, basin from TCN in stream sediment. This implies that large
errors in Dinf, basin in practice are more likely to arise from processes not represented in the model, such as deep-seated
landslides (Schide etal.,2022; Yanites etal.,2009) or sediment storage along or within the stream network (Grischott
etal.,2017). The inclusion of dynamic channels would increase the response time in the tectonic scenario since the
response would need to transit the channel network before propagating through the hillslopes (Hurst etal.,2012).
Figure 4. Landscape distributions of actual and inferred denudation rates Dact and Dinf (Equation11), and the difference Ddiff=Dinf−Dact for the climatic experiment
at 0, 10, 100, and 1,000 kyr after the perturbation. Topography shown with 2x vertical exaggeration. The model domain is 1.5km by 1.5km. Contours are every
15m. The maximum elevation is 135m at 1,000 kyr. A movie of the experiment is available as a supplement (MovieS2). The initial surfaces (t=0 kyr) illustrate the
abrupt initial change in all variables during this experiment. For this type of perturbation, Ddiff varies between negative and positive values depending on the time since
perturbation and hillslope position.
Journal of Geophysical Research: Earth Surface
REED ETAL.
10.1029/2023JF007201
12 of 20
4.2. Response Times
The time it takes a landscape to respond to external forcings is a fundamental property of the landscape. This
response time is useful because it can be compared to time scales of climatic or tectonic forcings and used to iden-
tify whether a landscape is likely to be close to steady state (e.g., Roering etal.,2001). Here we use the tectonic
and climatic experiments in Figures3 and4 to calculate response times of several characteristics of the landscape
(soil TCN concentrations, soil thickness, soil mineralogy, and denudation rates). To do this, we applied a modi-
fied version of the response time calculation in Ferrier and Perron(2020), adapted to accommodate variables with
Figure 5. (a) Map view of the initial topography for the tectonic and climatic perturbation experiments. Channels marked in blue. Numbers 1–4 indicate the hillslope
positions (1: ridge; 2: channel head; 3: sideslope; and 4: near-channel) that are tracked in panels (b–i). (b–e) Time-series plots of Ns, H, Chost,s, and Ddiff from tectonic
perturbation experiment. (f–i) Time series plots from climatic perturbation. Initial variability in Ns is due to hillslope soil transport. The variable responses show that
hillslope position determines how the underlying variables used to obtain Dinf evolve.
Journal of Geophysical Research: Earth Surface
REED ETAL.
10.1029/2023JF007201
13 of 20
similar initial and final values (e.g., for Dact and Dinf in the climatic experiment in Figure5). For instance, for Dinf,
we define the response time as the time it takes Dinf to get 95% of the way to its new steady-state value relative to
the maximum deviation of Dinf from its new steady-state value. In mathematical terms, this is the time at which
Dinf first satisfies the condition
|inf inf, final|<max(0.05|inf, inf, final |)
and continues to satisfy it for the
remainder of the simulation. Here, Dinf,t is the entire time series of Dinf for the modeling run, Dinf, final is the final
steady-state value of Dinf, and 0.05 is a 5% threshold level. The response time is defined this way for every point
in the landscape. We used this definition to calculate the response time, τ, for Ns, H, Chost,s, Dact, and Dinf, which
we denote τNs, τH, τChost,s, τDact, and τDinf, respectively.
When response times are mapped across the landscape, they reveal spatial patterns in the propagation of signals
throughout the domain. Figure7 shows that response times varied between the tectonic and climatic experiments
and among variables within each experiment. In the tectonic experiment, the response time of Dinf was greater
than the response times of all other variables. τDinf varied in space across the landscape from 368 kyr to 1.735
Myr and had a mean (± s.d.) of 1.304 Myr± 223 kyr (Figure7f). For the variables not associated with cosmo-
genic nuclides, areas near channels responded fastest (e.g., τH and τDact in Figures7b and7e). However, τDinf was
high along channels where hillslope sediment transport from ridges carried down parcels of soil with higher Ns
(Figure7f). Ridges responded slower than other parts of the landscape for every characteristic (Figure7a–7f),
mimicking the way the perturbation in bedrock uplift is translated upslope (Mudd & Furbish,2007). When this
experiment is viewed holistically, the redistribution of relict Ns downslope controls the response time of Dinf.
Response times in the climatic experiment were longer than those in the tectonic experiment for all variables
(Figure7g–7l). This differs from a similar experiment in Ferrier and Perron(2020) and is likely due to the intro-
duction of more complex topography and different parameter values. The mean (± s.d.) of τDinf in the climatic
experiment was 1.604 Myr± 276 kyr and varied from 784 kyr to 2.238 Myr. High τDinf in top-central area of
the modeling domain is indicative of topographic control on τDinf (Figure7l). This zone underwent the most
topographic adjustment during the perturbation. These response times show that relatively small perturbations
in the efficiencies of hillslope processes could lead to transient states lasting much longer than the timescale of
climatic fluctuations. Although this is the case, as shown in Figures5 and6, Ddiff on ridges is quite low for the vast
duration of modeling time. This means that even though τDinf is high for perturbations affecting climate-related
hillslope processes, Dinf values should reflect Dact at almost any point in time except for directly after the onset
of a step-change.
5. Discussion
5.1. Sensitivity of Response Times for Maximum Ddiff and Dinf to Hillslope Length
Hillslope length (Lh) is a fundamental landscape property and is defined as the distance from ridges to stream
channels (Grieve etal., 2016). Numerical modeling and experimental work have shown that Lh is set by the
competition between advective fluvial processes and diffusive hillslope processes (Perron etal.,2009; Sweeney
Figure 6. (a) Initial topography with selected basin outlined in black. (b) Difference between inferred and actual
basin-averaged denudation rates, Ddiff, basin, for the initial 2,000 kyr of modeling time in the tectonic experiment. During this
simulation, errors in Dinf, basin were no larger than 7mm kyr
−1, damped by the range of cosmogenic nuclide production rates
within the basin and through the collection of an integrated sediment sample. (c) The climatic experiment shows that Ddiff, basin
could be large directly after a perturbation but recover relatively rapidly given a constant tectonic regime.
Journal of Geophysical Research: Earth Surface
REED ETAL.
10.1029/2023JF007201
14 of 20
etal.,2015). In numerical simulations in Ferrier and Perron(2020), chemical erosion rates took longer to respond
to perturbations in landscapes with longer hillslopes. This matches previous modeling of topographic response
times of a single hillslope for both linear and nonlinear hillslope sediment transport to changes in bedrock uplift
(Fernandes & Dietrich,1997; Roering etal.,2001). These studies suggest that Lh should influence cosmogenic
nuclide behavior as well, including response times and the magnitude of Ddiff.
To test the sensitivity of the maximum Ddiff (Ddiff,max) and Dinf response time (τDinf) to Lh in our model, we gener-
ated four synthetic landscapes with a range of Lh values. We generated these using the previously mentioned
stream power incision model and applied the same parameters as in the numerical experiments in Figures35
except for Kf (effective erodibility) and the drainage area threshold for channel extraction. This yielded four land-
scapes that had average Lh values ranging from 34.5 to 93.2m. With these landscapes, we explored the sensitivity
of Ddiff,max and τDinf to Lh by applying the same tectonic-style perturbation to each of them, which involved a
step-change in bedrock uplift rate from 50mm kyr
−1 to 100 kyr
−1, as shown in Figure4. To demonstrate spatial
variations in the responses within each landscape, we computed Ddiff,max at multiple hillslope positions (ridge,
sideslope, channel head) and as a basin-averaged quantity in a small basin in each simulation.
The landscape position at which Dinf deviated the least from Dact was on ridges. Figure8a shows that this was the
case at all average hillslope lengths and that the magnitude of the deviation of Dinf from Dact decreased with Lh.
In contrast, the largest deviations of Dinf from Dact occurred at channel heads, where the largest deviations were
temporarily greater than the imposed change in uplift rate at the beginning of the simulation.
The patterns of Ddiff,max are instructive. Because ridges respond last to base-level perturbations, they are the parts
of the landscape where Dinf most closely tracks Dact, making them ideal sampling locations for measuring physical
and chemical erosion rates (Ferrier etal.,2016). This is particularly so above long hillslopes, where the differ-
ence between Dinf and Dact approaches zero (Figure8a). However, in landscapes with ridge-visiting landslides
(Campforts etal.,2022; Dahlquist etal.,2018), this agreement between Dinf and Dact could be violated. Similar
to the ridges, basin-averaged Dinf better depicts basin-averaged Dact as Lh increases, as the method considers
nuclide production rates throughout the basin, including slowly responding areas (Granger etal.,1996; Mudd
etal.,2016). Channel heads and sideslopes produced greater Ddiff,max with increasing Lh due to enhanced buffering
of Dinf by hillslope transport of relatively high Ns soil from the slowly responding ridges.
τDinf increased with Lh at all topographic positions and was largest at ridges (Figure8b). Landscapes with longer
hillslopes exhibited shorter basin-averaged τDinf relative to the sideslope and channel head. This is due to lower
final steady-state Dinf from downslope transport from ridges to these positions, which makes meeting the response
Figure 7. (a–l) Response times (τ) of soil cosmogenic nuclide concentration (Ns), soil thickness (H), cosmogenic nuclide host mineral concentration (Chost,s), actual
denudation rate (Dact), and inferred denudation rate (Dinf) for the tectonic (a–f) and climatic experiments (g–l) (Figures3 and4). The response time of Dinf is longest on
ridges due to the combined response times of all the underlying variables, but, as the text details, the deviation of Dinf from Dact is smallest on ridges.
Journal of Geophysical Research: Earth Surface
REED ETAL.
10.1029/2023JF007201
15 of 20
criterion more demanding. At the longest Lh, the channel head τDinf is lower than the sideslope because Dinf at the
channel head goes down by ∼5mm kyr
−1, easing the satisfaction of the response criterion. Response times were
also sensitive to U, with faster responses at higher U (Figure S3 in Supporting InformationS1).
5.2. Detection of Transience in a Small Catchment Directly With One Cosmogenic Nuclide
Recognition of ongoing topographic change at the scale of headwater streams can be challenging. Testing theoret-
ical concepts in fluvial and hillslope geomorphology often begins by assuming a condition of topographic steady
state (Dietrich etal.,2003). A test of transience using cosmogenic nuclides could eliminate concerns over this
confounding factor when formulating geomorphic transport laws relating mass fluxes to landscape properties.
Recently, Mudd(2017) highlighted that at the scale of a soil profile (experiencing no soil transport from upslope)
one cannot detect instantaneous removals of overlying material or step-changes in denudation rate using single
cosmogenic nuclide depth-profile sampling. This condition can be overcome by using a radionuclide pair such
as in situ
14C/
10Be, where one nuclide decays much more rapidly (Hippe,2017). Using this technique, changes
in the magnitude and timing of denudation rate can be identified for perturbations occurring in the last several
thousand years (e.g., anthropogenic change) (Hippe etal.,2019). Due to its fast decay rate, in situ-produced
14C
is only moderately responsive to changes in Dact in the 30mm kyr
−1 to 100mm kyr
−1 range (Skov etal.,2019),
which is a common range in soil-mantled uplands (Dixon & von Blanckenburg,2012).
Our results show that detection of basin-scale transience due to changes in U could be feasible through systematic
sampling using one nuclide. Divergent hillslopes close to the channel respond rapidly to the change (Figure9a).
If a low ridge and a higher ridge above it were sampled in the same basin, measured Dinf values on these ridges
would differ markedly during transience, potentially beyond commonly reported analytical uncertainties (10%).
In Figure9a, the difference between these Dinf curves is labeled as a “transience detection window.” The differ-
ence in Dinf values corresponds to the current magnitude of transience, but both the future steady-state Dact and the
time since the perturbation would be indeterminate. The maximum magnitude of the difference and the duration
of this transience detection window would scale with the length of the hillslope below the sampled ridge. This
suggests that measuring Ns on low and high ridges within a given basin may provide a constraint on the current
magnitude of basin-scale landscape transience. These could be paired with topographic analysis of knickpoints
(Neely etal.,2017) and stream network disequilibrium between neighboring basins (Willett etal.,2014) to obtain
a more complete picture of a landscape's deviation from steady state.
In contrast, no analogous transience detection window appeared in the climatic perturbation experiment due to
the synchronous response across the landscape (Figure9b). Instead, values of Dinf on different ridges agreed with
Figure 8. (a) Maximum difference between the cosmogenic
10Be-inferred and actual denudation rates (Ddiff,max) during
simulations in which rock uplift rate instantaneously increases. (b) Response times of inferred denudation rate (τDinf) to a
step-change in bedrock uplift rate from 50 to 100mm kyr
−1 on landscapes with different average hillslope lengths (Lh). The
ridge, sideslope, and channel head points are specific, manually selected hillslope positions similar to those in Figure5. The
basin-averaged points correspond to small basins similar to that in Figure6.
Journal of Geophysical Research: Earth Surface
REED ETAL.
10.1029/2023JF007201
16 of 20
one another within uncertainty throughout this simulation. This suggests that detecting long-lived changes in Dact
arising from changes in hillslope soil transport efficiency may require sedimentary record-based paleo-denudation
rates (Marshall etal.,2017).
6. Conclusions
We used our newly introduced landscape evolution model to compute changes in soil cosmogenic nuclide concen-
trations alongside simultaneous changes in topography, soil thickness, and soil composition stemming from ideal-
ized tectonic and climatic perturbations. Our simulations highlighted several key conclusions.
First, accounting for the effects of chemical erosion on soil quartz enrichment is necessary for accurately infer-
ring denudation rates in weathered soils (Figure2). This supports previous work by Small etal.(1999), Riebe
etal.(2001), and Riebe and Granger(2013). It also suggests that soil chemical erosion may explain some of
the scatter in compilations of TCN-based denudation rates (e.g., Portenga & Bierman, 2011) and thus may
help resolve climatic, lithologic, tectonic, and topographic controls on landscape evolution (e.g., Kirchner &
Ferrier,2013; Larsen etal.,2014; Perron,2017; Willenbring etal.,2013).
Second, tectonic perturbations produce spatiotemporal variations in Dinf that differ from those induced by climatic
perturbations (Figures35). This suggests that spatial and temporal patterns in TCN-based denudation rate esti-
mates are useful indicators of the drivers of past changes in topography and sediment fluxes.
Third, estimates of basin-averaged Dinf closely track actual basin-averaged denudation rates at most times during
these simulations, with Dinf differing from Dact by less than ∼10% for the entire tectonic experiment (Figure6).
The exception to this is a brief excursion after the climatic perturbation in Figure5, when Dact changed immedi-
ately everywhere across the landscape, while Dinf took ∼15 kyr to adjust. This implies that TCN-based estimates
of basin-averaged denudation rate should be a reliable reflection of actual denudation rates in transient land-
scapes, except shortly after perturbations that affect the entire landscape at once. This supports early studies that
validated the use of TCN to infer basin-averaged denudation rates (Bierman & Steig,1996; Brown etal.,1995;
Granger etal.,1996), which noted that mixing of stream sediment should strongly damp within-basin variations
in TCN, and therefore that well-mixed stream sediment should have TCN concentrations that closely match the
basin-average TCN concentration.
Fourth, the response times of Dinf to climatic and tectonic perturbations are long (∼10
5–10
6years) and increase
with hillslope length (Figures7 and8). This implies that the response times of Dinf should be influenced by the
Figure 9. (a) Cosmogenic
10Be-inferred denudation rate Dinf at two hillslope positions—one at a high-elevation ridge, the
other at a low-elevation ridge—in a simulation undergoing a step change in rock uplift rate from 50 to 100mm kyr
−1 at t=0
kyr. The uncertainty bounds are ±10% on Ns and are meant to show how Dinf values could be distinguished from one another
in practice. The transience detection window exists for around ∼700 kyr in this case. As both the high-ridge and low-ridge
Dinf are near Dact during the entire perturbation, the gap between the low ridge and high ridge would indicate the current
magnitude of transience but not the final Dact at steady state. (b) Same but for a step change in climatic parameters, which
does not exhibit a detection window.
Journal of Geophysical Research: Earth Surface
REED ETAL.
10.1029/2023JF007201
17 of 20
climatic, biological, and lithologic processes that control the efficiencies of river incision and soil transport and
hence hillslope length (Perron etal.,2009,2012; Richardson etal.,2019). This in turn raises the possibility that
the primary influences of life, climate, and rock type on denudation rate may not be on the magnitude of denuda-
tion rate, but rather on the duration and spatial pattern of its responses to perturbations.
Lastly, in landscapes that are still adjusting to a past perturbation, it may be possible to quantify the degree to
which the landscape deviates from steady state. For example, measuring a single TCN in two ridgetop soils within
a single basin can yield the current magnitude of topographic transience after a simple tectonic perturbation
(Figure9). Since ridges tend to have Dinf that most closely track Dact, sampling ridges rather than other landscape
positions may yield the clearest picture of this transient evolution.
Together, these results illustrate the ways in which this model can be used to investigate the coupled evolution of
topography, soil chemistry, and cosmogenic nuclide concentrations. The model will be a useful tool for exploring
the controls on TCN in soil and stream sediment in transient landscapes, and this in turn can help improve inter-
pretations of denudation rate estimates derived from field measurements of TCN.
Data Availability Statement
MATLAB code to run the model and initial conditions for featured model runs are available from the Zenodo
repository (Reed etal.,2023, https://doi.org/10.5281/zenodo.8256787). Model output files for the two numerical
experiments are large (>5GB) but are available upon request from the corresponding author.
References
Anderson, R. S. (2015). Particle trajectories on hillslopes: Implications for particle age and
10Be structure. Journal of Geophysical Research:
Earth Surface, 120(9), 1626–1644. https://doi.org/10.1002/2015JF003479
Balco, G. (2017). Production rate calculations for cosmic-ray−muon-produced
10Be and
26Al benchmarked against geological calibration data.
Quaternary Geochronology, 39, 150–173. https://doi.org/10.1016/j.quageo.2017.02.001
Balco, G., Stone, J. O., Lifton, N. A., & Dunai, T. J. (2008). A complete and easily accessible means of calculating surface exposure ages or
erosion rates from
10Be and
26Al measurements. Quaternary Geochronology, 3(3), 174–195. https://doi.org/10.1016/j.quageo.2007.12.001
Beeson, H. W., McCoy, S. W., & Keen-Zebert, A. (2017). Geometric disequilibrium of river basins produces long-lived transient landscapes.
Earth and Planetary Science Letters, 475, 34–43. https://doi.org/10.1016/j.epsl.2017.07.010
Bierman, P., & Steig, E. J. (1996). Estimating rates of denudation using cosmogenic isotope abundances in sediment. Earth Surface Processes and
Landforms, 21(2), 125–139. https://doi.org/10.1002/(sici)1096-9837(199602)21:2<125::aid-esp511>3.0.co;2-8
Blackburn, T., Ferrier, K. L., & Perron, J. T. (2018). Coupled feedbacks between mountain erosion rate, elevation, crustal temperature, and
density. Earth and Planetary Science Letters, 498, 377–386. https://doi.org/10.1016/j.epsl.2018.07.003
Borchers, B., Marrero, S., Balco, G., Caffee, M., Goehring, B., Lifton, N., etal. (2016). Geological calibration of spallation production rates in
the CRONUS-Earth project. Quaternary Geochronology, 31, 188–198. https://doi.org/10.1016/j.quageo.2015.01.009
Braucher, R., Bourlès, D., Merchel, S., Romani, J. V., Fernadez-Mosquera, D., Marti, K., etal. (2013). Determination of muon attenuation lengths
in depth profiles from in situ produced cosmogenic nuclides. Nuclear Instruments and Methods in Physics Research Section B: Beam Interac-
tions with Materials and Atoms, 294, 484–490. https://doi.org/10.1016/j.nimb.2012.05.023
Braun, J., Heimsath, A. M., & Chappell, J. (2001). Sediment transport mechanisms on soil-mantled hillslopes. Geology, 29(8), 683–686. https://
doi.org/10.1130/0091-7613(2001)029<0683:stmosm>2.0.co;2
Brown, E. T., Stallard, R. F., Larsen, M. C., Raisbeck, G. M., & Yiou, F. (1995). Denudation rates determined from the accumulation of in
situ-produced
10Be in the Luquillo Experimental Forest, Puerto Rico. Earth and Planetary Science Letters, 129(1–4), 193–202. https://doi.
org/10.1016/0012-821x(94)00249-x
Campforts, B., Shobe, C. M., Overeem, I., & Tucker, G. E. (2022). The art of landslides: How stochastic mass wasting shapes topography and influ-
ences landscape dynamics. Journal of Geophysical Research: Earth Surface, 127(8), e2022JF006745. https://doi.org/10.1029/2022jf006745
Clow, D. W., & Drever, J. I. (1996). Weathering rates as a function of f low through an alpine soil. Chemical Geology, 132(1–4), 131–141. https://
doi.org/10.1016/s0009-2541(96)00048-4
Codilean, A. T., Munack, H., Saktura, W. M., Cohen, T. J., Jacobs, Z., Ulm, S., etal. (2022). OCTOPUS database (v. 2). Earth System Science
Data, 14(8), 3695–3713. https://doi.org/10.5194/essd-14-3695-2022
Dahlquist, M. P., West, A. J., & Li, G. (2018). Landslide-driven drainage divide migration. Geology, 46(5), 403–406. https://doi.org/10.1130/
g39916.1
Dalca, A. V., Ferrier, K. L., Mitrovica, J. X., Perron, J. T., Milne, G. A., & Creveling, J. R. (2013). On postglacial sea level—III. Incorporating
sediment redistribution. Geophysical Journal International, 194(1), 45–60. https://doi.org/10.1093/gji/ggt089
Dietrich, W. E., Bellugi, D. G., Sklar, L. S., Stock, J. D., Heimsath, A. M., & Roering, J. J. (2003). Geomorphic transport laws for predicting
landscape form and dynamics. Geophysical Monograph-American Geophysical Union, 135, 103–132.
Dixon, J. L., & von Blanckenburg, F. (2012). Soils as pacemakers and limiters of global silicate weathering. Comptes Rendus Geoscience,
344(11–12), 597–609. https://doi.org/10.1016/j.crte.2012.10.012
Dunai, T. J. (2010). Cosmogenic nuclides: Principles, concepts and applications in the Earth surface sciences. Cambridge University Press.
https://doi.org/10.1017/CBO9780511804519
Fernandes, N. F., & Dietrich, W. E. (1997). Hillslope evolution by diffusive processes: The timescale for equilibrium adjustments. Water
Resources Research, 33(6), 1307–1318. https://doi.org/10.1029/97wr00534
Ferrier, K. L., Austermann, J., Mitrovica, J. X., & Pico, T. (2017). Incorporating sediment compaction into a gravitationally self-consistent model
for ice age sea-level change. Geophysical Journal International, 211(1), 663–672. https://doi.org/10.1093/gji/ggx293
Acknowledgments
We would like to thank Simon Mudd and
Sean Gallen for their reviews that greatly
improved the manuscript. We would
also like to thank Associate Editor Jon
Pelletier for his insightful comments. This
work was supported by NSF Grant EAR
2045433 to KLF.
Journal of Geophysical Research: Earth Surface
REED ETAL.
10.1029/2023JF007201
18 of 20
Ferrier, K. L., & Kirchner, J. W. (2008). Effects of physical erosion on chemical denudation rates: A numerical modeling study of soil-mantled
hillslopes. Earth and Planetary Science Letters, 272(3–4), 591–599. https://doi.org/10.1016/j.epsl.2008.05.024
Ferrier, K. L., Kirchner, J. W., & Finkel, R. C. (2012). Weak influences of climate and mineral supply rates on chemical erosion rates:
Measurements along two altitudinal transects in the Idaho Batholith. Journal of Geophysical Research, 117(F2), F02026. https://doi.
org/10.1029/2011jf002231
Ferrier, K. L., & Perron, J. T. (2020). The importance of hillslope scale in responses of chemical erosion rate to changes in tectonics and climate.
Journal of Geophysical Research: Earth Surface, 125(9), e2020JF005562. https://doi.org/10.1029/2020JF005562
Ferrier, K. L., Riebe, C. S., & Jesse Hahm, W. (2016). Testing for supply-limited and kinetic-limited chemical erosion in field measurements of
regolith production and chemical depletion. Geochemistry, Geophysics, Geosystems, 17(6), 2270–2285. https://doi.org/10.1002/2016gc006273
Ferrier, K. L., & West, N. (2017). Responses of chemical erosion rates to transient perturbations in physical erosion rates, and implications for
relationships between chemical and physical erosion rates in regolith-mantled hillslopes. Earth and Planetary Science Letters, 474, 447–456.
https://doi.org/10.1016/j.epsl.2017.07.002
Fletcher, S. J. (2019). Semi-Lagrangian advection methods and their applications in geoscience. Elsevier. https://doi.org/10.1016/C2018-0-02183-0
Forte, A. M., Yanites, B. J., & Whipple, K. X. (2016). Complexities of landscape evolution during incision through layered stratigraphy with
contrasts in rock strength. Earth Surface Processes and Landforms, 41(12), 1736–1757. https://doi.org/10.1002/esp.3947
Fritsch, F. N., & Carlson, R. E. (1980). Monotone piecewise cubic interpolation. SIAM Journal on Numerical Analysis, 17(2), 238–246. https://
doi.org/10.1137/0717021
Gilbert, G. K. (1877). Report on the Geology of the Henry Mountains. US Government Printing Office.
Godard, V., Hippolyte, J. C., Cushing, E., Espurt, N., Fleury, J., Bellier, O., etal. (2020). Hillslope denudation and morphologic response to a rock
uplift gradient. Earth Surface Dynamics, 8(2), 221–243. https://doi.org/10.5194/esurf-8-221-2020
Granger, D. E., Kirchner, J. W., & Finkel, R. (1996). Spatially averaged long-term erosion rates measured from in situ-produced cosmogenic
nuclides in alluvial sediment. The Journal of Geology, 104(3), 249–257. https://doi.org/10.1086/629823
Grieve, S. W., Mudd, S. M., & Hurst, M. D. (2016). How long is a hillslope? Earth Surface Processes and Landforms, 41(8), 1039–1054. https://
doi.org/10.1002/esp.3884
Grischott, R., Kober, F., Lupker, M., Hippe, K., Ivy-Ochs, S., Hajdas, I., etal. (2017). Constant denudation rates in a high alpine catchment for
the last 6 kyrs. Earth Surface Processes and Landforms, 42(7), 1065–1077. https://doi.org/10.1002/esp.4070
Heimsath, A. M. (2006). Eroding the land: Steady-state and stochastic rates and processes through a cosmogenic lens. Special Paper of the
Geological Society of America, 415, 111–129. https://doi.org/10.1130/2006.2415(07)
Heimsath, A. M., Chappell, J., Spooner, N. A., & Questiaux, D. G. (2002). Creeping soil. Geology, 30(2), 111–114. https://doi.
org/10.1130/0091-7613(2002)030<0111:cs>2.0.co;2
Heimsath, A. M., Dietrich, W. E., Nishiizumi, K., & Finkel, R. C. (1997). The soil production function and landscape equilibrium. Nature,
388(6640), 358–361. https://doi.org/10.1038/41056
Heimsath, A. M., Dietrich, W. E., Nishiizumi, K., & Finkel, R. C. (1999). Cosmogenic nuclides, topography, and the spatial variation of soil
depth. Geomorphology, 27(1–2), 151–172. https://doi.org/10.1016/S0169-555X(98)00095-6
Heimsath, A. M., Dietrich, W. E., Nishiizumi, K., & Finkel, R. C. (2001). Stochastic processes of soil production and transport: Erosion rates,
topographic variation and cosmogenic nuclides in the Oregon Coast Range. Earth Surface Processes and Landforms, 26(5), 531–552. https://
doi.org/10.1002/esp.209
Hippe, K. (2017). Constraining processes of landscape change with combined in situ cosmogenic
14C-
10Be analysis. Quaternary Science Reviews,
173, 1–19. https://doi.org/10.1016/j.quascirev.2017.07.020
Hippe, K., Gordijn, T., Picotti, V., Hajdas, I., Jansen, J. D., Christl, M., etal. (2019). Fluvial dynamics and
14C-
10Be disequilibrium on the Bolivian
Altiplano. Earth Surface Processes and Landforms, 44(3), 766–780. https://doi.org/10.1002/esp.4529
Hippe, K., Jansen, J. D., Skov, D. S., Lupker, M., Ivy-Ochs, S., Kober, F., etal. (2021). Cosmogenic in situ
14C-
10Be reveals abrupt Late Holocene
soil loss in the Andean Altiplano. Nature Communications, 12(1), 1–9. https://doi.org/10.1038/s41467-021-22825-6
Hurst, M. D., Grieve, S. W., Clubb, F. J., & Mudd, S. M. (2019). Detection of channel-hillslope coupling along a tectonic gradient. Earth and
Planetary Science Letters, 522, 30–39. https://doi.org/10.1016/j.epsl.2019.06.018
Hurst, M. D., Mudd, S. M., Walcott, R., Attal, M., & Yoo, K. (2012). Using hilltop curvature to derive the spatial distribution of erosion rates.
Journal of Geophysical Research, 117(F2), F02017. https://doi.org/10.1029/2011jf002057
Kirchner, J. W., & Ferrier, K. L. (2013). Mainly in the plain. Nature, 495(7441), 318–319. https://doi.org/10.1038/495318a
Knudsen, M. F., Egholm, D. L., & Jansen, J. D. (2019). Time-integrating cosmogenic nuclide inventories under the inf luence of variable erosion,
exposure, and sediment mixing. Quaternary Geochronology, 51, 110–119. https://doi.org/10.1016/j.quageo.2019.02.005
Lal, D. (1991). Cosmic ray labeling of erosion surfaces: In situ nuclide production rates and erosion models. Earth and Planetary Science Letters,
104(2–4), 424–439. https://doi.org/10.1016/0012-821X(91)90220-C
Larsen, I. J., Almond, P.C., Eger, A., Stone, J. O., Montgomery, D. R., & Malcolm, B. (2014). Rapid soil production and weathering in the
Southern Alps, New Zealand. Science, 343(6171), 637–640. https://doi.org/10.1126/science.1244908
Lifton, N., Sato, T., & Dunai, T. J. (2014). Scaling in situ cosmogenic nuclide production rates using analytical approximations to atmospheric
cosmic-ray fluxes. Earth and Planetary Science Letters, 386, 149–160. https://doi.org/10.1016/j.epsl.2013.10.052
Lucas, T. R. (1974). Error bounds for interpolating cubic splines under various end conditions. SIAM Journal on Numerical Analysis, 11(3),
569–584. https://doi.org/10.1137/0711049
Lucas, Y. (2001). The role of plants in controlling rates and products of weathering: Importance of biological pumping. Annual Review of Earth
and Planetary Sciences, 29(1), 135–163. https://doi.org/10.1146/annurev.earth.29.1.135
Marrero, S. M., Phillips, F. M., Borchers, B., Lifton, N., Aumer, R., & Balco, G. (2016). Cosmogenic nuclide systematics and the CRONUScalc
program. Quaternary Geochronology, 31, 160–187. https://doi.org/10.1016/j.quageo.2015.09.005
Marshall, J. A., Roering, J. J., Gavin, D. G., & Granger, D. E. (2017). Late Quaternary climatic controls on erosion rates and geomorphic
processes in western Oregon, USA. Bulletin, 129(5–6), 715–731. https://doi.org/10.1130/B31509.1
Mudd, S. M. (2017). Detection of transience in eroding landscapes. Earth Surface Processes and Landforms, 42(1), 24–41. https://doi.
org/10.1002/esp.3923
Mudd, S. M., & Furbish, D. J. (2007). Responses of soil-mantled hillslopes to transient channel incision rates. Journal of Geophysical Research,
112(F3), F03S18. https://doi.org/10.1029/2006JF000516
Mudd, S. M., Harel, M. A., Hurst, M. D., Grieve, S. W., & Marrero, S. M. (2016). The CAIRN method: Automated, reproducible calcula-
tion of catchment-averaged denudation rates from cosmogenic nuclide concentrations. Earth Surface Dynamics, 4(3), 655–674. https://doi.
org/10.5194/esurf-4-655-2016
Journal of Geophysical Research: Earth Surface
REED ETAL.
10.1029/2023JF007201
19 of 20
Neely, A. B., Bookhagen, B., & Burbank, D. W. (2017). An automated knickzone selection algorithm (KZ-Picker) to analyze transient landscapes:
Calibration and validation. Journal of Geophysical Research: Earth Surface, 122(6), 1236–1261. https://doi.org/10.1002/2017jf004250
Ott, R., Gallen, S. F., & Helman, D. (2023). Erosion and weathering in carbonate regions reveal climatic and tectonic drivers of carbonate land-
scape evolution. Earth Surface Dynamics, 11(2), 247–257. https://doi.org/10.5194/esurf-11-247-2023
Ott, R. F., Gallen, S. F., & Granger, D. E. (2022). Cosmogenic nuclide weathering biases: Corrections and potential for denudation and weathering
rate measurements. Geochronology, 4(2), 455–470. https://doi.org/10.5194/gchron-4-455-2022
Pelletier, J. D. (2012). Fluvial and slope-wash erosion of soil-mantled landscapes: Detachment-or transport-limited? Earth Surface Processes and
Landforms, 37(1), 37–51. https://doi.org/10.1002/esp.2187
Perron, J. T. (2011). Numerical methods for nonlinear hillslope transport laws. Journal of Geophysical Research, 116(F2). https://doi.
org/10.1029/2010JF001801
Perron, J. T. (2017). Climate and the pace of erosional landscape evolution. Annual Review of Earth and Planetary Sciences, 45(1), 561–591.
https://doi.org/10.1146/annurev-earth-060614-105405
Perron, J. T., Dietrich, W. E., & Kirchner, J. W. (2008). Controls on the spacing of first-order valleys. Journal of Geophysical Research, 113(F4),
F04016. https://doi.org/10.1029/2007jf000977
Perron, J. T., Kirchner, J. W., & Dietrich, W. E. (2009). Formation of evenly spaced ridges and valleys. Nature, 460(7254), 502–505. https://doi.
org/10.1038/nature08174
Perron, J. T., Richardson, P.W., Ferrier, K. L., & Lapôtre, M. (2012). The root of branching river networks. Nature, 492(7427), 100–103. https://
doi.org/10.1038/nature11672
Portenga, E. W., & Bierman, P.R. (2011). Understanding Earth’s eroding surface with
10Be. Geological Society of America Today, 21(8), 4–10.
https://doi.org/10.1130/G111A.1
Reed, M. M., Ferrier, K. L., & Perron, J. T. (2023). CosmoTadpole 1.0 [Software]. Zenodo. https://doi.org/10.5281/zenodo.8256787
Richardson, P.W., Perron, J. T., Miller, S. R., & Kirchner, J. W. (2020). Modeling the formation of topographic asymmetry by aspect-dependent
erosional processes and lateral channel migration. Journal of Geophysical Research: Earth Surface, 125(7), e2019JF005377. https://doi.
org/10.1029/2019JF005377
Richardson, P.W., Perron, J. T., & Schurr, N. D. (2019). Influences of climate and life on hillslope sediment transport. Geology, 47(5), 423–426.
https://doi.org/10.1130/G45305.1
Riebe, C. S., & Granger, D. E. (2013). Quantifying effects of deep and near-surface chemical erosion on cosmogenic nuclides in soils, saprolite,
and sediment. Earth Surface Processes and Landforms, 38(5), 523–533. https://doi.org/10.1002/esp.3339
Riebe, C. S., Kirchner, J. W., & Finkel, R. C. (2003). Long-term rates of chemical weathering and physical erosion from cosmogenic nuclides and
geochemical mass balance. Geochimica et Cosmochimica Acta, 67(22), 4411–4427. https://doi.org/10.1016/S0016-7037(03)00382-X
Riebe, C. S., Kirchner, J. W., & Finkel, R. C. (2004). Erosional and climatic effects on long-term chemical weathering rates in granitic landscapes
spanning diverse climate regimes. Earth and Planetary Science Letters, 224(3–4), 547–562. https://doi.org/10.1016/j.epsl.2004.05.019
Riebe, C. S., Kirchner, J. W., & Granger, D. E. (2001). Quantifying quartz enrichment and its consequences for cosmogenic measurements of
erosion rates from alluvial sediment and regolith. Geomorphology, 40(1–2), 15–19. https://doi.org/10.1016/s0169-555x(01)00031-9
Roering, J. J., Kirchner, J. W., & Dietrich, W. E. (1999). Evidence for nonlinear, diffusive sediment transport on hillslopes and implications for
landscape morphology. Water Resources Research, 35(3), 853–870. https://doi.org/10.1029/1998WR900090
Roering, J. J., Kirchner, J. W., & Dietrich, W. E. (2001). Hillslope evolution by nonlinear, slope-dependent transport: Steady state morphology
and equilibrium adjustment timescales. Journal of Geophysical Research, 106(B8), 16499–16513. https://doi.org/10.1029/2001JB000323
Schide, K., Gallen, S. F., & Lupker, M. (2022). Modelling the systematics of cosmogenic nuclide signals in fluvial sediments following extreme
events. Earth Surface Processes and Landforms, 47(9), 2325–2340. https://doi.org/10.1002/esp.5381
Skov, D. S., Egholm, D. L., Jansen, J. D., Sandiford, M., & Knudsen, M. F. (2019). Detecting landscape transience with in situ cosmogenic
14C
and
10Be. Quaternary Geochronology, 54, 101008. https://doi.org/10.1016/j.quageo.2019.101008
Small, E. E., Anderson, R. S., & Hancock, G. S. (1999). Estimates of the rate of regolith production using
10Be and
26Al from an alpine hillslope.
Geomorphology, 27(1–2), 131–150. https://doi.org/10.1016/S0169-555X(98)00094-4
Spiegelman, M., & Katz, R. F. (2006). A semi-Lagrangian Crank-Nicolson algorithm for the numerical solution of advection-diffusion problems.
Geochemistry, Geophysics, Geosystems, 7(4). https://doi.org/10.1029/2005gc001073
Sweeney, K. E., Roering, J. J., & Ellis, C. (2015). Experimental evidence for hillslope control of landscape scale. Science, 349(6243), 51–53.
https://doi.org/10.1126/science.aab0017
Theodoratos, N., & Kirchner, J. W. (2020). Dimensional analysis of a landscape evolution model with incision threshold. Earth Surface Dynam-
ics, 8(2), 505–526. https://doi.org/10.5194/esurf-8-505-2020
Torres, M. A., West, A. J., & Li, G. (2014). Sulphide oxidation and carbonate dissolution as a source of CO2 over geological timescales. Nature,
507(7492), 346–349. https://doi.org/10.1038/nature13030
von Blanckenburg, F. (2006). The control mechanisms of erosion and weathering at basin scale from cosmogenic nuclides in river sediment. Earth
and Planetary Science Letters, 242(3–4), 224–239. https://doi.org/10.1016/j.epsl.2005.11.017
Walker, J. C., Hays, P.B., & Kasting, J. F. (1981). A negative feedback mechanism for the long-term stabilization of Earth's surface temperature.
Journal of Geophysical Research, 86(C10), 9776–9782. https://doi.org/10.1029/jc086ic10p09776
West, A. J., Galy, A., & Bickle, M. (2005). Tectonic and climatic controls on silicate weathering. Earth and Planetary Science Letters, 235(1–2),
211–228. https://doi.org/10.1016/j.epsl.2005.03.020
Willenbring, J. K., Codilean, A. T., & McElroy, B. (2013). Earth is (mostly) flat: Appor tionment of the flux of continental sediment over millen-
nial time scales. Geology, 41(3), 343–346. https://doi.org/10.1130/g33918.1
Willett, S. D. (1999). Orogeny and orography: The effects of erosion on the structure of mountain belts. Journal of Geophysical Research,
104(B12), 28957–28981. https://doi.org/10.1029/1999jb900248
Willett, S. D., McCoy, S. W., Perron, J. T., Goren, L., & Chen, C. Y. (2014). Dynamic reorganization of river basins. Science, 343(6175), 1248765.
https://doi.org/10.1126/science.1248765
Yanites, B. J., Tucker, G. E., & Anderson, R. S. (2009). Numerical and analytical models of cosmogenic radionuclide dynamics in
landslide-dominated drainage basins. Journal of Geophysical Research, 114(F1), F01007. https://doi.org/10.1029/2008JF001088
Yoo, K., Amundson, R., Heimsath, A. M., Dietrich, W. E., & Brimhall, G. H. (2007). Integration of geochemical mass balance with sediment
transport to calculate rates of soil chemical weathering and transport on hillslopes. Journal of Geophysical Research, 112(F2), F02013. https://
doi.org/10.1029/2005JF000402
Zondervan, J. R., Stokes, M., Boulton, S. J., Telfer, M. W., & Mather, A. E. (2020). Rock strength and structural controls on fluvial erodibil-
ity: Implications for drainage divide mobility in a collisional mountain belt. Earth and Planetary Science Letters, 538, 116221. https://doi.
org/10.1016/j.epsl.2020.116221
Journal of Geophysical Research: Earth Surface
REED ETAL.
10.1029/2023JF007201
20 of 20
References From the Supporting Information
Barnhart, K. R., Glade, R. C., Shobe, C. M., & Tucker, G. E. (2019). Terrainbento 1.0: A Python package for multi-model analysis in long-term
drainage basin evolution. Geoscientific Model Development, 12(4), 1267–1297. https://doi.org/10.5194/gmd-12-1267-2019
Chamberlain, C. P., Waldbauer, J. R., & Jacobson, A. D. (2005). Strontium, hydrothermal systems and steady-state chemical weathering in active
mountain belts. Earth and Planetary Science Letters, 238(3–4), 351–366. https://doi.org/10.1016/j.epsl.2005.08.005
Fagherazzi, S., Howard, A. D., & Wiberg, P.L. (2002). An implicit finite difference method for drainage basin evolution. Water Resources
Research, 38(7), 21−1–21−5. https://doi.org/10.1029/2001WR000721
Howard, A. D. (1994). A detachment-limited model of drainage basin evolution. Water Resources Research, 30(7), 2261–2285. https://doi.
org/10.1029/94wr00757
Lague, D. (2014). The stream power river incision model: Evidence, theory and beyond. Earth Surface Processes and Landforms, 39(1), 38–61.
https://doi.org/10.1002/esp.3462
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
Carbonate rocks are highly reactive and can have higher ratios of chemical weathering to total denudation relative to most other rock types. Their chemical reactivity affects the first-order morphology of carbonate-dominated landscapes and their climate sensitivity to weathering. However, there have been few efforts to quantify the partitioning of denudation into mechanical erosion and chemical weathering in carbonate landscapes such that their sensitivity to changing climatic and tectonic conditions remains elusive. Here, we compile bedrock and catchment-averaged cosmogenic calcite-36 Cl denudation rates and compare them to weathering rates derived from stream water chemistry from the same regions. Local bedrock denudation and weathering rates are comparable, ∼ 20-40 mm ka −1 , whereas catchment-averaged denudation rates are ∼ 2.7 times higher. The discrepancy between bedrock and catchment-averaged denudation is 5 times lower compared to silicate-rich rocks, illustrating that elevated weathering rates make denudation more spatially uniform in carbonate-dominated landscapes. Catchment-averaged denudation rates correlate well with topographic relief and hillslope gradients, and moderate correlations with runoff can be explained by concurrent increases in weathering rates. Comparing denudation rates with weathering rates shows that mechanical erosion processes contribute ∼ 50 % of denudation in southern France and ∼ 70 % in Greece and Israel. Our results indicate that the partitioning between largely slope-independent chemical weathering and slope-dependent mechanical erosion varies based on climate and tectonics and impacts the landscape morphology. This leads us to propose a conceptual model whereby in humid, slowly uplifting regions, carbonates are associated with low-lying, flat topography because slope-independent chemical weathering dominates denudation. In contrast, in arid climates with rapid rock uplift rates, carbonate rocks form steep mountains that facilitate rapid, slope-dependent mechanical erosion required to compensate for inefficient chemical weathering and runoff loss to groundwater systems. This result suggests that carbonates represent an end member for interactions between climate, tectonics, and lithology.
Article
Full-text available
OCTOPUS v.2 is an Open Geospatial Consortium (OGC) compliant web-enabled database that allows users to visualise, query, and download cosmogenic radionuclide, luminescence, and radiocarbon ages and denudation rates associated with erosional landscapes, Quaternary depositional landforms, and archaeological records, along with ancillary geospatial (vector and raster) data layers. The database follows the FAIR (Findability, Accessibility, Interoperability, and Reuse) data principles and is based on open-source software deployed on the Google Cloud Platform. Data stored in the database can be accessed via a custom-built web interface and via desktop geographic information system (GIS) applications that support OGC data access protocols. OCTOPUS v.2 hosts five major data collections. CRN Denudation and ExpAge consist of published cosmogenic 10Be and 26Al measurements in modern fluvial sediment and glacial samples respectively. Both collections have a global extent; however, in addition to geospatial vector layers, CRN Denudation also incorporates raster layers, including a digital elevation model, gradient raster, flow direction and flow accumulation rasters, atmospheric pressure raster, and CRN production scaling and topographic shielding factor rasters. SahulSed consists of published optically stimulated luminescence (OSL) and thermoluminescence (TL) ages for fluvial, aeolian, and lacustrine sedimentary records across the Australian mainland and Tasmania. SahulArch consists of published OSL, TL, and radiocarbon ages for archaeological records, and FosSahul consists of published late-Quaternary records of direct and indirect non-human vertebrate (mega)fauna fossil ages that have been systematically quality rated. Supporting data are comprehensive and include bibliographic, contextual, and sample-preparation- and measurement-related information. In the case of cosmogenic radionuclide data, OCTOPUS also includes all necessary information and input files for the recalculation of denudation rates using the open-source program CAIRN. OCTOPUS v.2 and its associated data curation framework allow for valuable legacy data to be harnessed that would otherwise be lost to the research community. The database can be accessed at https://octopusdata.org (last access: 1 July 2022). The individual data collections can also be accessed via their respective digital object identifiers (DOIs) (see Table 1).
Article
Full-text available
Bedrock landslides shape topography and mobilize large volumes of sediment. Yet, interactions between landslide‐produced sediment and fluvial systems that together govern large‐scale landscape evolution are not well understood. To explain morphological patterns observed in steep, landslide‐prone terrain, we explicitly model stochastic landsliding and associated sediment dynamics. The model accounts for several common landscape features such as slope frequency distributions, which include values in excess of regional stability limits, quasi‐planar hillslopes decorated with straight, closely spaced channel‐like features, and accumulation of sediment in valley networks rather than on hillslopes. Stochastic landsliding strongly affects the magnitude and timing of sediment supply to the fluvial system. We show that intermittent sediment supply is ultimately reflected in topography. At dynamic equilibrium, landslide‐derived sediment pulses generate persistent landscape dynamism through the formation and breaching of landslide dams and epigenetic gorges as landslides force shifts in channel positions. Our work highlights the importance of interactions between landslides and sediment dynamics that ultimately control landscape‐scale response to environmental change.
Article
Full-text available
Cosmogenic radionuclides (CRNs) are the standard tool to derive centennial-to-millennial timescale denudation rates; however, it has been demonstrated that chemical weathering in some settings can bias CRNs as a proxy for landscape denudation. Currently, studies investigating CRN weathering biases have mostly focused on the largely insoluble target mineral quartz in felsic lithologies. Here, we examine the response of CRN build-up for both soluble and insoluble target minerals under different weathering scenarios. We assume a simple box model in which bedrock is converted to a well-mixed regolith at a constant rate, and denudation occurs by regolith erosion and weathering either in the regolith or along the regolith–bedrock interface, as is common in carbonate bedrock. We show that weathering along the regolith–bedrock interface increases CRN concentrations compared to a no-weathering case and how independently derived weathering rates or degrees can be used to correct for this bias. If weathering is concentrated within the regolith, insoluble target minerals will have a longer regolith residence time and higher nuclide concentration than soluble target minerals. This bias can be identified and corrected using paired-nuclide measurements of minerals with different solubility coupled with knowledge of either the bedrock or regolith mineralogy to derive denudation and long-term weathering rates. Similarly, single-nuclide measurements on soluble or insoluble minerals can be corrected to determine denudation rates if a weathering rate and compositional data are available. Our model highlights that for soluble target minerals, the relationship between nuclide accumulation and denudation is not monotonic. We use this understanding to map the conditions of regolith mass, weathering, and denudation rates at which weathering corrections for cosmogenic nuclides become large and ambiguous, as well as identify environments in which the bias is mostly negligible and CRN concentrations reliably reflect landscape denudation. We highlight how measurements of CRNs from soluble target minerals, coupled with bedrock and regolith mineralogy, can help to expand the range of landscapes for which centennial-to-millennial timescale denudation and weathering rates can be obtained.
Article
Full-text available
Soil sustainability is reflected in a long-term balance between soil production and erosion for a given climate and geology. Here we evaluate soil sustainability in the Andean Altiplano where accelerated erosion has been linked to wetter climate from 4.5 ka and the rise of Neolithic agropastoralism in the millennium that followed. We measure in situ cosmogenic 14C directly on cultivated hilltops to quantify late Holocene soil loss, which we compare with background soil production rates determined from cosmogenic 26Al and 10Be. Our Monte Carlo-based inversion method identifies two scenarios to account for our data: an increase in erosion rate by 1–2 orders of magnitude between ~2.6 and 1.1 ka, or a discrete event stripping ~1–2m of soil between ~1.9 and 1.1 ka. Coupled environmental and cultural factors in the Late Holocene signaled the onset of the pervasive human imprint in the Andean Altiplano seen today.
Article
Full-text available
Chemical erosion is of wide interest due to its influence on topography, nutrient supply to streams and soils, sediment composition, and Earth's climate. While controls on chemical erosion rate have been studied extensively in steady‐state models, few studies have explored the controls on chemical erosion rate during transient responses to external perturbations. Here we develop a numerical model for the coevolution of soil‐mantled topography, soil thickness, and soil mineralogy, and we use it to simulate responses to step changes in rates of rock uplift, soil production, soil transport, and mineral dissolution. These simulations suggest that tectonic and climatic perturbations can generate responses in soil chemical erosion rate that differ in speed, magnitude, and spatial pattern and that climatic and tectonic perturbations may impart distinct signatures on hillslope mass fluxes, soil chemistry, and sediment composition. The response time of chemical erosion rate is dominantly controlled by hillslope length and is secondarily modulated by rates of rock uplift, soil production, transport, and mineral dissolution. This strong dependence on drainage density implies that a landscape's chemical erosion response should depend on the relative efficiencies of river incision and soil transport and thus may be mediated by climatic and biological factors. The simulations further suggest that the timescale of the hillslope response may be long relative to that of river channel profiles, implying that chemical erosion response times may be limited more by the sluggishness of the hillslopes than by the rate of signal propagation through river channel profiles.
Article
Full-text available
The ability of erosional processes to incise into a topographic surface can be limited by a threshold. Incision thresholds affect the topography of landscapes and their scaling properties and can introduce nonlinear relations between climate and erosion with notable implications for long-term landscape evolution. Despite their potential importance, incision thresholds are often omitted from the incision terms of landscape evolution models (LEMs) to simplify analyses. Here, we present theoretical and numerical results from a dimensional analysis of an LEM that includes terms for threshold-limited stream-power incision, linear diffusion, and uplift. The LEM is parameterized by four parameters (incision coefficient and incision threshold, diffusion coefficient, and uplift rate). The LEM's governing equation can be greatly simplified by recasting it in a dimensionless form that depends on only one dimensionless parameter, the incision-threshold number Nθ. This dimensionless parameter is defined in terms of the incision threshold, the incision coefficient, and the uplift rate, and it quantifies the reduction in the rate of incision due to the incision threshold relative to the uplift rate. Being the only parameter in the dimensionless governing equation, Nθ is the only parameter controlling the evolution of landscapes in this LEM. Thus, landscapes with the same Nθ will evolve geometrically similarly, provided that their boundary and initial conditions are normalized according to appropriate scaling relationships, as we demonstrate using a numerical experiment. In contrast, landscapes with different Nθ values will be influenced to different degrees by their incision thresholds. Using results from a second set of numerical simulations, each with a different incision-threshold number, we qualitatively illustrate how the value of Nθ influences the topography, and we show that relief scales with the quantity Nθ+1 (except where the incision threshold reduces the rate of incision to zero).
Article
Full-text available
Some landscapes exhibit the intriguing characteristic that the steepness of hillslopes varies systematically with the direction they face, even where there is no bias introduced by bedrock structure. This topographic asymmetry has inspired numerous explanations. In the simplest scenario, insolation‐driven microclimatic differences lead to different erosion rates on opposing slopes and topographic asymmetry develops. Alternatively, lateral channel migration and the corresponding steepening of undercut slopes have also been suggested as a dominant cause of topographic asymmetry. To examine these proposed origins of asymmetric topography, we adapted a numerical landscape evolution model to include lateral channel migration as well as aspect‐induced variations in soil creep, regolith strength, and runoff. We compared the resulting topography and erosional response produced by each mechanism and found that the model with lateral channel migration produces unique signatures in erosion rates and the ridgetop Laplacian of elevation. To further investigate lateral channel migration, we developed a model of hillslope profile evolution in which sediment fluxes from opposing slopes control lateral channel migration rate and direction. We find that topographic asymmetry in this system can be self‐sustaining. All else being equal, we find that the aspect‐dependent models with variable regolith strength or runoff predict that less topographic asymmetry should develop as slope length decreases. These two models also predict that weaker asymmetry should develop as rock strength increases, whereas the lateral channel migration model predicts that greater rock strength should lead to weaker asymmetry only if lateral channel migration efficiency increases faster than fluvial incision efficiency.
Article
Full-text available
Documenting the spatial variability of tectonic processes from topography is routinely undertaken through the analysis of river profiles, since a direct relationship between fluvial gradient and rock uplift has been identified by incision models. Similarly, theoretical formulations of hillslope profiles predict a strong dependence on their base-level lowering rate, which in most situations is set by channel incision. However, the reduced sensitivity of near-threshold hillslopes and the limited availability of high-resolution topographic data has often been a major limitation for their use to investigate tectonic gradients. Here we combined high-resolution analysis of hillslope morphology and cosmogenic-nuclide-derived denudation rates to unravel the distribution of rock uplift across a blind thrust system at the southwestern Alpine front in France. Our study is located in the Mio-Pliocene Valensole molassic basin, where a series of folds and thrusts has deformed a plateau surface. We focused on a series of catchments aligned perpendicular to the main structures. Using a 1m lidar digital terrain model, we extracted hillslope topographic properties such as hilltop curvature CHT and nondimensional erosion rates E�. We observed systematic variation of these metrics coincident with the location of a major underlying thrust system identified by seismic surveys. Using a simple deformation model, the inversion of the E� pattern allows us to propose a location and dip for a blind thrust, which are consistent with available geological and geophysical data. We also sampled clasts from eroding conglomerates at several hilltop locations for 10Be and 26Al concentration measurements. Calculated hilltop denudation rates range from 40 to 120mmkyr􀀀1. These denudation rates appear to be correlated with E� and CHT that were extracted from the morphological analysis, and these rates are used to derive absolute estimates for the fault slip rate. This high-resolution hillslope analysis allows us to resolve short-wavelength variations in rock uplift that would not be possible to unravel using commonly used channel-profile-based methods. Our joint analysis of topography and geochronological data supports the interpretation of active thrusting at the southwestern Alpine front, and such approaches may bring crucial complementary constraints to morphotectonic analysis for the study of slowly slipping faults.
Article
The effect of punctuated mass wasting events on longer‐term erosion rates is not fully understood, and yet it is key to quantifying sediment generation, source‐to‐sink dynamics, and landscape evolution in active orogens. The measurement of terrestrial cosmogenic nuclides (TCN) in river sediments is a common method for determining basin‐averaged erosion rates over centennial‐to‐millennial timescales and is often used to compare erosional processes between catchments. However, these comparisons often overlook the role of landsliding rates and their spatial distribution in the measurement and potential variability of TCN signals. While it is widely accepted that basin‐scale perturbations should temporarily dilute TCN concentrations as landsliding mobilizes new, low concentration material, the impact of the catastrophic release of hillslope sediment caused by a single event on TCN signatures has not yet been systematically investigated. In this modeling study, we use a catchment in central Nepal to build upon previous modeling efforts to consider how TCNs are recorded in landscapes with varying erosion rates, landsliding rates, and spatial distribution of landslides. We then use the 25 April 2015 Mw 7.8 Gorkha earthquake, Nepal, as a case study to investigate how perturbations like earthquakes are recorded in TCN time series and transferred to and ultimately archived in the sedimentary record. We find that the likelihood of a perturbation to be measured by TCN dilution is based on a multitude of factors, including background erosion rates, long‐term landsliding rates, and the connectivity of newly released material to the fluvial system. Especially in landscapes like the central Himalaya with high background erosion and landsliding rates, changes in detrital TCN concentrations are not a reliable indicator of an upstream perturbation, nor should we expect a clear dilution signal following a major event. Our modeling results emphasize that TCN dilution is not a universal characteristic of high‐magnitude landslide‐triggering events.