Brady A. Flinchum's research while affiliated with Clemson University and other places

What is this page?


This page lists the scientific contributions of an author, who either does not have a ResearchGate profile, or has not yet added these contributions to their profile.

It was automatically created by ResearchGate to create a record of this author's body of work. We create such pages to advance our goal of creating and maintaining the most comprehensive scientific repository possible. In doing so, we process publicly available (personal) data relating to the author as a member of the scientific community.

If you're a ResearchGate member, you can follow this page to keep up with this author's work.

If you are this author, and you don't want us to display this page anymore, please let us know.

Publications (30)


Sensing a Connection: Tree Distribution is Influenced by Deep Critical Zone Structure
  • Preprint

May 2024

·

5 Reads

Brady A Flinchum

·

·

This study explores the impact of deep ( >5 m) critical zone (CZ) architecture on vegetation distribution in a semi-arid snow-dominated climate. Utilizing seismic refraction surveys, we identified a significant correlation between saprolite thickness and LiDAR-derived canopy heights (R²=0.47). We argue that CZ structure, specifically shallow fractured bedrock under valley bottoms, redirects groundwater to locations where trees are established—suggesting they are located in specific locations with access to nutrients and water. This work provides a unique spatially exhaustive perspective and adds to growing evidence that in addition to other factors such as slope, aspect, and climate, deep CZ structure plays a vital role in ecosystem development and resilience.

Share

Sensing a Connection: Tree Distribution is Influenced by Deep Critical Zone Structure

March 2024

·

20 Reads

This study explores the impact of deep ( >5 m) critical zone (CZ) architecture on vegetation distribution in a semi-arid snow-dominated climate. Utilizing seismic refraction surveys, we identified a significant correlation between saprolite thickness and LiDAR-derived canopy heights (R²=0.47). We argue that CZ structure, specifically shallow fractured bedrock under valley bottoms, redirects groundwater to locations where trees are established—suggesting they are located in specific locations with access to nutrients and water. This work provides a unique spatially exhaustive perspective and adds to growing evidence that in addition to other factors such as slope, aspect, and climate, deep CZ structure plays a vital role in ecosystem development and resilience.


Hill shade map of the BW study site modified from Flinchum et al. (2022), where various seismic refraction profiles collected by WyCEHG are demarcated with black lines. The seismic refraction profile used in this study is L29. The yellow dot indicates where x = 0 along the transect. The red stars show the locations of boreholes that were drilled and logged. In this work, we show borehole logs from BW1 and BW4, which are located directly on L29.
Left panel: a filtered (5–56 Hz) shot gather from a source located 0 m along the transect. Upper right panel: the geophone recording for the instrument located 120 m along the transect. The first arrival, p‐wave coda, higher mode surface wave, and fundamental mode surface wave are highlighted. Each of these four phases is back‐projected to construct two of the sensitivity kernels shown in Figure 3 (eight in total). One sensitivity kernel is with respect to p‐wave velocity (Vp) while the other is with respect to shear‐wave velocity (Vs). Bottom right panel: a spectrogram of the trace in the upper right panel, showing higher frequencies in the p‐wave first arrival and lower frequencies in the Rayleigh wave.
Sensitivity kernels with respect to each model parameter (Vp and Vs) for each of the four highlighted phases in the upper right panel of Figure 2. The first column corresponds to sensitivity with respect to Vp, while the second column corresponds to sensitivity with respect to Vs. The first row shows the sensitivity of the first arrival, the second row shows the sensitivity of the p‐wave coda, the third row shows the sensitivity of the higher mode surface wave, and the last row shows the sensitivity of the fundamental mode surface wave.
A flow chart of our FWT strategy showing both preliminary steps and FWT stages. The ray‐based travel time tomography step was performed by Flinchum et al., 2022. We then go onto estimate source‐time functions using the method described in subsection 3.3. The wave equation dispersion inversion step is discussed in detail in Section 3.5. FWT steps are explained in detail in Section 3.6. Please note that regularization during FWT is enforced by smoothing gradients with a Gaussian filter. During the surface wave FWT step, we allow model updates in Vp and Vs, although it should be noted that the model updates for Vp are relatively small. During the body wave FWT step, we only allow Vp to be updated and decrease the smoothing radius of the Gaussian filter every time higher frequency data is included. For the surface wave step, we use 15 LBFGS iterations for each frequency band. For the body wave step, we use 20 iterations for the first frequency band, and 40 iterations each for the last two bands.
Results from the synthetic FWT experiment. The left column has Vs models and the right column has Vp models. The first row shows the starting models, the second row shows the target models, and last row shows the inverted models. Velocity contours on the Vs and Vp models have intervals of 250 m/s and 500 m/s respectively.

+6

2D Near‐Surface Full‐Waveform Tomography Reveals Bedrock Controls on Critical Zone Architecture
  • Article
  • Full-text available

February 2024

·

201 Reads

·

2 Citations

Earth and Space Science

·

·

·

[...]

·

For decades, seismic imaging methods have been used to study the critical zone, Earth's thin, life‐supporting skin. The vast majority of critical zone seismic studies use traveltime tomography, which poorly resolves heterogeneity at many scales relevant to near‐surface processes, therefore limiting progress in critical zone science. Full‐waveform tomography can overcome this limitation by leveraging more seismic data and enhancing the resolution of geophysical imaging. In this study, we apply 2D full‐waveform tomography to match the phases of observed seismograms and elucidate previously undetected heterogeneity in the critical zone at a well‐studied catchment in the Laramie Range, Wyoming. In contrast to traveltime tomograms from the same data set, our results show variations in depth to bedrock ranging from 5 to 60 m over lateral scales of just tens of meters and image steep low‐velocity anomalies suggesting hydrologic pathways into the deep critical zone. Our results also show that areas with thick fractured bedrock layers correspond to zones of slightly lower velocities in the deep bedrock, while zones of high bedrock velocity correspond to sharp vertical transitions from bedrock to saprolite. By corroborating these findings with borehole imagery, we hypothesize that lateral changes in bedrock fracture density majorly impact critical zone architecture. Borehole data also show that our full‐waveform tomography results agree significantly better with velocity logs than previously published traveltime tomography models. Full‐waveform tomography thus appears unprecedentedly capable of imaging the spatially complex porosity structure crucial to critical zone hydrology and processes.

Download

(a) Hillshade LiDAR map showing the two profiles and borehole locations for the South Carolina Piedmont (SCP) site. (b) Hillshade LIDAR map showing the BW profile and borehole locations. Shot gathers and traces in panels c–h are from BW. (c)–(h) Examples of our P‐wave and S‐wave picks and data from vertical and horizontal shots. The red symbols are the P‐wave travel time picks, and the cyan symbols are the Sh‐wave travel time picks. (c) Stacked shot gather using vertical geophones and vertical source. (d) Traces are from 10 m. The left trace is from the vertical shot gather (panel c) and the right trace is from the horizontal shot gather (panel f). (e) Pre‐stacked traces from 10 m in the horizontal gather. The black trace represents the recording when swinging in direction 1 and the gray trace represents swinging in direction 2. The first arrival polarity is swapped indicating the S‐wave arrival. When these two traces are subtracted it produces the second trace in panel d. (f) Stacked (by subtracting opposite direction swings) shot gather using horizontal geophones and the horizontal source. (g) Same as panel d but the traces are extracted from 130 m. The left trace is from the vertical shot gather (panel c) and the right trace is from the horizontal shot gather (panel f). (h) Same as panel e but for the trace at 130 m.
Seismic refraction results. The left column (a)–(c) are from SCP L2, the middle column (d)–(f) are from SCP L1, and the right column (g)–(i) are from BW. The top row (a, d, and g) are final Vp models masked by ray coverage with the RMS fits shown in the lower right‐hand corner. The solid black contour is 1,200 m/s, and the dashed black contours are 1,100 and 1,300 m/s. The solid gray contour is 4,000 m/s while the dashed gray contours are 3,500 and 4,500 m/s. The Vs model results are the middle row (b, e, and h). The colormap has been rescaled, assuming a constant Vp/Vs value of 1.5. The velocity contours are from the Vp model. The bottom row is the Vp/Vs values. The colormap was chosen so that values with Vp/Vs values less than 1.4 (υ $\upsilon $ < 0) are shown as cooler colors and Vp/Vs values greater than 1.4 (υ $\upsilon $ > 0) are shown in warmer colors. Poisson's ratios are shown in the color scale using Equation 1.
(a) A comparison of SCP L2 and the sonic velocity logs. The thin black line is the sonic Vp. The thin gray line is the sonic Vs. The dashed blue line and corresponding envelope are the extracted Vp ± 1σ (Figure S5 in Supporting Information S1). The dashed orange line and corresponding envelope are the extracted Vs ± 1σ. The surface profiles only show where at least one ray passes through the Vp and Vs models. (b) The black curve is the Vp/Vs profile computed from the sonic logs. The dashed purple line is the Vp/Vs value extracted from the tomography model. The envelope is the uncertainty associated by taking the ratio using (Vp + 1σ)/(Vs − 1σ) and (Vp − 1σ)/(Vs + 1σ). (c) Sonic velocities, where Vs is plotted as a function of Vp for all available sonic logs. The yellow color is for SCP, and the brown color is for BW. Normalized Gaussian PDFs for Vs are shown on the right of the plot and Vp are plotted on top. Dashed lines are constant Vp/Vs values. (d) Normalized Gaussian PDFs of the distributions of Vp/Vs. Poisson's ratios are plotted on the top using Equation 1. The means and standard errors are reported in each plot. Vp/Vs values less than 1.4 (υ $\upsilon $ < 0) are marked by a gray box.
(a) SEM sections from a borehole ∼3 km Southwest of our location (Holbrook et al., 2019). The top panel shows an unweathered bedrock sample from 50 m depth and the bottom shows a weathered bedrock from 32 m depth. (b) SEM sections from a thermally treated granite from Griffiths et al. (2018). The top is the untreated sample, and the bottom plot is an SEM section after being heat treated to 450 C to induce microcracks. The samples' Vp, Vs, Vp/Vs values, and Poisson's ratios are reported on the panels. (c) Probability density functions of Vp for all data with Vp/Vs values less than 1.4 (υ $\upsilon $ < 0). The percentage data with values less than 1.4 are shown in each plot. Background colors are CZ structure based on Vp. White is saprolite (Vp ≤ 1,200 m/s), gray is fractured bedrock (1,200 m/s < Vp ≤ 4,000 m/s), and dark gray is unfractured bedrock (Vp > 4,000 m/s). (d) Vp/Vs values plotted as a function of Vp and colored by Vs. Gray shaded regions are the same as panel d with an additional gray rectangle to highlight Vp/Vs values less than 1.4. The top panel is for BW. The bottom panel is for SCP, which includes both SCP L1 and SCP L2. Squares are data from seismic refraction surveys, and circles are from sonic logs. The dashed black line is the theoretical Vp/Vs values assuming no porosity based on the mineralogic compositions from Flinchum et al. (2018a) and Flinchum et al. (2018b) for BW and Holbrook et al. (2019) for SCP (Supporting Information S1). The solid black line is the Vp/Vs value from the thermally treated granite shown in panel c. The dashed white line is our interpretation of the Vp/Vs value behavior as physical and chemical weathering creates porosity. The vertical white lines are approximate boundaries of change. The line at the highest P‐wave velocities marks the onset of weathering, and the line around 1,200 m/s marks the location where Vp/Vs values increase.
Low Vp/Vs Values as an Indicator for Fractures in the Critical Zone

January 2024

·

280 Reads

Plain Language Summary When a material is squeezed, the ratio between the change in height and width is described by an elastic parameter called Poisson's ratio. Most earth materials have a positive Poisson ratio, meaning the material will expand when squeezed (e.g., Playdough or wet sand). Materials with a negative Poisson's ratio rarely occurs naturally and will shrink in all directions when squeezed. Cork is a common material with a Poisson's ratio of approximately zero. Cork is ideal for bottling wine because its width does not change when pushing it into the bottle's narrow neck. Here we use surface‐based measurements to quantify Poisson's ratio from P‐wave (Vp) and S‐wave (vs.) velocities in the top 50 m of Earth's surface. Our results show an unexpected result—material in the CZ has a negative Poisson's ratio. We believe this unexpected behavior is caused by the combination of low effective pressures and small and irregular cracks created during rocks' transformation into soil. The cracks have a greater impact on the material's ability to resist compression. At the same time, most of the rock is still coherent and thus only experiences a minimal loss of shear strength.


Borehole Nuclear Magnetic Resonance Estimation of Specific Yield in a Fractured Granite Aquifer

November 2023

·

25 Reads

Ground Water

In this study, we introduce a novel field‐based method to estimate specific yield (S y ) in fractured, low‐porosity granite aquifers using borehole nuclear magnetic resonance (bNMR). This method requires collecting a bNMR survey immediately following a pump test, which dewaters the near‐borehole fractures. The residual water content measured from bNMR is interpreted as “bound” and represents the specific retention (S r ) while the water drained by the pump is the S y . The transverse relaxation cutoff time (T 2C ) is the length of time that partitions the total porosity measured by bNMR into S r and S y . When applying a calibrated T 2C , S y equals the bNMR total porosity minus S r ; thus, a calibrated T 2C is required to determine S y directly from NMR results. Based on laboratory experiments on sandstone cores, the default T 2C is 33 ms; however, its applicability to fractured granite aquifers is uncertain. The optimal T 2C based on our pumping test is 110 ± 25 ms. Applying this calibrated T 2C on a saturated, A‐type granite at our field site, we estimate the S y to be 0.012 ± 0.005 m ³ m ‐3 which is significantly different from the S y (0.021 ± 0.005 m ³ m ‐3 ) estimate using the default T 2C of 33 ms. This S y estimate falls within a range determined using traditional hydraulic testing at the same site. Using the conventional T 2C (33 ms) for fractured granite leads to an inaccurate S y ; therefore, it is essential to calibrate the bNMR T 2C for the local site conditions prior to estimating S y . This article is protected by copyright. All rights reserved.


Integrated hydrological modeling of the No-Name watershed, Medicine Bow Mountains, Wyoming (USA)Modélisation hydrologique intégrée du bassin versant de No-Name, Montagnes de Medicine Bow, Wyoming (Etats-Unis d’Amérique)Modelización hidrológica integrada de una cuenca en Medicine Bow Mountains, Wyoming (EE.UU.)美国怀俄明州Medicine Bow山脉无名流域的综合水文模拟Modelagem hidrológica integrada da bacia hidrográfica No-Name, Montanhas Medicine Bow, Wyoming (EUA)

September 2023

·

21 Reads

Hydrogeology Journal

Integrated modeling of headwater watersheds in mountain environments is often limited by the lack of hydrological characterization and monitoring data. For the No-Name watershed in the Medicine Bow Mountains in Wyoming (USA), this research integrates regional surface and subsurface hydrological and geophysical measurements to create three-dimensional integrated hydrological models with which interactions between surface water, soil water, and groundwater are elucidated. Data used to build and calibrate the integrated model include a digital elevation model (DEM), stream discharge at the outlet of the watershed, soil-moisture data, weather data, and geophysical surveys including seismic refraction, airborne resistivity, and nuclear magnetic resonance (NMR). Based on interpretation of geophysical measurements, subsurface hydrostratigraphy consists of a top unconsolidated layer, a middle layer of fractured granite and metamorphic bedrock, and a lower protolith. Given that both measurements and interpretations have uncertainty, a sensitivity analysis was carried out to evaluate conceptual model uncertainty, which suggests the following: (1) for predicting stream discharge at the No-Name outlet, the most influential parameters are the Manning coefficient, DEM, hydrostratigraphy and hydraulic conductivity, and land cover. Compared to a lower-resolution DEM, a LiDAR DEM can lead to more accurate predictions of the stream discharge and stream elevation profile. (2) For predicting soil moisture, the most influential parameters are hydrostratigraphy and the associated hydraulic conductivities and porosities. (3) Based on a calibration exercise, the likely values for subsurface hydraulic conductivity at No-Name are ~10–5 m/s (the unconsolidated layer), ~10–6 m/s (fractured bedrock), and ~10–6 m/s or lower (protolith).


Near-Surface Full-Waveform Inversion Reveals Bedrock Controls on Critical Zone Architecture

August 2023

·

388 Reads

For decades, seismic imaging methods have been used to study the critical zone, Earth’s thin, life-supporting skin. The vast majority of critical zone seismic studies use traveltime tomography, which poorly resolves heterogeneity at many scales relevant to near-surface processes, therefore, limiting progress in critical zone science. Full-waveform inversion can overcome this limitation by leveraging more of the seismic waveform and enhancing the resolution of geophysical imaging. In this study, we apply full-waveform inversion to elucidate previously undetected heterogeneity in the critical zone at a well-studied catchment in the Laramie Range, Wyoming. In contrast to traveltime tomograms from the same data set, our results show variations in depth to bedrock ranging from 5 to 60 meters over lateral scales of just tens of meters and image steep low-velocity anomalies suggesting hydrologic pathways into the deep critical zone. Our results also show that areas with thick fractured bedrock layers correspond to zones of slightly lower velocities in the deep bedrock, while zones of high bedrock velocity correspond to sharp vertical transitions from bedrock to saprolite. By corroborating these findings with borehole imagery, we hypothesize that lateral changes in bedrock fracture density majorly impact critical zone architecture. Borehole data also show that our full-waveform inversion results agree significantly better with velocity logs than previously published traveltime tomography models. Full-waveform inversion thus appears unprecedently capable of imaging the spatially complex porosity structure crucial to critical zone hydrology and processes.


FIGURE (A) Fence diagram displaying all l.. kilometers of inverted seismic velocity data. (B) Seismic results from line e. This profile crosses both ridges and the central valley. One thousand two hundred and d,,,, m/s velocity contours are shown in solid white and dotted white lines, respectively. (C) Vertical velocity gradient from line e. (D) Seismic results from line e which lies on the southern side of the northern ridge. This profile did not reach high velocities (>,,,, m/s), so only the e,,,, m/s velocity contour is visible. (E) Vertical velocity gradient from line e. This profile shows a sharp vertical velocity gradient at the soil/saprolite boundary.
FIGURE (A) Histogram results of slope analysis with the north-facing slopes separated by vegetation height along with the median of those values shown by the solid blue line. (A) shows north-facing slopes with vegetation > meter. (B) Histogram results of north-facing slopes with vegetation < meter and a median of those values shown by the dashed blue line. (C) Histogram results of south-facing slopes and a median of those values shown by the solid orange line. (D) Median velocity values of the north-facing slope with and without vegetation and the south-facing slope. Bifurcation of the north facing slope values indicated with gray stripes.
Mapping bedrock topography: a seismic refraction survey and landscape analysis in the Laramie Range, Wyoming

July 2023

·

114 Reads

·

4 Citations

Frontiers in Water

Physical, chemical, and biological processes create and maintain the critical zone (CZ). In weathered and crystalline rocks, these processes occur over 10–100 s of meters and transform bedrock into soil. The CZ provides pore space and flow paths for groundwater, supplies nutrients for ecosystems, and provides the foundation for life. Vegetation in the aboveground CZ depends on these components and actively mediates Earth system processes like evapotranspiration, nutrient and water cycling, and hill slope erosion. Therefore, the vertical and lateral extent of the CZ can provide insight into the important chemical and physical processes that link life on the surface with geology 10–100 s meters below. In this study, we present 3.9 km of seismic refraction data in a weathered and crystalline granite in the Laramie Range, Wyoming. The refraction data were collected to investigate two ridges with clear contrasts in vegetation and slope. Given the large contrasts in slope, aspect, and vegetation cover, we expected large differences in CZ structure. However, our results suggest no significant differences in large-scale (>10 s of m) CZ structure as a function of slope or aspect. Our data appears to suggest a relationship between LiDAR-derived canopy height and depth to fractured bedrock where the tallest trees are located over regions with the shallowest depth to fractured bedrock. After separating our data by the presence or lack of vegetation, higher P-wave velocities under vegetation is likely a result of higher saturation.


Differences in dieback, forest cover and bedrock composition
a, Bald Mountain (BM), Dinkey Dome (DD) and Duff Creek (DC) have differing bedrock (black outlines)³⁶ but similar elevation, climate (Extended Data Table 1) and erosion rates³⁷. b, Biotite concentration correlates with phosphorus concentration in bedrock ([P]). c, Yearly average evapotranspiration (lines show 95% confidence interval) from Landsat-based NDVI (Methods). Dieback lagged start of 2011–2017 drought (grey box) by three years. d, Aridity index (potential evapotranspiration divided by precipitation) varies from year to year but is similar across the sites each year. e–j, National Agricultural Imagery Program imagery from before (e–g) and after (h-j) widespread dieback at Bald Mountain (e,h), Dinkey Dome (f,i) and Duff Creek (g,j). Circles highlight dieback of individual trees from 2014 to 2016. k–m, Spatial variations in percentage change in evapotranspiration from 2014 to 2016 (Methods) at Bald Mountain (k), Dinkey Dome (l) and Duff Creek (m) match observations of leaf browning and tree mortality from aerial imagery, supporting reduction in evapotranspiration as a measure of dieback. n, Distribution of dieback differs markedly between sites (P < 0.0001; Kolmogorov–Smirnov test with Bonferroni correction for three samples) and collectively spans 94% of range in dieback at similar elevations in surrounding region (black line; Extended Data Fig. 1). Colours throughout figure correspond to site labels in a.
Source data
Porosity and water-storage capacity controlled by bedrock composition
a–c, Joint distribution of porosity and depth from rock physics model of seismic refraction data and condensed into site-wide porosity profiles²⁶ (Methods) for Bald Mountain (a), Dinkey Dome (b) and Duff Creek (c). d–f, Frequency distributions of mean porosity in top 2 m (d), unit pore volume (the depth of water required to fill pores in soil and saprolite) (Methods) (e) and combined thickness of soil and saprolite (material with P-wave velocity < 1.1 km s–1 (refs. 26,38)) (f). Circles represent averages over individual Landsat pixels. Box plots show median (centre line) and first and third quartiles (box); whiskers extend up to 1.5 times the interquartile range past the first and third quartiles. g,h, Average (±1 s.d.) pore volume per unit area increases with average bedrock concentration of biotite (g) and phosphorus (h).
Source data
Plant water demand, water-storage potential and drought response
a, Pre-drought (2007–2011 average) evapotranspiration versus unit pore volume, showing averages (±1 s.e.m.) by site (squares) and pixel (circles) with linear-log fit to all data (dashed; r² = 0.75) and statistically distinct linear relationships for Duff Creek (green), Dinkey Dome (blue) and Bald Mountain (red). b, Dieback increases with unit pore volume (r² = 0.51). c, Evapotranspiration in 2016 along survey lines was generally lower than evapotranspiration in 2014 and thus plots below dashed 1/1 line, reflecting dieback during the drought. A three-parameter log–log regression, with a slope, intercept and offset between Bald Mountain and the other two sites (Methods and Extended Data Table 3), yields slope = 0.88 ± 0.02, indicative of structural overshoot, in which locations with higher pre-dieback evapotranspiration (and higher forest productivity) had disproportionately higher dieback (colour bar).
Source data
Bedrock control on drought response in the Sierra Nevada
a, Bedrock sample locations from ref. ¹⁷ (black) and ref. ³² (grey) within the Sierra Nevada Batholith (blue outline). b, Dieback increased disproportionately with average evapotranspiration in 2014 (P < 0.001), implying widespread bedrock control on dieback. c,d, This is confirmed by positive correlations between [P] and both pre-dieback evapotranspiration (c) and dieback (d) (average ± 1 s.e.m.). e, Mafic mineral concentrations in bedrock (colour index, in volume %) increase with [P] after ref. ³². f, Systems diagram showing links between bedrock composition and forest dieback in the critical zone (conceptualized in background image). Positive and negative couplings are arrows and lines terminated by circles, respectively. Labels refer to supporting results. Dotted connector near top shows hypothesized buffering of dieback by storage capacity, an effect that was overwhelmed here by the positive coupling between forest vulnerability and water demand. On the basis of the regression offset in Fig. 3c, we hypothesize that the strength of the coupling between forest productivity and dieback depends on species distribution, which in turn depends on bedrock composition.
Source data
Forest vulnerability to drought controlled by bedrock composition

September 2022

·

515 Reads

·

33 Citations

Nature Geoscience

Forests are increasingly threatened by climate-change-fuelled cycles of drought, dieback and wildfires. However, for reasons that remain incompletely understood, some forest stands are more vulnerable than others, leaving a patchwork of varying dieback and wildfire risk after drought. Here, we show that spatial variability in forest drought response can be explained by differences in underlying bedrock. Our analysis links geochemical measurements of bedrock composition, geophysical measurements of subsurface weathering and remotely sensed changes in evapotranspiration during the 2011–2017 drought in California. We find that evapotranspiration plummeted in dense forest stands rooted in weathered, nutrient-rich bedrock. By contrast, relatively unweathered, nutrient-poor bedrock supported thin forest stands that emerged unscathed from the drought. By influencing both subsurface weathering and nutrient supply, bedrock composition regulates the balance of water storage and demand in mountain ecosystems. However, rather than enhancing forest resilience to drought by providing more water-storage capacity, bedrock with more weatherable and nutrient-rich minerals induced greater vulnerability by enabling a boom–bust cycle in which higher ecosystem productivity during wet years drives excess plant water demand during droughts.



Citations (19)


... Despite the apparent asymmetrical contrast, Uecker et al. (2023) collected 3.9 km of seismic 108 refraction data and showed that the depth to bedrock could be greater than 60 m and argued that 109 trees on the north-facing slopes are preferentially located to obtain more water based on 110 differences in P-wave velocity (Uecker et al., 2023). Another well-studied site is the Blair-Wallis 111 (BW) located 12 km south (Flinchum, Holbrook, Grana, et bedrock topography was observed and fits well with a regional stress model under regional 115 compressive stress (Flinchum, Holbrook, Rempe, et al., 2018). ...

Reference:

Sensing a Connection: Tree Distribution is Influenced by Deep Critical Zone Structure
Mapping bedrock topography: a seismic refraction survey and landscape analysis in the Laramie Range, Wyoming

Frontiers in Water

... One recent hypothesis is the role of rock moisture in forest resilience to drought. Callahan et al. (2022) used seismic refraction surveys in three geochemically distinct plutonic bedrock systems with similar climate in California. They suggested higher dieback was linked to weatherable material and nutrient-rich rock; trees in nutrient-poor bedrock were more resilient. ...

Forest vulnerability to drought controlled by bedrock composition

Nature Geoscience

... However, the estimation of petrophysical properties-such as porosity, lithology, and fluid saturations-can be formulated as an inverse problem. In this case, they are usually referred to as petrophysical inversions (in seismic exploration, they are sometimes also known as rock-physics inversion-e.g., Grana, Parsekian, et al., 2022). In the deterministic frameworks, a physical relationship mapping the rock properties to the geophysical parameters is often necessary (Cao et al., 2023;Foged et al., 2014;Mastrocicco et al., 2010). ...

Geostatistical Rock Physics Inversion for Predicting the Spatial Distribution of Porosity and Saturation in the Critical Zone
  • Citing Article
  • June 2022

Mathematical Geosciences

... In critical zone studies, bulk density is an important physical property as an indicator for soil quality and compaction (Suuster et al. 2011 ). Lateral density variations can reveal information about changes in porosity, fracture distribution and soil weathering (Flinchum et al. 2022 ). Density is used to inform studies of root growth (Brimhall et al. 1992 ;Dexter 2004 ), water movement and retention (Huang et al. 2011 ;Flinchum et al. 2018 ), as well as carbon and nutrient content in soil layers (Nanko et al. 2014 ). ...

What Do P-Wave Velocities Tell Us About the Critical Zone?

Frontiers in Water

... Pasquet et al., 2021). Therefore, if the meaning of the phase velocity determined by the multi-station method is understood in laterally varying sites, using and inverting the dispersion data obtained by multi-size spatial windows may help to achieve a more comprehensive understanding of the subsurface.The goal of this paper is to develop a new method, which is both robust in dispersion measurement and applicable for laterally varying models, for analyzing 2D active-source surface wave data. ...

Multi-Window Weighted Stacking of Surface-Wave Dispersion

Geophysics

... Such recharge pathways are known to be controlled by the structure and hydraulic properties of the surface geology (Scanlon et al., 2006;Wheater et al., 2010). For example, it has been widely demonstrated that the recharge rates linked to transmission losses of rivers are less influenced by river stage height than the lithology and the hydraulic conductivity of the riverbed and the unsaturated zone (Carter and Alkali, 1996;Dahan et al., 2008;Costa et al., 2012;Flinchum et al., 2020;Zarate et al., 2021). ...

Identifying recharge under subtle ephemeral features in a flat-lying semi-arid region using a combined geophysical approach

Hydrology and Earth System Sciences

... Finding adequate supply, of acceptable quality, with a sufficient yield within an economically viable distance from a community, is often a challenge. There has been previous research to improve the probability of finding water in these regions (Munday et al., 2020;Peeters et al., 2020), but water security remains a challenge in arid communities in central Australia due to the predominance of local-scale groundwater systems hosted in fractured and weathered bedrock (Jacobson and Lau, 1987). ...

Groundwater Knowledge Integration System (GKIS): Probabilistic groundwater prospectivity mapping with iterative updating of conceptualisation

... Groundwater impractical at large scales (Zell and Sanford 2020;Condon et al. 2021;Gleeson et al. 2021). Developing an accurate subsurface architecture becomes even more important given the uncertainties and that alternate subsurface representations are rarely explored (Enemark et al. 2019(Enemark et al. , 2020. Integrated hydrologic models simulate surface and subsurface flow simultaneously. ...

A Systematic Approach to Hydrogeological Conceptual Model Testing, Combining Remote Sensing and Geophysical Data

... Because direct observations of the subsurface portion of the CZ are difficult, requiring trenches, soil pits, or boreholes, geophysical imaging is often used to study the shallow subsurface (e.g., Parsekian et al., 2015). Seismic imaging has the advantage of being primarily sensitive to porosity (e.g., Callahan et al., 2020;Flinchum et al., 2018;Hayes et al., 2019;Holbrook et al., 2014), which determines subsurface water storage capacity and reflects chemical and physical weathering in eroding landscapes. In the near-surface, the seismic methods most used are first-arrival traveltime tomography (FATT) and multichannel analysis of surface waves (MASW). ...

Subsurface Weathering Revealed in Hillslope‐Integrated Porosity Distributions

... ratio in elastic isotropic material (Flinchum et al., 2020;Pasquet et al., 2016). Unlike the other elastic moduli, Poisson's ratio ( ) does not require the density only the shear (G) and bulk (K) moduli (Equation 1): ...

Identifying recharge under subtle ephemeral features in flat-lying semi-arid region using a combined geophysical approach