ACS Publications. Most Trusted. Most Cited. Most Read
My Activity
CONTENT TYPES

Spatial Microanalysis of Natural 13C/12C Abundance in Environmental Samples Using Laser Ablation-Isotope Ratio Mass Spectrometry

  • Andrei Rodionov*
    Andrei Rodionov
    Institute of Crop Science and Resource Conservation (INRES), Soil Science and Soil Ecology, University of Bonn, Nussallee 13, Bonn 53115, Germany
    *E-mail: [email protected]
  • Eva Lehndorff
    Eva Lehndorff
    Soil Ecology, University of Bayreuth, Dr.-Hans-Frisch-Str. 1-3, Bayreuth 95448, Germany
  • Ciprian C. Stremtan
    Ciprian C. Stremtan
    Teledyne CETAC Technologies, 14306 Industrial Road, Omaha, Nebraska 68144, United States
  • Willi A. Brand
    Willi A. Brand
    Max-Planck-Institute for Biogeochemistry, Beutenberg Campus, P.O. Box 100164, Jena 07701, Germany
  • Heinz-Peter Königshoven
    Heinz-Peter Königshoven
    Feinmechanische Werkstatt, Institute of Physical and Theoretical Chemistry, University of Bonn, Wegeler Str. 12, Bonn 53115, Germany
  • , and 
  • Wulf Amelung
    Wulf Amelung
    Institute of Crop Science and Resource Conservation (INRES), Soil Science and Soil Ecology, University of Bonn, Nussallee 13, Bonn 53115, Germany
    More by Wulf Amelung
Cite this: Anal. Chem. 2019, 91, 9, 6225–6232
Publication Date (Web):April 1, 2019
https://doi.org/10.1021/acs.analchem.9b00892

Copyright © 2019 American Chemical Society. This publication is licensed under these Terms of Use.

  • Open Access

Article Views

2113

Altmetric

-

Citations

LEARN ABOUT THESE METRICS
PDF (5 MB)
Supporting Info (1)»

Abstract

The stable 13C/12C isotope composition usually varies among different organic materials due to isotope fractionation during biochemical synthesis and degradation processes. Here, we introduce a novel laser ablation-isotope ratio mass spectrometry (LA-IRMS) methodology that allows highly resolved spatial analysis of carbon isotope signatures in solid samples down to a spatial resolution of 10 μm. The presented instrumental setup includes in-house-designed exchangeable ablation cells (3.8 and 0.4 mL, respectively) and an improved sample gas transfer, which allow accurate δ13C measurements of an acryl plate standard down to 0.6 and 0.4 ng of ablated carbon, respectively (standard deviation 0.25‰). Initial testing on plant and soil samples confirmed that microheterogeneity of their natural 13C/12C abundance can now be mapped at a spatial resolution down to 10 μm. The respective δ13C values in soils with C3/C4 crop sequence history varied by up to 14‰ across a distance of less than 100 μm in soil aggregates, while being partly sorted along rhizosphere gradients of <300 μm from Miscanthus plant roots into the surrounding soil. These very first demonstrations point to the appearance of very small metabolic hotspots originating from different natural isotope discrimination processes, now traceable via LA-IRMS.

The stable 13C/12C isotope composition of environmental samples has become a unique tracer of carbon assimilation pathways. In plants, its analysis provided key information about the assimilation of fossil carbon sources (Suess effect) (1) and about differentiated C3 and C4 photosynthetic pathways (2,3) and gave hints regarding water stress adaptation pathways. (4−6) In marine and lacustrine environments, the analysis helped to track terrestrial C input and was useful for paleoclimate reconstruction. (7−9) In terrestrial environments, δ13C analyses were indispensable for assessing the mean residence time of soil carbon after C3/C4 vegetation changes, (10,11) for elucidating microbial and faunal food webs, (12) and for paleo-environmental reconstructions. (13−16) However, spatially homogeneous C isotope signals were assumed for all these studies because it has not been possible to map microscale variation of the 13C/12C isotope composition in routine lab analyses down to the natural abundance level. The present contribution provides a viable solution to this problem.
Different instrumental setups using laser ablation (LA) combined with other detectors have already been applied on a micrometer scale for measuring bacterial cells and plant and animal tissue, as well as to other substances in geological and biological studies. (17−23) For ablation of organic matter, cold Nd:YAG laser ablation systems are recommended. (24,25) Lasers operating at 213 nm ensure cold ablation as more traditional 266 nm systems (24,25) and thus avoid isotope fractionation effects during the ablation process. All previous laser ablation studies relied on obtaining spatially resolved signatures from organic solids with relatively large and homogeneous organic C contents compared to soil (wood, single hair samples, or pollen grains of C3 and C4 plants), yielding ablation spot sizes between 150 and 50 μm (22,24,25) and a sensitivity between 65 and 24 ng C. (25,26) Higher spatial resolution and sensitivity are crucial for many environmental studies, which have aims such as identifying microbial hotspots in soils and sediments, or patchy sequestration of carbon.
In this contribution, we introduce a laser ablation-isotope ratio mass spectrometry (LA-IRMS) system using a frequency quintupled 213 nm Q-switched Nd:YAG laser with <5 ns pulse width and flat-top energy profile (>4 mJ pulse–1; Teledyne CETAC Technologies, Omaha, USA). Laser ablation of particulate matter was done in newly designed low-volume cells; the ablated particles were collected and transported using helium as the carrier gas. Particles were oxidized to CO2 in an oven; then, sample gas was transferred to a system of two cryo-traps for gas cleanup and peak focus (modified PreCon system; see ref (27)). Sample gas was then led through a GC column before it entered the IRMS DeltaVplus via a ConFloIV interface (both from ThermoFisher Scientific) for analysis of δ13C. The challenges were to increase the spatial resolution of the δ13C signal by an order of magnitude by decreasing the area of ablated spots down to 10 μm, thus responding to a growing interest in isotopical characterization of microbial hot spots, organic exudates of roots, and residues of plants and microbes. Doing so would enable us to reach into areas where the utilization of organic materials is physically hindered by, e.g., encrustation processes and occlusion within aggregates and, thus, limited bioaccessibility.

Experimental Approach

ARTICLE SECTIONS
Jump To

Ablation System and Isotope Measurement

For laser ablation, highly focused laser light pulses are used to remove (ablate) material at a spatial resolution of a few micrometers in diameter from discrete areas of virtually any solid sample (e.g., ref (28) and references therein). For the present study, we used a Teledyne LSX-213G2+ solid state Nd:YAG laser system (CETAC, Omaha, NE, USA) that produces deep UV laser radiation at a wavelength of 213 nm. The laser pulse frequency can be variably adjusted; frequencies from 10 to 20 Hz proved to be optimal for this application. Furthermore, the laser allowed adjustment of the energy output, the ablated spot size, and the number of laser shots.
The standard cell of the LSX213 G2+ laser (two volume active cell HelEx II) was replaced with an in-house-designed single-volume cell (see Figure 1; the cell was constructed with an active volume of 3.8 cm3 and a total internal volume of 17.2 cm3; for details see Figure S-1). The cell was mounted on an adapted plane mount aligned to the laser’s original movable stages, which allowed the sample to be monitored by a computer-controlled video microscope and internal LED illumination (coaxial, ring, and transmitted light sources). As we expected, the samples to have very low amounts of C, we also produced a second cell with the same design, but with an even smaller active volume of 0.4 cm3 and, therefore, with reduced background signal. Both cells are easy to handle and are interchangeable.

Figure 1

Figure 1. Setup of the LA-IRMS system starting with a newly constructed low-volume cell. Only the upper part of the cell is the active operating volume for laser ablation (marked in red; more details in the Supporting Information).

The upper and lower windows of the single-volume cell were made of quartz glass and sealed using O-rings (Ø30 × 2 mm; form error: 3/0.5 wedge: <5′; Laseroptik corp., Garbsen, Germany). The top window was specially coated (AR213 nm/0°; R < 0.25%) to reduce reflection of the 213 nm wavelength.
The system was continuously flushed with high purity helium (BIP5.0), which ran through a stainless steel line of 1/16 in., directly connected to a mass flow controller (MFC adjusted to constant flow rate of 10 mL min–1; Alicat, Cole-Palmer, Vernon-Hills, IL, USA) and a PEEK 1/16 in. line between the MFC and the cell (Figure 1).
The resin-embedded sample was positioned on a spring inside the cell to reduce the cell’s active volume (see Figure S-1; sample was embedded in Araldite 2020, Huntsman, δ13C −29.98 ± 0.5‰, and then cut and polished with corundum to avoid C contamination on the sample surface). Helium was purged with 10 mL/min into the passive volume of the cell and then directed via four channels to the active cell volume above the sample. These four channel outlets were tilted (with 45° to the cell radius) to ensure gas rotation on the sample surface, thus increasing flow dynamics and particle suspension in the gas phase. The maximum sample diameters were 25 and 10 mm (3.8 and 0.4 cm3 cell volume), respectively, which corresponds to sample sizes used in other applications, such as nanoscale secondary ion mass spectrometry (NanoSIMS).
The output port was laterally aligned close to the sample surface to increase particle collection efficiency (Figure S-1). This port was coupled to an IRMS via a PreCon automated interface, equipped with a continuous gas flow combustion furnace. The furnace was used to convert the carbon in the ablated sample to CO2 at 900 °C. Low cell volumes allowed the oxidation of the complete sample in the oven. A Nafion trap was installed after the furnace to eliminate remaining traces of water. This was followed by cryo-trapping of the evolved CO2 in a first automated trap using liquid nitrogen (Figure 1). The second automated trap focused the CO2. After switching the Valco valve, both traps were purged with a reduced He flow, now received from the ConFlo interface (He flow of approximately 0.3 mL min–1). Sample gas was then led through a capillary column for further cleanup and peak shaping (PoraPLOT Q, 25 m length, 0.25 mm diameter, 8 μm film; Agilent Technologies, Santa Clara, CA, USA) to a ConFlow and Delta V plus IRMS detector (Figure 1; ThermoFisher Scientific, Bremen, Germany). During sample exchange and stand-by, an additional He flow was added through the PreCon system (He purge at 10 psi, Figure 1) to ensure continuous He flow through the combustion furnace and lines (for stand-by, helium flow from the laser MFC had to be reduced to 3 mL min–1 to avoid overpressure at the laser (max. 19 psi)). To maintain the oxidizing capacity of the combustion, the furnace was periodically oxidized by using a Crystal-mix (Helium N46 mixed with 5.0 ± 0.1 Vol% Oxygen N45).
For accurate isotope ratio assessment, we used the Isodat Software (ThermoFisher Scientific), which basically starts each run with a peak center adjustment. Then, three peaks of CO2 reference gas were introduced before sample ablation started. Directly before sample gas analysis, the IRMS performance was again checked by two CO2 reference peaks (reference gas switched in through ConFloIV after cryo-trapping). The complete analytical procedure was programmed via an ISL protocol embedded in the Isodat software, which enables timing of valve position change, ConFloIV action (CO2 reference gas inlet/sample gas inlet), and positioning of cryo-traps. The laser was controlled via the Chromium2.2 software (Laser Ablation System V.2017.1.31.0), and the ablation was started manually to work in time with the IRMS part of the LA-IRMS system. Each sample run needed about 10 min.

Standards and Reference Samples

For calibration of the CO2 gas tank and reference samples, we used isotope reference standards reported versus Vienna PeeDee Belemnite (vs VPDB; acetanilide −29.53‰; cellulose −24.72‰; sucrose −10.45‰; National Institute of Standards and Technology, Gaithersburg, MD, USA), measured on a FlashEA 1112 elemental analyzer connected to a Delta V advantage IRMS (ThermoFisher Scientific, Bremen, Germany). In-house EA-IRMS referenced samples, which were chosen to resemble different types of organic matter occurring in soil, were applied to test the LA-IRMS system (beech wood −27.09 ± 0.02‰; einkorn wheat −28.71 ± 0.03‰; acryl plate −29.75 ± 0.06‰; IAEA-CH-3–24.77 ± 0.06‰; sucrose −26.93 ± 0.05‰; Gordonia alcanivorans −23.07 ± 0.00‰).
For the ablation tests, we produced press-pellets from each separate and in-house-referenced sample. With the exception of the IAEA-CH-3 standard (29) (which was cut into pieces), all samples were thoroughly homogenized to avoid any inconsistencies. The standards were placed in stainless steel sample holders in order to avoid any cross contamination (Figure S2). Each sample was ablated several times to test the effects of laser energy efficiency, as well as to try and correlate ablation depth to the number of shots.

Environmental Samples

To illustrate the performance of our LA-IRMS system, we used the following sample sets: (1) We assessed the carbon isotope composition in so-called occluded large microaggregates (53–250 μm), isolated from wet fractionated and ultrasonically treated macroaggregates (250–8000 μm). Such aggregates are considered to play a key role in C sequestration processes in soil because they protect physically occluded C from decay. (2) We assessed the carbon isotope signature along gradients from C4 roots (Miscanthus) into the surrounding soil (previously under C3 vegetation), with the knowledge that the living area around roots (the so-called rhizosphere) is a hot spot for microbial activity, thus inducing steep gradients in carbon concentration and its turnover. (3) We assessed the spatial variability of carbon isotope composition in soil particles <2 mm, which have received C inputs by C3 and C4 vegetation (two years after C3 to C4 vegetation change).
Soil microaggregates (set 1) were formed in arable loamy Haplic Luvisols (35% clay; 5–15 cm sampling depth, Scheyern experiment station, Southern Germany). (30)
Topsoil samples from a Miscanthus field (set 2) were collected from the experimental research station of Bonn University in Klein-Altendorf, Germany, continuously cropped with Miscanthus since 2007. Coordinates were 50°35′37.18 N/6°59′10.24 E at an altitude of 240 m a.s.l. The mean annual temperature and precipitation were 9.5 °C and 605 mm. The main soil type was a Luvic Stagnosol (Episiltic) (IUSS Working Group WRB, 2015), with a silty-loam texture (63% silt, 23% clay, and 14% sand). (31)
For set 3, soil particles <2 mm from a C3/C4 vegetation change experiment at the Terrestrial Observatory TERENO, Selhausen, Germany, were sampled in 2014 (0–30 cm soil depth). Soil had been continuously cropped with maize since 2012 (Zea mays L.). (32) The mean annual temperature and precipitation were 9.8 °C and 693 mm, respectively. The main soil type was a Stagnic Cambisol (Skeletic) (IUSS Working Group WRB, 2015) with a silty-loam texture (64% silt, 15% clay, and 18% sand). (32)
Environmental samples were taken as large blocks with a spade. They were manually broken to undisturbed aggregates of approximately 15 mm in diameter and air-dried (set 2), further processed using wet sieving into different aggregate size classes and freeze-dried (set 1), (30) or simply dry-sieved to 2 mm (set 3). (32) The dried subsamples were then embedded in Araldite before analysis.

Surface Mapping

To determine the ablation quality and to assess the crater depth and amount of ablated material, noncontact 2D and 3D surface mappings were applied to the acryl plate in-house standard. The measurements were carried out using an Omniscan MicroXAM white lite roughness microstitching interferometer (ADE Phase Shift, Tucson, AZ, USA) at Trinity College Dublin Center for Microscopy and Analysis. The repeated measurements of the crater depths were performed on combinations of ablation spot size with number of shots (Figure S3; Table S1; n = 5).

Results and Discussion

ARTICLE SECTIONS
Jump To

Blank Minimizing through Sample Aerosol Transport Optimization

There is a general risk that CO2 from sources other than the sample may contribute to the assessment of δ13C values by LA-IRMS. We call this the blank signal, originating from small amounts of CO2 already present in the cell, entering through the Valco-6-port connector, or resulting from other tiny leakages in the system. Thus, the isotope discrimination tests were performed at C concentrations far above the blank signal. However, to perform accurate LA-IRMS analyses at very low C concentrations at high spatial resolution, minimization of the blank signal was crucial. We observed that its contribution to the total signal increased with increasing loading time on the cryo-trap (Figure 2). However, the isotope composition of the blank was not affected by the loading time. It averaged −18.78 ± 0.54‰, ranging from −17.93‰ to −19.27‰ on different days (n = 34; individual data not shown). With the 3.8 mL cell and a trap loading time of 7 s, the blank fell below the routine detection limit of the IRMS system (which was preset to 50 mV; that is, close to the baseline). With a loading time of 20 s, the blank signal remained stable with a peak area less than 0.6 Vs; this means that any signal above this value must have resulted from sample ablation. Thus, very clear ablation signals from the in-house acryl standard were produced with spots of 30 μm size and 40 shots in a 3.8 mL cell volume (Figure S-4). By minimizing the time window for cryo-trapping, that is, finding the optimum delay time between particle ablation in the cell and particle arrival at trap 1 (loading time), the blank signal could be reduced significantly as needed for improved sensitivity. All these steps were controlled by the Valco-6-port valve in the PreCon settings (modified ISL method of the PreCon, loaded in the Isodat Software, ThermoFisher Scientific). The method was programmed with a delay time of 2 s until switching to cryo-trapping with a loading time of 20 s for the 3.8 mL cell (see details for the 0.4 mL cell below).

Figure 2

Figure 2. Impact of loading time in the first cryo-trap on the blank peak area in cells with 3.8 and 0.4 mL active volume (without laser ablation). If the loading time was <7 s, the blank was at the baseline level of the IRMS detector for both cells. The isotopic signature of the blank signal was not affected by the loading time.

Energy Efficiency and Limitation of the Fractionation

First, we tested the optimal efficiency of the LA-IRMS system under different ablation settings and for different in-house standardized organic materials with known δ13C value, such as beech wood, sucrose, einkorn grass, the actinomycete Gordonia alcanivorans, an acryl plate, and the certified IAEA-CH-3 standard. The question was whether these organic materials, which commonly occur in soil, show different ablation behavior. To perform these tests, we produced press-pellets from milled standard materials and placed them in the metal sample holders (Figure S-1a,b). According to Jeong et al., (33) the entrainment efficiency of particle transport decreases with increasing laser power density and is strongly affected by laser beam diameter at constant laser pulse energy (that is, removing more sample than can be efficiently collected and transported) or at variable laser pulse energy with fixed spot size. We tested the appropriate laser pulse energy using a fixed spot size and a fixed number of shots in increments, starting from the configuration of 40% energy output to 20%, 15%, 10%, 5%, and finally 3.5% energy output (Figure 3). Using a constant spot size of 30 μm and 40 pulses, the results of different in-house-referenced samples clearly showed that the 3.5% energy output setting of the laser (corresponding to a fluence of 0.71 J/cm2) was appropriate for organic matter ablation independent of matrix. If higher laser energies were set, δ13C decreased from the referenced value, probably due to partial heating of the sample inducing isotopic discrimination (Figure 3). Also, the blank contribution was significantly reduced and, since this was constant during the day, it did not interfere with the sample measurements. Across different days, however, the isotope composition of the blank varied, which must be accounted for when analyzing a sample at different days (Figure 3).

Figure 3

Figure 3. Discrimination of stable carbon isotopes from different organic matrices by ablation with increasing laser energy.

Sensitivity

The sensitivity of the LA-IRMS system was tested by applying variable spot sizes (40, 30, 25, 20, 15, and 10 μm) and numbers of pulses (40, 30, 20, and 10 shots) to our organic matrices (cell size 3.8 mL, Figure S-4a,b). As expected, the results showed that decreasing the spot size impacted the resulting signal intensity more than a decreasing number of shots. The combination of both, that is, the dilution of the blank signal in a sample by increasing the sample volume, resulted in a linear decrease of the measured isotope signal for all tested in-house-referenced samples (Δδ13C = δ13Ctotal – δ13Creference; Figure S-4a; Table S1). In the 3.8 mL cell, the isotope signal was stable for 30 μm spots due to a negligible blank signal, and the resulting isotope signatures were close to the reference values (Figure S-4b).
Here, we directly measured the blank value and subtracted the respective contribution of blank carbon from sample carbon. In detail, this means that, when ablating soil samples, we need to compare the resulting carbon signature with that of the reference peak from standard material that was homogeneous in carbon content. The subtraction then refines the resulting signals by a linearity correction, which involves using measured and true values for the CO2 reference gas.
A blank correction for isotope ratio measurement with small amounts of samples can be performed by regression or by subtraction (34,35) and calculated as described by Werner et al.: (36)
where A is the area of the m/z 44 peak.
The trapping time of 20 s needed for the complete transport of the sample gas suspension from a 3.8 mL cell to the first cryo-trap poses a constraint for sampling resolution; namely, the signal intensity from ablated carbon material had to be in excess of 0.3 Vs for it to be reproducible (Figure S-4b).
To calculate the ablated C amount, we used an acryl plate and measured the total ablated volume (Figure S-3a; C content 60%; density 1.19 g cm3; δ13CAcryl = −29.75 ± 0.06‰, relative to VPDB using EA-IRMS). We determined the ablated crater volume for different combinations of laser spot size and shot numbers (Figure S-3b; Table S1). The crater geometry and overall symmetry may influence the depth resolution and fractionation effects, which might not be strictly related only to the beam profile. The Nd:YAG laser was able to produce flat-bottom and straight-wall craters. (37,38) The crater geometry (Figure S-3b,c) maintained flat-bottom laser beam profiles during repeated ablations at a single spot, using focused laser light pulses with 10 and 40 μm operative diameters. Because of this crater assessment, we were able to calibrate the LA-IRMS signal to C amount and, thus, also determine sample size requirements for the reference measurements comparable to Moran et al. (25) The final limit of determination corresponded to a carbon amount of 0.6 ng for the 3.8 mL cell, which corresponded to a minimum of 20 μm ablated spot size (Figure 4a).

Figure 4

Figure 4. Detection limit of the LA-IRMS system given in C amount (a) for the cell with 3.8 mL active volume and (b) for the cell with 0.4 mL active volume (detected with an acryl plate with known ablated crater volume; see also Supporting Information).

Sensitivity Enhancement

In order to be able to detect carbon when using spots <20 μm in diameter, we built a second sample cell with reduced active volume (0.4 mL). We used it with decreased loading time for cryo-trapping in order to minimize blank signal accumulation in the cryo-traps (Figure 2). The isotopic signature of the blank was comparable to that of the 3.8 mL cell, but the contribution of the blank to the ablated signal was significantly smaller. The trapping time needed for the evacuation (transport efficiency) of the sample was best at 10 s (Figure S-5). A saturation of raw isotopic signal from the acryl plate was reached at a δ-value of −37.42 ± 0.37‰, irrespective of the loading time; this indicates that the blank contribution could then be neglected (Figure S-5a). The smaller cell finally yielded the same δ13C values as the larger cell after blank correction (Table 1; Table S-1). We conclude that operation with the small cell is feasible, which enables us to reduce the combination of spot size for laser ablation to 15 μm with 20 shots and to 10 μm with 40 shots. The limit of determination was calculated to be 0.3 ng C for the 0.4 mL cell, which corresponded to a minimum ablated spot size of 10 μm (Figure 5b).

Figure 5

Figure 5. Spatially resolved microanalysis of δ13C heterogeneity in a soil microaggregate structure from a cropped Luvisol field of 33% clay (53–250 μm aggregate fraction, 20 μm spot size, 40 shots per ablation). Raw δ13C values shown in red and blank corrected δ13C values, in black.

Table 1. Accuracy Tests of the LA-IRMS System with 3.8 and 0.4 mL Active Cell Volume
 laser ablation-IRMS 
 active volume 3.8 mLactive volume 0.4 mLEA-IRMS
sampleδ13C ± st.dev. (‰)offset (‰)nδ13C ± st.dev. (‰)offset (‰)nδ13C ± st.dev. (‰)n
acryl platea–29.68 ± 0.25–0.0739–29.77 ± 0.250.0222–29.75 ± 0.065
IAEA-CH-3–24.72 ± 0.780.00b30–24.80 ± 0.570.08b21–24.77 ± 0.065
Beech–27.05 ± 0.64–0.0427   –27.09 ± 0.025
a

Acryl plate sampling corresponded to 3.4 ng C in the 3.8 mL cell and 0.7 ng C in the 0.4 mL cell (see text for calculation).

b

IUPAC reference for δ13C value IAEA-CH-3 = −24.72 ± 0.04‰. (29)

Analyses of Environmental Samples

We first applied our methodology to soil microaggregates (sample set 1, long-term arable experiment station at Scheyern, Germany). We fixed the occluded microaggregate fraction in Araldite and detected spotwise the δ13C using the 3.8 mL cell (Figure 5). During the analyses of the embedded material, we randomly repeated the acryl plate standard measurement in order to have the reference measurements in the same sample run (Figure S-2c). The bulk soil and microaggregate fraction (53–250 μm) were characterized for their C isotope composition before using EA-IRMS and showed an average δ13C value of −25.10 ± 0.07‰ and −25.36 ± 0.42‰, respectively. However, the single isotopic signals in the ablated material showed a much larger variability, which reached a data spread >1‰ independent of ablation spot size (Table 2). In soil, such a shift in δ13C values corresponds to organic matter degradation along a depth interval of a few dm in native sites not experiencing a C3/C4 vegetation change, which may well comprise >1000 years of organic matter turnover and genesis. (39) Therefore, the accuracy of δ13C analyses for bulk soil or even a soil aggregate fraction in no way represents the true heterogeneity of processes involved in soil organic matter transformation and genesis. Instead, we have to conclude that there are hot spots of C accrual, even at microaggregate level, that originate from different isotope discrimination processes and, thus, probably from different organic molecules and, respectively, from different plant or microbial debris.
Table 2. δ13C in Spots Ablated with 20 and 30 μm Spot Size from Soil Microaggregatesa
fraction [μm]spot size [μm]LA-IRMS δ13C [‰]EA-IRMS δ13C ± s.d. [‰]
<2000  –25.10 ± 0.07
250–53 occluded20–25.28–25.36 ± 0.42
  –23.59 
  –26.83 
  –25.28 
  –24.23 
  –25.36 
  –25.21 
250–53 occluded30–27.36 
  –26.68 
  –23.90 
  –25.60 
  –24.98 
  –25.13 
a

Laser ablation revealed high spatial variability, which is not visible from standard EA-IRMS analysis. The δ13C was not influenced by spot size (3.8 mL cell active volume, mean isotope signal from 13 LA-IRMS measurements −25.34 ± 1.11‰ compared well to bulk sample EA-IRMS analysis).

Heterogeneous δ13C distribution not only is to be expected in soils but also has previously been found for tree rings (24) and may be expected as well for root exudation of C into soil as for nutrient uptake and related feeding of the microbial community. We detected the microscale shift in C isotope composition in a gradient ablated from the root to the root-to-soil contact zone (rhizosphere) and the soil matrix (Figures 6a,b and S-6). The root, which was from Miscanthus (sample set 2), clearly consisted of C4-photosynthesis derived C (with an average δ13C of −19‰, Figure 6a), while soil in about 300 μm distance did not yet show signs of C4-carbon input. However, signals had already mixed in less than a 300 μm distance, indicating that 25–50% of the C was gradually exchanged in this region. To our knowledge, this is the very first in situ detection of rhizosphere gradients using stable-isotope screening. Rhizosphere gradients evolve because of water and related nutrient uptake, which eventually cost carbon to feed the nutrient-mobilizing community and are frequently assumed to extend into a gradient of up to 1–4 mm. Figure 6a suggests that rhizosphere extension in terms of C exudation from roots to soil may even be limited to 100 μm, that is, very close to the rhizoplane. These findings shed new light on rhizosphere biochemistry, since getting hold of its extension has challenged research until now (e.g., ref (40)).

Figure 6

Figure 6. Spatial differences of isotopic signatures inside the root, rhizosphere area, and soil matrix from a Miscanthus root ablated with (a) constant spot size of 30 μm and 30 shot numbers and with (b) variable spot size of 15 μm and 40 shot numbers (white-colored signatures) and spot size of 10 μm and 80 shot numbers (black colored signatures), both performed on a 60 μm cross section; in-house reference acryl = −29.75 ± 0.04‰.

When using the same embedded material, but decreasing spot size to 10 μm, even a larger range of δ13C isotope values was found (Figure 6b). This range reached 13.76‰, which was slightly larger than the range detected with a spot size of 15 μm (Δδ13C = 13.25‰). While this may still be due to random hits of different hot spots in soil, none of the δ-values above −17‰ (Figure 6b) had been found using 30 μm-sized spots (Figure 6a). We attribute this to the improvement of our cell: only a spot size of 10 μm is able to discover the full range of isotope values in the environment (here, soils with C3/C4 vegetation change), while a spot size of 30 μm already includes the risk of averaging across a sampling area that is too large to detect realistic hot spots. As former attempts to track δ13C values in plant and soil had to rely on a spot size of at least 100–150 μm (e.g., ref (23)), we see our method developments as a clear step toward a now upcoming, in situ mapping of δ13C isotope values and discrimination processes in the environment.
In order to obtain an independent field replicate, we took another sample from 2 years of the C3/C4 vegetation change from the Terrestrial Observatory TERENO Selhausen, Germany. (3,32) Even though we only performed laser shots on sieved soil aggregates, we detected the full range of possible δ13C values, reflecting hotspots of both C3 and C4 plant origin (Figure 7). Again, this discovered range was almost as large as the difference between C3 and C4 plant species (Δδ13C = 12.44‰), even though the hot spots looked macroscopically similar. The results have two implications. From an analytical point of view, these data show that it is possible to resolve hot spots of C accrual with different 13C isotope discrimination in an environment even down to a spatial scale of 10 μm. From the environmental point of view, the data implies that it is possible to find microsites at which original C3 plant material had remained untouched in soil, while other places were comprised solely of the newly C4 carbon introduced; in other words, there are microsites with more or less 0–100% of C sequestration, separated by a distance of only a few micrometers.

Figure 7

Figure 7. Spatial differences of isotopic signatures ablated from 2 mm sieved soil with constant spot size of 10 μm and 60 shot numbers after 2 years of C3 to C4 vegetation change (Zea mais L., Selhausen; in-house reference acryl = −29.57 ± 0.32‰).

Conclusions and Implications

ARTICLE SECTIONS
Jump To

By adapting laser ablation isotope ratio monitoring to δ13C analysis in soil organic matter on the microscale, we found significant heterogeneity in soil microaggregates after C3/C4 vegetation change (>10‰), as well as in microaggregates that did not receive a vegetation change (3‰), and we were able to detect C4–C inputs from roots into soil. Hence, this technique allows us to spatially recognize hot spots of organic C turnover in soil, as well as of input dimensions of rhizodeposits and root exudates. The same will be possible in other environmental samples like sediments, geological and geoarcheological archives, or biological tissues. The achieved current spot size of 10 μm allows detection of natural δ13C abundance far below conventional IRMS routines, which is achieved here by minimizing the split of sample gas fluxes and by reducing active cell volume to 3.8 and 0.4 mL. Future challenges might be the automatization of the ablation procedure (communication between laser and IRMS equipment) and the extension to δ15N analysis, which requires a modified analytical setup for focusing on the N2 peak.

Supporting Information

ARTICLE SECTIONS
Jump To

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.analchem.9b00892.

  • One table and six figures showing the determination of carbon amount by measurements of crater depth on an acryl plate (Table S-1), configuration of the newly developed laser cells (Figure S-1), images of press-pellets from the ground and mixed standard substances (Figure S-2), images of spot depth and surface mapping of ablated spots (Figure S-3), deviation of the δ13C signal from referenced values due to the contribution of the blank signal (Figure S-4), adjustment of loading times in the cryo-trap (Figure S-5), and stable C isotope signatures of a Miscanthus rhizome and the surrounding soil (Figure S-6) (PDF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

ARTICLE SECTIONS
Jump To

  • Corresponding Author
    • Andrei Rodionov - Institute of Crop Science and Resource Conservation (INRES), Soil Science and Soil Ecology, University of Bonn, Nussallee 13, Bonn 53115, GermanyPresent Address: A.R.: Soil Ecology, University of Bayreuth, Dr.-Hans-Frisch-Str. 1-3, 95448 Bayreuth, GermanyOrcidhttp://orcid.org/0000-0001-5971-1948 Email: [email protected]
  • Authors
    • Eva Lehndorff - Soil Ecology, University of Bayreuth, Dr.-Hans-Frisch-Str. 1-3, Bayreuth 95448, Germany
    • Ciprian C. Stremtan - Teledyne CETAC Technologies, 14306 Industrial Road, Omaha, Nebraska 68144, United States
    • Willi A. Brand - Max-Planck-Institute for Biogeochemistry, Beutenberg Campus, P.O. Box 100164, Jena 07701, Germany
    • Heinz-Peter Königshoven - Feinmechanische Werkstatt, Institute of Physical and Theoretical Chemistry, University of Bonn, Wegeler Str. 12, Bonn 53115, Germany
    • Wulf Amelung - Institute of Crop Science and Resource Conservation (INRES), Soil Science and Soil Ecology, University of Bonn, Nussallee 13, Bonn 53115, Germany
  • Notes
    The authors declare no competing financial interest.

Acknowledgments

ARTICLE SECTIONS
Jump To

We acknowledge the technician, Leona O’Connor, at Trinity College Dublin Center for Microscopy and Analysis, for performing the surface measurements. We also wish to thank members of the staff at Steinmann Institute, Nils Jung and Thomas Schulz at the University of Bonn, and Raphael Njul at the Bavarian Research Institute of Experimental Geochemistry and Geophysics (BGI), University of Bayreuth, for their support of the sample embedding and preparation. This work is associated with the MAD Soil project, which was funded by the DFG (Deutsche Forschungsgemeinschaft, Research Unit 2179). We very much acknowledge the help of two anonymous reviewers.

References

ARTICLE SECTIONS
Jump To

This article references 40 other publications.

  1. 1
    Köhler, P. Using the Suess effect on the stabile carbon isotope to distinguish the future from the past in radiocarbon. Environ. Res. Lett. 2016, 11, 124016,  DOI: 10.1088/1748-9326/11/12/124016
  2. 2
    Beer, K. D.; Orellana, M. V.; Baliga, N. S. Modeling the evolution of C4 photosynthesis. Cell 2013, 153, 14271429,  DOI: 10.1016/j.cell.2013.05.058
  3. 3
    Kellogg, E. A. C4 photosynthesis. Curr. Biol. 2013, 23, R5949,  DOI: 10.1016/j.cub.2013.04.066
  4. 4
    Moreno-Gutiérrez, C.; Dawson, T. E.; Nicolás, E.; Querejeta, J. I. Isotopes reveal contrasting water use strategies among coexisting plant species in a Mediterranean ecosystem. New Phytol. 2012, 196, 489496,  DOI: 10.1111/j.1469-8137.2012.04276.x
  5. 5
    Gebrekirstos, A.; van Noordwijk, M.; Neufeldt, H.; Mitlöhner, R. Relationships of stable carbon isotopes, plant water potential and growth: an approach to asses water use efficiency and growth strategies of dry land agroforestry species. Trees 2011, 25, 95102,  DOI: 10.1007/s00468-010-0467-0
  6. 6
    Hartman, G.; Danin, A. Isotopic values of plants in relation to water availability in the Eastern Mediterranean region. Oecologia 2010, 162, 837852,  DOI: 10.1007/s00442-009-1514-7
  7. 7
    Hudson, A. M.; Quade, J.; Ali, G.; Boyle, D.; Bassett, S.; Huntington, K. W.; De los Santos, M. G.; Cohen, A. S.; Lin, K.; Wang, X. Stable C, O and clumped isotope systematics and 14C geochronology of carbonates from the Quaternary Chewaucan closed-basin lake system, Great Basin, USA: Implications for paleoenvironmental reconstructions using carbonates. Geochim. Cosmochim. Ac. 2017, 212, 274302,  DOI: 10.1016/j.gca.2017.06.024
  8. 8
    Wilson, K. E.; Maslin, M. A.; Leng, M. J.; Kingston, J. D.; Deino, A. L.; Edgar, R. K.; Mackay, A. W. East African lake evidence for Pliocene millennial-scale climate variability. Geology 2014, 42, 955958,  DOI: 10.1130/G35915.1
  9. 9
    Laskov, C.; Amelung, W.; Peiffer, S. Organic Matter Preservation in the Sediment of an Acidic Mining Lake. Environ. Sci. Technol. 2002, 36, 42184223,  DOI: 10.1021/es010320y
  10. 10
    Derrien, D.; Amelung, W. Computing the mean residence time of soil carbon fractions using stable isotopes: impacts of the model framework. Europ. J. Soil Sci. 2011, 62, 237252,  DOI: 10.1111/j.1365-2389.2010.01333.x
  11. 11
    Still, C. J.; Berry, J. A.; Collatz, G. J.; DeFries, R. S. Global distribution of C 3 and C 4 vegetation: Carbon cycle implications. Global Biogeochem. Cycles 2003, 17, 6-16-14,  DOI: 10.1029/2001GB001807
  12. 12
    Blagodatskaya, E.; Yuyukina, T.; Blagodatsky, S.; Kuzyakov, Y. Turnover of soil organic matter and of microbial biomass under C3–C4 vegetation change: Consideration of 13C fractionation and preferential substrate utilization. Soil Biol. Biochem. 2011, 43, 159166,  DOI: 10.1016/j.soilbio.2010.09.028
  13. 13
    Schemmel, F.; Mikes, T.; Rojay, B.; Mulch, A. The impact of topography on isotopes in precipitation across the Central Anatolian Plateau (Turkey). Am. J. Sci. 2013, 313, 6180,  DOI: 10.2475/02.2013.01
  14. 14
    Leier, A.; Quade, J.; DeCelles, P.; Kapp, P. Stable isotopic results from paleosol carbonate in South Asia: Paleoenvironmental reconstructions and selective alteration. Earth Planet. Sci. Lett. 2009, 279, 242254,  DOI: 10.1016/j.epsl.2008.12.044
  15. 15
    Sikes, N. E.; Ashley, G. M. Stable isotopes of pedogenic carbonates as indicators of paleoecology in the Plio-Pleistocene (upper Bed I), western margin of the Olduvai Basin, Tanzania. J. Hum. Evol. 2007, 53, 574594,  DOI: 10.1016/j.jhevol.2006.12.008
  16. 16
    Sankaran, M.; Hanan, N. P.; Scholes, R. J.; Ratnam, J.; Augustine, D. J.; Cade, B. S.; Gignoux, J.; Higgins, S. I.; Le Roux, X.; Ludwig, F.; Ardo, J.; Banyikwa, F.; Bronn, A.; Bucini, G.; Caylor, K. K.; Coughenour, M. B.; Diouf, A.; Ekaya, W.; Feral, C. J.; February, E. C.; Frost, P. G. H.; Hiernaux, P.; Hrabar, H.; Metzger, K. L.; Prins, H. H. T.; Ringrose, S.; Sea, W.; Tews, J.; Worden, J.; Zambatis, N. Determinants of woody cover in African savannas. Nature 2005, 438, 846849,  DOI: 10.1038/nature04070
  17. 17
    van Malderen, S. J. M.; Laforce, B.; van Acker, T.; Vincze, L.; Vanhaecke, F. Imaging the 3D trace metal and metalloid distribution in mature wheat and rye grains via laser ablation-ICP-mass spectrometry and micro-X-ray fluorescence spectrometry. J. Anal. At. Spectrom. 2017, 32, 289298,  DOI: 10.1039/C6JA00357E
  18. 18
    van Malderen, S. J. M.; Vergucht, E.; De Rijcke, M.; Janssen, C.; Vincze, L.; Vanhaecke, F. Quantitative Determination and Subcellular Imaging of Cu in Single Cells via Laser Ablation-ICP-Mass Spectrometry Using High-Density Microarray Gelatin Standards. Anal. Chem. 2016, 88, 57835789,  DOI: 10.1021/acs.analchem.6b00334
  19. 19
    Wu, B.; Becker, J. S. Imaging of elements and molecules in biological tissues and cells in the low-micrometer and nanometer range. Int. J. Mass Spectrom. 2011, 307, 112122,  DOI: 10.1016/j.ijms.2011.01.019
  20. 20
    Limbeck, A.; Galler, P.; Bonta, M.; Bauer, G.; Nischkauer, W.; Vanhaecke, F. Recent advances in quantitative LA-ICP-MS analysis: challenges and solutions in the life sciences and environmental chemistry. Anal. Bioanal. Chem. 2015, 407, 65936617,  DOI: 10.1007/s00216-015-8858-0
  21. 21
    Scarciglia, F.; Barca, D.; De Rosa, R.; Pulice, I. Application of laser ablation ICP-MS and traditional micromorphological techniques to the study of an Alfisol (Sardinia, Italy) in thin sections: Insights into trace element distribution. Geoderma 2009, 152, 113126,  DOI: 10.1016/j.geoderma.2009.05.021
  22. 22
    Spötl, C.; Mattey, D. Stable isotope microsampling of speleothems for palaeoenvironmental studies: A comparison of microdrill, micromill and laser ablation techniques. Chem. Geol. 2006, 235, 4858,  DOI: 10.1016/j.chemgeo.2006.06.003
  23. 23
    Bruneau, P. M. C.; Ostle, N.; Davidson, D. A.; Grieve, I. C.; Fallick, A. E. Determination of rhizosphere 13C pulse signals in soil thin sections by laser ablation isotope ratio mass spectrometry. Rapid Commun. Mass Spectrom. 2002, 16, 21902194,  DOI: 10.1002/rcm.740
  24. 24
    Schulze, B.; Wirth, C.; Linke, P.; Brand, W. A.; Kuhlmann, I.; Horna, V.; Schulze, E.-D. Laser ablation-combustion-GC-IRMS--a new method for online analysis of intra-annual variation of delta13C in tree rings. Tree Physiol. 2004, 24, 11931201,  DOI: 10.1093/treephys/24.11.1193
  25. 25
    Moran, J. J.; Newburn, M. K.; Alexander, M. L.; Sams, R. L.; Kelly, J. F.; Kreuzer, H. W. Laser ablation isotope ratio mass spectrometry for enhanced sensitivity and spatial resolution in stable isotope analysis. Rapid Commun. Mass Spectrom. 2011, 25, 12821290,  DOI: 10.1002/rcm.4985
  26. 26
    van Roij, L.; Sluijs, A.; Laks, J. J.; Reichart, G.-J. Stable carbon isotope analyses of nanogram quantities of particulate organic carbon (pollen) with laser ablation nano combustion gas chromatography/isotope ratio mass spectrometry. Rapid Commun. Mass Spectrom. 2017, 31, 4758,  DOI: 10.1002/rcm.7769
  27. 27
    Brand, W. A. PreCon: A Fully Automated Interface for the Pre-Gc Concentration of Trace Gases on Air for Isotopic Analysis. Isot. Environ. Health Stud. 1995, 31, 277284,  DOI: 10.1080/10256019508036271
  28. 28
    Günther, D.; Hattendorf, B. Solid sample analysis using laser ablation inductively coupled plasma mass spectrometry. TrAC, Trends Anal. Chem. 2005, 24, 255265,  DOI: 10.1016/j.trac.2004.11.017
  29. 29
    Brand, W. A.; Coplen, T. B.; Vogl, J.; Rosner, M.; Prohaska, T. Assessment of international reference materials for isotope-ratio analysis (IUPAC Technical Report). Pure Appl. Chem. 2014, 86, 425467,  DOI: 10.1515/pac-2013-1023
  30. 30
    Krause, L.; Rodionov, A.; Schweizer, S. A.; Siebers, N.; Lehndorff, E.; Klumpp, E.; Amelung, W. Microaggregate stability and storage of organic carbon is affected by clay content in arable Luvisols. Soil Tillage Res. 2018, 182, 123129,  DOI: 10.1016/j.still.2018.05.003
  31. 31
    Glisczynski, F. v.; Pude, R.; Amelung, W.; Sandhage-Hofmann, A. Biochar-compost substrates in short-rotation coppice: Effects on soil and trees in a three-year field experiment. J. Plant Nutr. Soil Sci. 2016, 179, 574583,  DOI: 10.1002/jpln.201500545
  32. 32
    Schiedung, H.; Tilly, N.; Hütt, C.; Welp, G.; Brüggemann, N.; Amelung, W. Spatial controls of topsoil and subsoil organic carbon turnover under C3–C4 vegetation change. Geoderma 2017, 303, 4451,  DOI: 10.1016/j.geoderma.2017.05.006
  33. 33
    Jeong, S. H.; Borisov, O. V.; Yoo, J. H.; Mao, X. L.; Russo, R. E. Effects of Particle Size Distribution on Inductively Coupled Plasma Mass Spectrometry Signal Intensity during Laser Ablation of Glass Samples. Anal. Chem. 1999, 71, 51235130,  DOI: 10.1021/ac990455a
  34. 34
    Ohlsson, K. E. A. Uncertainty of blank correction in isotope ratio measurement. Anal. Chem. 2013, 85, 53265329,  DOI: 10.1021/ac4003968
  35. 35
    Werner, R. A.; Brand, W. A. Referencing strategies and techniques in stable isotope ratio analysis. Rapid Commun. Mass Spectrom. 2001, 15, 501519,  DOI: 10.1002/rcm.258
  36. 36
    Werner, R. A.; Bruch, B. A.; Brand, W. A. ConFlo III - an interface for high precision δ13C and δ15N analysis with an extended dynamic range. Rapid Commun. Mass Spectrom. 1999, 13, 12371241,  DOI: 10.1002/(SICI)1097-0231(19990715)13:13<1237::AID-RCM633>3.0.CO;2-C
  37. 37
    Borisov, O. V.; Mao, X.; Russo, R. E. Effects of crater development on fractionation and signal intensity during laser ablation inductively coupled plasma mass spectrometry. Spectrochim. Acta, Part B 2000, 55, 16931704,  DOI: 10.1016/S0584-8547(00)00272-X
  38. 38
    Bleiner, D.; Plotnikov, A.; Vogt, C.; Wetzig, K.; Günther, D. Depth profile analysis of various titanium based coatings on steel and tungsten carbide using laser ablation inductively coupled plasma--″time of flight″ mass spectrometry. Fresenius' J. Anal. Chem. 2000, 368, 221226,  DOI: 10.1007/s002160000417
  39. 39
    Paul, E. A.; Follett, R. F.; Leavitt, S. W.; Halvorson, A.; Peterson, G. A.; Lyon, D. J. Radiocarbon dating for determination of soil organic matter pool sizes and dynamics. Soil Sci. Soc. Am. J. 1997, 61, 10581067,  DOI: 10.2136/sssaj1997.03615995006100040011x
  40. 40
    Holz, M.; Zarebanadkouki, M.; Kuzyakov, Y.; Pausch, J.; Carminati, A. Root hairs increase rhizosphere extension and carbon input to soil. Ann. Bot. 2018, 121, 6169,  DOI: 10.1093/aob/mcx127

Cited By

ARTICLE SECTIONS
Jump To

This article is cited by 22 publications.

  1. Eva Lippold, Steffen Schlüter, Carsten W. Mueller, Carmen Höschen, Gertraud Harrington, Rüdiger Kilian, Martina I. Gocke, Eva Lehndorff, Robert Mikutta, Doris Vetterlein. Correlative Imaging of the Rhizosphere─A Multimethod Workflow for Targeted Mapping of Chemical Gradients. Environmental Science & Technology 2023, 57 (3) , 1538-1549. https://doi.org/10.1021/acs.est.2c07340
  2. Timothy Vazquez, Colette Taylor, Maverick Knowlton, Sean Williams, Theresa Evans-Nguyen. Enabling Isotope Ratio Measurements on an Ion Trap Mass Spectrometer. Journal of the American Society for Mass Spectrometry 2020, 31 (8) , 1722-1729. https://doi.org/10.1021/jasms.0c00143
  3. Amir H. Ahkami, Odeta Qafoku, Tiina Roose, Quanbing Mou, Yi Lu, Zoe G. Cardon, Yuxin Wu, Chunwei Chou, Joshua B. Fisher, Tamas Varga, Pubudu Handakumbura, Jayde A. Aufrecht, Arunima Bhattacharjee, James J. Moran. Emerging sensing, imaging, and computational technologies to scale nano-to macroscale rhizosphere dynamics – Review and research perspectives. Soil Biology and Biochemistry 2024, 189 , 109253. https://doi.org/10.1016/j.soilbio.2023.109253
  4. Daniel M. Cleary, Timothy J. Linley, Jason M. Kriesel, Andrew Fahrland, James F. Kelly, James J. Moran. Laser Ablation-Capillary Absorption Spectroscopy: A novel approach for high throughput and increased spatial resolution measurements of δ13C in plant-soil systems. Soil Biology and Biochemistry 2023, 187 , 109208. https://doi.org/10.1016/j.soilbio.2023.109208
  5. Chongkai Hao, Ping Du, Jie Ren, Lijuan Hu, Zongpeng Zhang. Halophyte Elymus dahuricus colonization regulates microbial community succession by mediating saline-alkaline and biogenic organic matter in bauxite residue. Science of The Total Environment 2023, 905 , 167140. https://doi.org/10.1016/j.scitotenv.2023.167140
  6. Franziska E. Zahn, Erik Söll, Thomas K. Chapin, Deyi Wang, Sofia I. F. Gomes, Nicole A. Hynson, Johanna Pausch, Gerhard Gebauer. Novel insights into orchid mycorrhiza functioning from stable isotope signatures of fungal pelotons. New Phytologist 2023, 239 (4) , 1449-1463. https://doi.org/10.1111/nph.18991
  7. Wulf Amelung, Nele Meyer, Andrey Rodionov, Claudia Knief, Michaela Aehnelt, Sara L. Bauke, Danh Biesgen, Stefan Dultz, Georg Guggenberger, Maguy Jaber, Erwin Klumpp, Ingrid Kögel-Knabner, Volker Nischwitz, Steffen A. Schweizer, Bei Wu, Kai U. Totsche, Eva Lehndorff. Process sequence of soil aggregate formation disentangled through multi-isotope labelling. Geoderma 2023, 429 , 116226. https://doi.org/10.1016/j.geoderma.2022.116226
  8. Charlotte Védère, Laure Vieublé Gonod, Naoise Nunan, Claire Chenu. Opportunities and limits in imaging microorganisms and their activities in soil microhabitats. Soil Biology and Biochemistry 2022, 174 , 108807. https://doi.org/10.1016/j.soilbio.2022.108807
  9. A. Schnepf, A. Carminati, M. A. Ahmed, M. Ani, P. Benard, J. Bentz, M. Bonkowski, M. Knott, D. Diehl, P. Duddek, E. Kröner, M. Javaux, M. Landl, E. Lehndorff, E. Lippold, A. Lieu, C. W. Mueller, E. Oburger, W. Otten, X. Portell, M. Phalempin, A. Prechtel, R. Schulz, J. Vanderborght, D. Vetterlein. Linking rhizosphere processes across scales: Opinion. Plant and Soil 2022, 478 (1-2) , 5-42. https://doi.org/10.1007/s11104-022-05306-7
  10. James J. Moran, Timothy J. Linley, Camille N. Makarem, James F. Kelly, Eric D. Wilcox Freeburg, Daniel M. Cleary, M. Lizabeth Alexander, Jason M. Kriesel. Spectroscopy-based isotopic (δ13C) analysis for high spatial resolution of carbon exchange in the rhizosphere. Rhizosphere 2022, 23 , 100564. https://doi.org/10.1016/j.rhisph.2022.100564
  11. Xue-Ling Chang, Lingyun Chen, Boning Liu, Sheng-Tao Yang, Haifang Wang, Aoneng Cao, Chunying Chen. Stable isotope labeling of nanomaterials for biosafety evaluation and drug development. Chinese Chemical Letters 2022, 33 (7) , 3303-3314. https://doi.org/10.1016/j.cclet.2022.03.057
  12. Anna S. Wendel, Sara L. Bauke, Wulf Amelung, Claudia Knief. Root-rhizosphere-soil interactions in biopores. Plant and Soil 2022, 475 (1-2) , 253-277. https://doi.org/10.1007/s11104-022-05406-4
  13. Riccardo Finotello, Mohamed Tamaazousti, Jean-Baptiste Sirven. HyperPCA: A powerful tool to extract elemental maps from noisy data obtained in LIBS mapping of materials. Spectrochimica Acta Part B: Atomic Spectroscopy 2022, 192 , 106418. https://doi.org/10.1016/j.sab.2022.106418
  14. Qiuyue Shi, Cheng Fang, Changzhou Yan, Xue-Ling Chang, Xian Zhang, Han Zhang. Visualization of fullerenol nanoparticles distribution in Daphnia magna using Laser Ablation-isotope Ratio Mass (LA-IRMS) and Matrix-assisted Laser Desorption/Ionization Imaging Mass Spectrometry (MALDI-IMS). Ecotoxicology and Environmental Safety 2022, 232 , 113226. https://doi.org/10.1016/j.ecoenv.2022.113226
  15. Peng Yao, Haiyan Ni, Dipayan Paul, Agne Masalaite, Ru-Jin Huang, Harro A.J. Meijer, Ulrike Dusek. An automated method for thermal-optical separation of aerosol organic/elemental carbon for 13C analysis at the sub-μgC level: A comprehensive assessment. Science of The Total Environment 2022, 804 , 150031. https://doi.org/10.1016/j.scitotenv.2021.150031
  16. Tom Guhra, Katharina Stolze, Kai Uwe Totsche. Pathways of biogenically excreted organic matter into soil aggregates. Soil Biology and Biochemistry 2022, 164 , 108483. https://doi.org/10.1016/j.soilbio.2021.108483
  17. Manuel Vergara Sosa, Eva Lehndorff, Andrei Rodionov, Martina Gocke, Alexandra Sandhage-Hofmann, Wulf Amelung. Micro-scale resolution of carbon turnover in soil - Insights from laser ablation isotope ratio mass spectrometry on water-glass embedded aggregates. Soil Biology and Biochemistry 2021, 159 , 108279. https://doi.org/10.1016/j.soilbio.2021.108279
  18. Carsten W. Mueller, Markus Steffens, Henning Buddenbaum. Permafrost soil complexity evaluated by laboratory imaging Vis ‐ NIR spectroscopy. European Journal of Soil Science 2021, 72 (1) , 114-119. https://doi.org/10.1111/ejss.12927
  19. Doris Vetterlein, Andrea Carminati, Ingrid Kögel-Knabner, Gerd Patrick Bienert, Kornelia Smalla, Eva Oburger, Andrea Schnepf, Thomas Banitz, Mika Tapio Tarkka, Steffen Schlüter. Rhizosphere Spatiotemporal Organization–A Key to Rhizosphere Functions. Frontiers in Agronomy 2020, 2 https://doi.org/10.3389/fagro.2020.00008
  20. Han ZHANG, Lai WEI, Xu LIAO, Jia-Ni WANG, Qiu-Yue SHI, Xian ZHANG. In Situ Carbon Stable isotope Analysis of Organic Carbon by Laser Ablation-Isotope Ratio Mass Spectrometry. Chinese Journal of Analytical Chemistry 2020, 48 (6) , 774-779. https://doi.org/10.1016/S1872-2040(20)60028-7
  21. E. Hywel Evans, Jorge Pisonero, Clare M. M. Smith, Rex N. Taylor. Atomic spectrometry update: review of advances in atomic spectrometry and related techniques. Journal of Analytical Atomic Spectrometry 2020, 35 (5) , 830-851. https://doi.org/10.1039/D0JA90015J
  22. Jeffrey R. Bacon, Owen T. Butler, Warren R. L. Cairns, Jennifer M. Cook, Christine M. Davidson, Olga Cavoura, Regina Mertz-Kraus. Atomic spectrometry update – a review of advances in environmental analysis. Journal of Analytical Atomic Spectrometry 2020, 35 (1) , 9-53. https://doi.org/10.1039/C9JA90060H
  • Abstract

    Figure 1

    Figure 1. Setup of the LA-IRMS system starting with a newly constructed low-volume cell. Only the upper part of the cell is the active operating volume for laser ablation (marked in red; more details in the Supporting Information).

    Figure 2

    Figure 2. Impact of loading time in the first cryo-trap on the blank peak area in cells with 3.8 and 0.4 mL active volume (without laser ablation). If the loading time was <7 s, the blank was at the baseline level of the IRMS detector for both cells. The isotopic signature of the blank signal was not affected by the loading time.

    Figure 3

    Figure 3. Discrimination of stable carbon isotopes from different organic matrices by ablation with increasing laser energy.

    Figure 4

    Figure 4. Detection limit of the LA-IRMS system given in C amount (a) for the cell with 3.8 mL active volume and (b) for the cell with 0.4 mL active volume (detected with an acryl plate with known ablated crater volume; see also Supporting Information).

    Figure 5

    Figure 5. Spatially resolved microanalysis of δ13C heterogeneity in a soil microaggregate structure from a cropped Luvisol field of 33% clay (53–250 μm aggregate fraction, 20 μm spot size, 40 shots per ablation). Raw δ13C values shown in red and blank corrected δ13C values, in black.

    Figure 6

    Figure 6. Spatial differences of isotopic signatures inside the root, rhizosphere area, and soil matrix from a Miscanthus root ablated with (a) constant spot size of 30 μm and 30 shot numbers and with (b) variable spot size of 15 μm and 40 shot numbers (white-colored signatures) and spot size of 10 μm and 80 shot numbers (black colored signatures), both performed on a 60 μm cross section; in-house reference acryl = −29.75 ± 0.04‰.

    Figure 7

    Figure 7. Spatial differences of isotopic signatures ablated from 2 mm sieved soil with constant spot size of 10 μm and 60 shot numbers after 2 years of C3 to C4 vegetation change (Zea mais L., Selhausen; in-house reference acryl = −29.57 ± 0.32‰).

  • References

    ARTICLE SECTIONS
    Jump To

    This article references 40 other publications.

    1. 1
      Köhler, P. Using the Suess effect on the stabile carbon isotope to distinguish the future from the past in radiocarbon. Environ. Res. Lett. 2016, 11, 124016,  DOI: 10.1088/1748-9326/11/12/124016
    2. 2
      Beer, K. D.; Orellana, M. V.; Baliga, N. S. Modeling the evolution of C4 photosynthesis. Cell 2013, 153, 14271429,  DOI: 10.1016/j.cell.2013.05.058
    3. 3
      Kellogg, E. A. C4 photosynthesis. Curr. Biol. 2013, 23, R5949,  DOI: 10.1016/j.cub.2013.04.066
    4. 4
      Moreno-Gutiérrez, C.; Dawson, T. E.; Nicolás, E.; Querejeta, J. I. Isotopes reveal contrasting water use strategies among coexisting plant species in a Mediterranean ecosystem. New Phytol. 2012, 196, 489496,  DOI: 10.1111/j.1469-8137.2012.04276.x
    5. 5
      Gebrekirstos, A.; van Noordwijk, M.; Neufeldt, H.; Mitlöhner, R. Relationships of stable carbon isotopes, plant water potential and growth: an approach to asses water use efficiency and growth strategies of dry land agroforestry species. Trees 2011, 25, 95102,  DOI: 10.1007/s00468-010-0467-0
    6. 6
      Hartman, G.; Danin, A. Isotopic values of plants in relation to water availability in the Eastern Mediterranean region. Oecologia 2010, 162, 837852,  DOI: 10.1007/s00442-009-1514-7
    7. 7
      Hudson, A. M.; Quade, J.; Ali, G.; Boyle, D.; Bassett, S.; Huntington, K. W.; De los Santos, M. G.; Cohen, A. S.; Lin, K.; Wang, X. Stable C, O and clumped isotope systematics and 14C geochronology of carbonates from the Quaternary Chewaucan closed-basin lake system, Great Basin, USA: Implications for paleoenvironmental reconstructions using carbonates. Geochim. Cosmochim. Ac. 2017, 212, 274302,  DOI: 10.1016/j.gca.2017.06.024
    8. 8
      Wilson, K. E.; Maslin, M. A.; Leng, M. J.; Kingston, J. D.; Deino, A. L.; Edgar, R. K.; Mackay, A. W. East African lake evidence for Pliocene millennial-scale climate variability. Geology 2014, 42, 955958,  DOI: 10.1130/G35915.1
    9. 9
      Laskov, C.; Amelung, W.; Peiffer, S. Organic Matter Preservation in the Sediment of an Acidic Mining Lake. Environ. Sci. Technol. 2002, 36, 42184223,  DOI: 10.1021/es010320y
    10. 10
      Derrien, D.; Amelung, W. Computing the mean residence time of soil carbon fractions using stable isotopes: impacts of the model framework. Europ. J. Soil Sci. 2011, 62, 237252,  DOI: 10.1111/j.1365-2389.2010.01333.x
    11. 11
      Still, C. J.; Berry, J. A.; Collatz, G. J.; DeFries, R. S. Global distribution of C 3 and C 4 vegetation: Carbon cycle implications. Global Biogeochem. Cycles 2003, 17, 6-16-14,  DOI: 10.1029/2001GB001807
    12. 12
      Blagodatskaya, E.; Yuyukina, T.; Blagodatsky, S.; Kuzyakov, Y. Turnover of soil organic matter and of microbial biomass under C3–C4 vegetation change: Consideration of 13C fractionation and preferential substrate utilization. Soil Biol. Biochem. 2011, 43, 159166,  DOI: 10.1016/j.soilbio.2010.09.028
    13. 13
      Schemmel, F.; Mikes, T.; Rojay, B.; Mulch, A. The impact of topography on isotopes in precipitation across the Central Anatolian Plateau (Turkey). Am. J. Sci. 2013, 313, 6180,  DOI: 10.2475/02.2013.01
    14. 14
      Leier, A.; Quade, J.; DeCelles, P.; Kapp, P. Stable isotopic results from paleosol carbonate in South Asia: Paleoenvironmental reconstructions and selective alteration. Earth Planet. Sci. Lett. 2009, 279, 242254,  DOI: 10.1016/j.epsl.2008.12.044
    15. 15
      Sikes, N. E.; Ashley, G. M. Stable isotopes of pedogenic carbonates as indicators of paleoecology in the Plio-Pleistocene (upper Bed I), western margin of the Olduvai Basin, Tanzania. J. Hum. Evol. 2007, 53, 574594,  DOI: 10.1016/j.jhevol.2006.12.008
    16. 16
      Sankaran, M.; Hanan, N. P.; Scholes, R. J.; Ratnam, J.; Augustine, D. J.; Cade, B. S.; Gignoux, J.; Higgins, S. I.; Le Roux, X.; Ludwig, F.; Ardo, J.; Banyikwa, F.; Bronn, A.; Bucini, G.; Caylor, K. K.; Coughenour, M. B.; Diouf, A.; Ekaya, W.; Feral, C. J.; February, E. C.; Frost, P. G. H.; Hiernaux, P.; Hrabar, H.; Metzger, K. L.; Prins, H. H. T.; Ringrose, S.; Sea, W.; Tews, J.; Worden, J.; Zambatis, N. Determinants of woody cover in African savannas. Nature 2005, 438, 846849,  DOI: 10.1038/nature04070
    17. 17
      van Malderen, S. J. M.; Laforce, B.; van Acker, T.; Vincze, L.; Vanhaecke, F. Imaging the 3D trace metal and metalloid distribution in mature wheat and rye grains via laser ablation-ICP-mass spectrometry and micro-X-ray fluorescence spectrometry. J. Anal. At. Spectrom. 2017, 32, 289298,  DOI: 10.1039/C6JA00357E
    18. 18
      van Malderen, S. J. M.; Vergucht, E.; De Rijcke, M.; Janssen, C.; Vincze, L.; Vanhaecke, F. Quantitative Determination and Subcellular Imaging of Cu in Single Cells via Laser Ablation-ICP-Mass Spectrometry Using High-Density Microarray Gelatin Standards. Anal. Chem. 2016, 88, 57835789,  DOI: 10.1021/acs.analchem.6b00334
    19. 19
      Wu, B.; Becker, J. S. Imaging of elements and molecules in biological tissues and cells in the low-micrometer and nanometer range. Int. J. Mass Spectrom. 2011, 307, 112122,  DOI: 10.1016/j.ijms.2011.01.019
    20. 20
      Limbeck, A.; Galler, P.; Bonta, M.; Bauer, G.; Nischkauer, W.; Vanhaecke, F. Recent advances in quantitative LA-ICP-MS analysis: challenges and solutions in the life sciences and environmental chemistry. Anal. Bioanal. Chem. 2015, 407, 65936617,  DOI: 10.1007/s00216-015-8858-0
    21. 21
      Scarciglia, F.; Barca, D.; De Rosa, R.; Pulice, I. Application of laser ablation ICP-MS and traditional micromorphological techniques to the study of an Alfisol (Sardinia, Italy) in thin sections: Insights into trace element distribution. Geoderma 2009, 152, 113126,  DOI: 10.1016/j.geoderma.2009.05.021
    22. 22
      Spötl, C.; Mattey, D. Stable isotope microsampling of speleothems for palaeoenvironmental studies: A comparison of microdrill, micromill and laser ablation techniques. Chem. Geol. 2006, 235, 4858,  DOI: 10.1016/j.chemgeo.2006.06.003
    23. 23
      Bruneau, P. M. C.; Ostle, N.; Davidson, D. A.; Grieve, I. C.; Fallick, A. E. Determination of rhizosphere 13C pulse signals in soil thin sections by laser ablation isotope ratio mass spectrometry. Rapid Commun. Mass Spectrom. 2002, 16, 21902194,  DOI: 10.1002/rcm.740
    24. 24
      Schulze, B.; Wirth, C.; Linke, P.; Brand, W. A.; Kuhlmann, I.; Horna, V.; Schulze, E.-D. Laser ablation-combustion-GC-IRMS--a new method for online analysis of intra-annual variation of delta13C in tree rings. Tree Physiol. 2004, 24, 11931201,  DOI: 10.1093/treephys/24.11.1193
    25. 25
      Moran, J. J.; Newburn, M. K.; Alexander, M. L.; Sams, R. L.; Kelly, J. F.; Kreuzer, H. W. Laser ablation isotope ratio mass spectrometry for enhanced sensitivity and spatial resolution in stable isotope analysis. Rapid Commun. Mass Spectrom. 2011, 25, 12821290,  DOI: 10.1002/rcm.4985
    26. 26
      van Roij, L.; Sluijs, A.; Laks, J. J.; Reichart, G.-J. Stable carbon isotope analyses of nanogram quantities of particulate organic carbon (pollen) with laser ablation nano combustion gas chromatography/isotope ratio mass spectrometry. Rapid Commun. Mass Spectrom. 2017, 31, 4758,  DOI: 10.1002/rcm.7769
    27. 27
      Brand, W. A. PreCon: A Fully Automated Interface for the Pre-Gc Concentration of Trace Gases on Air for Isotopic Analysis. Isot. Environ. Health Stud. 1995, 31, 277284,  DOI: 10.1080/10256019508036271
    28. 28
      Günther, D.; Hattendorf, B. Solid sample analysis using laser ablation inductively coupled plasma mass spectrometry. TrAC, Trends Anal. Chem. 2005, 24, 255265,  DOI: 10.1016/j.trac.2004.11.017
    29. 29
      Brand, W. A.; Coplen, T. B.; Vogl, J.; Rosner, M.; Prohaska, T. Assessment of international reference materials for isotope-ratio analysis (IUPAC Technical Report). Pure Appl. Chem. 2014, 86, 425467,  DOI: 10.1515/pac-2013-1023
    30. 30
      Krause, L.; Rodionov, A.; Schweizer, S. A.; Siebers, N.; Lehndorff, E.; Klumpp, E.; Amelung, W. Microaggregate stability and storage of organic carbon is affected by clay content in arable Luvisols. Soil Tillage Res. 2018, 182, 123129,  DOI: 10.1016/j.still.2018.05.003
    31. 31
      Glisczynski, F. v.; Pude, R.; Amelung, W.; Sandhage-Hofmann, A. Biochar-compost substrates in short-rotation coppice: Effects on soil and trees in a three-year field experiment. J. Plant Nutr. Soil Sci. 2016, 179, 574583,  DOI: 10.1002/jpln.201500545
    32. 32
      Schiedung, H.; Tilly, N.; Hütt, C.; Welp, G.; Brüggemann, N.; Amelung, W. Spatial controls of topsoil and subsoil organic carbon turnover under C3–C4 vegetation change. Geoderma 2017, 303, 4451,  DOI: 10.1016/j.geoderma.2017.05.006
    33. 33
      Jeong, S. H.; Borisov, O. V.; Yoo, J. H.; Mao, X. L.; Russo, R. E. Effects of Particle Size Distribution on Inductively Coupled Plasma Mass Spectrometry Signal Intensity during Laser Ablation of Glass Samples. Anal. Chem. 1999, 71, 51235130,  DOI: 10.1021/ac990455a
    34. 34
      Ohlsson, K. E. A. Uncertainty of blank correction in isotope ratio measurement. Anal. Chem. 2013, 85, 53265329,  DOI: 10.1021/ac4003968
    35. 35
      Werner, R. A.; Brand, W. A. Referencing strategies and techniques in stable isotope ratio analysis. Rapid Commun. Mass Spectrom. 2001, 15, 501519,  DOI: 10.1002/rcm.258
    36. 36
      Werner, R. A.; Bruch, B. A.; Brand, W. A. ConFlo III - an interface for high precision δ13C and δ15N analysis with an extended dynamic range. Rapid Commun. Mass Spectrom. 1999, 13, 12371241,  DOI: 10.1002/(SICI)1097-0231(19990715)13:13<1237::AID-RCM633>3.0.CO;2-C
    37. 37
      Borisov, O. V.; Mao, X.; Russo, R. E. Effects of crater development on fractionation and signal intensity during laser ablation inductively coupled plasma mass spectrometry. Spectrochim. Acta, Part B 2000, 55, 16931704,  DOI: 10.1016/S0584-8547(00)00272-X
    38. 38
      Bleiner, D.; Plotnikov, A.; Vogt, C.; Wetzig, K.; Günther, D. Depth profile analysis of various titanium based coatings on steel and tungsten carbide using laser ablation inductively coupled plasma--″time of flight″ mass spectrometry. Fresenius' J. Anal. Chem. 2000, 368, 221226,  DOI: 10.1007/s002160000417
    39. 39
      Paul, E. A.; Follett, R. F.; Leavitt, S. W.; Halvorson, A.; Peterson, G. A.; Lyon, D. J. Radiocarbon dating for determination of soil organic matter pool sizes and dynamics. Soil Sci. Soc. Am. J. 1997, 61, 10581067,  DOI: 10.2136/sssaj1997.03615995006100040011x
    40. 40
      Holz, M.; Zarebanadkouki, M.; Kuzyakov, Y.; Pausch, J.; Carminati, A. Root hairs increase rhizosphere extension and carbon input to soil. Ann. Bot. 2018, 121, 6169,  DOI: 10.1093/aob/mcx127
  • Supporting Information

    Supporting Information

    ARTICLE SECTIONS
    Jump To

    The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.analchem.9b00892.

    • One table and six figures showing the determination of carbon amount by measurements of crater depth on an acryl plate (Table S-1), configuration of the newly developed laser cells (Figure S-1), images of press-pellets from the ground and mixed standard substances (Figure S-2), images of spot depth and surface mapping of ablated spots (Figure S-3), deviation of the δ13C signal from referenced values due to the contribution of the blank signal (Figure S-4), adjustment of loading times in the cryo-trap (Figure S-5), and stable C isotope signatures of a Miscanthus rhizome and the surrounding soil (Figure S-6) (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

You’ve supercharged your research process with ACS and Mendeley!

STEP 1:
Click to create an ACS ID

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

MENDELEY PAIRING EXPIRED
Your Mendeley pairing has expired. Please reconnect