127
The role of higher molecular weight dissolved organic nitrogen in the plant- soil nitrogen cycle Kirsten Lønne Enggrob Phd Thesis, Science and Technology, 2019 Department of Agroecology Faculty of Science and Technology Aarhus University, Foulum Blichers Allé 20 P.O. Box 50 8830 Tjele Denmark

The role of higher molecular weight dissolved organic ... · nitrogen compounds contribute directly to bacterial tissue build-up. Thus, when large organic nitrogen compounds are dissolved,

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

The role of higher molecular weight

dissolved organic nitrogen in the plant-

soil nitrogen cycle

Kirsten Lønne Enggrob

Phd Thesis, Science and Technology, 2019

Department of Agroecology

Faculty of Science and Technology

Aarhus University, Foulum

Blichers Allé 20

P.O. Box 50

8830 Tjele

Denmark

Main supervisor

Senior Researcher Jim Rasmussen

Department of Agroecology

Aarhus University, Denmark

Co-supervisor

Associate Professor Lars Elsgaard

Department of Agroecology

Aarhus University, Denmark

Assessment Committee

Professor Mathias Neumann Andersen (Chairman)

Department of Agroecology

Aarhus University, Denmark

Associated Professor Anke M. Herrmann

Department of Soil & Environment

Swedish University of Agricultural Sciences, Sweden

Senior Lecturer Paul W. Hill

School of Natural Sciences

Bangor University, United Kingdom

i

Preface

This thesis entitled “The role of higher molecular weight dissolved organic nitrogen in the plant-soil

nitrogen cycle” is submitted in fulfilment of the requirement for the Doctor of Philosophy (PhD) degree

at Faculty of Science and Technology, Aarhus University, Denmark. This PhD project was supervised

by Senior researcher Jim Rasmussen and Associate Professor Lars Elsgaard.

This thesis is a result of work conducted from January 2015 to January 2019 at Department of

Agroecology, Aarhus University. This project was financially supported by The Independent Research

Fund Denmark – Technology and Production (Project no. 1335-00760B).

This thesis is based on the work presented in one published paper and two prepared for submission:

Paper 1:

Enggrob, K.L., Larsen, T., Larsen, M., Elsgaard, L., Rasmussen, J., 2019. The influence of hydrolysis and

derivatization on the determination of amino acid contentand isotopic ratios in dual‐labeled (13C,15N)

white clover. Rapid Commun Mass Spectrom 33, 21-30. DOI: 10.1002/rcm.8300.

Paper 2:

Enggrob, K.L., Larsen, T., Rasmussen, J. Molecular size doesn't matter for turning over large organic N

in soil. (Prepared for submission to Nature)

Paper 3:

Enggrob, K.L., Jakobsen, C.M., Pedersen, I.F., Rasmussen, J. Newly depolymerized large organic N

contributes directly to maize amino acid uptake. (Prepared for submission to New Phytologist)

ii

Acknowledgements

I would like to acknowledge my supervisor Jim Rasmussen, whom granted me the opportunity to

conduct this PhD project, without his contribution and support this work would not have been

possible. I would also like to acknowledge my co-supervisors Lars Elsgaard, whom through

discussions and guidance steered me in the right direction.

I would like to acknowledge Senior Scientist Mogens Larsen for giving me the opportunity to work

with the GC-C-IRMS, and a special thanks to the lab technicians Anne Krustrup and Birgit Hørdum Løth

for help and support with the laboratory work during my stay at Department of Animal Science. Also a

special thanks to lab technician Cecilie Kokholm and scientific assistant Janni Hansen for their help in

the laboratory, and the technicians at the Soil Fertility section in general.

Thanks to my office mates Betina Nørgaard Pedersen and Julie Therese Christensen for both moral and

work related support and a special thanks to all the other PhD students in Soil Fertility for the work

related discussions and social related debates. Also, I would like to thank all my coworkers in Soil

Fertility for provided a great work environment.

Last but not least, a special thanks to by beloved Husband Knud Erik and sons, Benjamin and

Alexander, and family and friends for their love, encouragement and continuous support.

Kirsten Lønne Enggrob

February 2019

Aarhus University, Foulum, Denmark

iii

Summary in English

Nitrogen (N) is an essential nutrient for plant growth required in large amounts. Efficient use of N in

agricultural systems is essential in the strive for sustainability in crop production and to counter the

environmental and climate change challenges related to food production. A key issue in predicting

plant available N is the turnover of complex higher molecular weight (Mw) organic N, like proteins and

peptides, to lower Mw organic N, available for direct plant and microbial uptake. However, there is a

lack of knowledge of the mechanisms controlling the fate of DON pools in soil.

In this project I investigated the role of higher Mw DON compounds in the plant-soil nitrogen cycle

with a specific focus on amino acids bound in peptides and proteins. The work was divided into three

objective namely: (i) to set up a compound specific isotope analysis (CSIA) to characterize and

quantify amino acids in 15N and 13C-labeled high Mw DON derived from white clover, (ii) to investigate

the turnover of higher Mw DON in soil with different management histories, and (iii) to investigate the

turnover of higher Mw DON in a soil with plant growth to determine the influences of the competition

between plants and soil microorganisms on the turnover of higher Mw DON and the uptake of N by

plants.

Firstly, I examined the efficiency of a standard acidic hydrolysis (6 M HCl, 20 h at 110°C) and a fast

acidic hydrolysis (6 M HCl, 70 min at 150°C) on the recovery of amino acids from a protein standard

Bovine Serum Albumin (BSA). I tested two derivaterization methods, N-acetyl methyl esterification

(NACME) and N-acetyl isopropyl esterification (NAIP), for the gas chromatography combustion

isotopic ratio mass spectrometry (GC-C-IRMS) analysis of amino acid standards. The best methods

were tested on dual-labeled (13C and 15N) clover shoot and root juice, divided in four Mw sized

fractions. The NAIP derivatization successfully resulted in higher recovery compared to the NACME

derivatization method. The NAIP derivaterization gave very low limit of detection (LOD) < 2 pmol and

limit of quantification (LOQ) ranging from 0.55-4.89 pmol across amino acids. Comparing

concentrations of individual amino acids in hydrolyzed versus un-hydrolyzed samples of the low Mw

sized fraction (< 1 kDa) showed a significant decline in concentration for seven amino acids after

hydrolysis. Despite the decline in amino acid concentration, I found a linear connection between the

obtained atomic fraction (13C and 15N) for individual amino acids of hydrolyzed versus un-hydrolyzed

samples for the <1 kDa fraction. The methodology distinguished differences in atomic fractions across

amino acid, in individual amino acid in Mw size fractions, and between shoot and root samples of

experimentally labeled white clover. Uniquely, the method separate glutamate and glutamine, which is

usually hard to achieve. Thus, the first part of my study presented an optimized methodology for GC-C-

IRMS analysis of amino acids in enriched organic N samples for 13C and 15N amino acid stable isotopic

probing (SIP).

iv

Secondly, I investigated the fate of peptide-sized and protein-sized organic N fractions in soils from

two long-term field experiments (LTE) markedly differing in condition for microorganisms. Contrary

to the present paradigm, the results showed that for all soils the exo-enzymatic depolymerization was

not per se the rate-limiting step in the turnover of these compounds nor was protection via strong

sorption to the soil mineral phase. Instead, strong evidence pointed to that gram-positive bacteria are

the key actors in the decomposition of protein-sized nitrogen compounds and that large organic

nitrogen compounds contribute directly to bacterial tissue build-up. Thus, when large organic nitrogen

compounds are dissolved, turnover occurs rapidly, irrespective of molecular size, and the bacterial

incorporation of these rapid cycling compounds potentially make an important contribution to soil

organic matter formation.

Thirdly, LTE soils with and without maize were added >100 kDa organic N, to investigate the

contribution of large Mw dissolved organic N to microbial and plant carbon (C) and N nutrition.

Mineralization of >100 kDa organic N increased with soil pH in soil without maize, but no effect of soil

pH was seen for soil with maize. The >100 kDa organic N disappeared rapidly in soils with and without

maize, but surprisingly more >100 kDa organic N derived amino acids remained in soil with than

without maize – most likely in the microbial biomass. Total 15N uptake in maize increased with higher

soil pH and the organic N uptake estimated to account for 20-30% of the total 15N uptake across the

soil pH gradient. Organic N uptake was confirmed by presence of 13C-labeled amino acids in the maize

roots. The study shows that when bio-available N is derived from large molecular sized organic N then

the importance of plant organic N uptake increases, and that rhizosphere microorganisms increase

anabolic utilization of organic N compared to bulk soil microorganisms.

v

Dansk sammendrag

Kvælstof (N) er et vigtigt plantenæringsstof, der kræves i store mængder. Effektiv anvendelse af N i

landbruget er afgørende for at øge bæredygtigheden i planteproduktionen og derved mindske de

miljømæssige og klimamæssige problemer, der er forbundet med fødevareproduktion. For at

forudsige tilgængeligheden af N skal vi forstå omdannelsen af komplekse organiske N forbindelser,

som proteiner og peptider af høj molekylvægt, til mindre organiske N forbindelser af lavere

molekylevægt. Idet mindre organiske N forbindelser er direkte tilgængelige for plante- og mikrobiel

optagelse. Vi mangler dog viden om de mekanismer, der styrer DON-puljernes skæbne i jorden.

I dette projekt undersøger jeg stort DONs rolle i plante-jord N kredsløbet, med fokus på aminosyrer

bundet i peptider og proteiner. Arbejdet var inddelt i tre delmål: (i) optimering af en stof specifik

isotop analyse (CSIA) til karakterisering og kvantificeringen af aminosyrer i 15N og 13C-mærket

højmolekylært DON fra hvidkløver, (ii) undersøgelse af omdannelsen af højmolekylært DON i jord fra

langvarige forsøg med forskellig historik og, (iii) undersøgelse af omdannelsen af højmolekylært DON i

jord med tilstedeværelsen af planter for at bestemme indflydelsen konkurrencen mellem plante og

mikroorganismer på omsætningen af højmolekylært DON, samt plantens optag af N.

Først undersøgte jeg effektiviteten af en standard hydrolyse og en hurtig hydrolyse på genfindelsen af

aminosyrer fra et standard protein Bovine Serum Albumin (BSA). Jeg testede to derivatiserings

metoder, N-acetyl methyl esterificering (NACME) and N-acetyl isopropyl esterificering (NAIP), til

analyse af aminosyrestandarder på gaskromatografisk isotop-ratio massespektrometrisk (GC-C-IRMS)

analyse. De bedste fremgangsmåder blev derefter yderligere testet på dobbelt mærket (13C og 15N)

hvidkløverblad- og rodsaft, opdelt i fire molekylevægt fraktioner. NAIP derivatiseringen resulterede i

meget lave detektions- (LOD) < 2pmol og kvantificeringsgrænser (LOQ) liggende mellem 0.55 – 4.89

pmol på tværs af aminosyrer. Sammenligningen af koncentrationen af individuelle aminosyrer fra

henholdsvis hydrolyserede versus ikke hydrolyserede prøver af lavmolekylært organisk N (<1 kDa)

viste et signifikant fald i koncentrationen fra syv aminosyrer i de hydrolyserede prøver. På trods af

nedgangen i koncentrationen af aminosyrer fandt jeg en lineær sammenhæng mellem de målte

atomfraktioner (13C og 15N) for individuelle aminosyrer fra hydrolyserede versus ikke hydrolyserede

prøver fra <1 kDa fraktionen. Fremgangsmåden kunne adskille forskellene i atomfraktionerne på

tværs af aminosyrerne, for individuelle aminosyrer på tværs af molekylærvægt fraktionerne og

mellem rod- og bladsaftsprøver fra eksperimentelt mærket hvidkløver. Ret enestående kunne

fremgangsmåden adskille glutamat og glutamin, hvilke normal er svært at opnå. Således opnåede jeg i

den første del af mit studie at lave en optimeret metode til GC-C-IRMS analyse af aminosyrer fra

organiske N prøver beriget med stabile isotoper .

vi

For det andet, undersøgte jeg omdannelsen af organiske N forbindelser af peptid- og proteinstørrelse i

jord fra to langvarige markforsøg (LTE) med markante forskelle i dyrkningshistorik og dermed de

mikrobielle miljøer. I modsætning til det nuværende paradigme, viste resultaterne for alle jorderne, at

hverken exo-enzymatiske depolymerisering eller beskyttelse højmolekylært organisk N via stærk

sorption til jordens mineralfase var begrænsende for omsætningen. I stedet for fandt jeg, at gram-

positive bakterier er nøgleaktørerne i nedbrydning af kvælstofforbindelser i proteinstørrelse, og at

store organiske nitrogenforbindelser bidrager direkte til bakteriel vævsopbygning. Forsøget viste, at

når først organiske N forbindelser er opløst, sker omsætningen hurtigt uanset molekylstørrelse, og at

bakteriel indbygning af disse stoffer potentielt udgør et vigtigt bidrag til dannelse af organisk stof i

jorden.

For det tredje, undersøgte jeg bidraget fra stort organisk N (>100 kDa) til plante og mikroorganismers

optag af kulstof (C) og N. Mineralisering af >100 kDa organisk N steg med jord pH i jord uden majs,

mens der ingen virkning var af jordens pH ved tilstedeværelse af majs. Aminosyrer fra det stor

organiske N forsvandt hurtigt både i jord med og uden majs, men overraskende genfandt jeg en større

andel af organisk N afledte aminosyrer i jord med majs end i jord uden majs - sandsynligvis fordi

aminosyrerne var indbygget i den mikrobielle biomasse. Det totale 15N optag i majs steg med jordens

pH, og N optaget i organisk form blev anslået til at udgøre 20-30% af det samlede 15N optagelse på

tværs af den undersøgte pH gradient. Det direkte optag af organiske N forbindelser blev bekræftet ved

tilstedeværelse af 13C-mærkede aminosyrer i majsrødderne.

Studiet viste, at vigtigheden af plantens optag af organisk N var større for højmolekylært organisk N

end det er blevet fundet undersøgelser af lavmolekylært organisk N, og at mikroorganismer i

rhizosfæren øger den anabolske udnyttelse af det højmolekylære organiske N sammenlignet med

mikroorganismer i jord uden planter.

vii

Abbreviations AA Amino acid

BSA Bovine Serum Albumin

C Carbon

CSIA Compound specific isotopic analysis

DIN Dissolved inorganic nitrogen

DON Dissolved organic nitrogen

GC-C-IRMS Gas chromatography combustion isotopic ratio mass spectrometry

HpH High pH soil

ISTD Internal standard

LOD Limit of detection

LOQ Limit of quantification

LpH Low pH soil

LSC Liquid Scintillation Counting

MpH Medium pH soil

Mw Molecular weight

N Nitrogen

NACME N-acetyl methyl esterification

NAIP N-acetyl isopropyl esterification

NH4+ Ammonium

NO3- Nitrate

PLFA Phospholipid fatty acid

SIP Stable isotopic probing

SMB Soil microbial biomass

Amino acids abbreviations Ala Alanine

Val Valine

Gly Glycine

Leu Leucine

Ile iso Leucine

Nle nor Leucine

Pro Proline

Thr Threonine

Asn Aspargine

Asp Aspartic acid

Ser Serine

Glu Glutamate

Gln Glutamine

Phe Phenylalanine

Tyr Tyrosine

Lys Lysine

Asx Aspargine + Aspartate

Glx Glutamate + Glutamine

Pro/Thr Proline + Threonin

viii

Contents

Preface ............................................................................................................................................................... i

Acknowledgements ........................................................................................................................................ ii

Summary in English ...................................................................................................................................... iii

Dansk sammendrag ........................................................................................................................................ v

Abbreviations ................................................................................................................................................ vii

1. General introduction ............................................................................................................................ 1

1.1. Studying the plant-soil nitrogen cycling ....................................................................................... 1

1.2. Aim and hypothesis ........................................................................................................................... 3

2. Method theory ........................................................................................................................................ 5

2.1. Stable isotopic probing ..................................................................................................................... 5

2.2. Bulk isotopic analysis ....................................................................................................................... 6

2.2.1. 14CO2 analysis by Liquid Scintillation Counter ......................................................................... 6

2.2.2. 13C and 15N analysis by Flash Elemental Analyzer Isotopic ratio mass spectrometer ...... 7

2.3. Compound specific isotopic analysis ............................................................................................. 8

2.3.1. Amino acid analysis ....................................................................................................................... 8

2.3.2. Phospholipid fatty acid (PLFA) analysis ................................................................................... 9

3. Experimental work .............................................................................................................................. 10

3.1. Developing the CSIA for amino acid (paper 1) ........................................................................... 10

3.1.1. Tuning the GC-C-IRMS analysis ................................................................................................. 11

3.1.2. Key findings .................................................................................................................................. 13

3.2. Turnover of higher Mw organic nitrogen (paper 2) ................................................................. 15

3.2.1. Incubation experiment ............................................................................................................... 16

3.2.2. Three extra sized fractions ........................................................................................................ 19

3.2.3. The amino acid CSIA of the soil samples ................................................................................. 22

3.2.4. Evaluating the amino acid CSIA results ................................................................................... 22

3.2.5. Confirming the results using different LTE soils ................................................................... 23

3.2.6. The PLFA CSIA of the soil samples ............................................................................................ 25

3.3. Plant N uptake from higher Mw DON (paper 3) ........................................................................ 27

3.3.1. Key findings .................................................................................................................................. 31

4. General discussion .............................................................................................................................. 34

4.1. The analytical method for amino acid CSIA ................................................................................ 34

4.2. Mineralization and sorption of organic N in soil with or without plants ............................. 35

4.3. Organic nitrogen in soil with and without plants ..................................................................... 36

ix

4.4. Plant N uptake .................................................................................................................................. 37

5. Conclusion ............................................................................................................................................. 38

6. References ............................................................................................................................................. 40

7. Appendices ............................................................................................................................................ 44

1

1. General introduction

1.1. Studying the plant-soil nitrogen cycling

The traditional understanding of the nitrogen (N) cycling in the plant-soil system is that the soil

microbiota has to fully decompose organic bound N to inorganic N in order to make N available for

plant uptake. This view is dating back to Liebig (1842), and was supported by the recognition that

microbial mediated decomposition of organic N resulted in ammonium (NH4+) as an end product

(Waksman, 1932). Hence mineralization of organic N became a central element in the perception of

the N cycle in the plant-soil system. This, and further observations described in a reviewing paper by

Schimel and Bennett (2004), highlighted two core assumptions in relation to studying plant-soil N

cycling: 1) plants only use IN and 2) plants are poor competitors for available soil N relative to

microbes. In the 1980’s and 1990’s studies began to find evidence that plants can use not only

inorganic N, but (among others) also amino acids as N sources at least in N limited ecosystems

(Nasholm et al., 1998; Jones and Kielland, 2002). In the competition for soil nitrogen between plants

and soil microorganisms, soil microorganisms have a number of advantages, such as high substrate

affinities, high surface to volume ratio, and fast growth rates Hodge et al. (2000). Yet, it is most often

found that plants are able to compete successfully for uptake of N. It has been speculated that this may

be because of the cooperation between mycorrhizal fungi and the roots. However, in their seminal

paper Hodge et al. (2000) states in their concluding remarks that: “For most plant species, both the

direct uptake of simple organic compounds and arbuscular mycorrhizal assistance appear to be

unimportant in N capture.” Instead, they suggest that the reason why plants eventually capture most N

is because of their longer life span than soil microorganisms. However, for the last few decades several

studies have documented the ability of plants, including major crops (Nasholm et al., 2001), to directly

utilize lower molecular weight (Mw) dissolved organic nitrogen (DON) in the form of amino acids and

small peptides (Owen and Jones, 2001; Jones and Murphy, 2007; Ge et al., 2009). This calls for further

studies investigating the importance of organic N uptake in plants, figure 1.

Soil organic N exists predominantly as proteinaceous material, about 40%, (Jan et al., 2009). Proteins

are linear polymers build of monomer units of amino acids, which are linked end to end in peptide

bonds, thereby forming polypeptide chains. Most natural proteins consist of between 50 and 2000

amino acids, typically corresponding to Mw’s between 4 and 544 kDa (Berg et al., 2006). Warren

(2014) illustrated the distribution of DON among size classes and the distribution of DON monomers

among main classes with protein amino acids being a key component of dissolved monomers. It was

pointed out that we have a good understanding of the lower Mw sized class (< 1 kDa), but a poor

understanding (less than five studies) of higher Mw sized classes (> 100 kDa). We know that higher

2

Figure 1. Conceptual figure showing the routes of N flow from higher molecular weight (Mw) dissolved organic

N (DON) into bio-available N. Initially, the large organic N needs to be depolymerized to lower Mw DON, which

either can be directly taken up by plants or be mineralized to inorganic N forms. The lower box shows that we

presently lack knowledge of the proportion of total N uptake occurring in organic form.

Mw organic N constitute a major part of DON (Jones et al., 2012; Warren, 2014), where bound amino

acids are an important component (Jamtgard et al., 2010). Higher Mw DON needs to undergo

depolymerization (Schimel and Bennett, 2004), the process of converting polymers, such as protein

and peptides, into monomers, such as amino acids, in order to make the DON plant available. The

proteolytic activity for large organic N depolymerization is known to be affected by among other soil

pH, active microbial communities, and presence of plants (Godlewski and Adamczyk, 2007;

Sinsabaugh et al., 2008; Vranova et al., 2013). DON may also undergo mineralization, the process of

converting organic N, both polymers and monomers, into CO2 and inorganic N (NH4+ and NO3

-). The

turnover of free amino acids in soil solution ranges from 1 to 12 hours (Jones et al., 2005), whereas the

mineralization of a protein solution, containing compounds with Mw of 65, 75 and 120 kDa, was

approximately 20 fold slower than the mineralization rate of amino acids (Jan et al., 2009). Thus the

present understanding is that depolymerization of higher Mw organic N (Mw > 1 kDa) to lower Mw

organic N (< 1 kDa) is the bottleneck in soil N cycling (Schimel and Bennett, 2004). In the soil, the

depolymerization is mediated by the release of extracellular enzymes from the soil microbial biomass

(SMB) (Burns et al., 2013), and plants (Godlewski and Adamczyk, 2007). The cleavage of the higher

Mw organic N release lower Mw organic N such as amino acids, and short peptides, which to a large

extent are bioavailable and can be used directly by plants and microorganisms (Figure 2).

The interaction between plants, soil and SMB, is a complex system of release and uptake of nutrients.

Each of the nutrient pools can be measured as a concentration, as illustrated by Warren (2014), but

the concentrations itself cannot give us the full story of what is going on in the soil, during

HigherMw DON

LowerMw DON

Plant organic N uptake

Depolymerization

CO2 [g]

Inorganic NNH4

+, NO3-Mineralization

Bio-available N

Plant inorganic N uptake

Organic N

uptake

Inorganic N

uptake

??

3

Figure 2: Conceptual figure showing the overall plant-soil N cycle with the present study focusing on the organic

side to the left where the retention of higher Mw DON (>1 kDa) and depolymerization of this DON to lower Mw

DON (<1 kDa) is investigated.

depolymerization and mineralization of organic matter. The rate at which higher Mw organic N is

turned over in the soil plant system is needed in order to determine the bottleneck (if any) in the

transformation of higher Mw organic N into bioavailable lower Mw DON. When investigating the

turnover of organic N in soil, it is beneficial to use amino acids as the target compounds group due to

the soils high content of proteinaceous material. The best way to follow the turnover of protein is

through stable isotopic probing (SIP) (Dumont and Murrell, 2005). When using SIP a stable isotope is

introduced, in excess amounts, to the system. Typical isotopes used are 13C, 14C and 15N. The movement

of the isotopes can then be followed through the processes by either bulk isotopic analysis or

compound specific isotopic analysis (CSIA). There is a need for developing a suitable amino acid CSIA

protocol, to tackle the challenges associated with hydrolysis and purification of plant and soil samples.

Importantly is also the derivatization, which can influence which amino acids can be analyzed.

1.2. Aim and hypothesis

The overall aim of my PhD project was to investigate the fate of higher Mw DON when it enters a soil

with and without plants. The first objective was to set up a compound specific isotope analysis (CSIA)

to characterize and quantify amino acids in double labeled (15N and 13C) high Mw DON derived from

Soil (micro)

organisms

Soilorganic N

HigherMw DON

LowerMw DON

NO3-

NH4+

Pla

nt N

N2 [g]

N2O [g]

N2O [g]

Pla

nt N

Mobilizing

Immobilizing

Plant flows

Exo-enzymatic

Gasseous emissions

CO2 [g]

4

white clover. The second objective was to investigate the turnover of higher Mw DON in soil with

different management histories. The third objective was to investigate the turnover of higher Mw DON

in a soil with plant growth to determine the influences of the competition between plants and SMB on

the turnover of higher Mw DON and the uptake of N by plants.

The corresponding hypothesis are

1. In cultivated soil the pool of Higher Mw DON represents the bottleneck in the production of plant

available N from soil organic N (Figure 3) (Schimel and Bennett, 2004).

2. Differences in soil pH will affect which microbial communities dominate the decomposition of

DON; at low pH fungi is expected to dominate and at increasing pH the dominating microbial

communities will shift towards bacteria (Rousk and Baath, 2011).

3. The competition between plants and the microbial communities for plant available DON will

increase the turnover of high Mw DON compared to a soil without plants (Godlewski and

Adamczyk, 2007).

4. At low soil pH, organic N turnover is expected to be slower and hence there will be a greater

chance of direct organic N plant uptake as indicated by 13C presence in roots, whereas at higher pH

mineralization will be greater and so will dissolved inorganic 15N (DI15N).

Figure 3: Schematic representation of the amino acid-based constitutes of higher Mw DON, showing the relation

to Mw sizes of free amino acids and amino acids bound in peptides and proteins.

5

2. Method theory

In the following sections, an introduction to the method theory used during the experimental work of

this PhD project is presented. Each section will contain a description on the given analysis technique

and critical reflections on their limitations.

2.1. Stable isotopic probing

We wanted to grasp a picture of the plant derived DON turnover in the plant soil system.

Stable isotopic probing (SIP) techniques were applied in studies of white clover derived dissolved

organic nitrogen (DON) produced from screw pressing triple labeled (13C, 14C, 15N) white clover into

juice. The production of the triple labeled white clover juice is fully described in paper 1 (Enggrob et

al., 2019).

We firstly needed to enrich white clover plants with 13C, 14C and 15N and there are several techniques

that can be used to induce labeling into plants (Wichern et al., 2008), some of which are illustrated in

Figure 4. Atmospheric labeling can be used when it is possible to contain the air around the plants to

be enriched, such as in laboratory or pot experiments, and when the enriched compounds can be made

airborne and available for plant uptake, such as 13CO2 and 14CO2. Soil labeling is another option, in

which the enriched compound is either mixed with the soil before planting or seeding, or the enriched

compound is dissolved and added during irrigation. Soil labeling can be used both in field and in pot

experiments. Any enriched compounds, expected to be available for plant uptake, can be used in soil-

labeling experiments.

Figure 4: Introduction of labeling into white clover occurred for C-tracers via CO2-labeling, and for the N-tracer

via soil N labeling.

6

For the labeling of white clover, we wanted the enrichment to enter the plant as naturally as possible

to ensure a natural distribution throughout the plant. The best way to do that is by continuous labeling

throughout the growth period. I therefore used repeatable atmospheric labeling for the enrichment in

13C and 14C, as 13CO2 and 14CO2 and soil labeling by irrigation with each irrigation to create the

enrichment in 15N with a 3 at% 15N-(NH4)2SO4 solution; details of the labelling procedure are described

in paper 1 (Enggrob et al., 2019). Briefly, from day one, the water used for irrigation contained 3 at%

15N-PK fertilizer. From week 8 and onward, the 13/14CO2 was introduced to the white clover as descried

by (Rasmussen et al., 2008). Within each pot of clover, a beaker containing 5 ml of a saturated solution

of 13C and 14C labeled sodium bicarbonate dissolved in 1M NaOH was placed. The pot was then covered

by a transparent plastic bag and the 13CO2 and 14CO2 was made air born by the addition of 5 ml 2M HCl

to the beaker. After 2 h, the labeling was stopped by removing the plastic bag and discarding the

beaker.

2.2. Bulk isotopic analysis

Bulk isotopic analysis are good to give an overview of the total amount of the isotope in question, in

the particular sample. It is a fast and efficient way to follow the fluxes and pools of enriched

compounds. The analysis itself is not time consuming and it is therefore possible to make time series.

But it also have its limitation, for instance, when analyzing 15N it is not possible to get any information

on the distribution in inorganic an organic compounds without pre-treatment. Neither is it possible to

get any information on whether the compounds of interest are in their original form or to what extent

the compounds are turned over.

When studying a complex system such as the plant soil N system, in which a certain amount of DON is

added to the soil, and the movement of DON through the system over time is what we want to

highlight, then it is not enough just to take a soil sample at the end of the experiment. A time series is

needed, as is a division of the system into sample type such as gas, soil solution and soil samples. Then

it is possible to trace some of the routes the DON goes through in the soil system.

In this study, we used two types of bulk isotopic analysis (1) Liquid Scintillation Counting (LSC) for the

analysis of 14CO2, and (2) Flash Elemental analysis Isotope Ratio Mass Spectrometry for the analysis of

13C and 15N in both soil solution and soil.

2.2.1. 14CO2 analysis by Liquid Scintillation Counter

The mineralization of organic compounds is analyzed as the production of 14CO2 after the addition of a

given 14C-organic compound to soil (Jones, 1999; Owen and Jones, 2001; Jones and Kielland, 2002;

Kemmitt et al., 2008). The microbial mineralization of any organic compound results in the production

of CO2. The production of CO2 can therefore be seen as an indicator of the activity in that given soil. The

7

CO2 production can follow two principal time courses, i.e., with or without a lag phase in the beginning

before the mineralization takes off. A lag phase indicates that, before the microorganisms can utilize

the organic compounds, there needs to be a proliferation (growth) of an initially small population or

the necessary enzymes have to be induced in an existing larger population. No lag phase indicates that

the SMB is immediately able to utilize the organic compound.

In my study, the mineralization of dissolved organic compounds to 14CO2 follows a first order kinetic

decay model (Boddy et al., 2007):

Equation 1: 𝑆 = 𝑆𝑈 + [𝑎 × 𝑒𝑥𝑝(−𝑘 × 𝑡)]

Where S is the 14C label remaining in the soil, SU is the amount of unrecovered 14CO2, k is the

exponential coefficient, the production, a is the respiration constants for the given system and t is time.

This function makes it possible to calculate a half-life time (𝑡½) for the given pool of dissolved organic

compounds:

Equation 2: 𝑡½ =𝑙𝑛(2)

𝑘

It has, however, been suggested by (Boddy et al., 2007; Boddy et al., 2008) that the mineralization

should be described by a double first order kinetic decay model, arguing that the first part describing

the turnover of the added compound to be investigated, and that the second part refers to the turnover

of storage or anabolic procuct from microbial carbon uptake the SMB itself:

Equation 3: 𝑆 = 𝑆𝑈 + [𝑎1 × 𝑒𝑥𝑝(−𝑘1𝑡)] + [𝑎2 × 𝑒𝑥𝑝(−𝑘2𝑡)]

Where a1 is the respiration constant and a2 is the immobilization constant in the SMB (Farrell et al.,

2011).

To analyze the produced 14CO2, it first have to be collected in a liquid form, this is done by introducing

a base trap, containing NaOH, to the system. Once the 14CO2 is trapped, a scintillation cocktail is added

to the system. The Liquid Scintillation Counter (LSC) works by the 14C in the sample sending out

radioactive β particles, which are picked up by the Scintillation cocktail and transformed into a flash of

light, which then are picked up by a counter.

2.2.2. 13C and 15N analysis by Flash Elemental Analyzer

Isotopic ratio mass spectrometer

An Elemental Analyzer can analyze a variety of solid or liquid sample types, and it gives the elements

or isotopic composition of the given sample. Depending on the detector used, it can give either a

qualitative or a quantitative view of the elements or isotopic ratio in the sample. As a detector, we used

both a thermal conductivity detector (TCD) and an isotopic ratio mass spectrometer (IRMS), enabling

us to get both the total amount of N and C, and the delta values δ13C and δ15N. From these data we were

able to calculating the amount of 13C and 15N in the sample.

8

2.3. Compound specific isotopic analysis

The ability of tracking a single compound or compound group through the cycling of organic N in the

plant soil N cycle is a valuable technique when studying the uptake and turnover of specific

compounds. CSIA provides information of both the concentration and the enrichment of the compound

of interest. This makes it possible to follow the changes both in concentration and in enrichment,

whether the compound is turned over or taken up. When doing CSIA, firstly the compound group of

interest must be isolated, doing a purification, secondly the group of compounds must be separated

into individual compounds by chromatography, either liquid chromatography (LC) or gas

chromatography (GC), and thirdly the compound and the isotopes are detected by either a time of

flight mass spectrometer (TOF-MS) or an isotopic ratio mass spectrometer (IRMS). In this project we

analyzed two compound groups, amino acids and phospholipid fatty acid (PLFA) using gas

chromatography combustion isotopic ratio mass spectrometry (GC-C-IRMS).

The GC analysis implies that the compound of interest must be gaseous. For the analysis of amino

acids, this is ensured by derivating the amino acids (see section 3.1) before the analysis and by

adjusting the temperature of the inlet to the GC column. The separation on the GC column is very

important in the GC-C-IRMS analysis because the combustion oven oxidize everything into CO2 and N2.

The separation can be controlled by selecting the right column and adjusting a temperature gradient

over the time of the separation of the compounds of interest. Hereafter the ratio of the 13C/12C and the

ratio of 14N/15N are detected by the IRMS. To be able to detect the N2 resulting from the combustion of

the compound in question, the CO2 must first be removed by leading the gas through a liquid nitrogen

freeze trap, thereby freezing the CO2 solid, and allowing the N2 to be detected. Due to the complete

oxidation of the compounds in the combustion oven, standards for each compound of interest are

necessary for the identification and concentration calculation. Also, to monitor the efficiency of the

sample treatment and the analysis, an internal standard, not representing one of the compounds of

interest, must also be added.

2.3.1. Amino acid analysis

As previously stated, is 40% of soil organic nitrogen bound in protein (Jan et al., 2009), and all proteins

are build from a repertoire of 20 different amino acids. Proteins are too complex to analyze directly,

but rather they are hydrolyzed, thereby breaking the peptide bond, separating them into the building

blocks (i.e., amino acids). Amino acids are relative simple compounds that consist of a central C atom

linked to an amino group, a carboxylic acid group, a hydrogen and a distinctive side chain, which

determine the function of the amino acids (Berg et al., 2006). It therefore stand to reason that to follow

the turnover of higher Mw soil organic N the target compound is amino acid. Therefore, the first

9

objective of my PhD project was to develop a CSIA for the analysis of amino acids bound in plant and

soil samples (see section 3.1 and paper 1 for further details).

2.3.2. Phospholipid fatty acid (PLFA) analysis

An important building block of all cell membranes is phospholipid fatty acid (PLFA). PLFA are the

primary lipids of cellular membranes, consists of hydrophilic head and a hydrophobic tail (Berg et al.,

2006). PLFAs are widely being used as biomarkers for different microbial groups (Frostegard et al.,

1993; Fierer et al., 2003; Stromberger et al., 2012), but one have to be aware that some of the same

biomarkers can be an indicator for different effects (Frostegard et al., 2011). In my PhD project, I was

looking for biomarkers for gram-positive bacteria, gram- negative bacteria and fungi. The soil

microorganisms are crucial for the soil function, however only the active microorganisms are involved

in the ongoing processes (Blagodatskaya and Kuzyakov, 2013). By combining SIP with then PLFA

methods, it ensured the measurement of the activity of the target microbial groups (Knief et al., 2003;

Boschker et al., 2014; Kusliene et al., 2014). The extraction and analysis of the PLFAs of the soil

samples were done as described by Petersen et al. (2002).

10

3. Experimental work

In the following section an experimental overview is given followed by presentation and discussion of key

results.

In order to investigate the fate of higher Mw DON in soils with different management histories, a series

of experiments were carried out. White clover was grown in pots and triple labeled by soil labeling

(15N) and atmospheric labeling (13C, 14C). After harvest, the white clover was juiced by screw pressing,

and the juice was fractionated into Mw sized fractions. Fractions of < 1 kDa, 1-10 kDa, 10-100 kDa, and

> 100 kDa were used for the development of the CSIA method. Fractions of 1-10 kDa, > 10 kDa, 10-30

kDa, 30-100 kDa and >100 kDa were used in soil incubation experiments without plants, whereas only

the >100 kDa sized fraction was used for the experiment with plants.

The experimental work represented three lines of experiments, one for each objectives, eventually

resulting out in three publications.

During the first line of experiments (paper 1), white clover were grown in pots, simultaneously

enriched in 13C, 14C and 15N, as descried in section 2.1. After harvest, both the shoots (including the

stolen) and the roots were screw pressed into shoot and root juice and subsequently fractionated into

four Mw size classes. A CSIA method was developed to analyze the DON for the content and

distribution of amino acids along with the atomic fraction of 13C and 15N. Details of the experimental

work, not described in the paper 1, are detailed in the following sections (3.1).

During the second line of experiments, Mw sized fractions of DON solutions were incubated with soils

of different management histories, in order to investigate the influences of different soil and microbial

properties on the turnover of DON. Different Mw sized fractions of DON solutions were used to

investigate the bottleneck of the organic nitrogen turnover described by Jan et al. (2009), and

illustrated in Figure 3. Doing the experiment, the first results, led us to test not only two Mw sized

fractions, 1-10 kDa and > 10 kDa, but a total of five Mw sized fractions: 1-10 kDa, >10 kDa, 10-30 kDa,

30-100 kDa and >100 kDa.

During the third line of experiments the highest Mw DON fraction (>100 kDa) was incubated with soil

in which maize were growing, i.e., to investigate how the competition between plants and the SMB will

affect the turnover of higher Mw DON and the plant uptake of ON.

3.1. Developing the CSIA for amino acid (paper 1)

Amino acids, as mentioned in section 2.3.1, consists of a central C atom linked to an amino group, a

carboxylic acid group, a hydrogen and a distinctive side chain. The most common method used for the

analysis of amino acids is GC-C-IRMS (Fountoulakis and Lahm, 1998; Corr et al., 2007; Larsen et al.,

2013; Yarnes and Herszage, 2017). Before the amino acids can be analyzed by gas chromatography

11

they have to be made more volatile, so they become airborne and available for gas separation. The

transformation to a more volatile compound is done by the addition of a secondary functional group to

the amino group and the carboxylic acid group of the amino acid. This process is called derivatization.

To ensure a proper separation on the gas chromatograph, a proper column and temperature gradient

must be adjusted. Finally, analyzing proteinaceous material in a natural sample requires hydrolysis to

release the amino acids from the peptide bonds, and purification of the sample to eliminate

contaminators. The choice of hydrolysis method and the purification of the natural samples are fully

described in paper 1 (Enggrob et al., 2019).

3.1.1. Tuning the GC-C-IRMS analysis

The goal was to achieve the best possible separation of multiple amino acids in the shortest possible

time. Several parameters can influence the efficiency of the gas chromatograph performances, most of

which are controlled by the GC software (Isodat 3.0). Two important parameters are always adjusted

to fit the particular analysis: the column and the temperature gradient controlling the temperature of

the column. Based on the literature (Corr et al., 2007), the VF‐23m capillary column (60 m× 0.25 mm

i.d. × 0.25 μm film thickness; AgilentTechnologies, Amstelveen, The Netherlands) was chosen. Corr et

al. (2007) also inspired the initial temperature gradient which starts at a temperature of 40°C; then

the temperature was first raised to 120°C over 4 min, secondly to 190°C over 23 min, and finally to

250°C over 12 min and held for 20 min.

Amino acid standards, both single standards and mixed standards containing 21 amino acids, were

used to test the temperature gradient.

Two derivatization methods, based on (Corr et al., 2007; Larsen et al., 2013), were evaluated for the

analysis of amino acids. These were N-acetyl methyl esterification (NACME) and N-acetyl isopropyl

esterification (NAIP) as depicted in Figure 5. The derivatization procedures are described in paper 1

Figure 5. Step by step structural information of the two amino acid derivatization methods, i.e. (A) the N-acetyl

methyl esterification, and (B) N-acetyl isopropyl esterification.

12

(Enggrob et al., 2019) but in brief, the two derivatization methods differ only in the use of methanol in

the NACME method and the use of isopropanol in the NAIP method.

Two series of standards, one for each derivatization method, were subject to GC-C-IRSM analysis to

help optimizing the temperature gradient to improve separation of the amino acids. The final

temperature gradient was as follows: initial temperature were set to 90°C and held for 1 min, secondly

the temperature were raised to 120°C over 2 min, thirdly the temperature were raised to 250°C over

43 min and held at 250°C for 45 min.

With the NACME derivatization method we were able to obtain repeatable signals for 10 out of 21

amino acids (Table 1) namely nor valine (ISTD), nor leucine (ISTD), threonine, aspartic acid, Serine,

glutamate, phenylalanine, hydroxyproline, tyrosine, and lysine; with lysine eluated as the last after

5404 s. Despite repeated attempts, we were not able to obtain repeatable stable derivatives from the

NACME derivatization of alanine, valine, glycine, leucine, iso leucine, proline, aspargine, glutamine,

methionine, cysteine or tryptophan in single or mixed standards.

Table 1: Retention times of amino acids obtained with the NACME or NAIP derivatization methods, respectively

(ISTD = internal standard).

In contrast, were we able to obtain separation and stable retention time for all 21 amino acids in single

standards with the NAIP method (Table 1). Again, lysine eluted as the last with a retention time of

5181 s. However, in mixed standards, methionine and cysteine disappeared, whereas proline +

NACME Amino

acid

Retention time individual standards

[s] NAIP

Amino acid

Retention time individual standards

[s]

Retention time mixed standards [s]

1 Ala 1232,3 1230

2 Val 1360,2 1360

1 AvlISTD 1422 and 1660 3 AvlISTD 1438,5 1444

4 Gly 1443,8 1444

5 Leu 1465,3 1464

6 Ile 1477,6 1478

2 Nle ISTD 1783 7 NleISTD 1556,2 1554

8 Pro 1874 1886

3 Thr 2069 9 Thr 1883 1886

10 Asn 1996,8 2000

4 Asp 2207 11 Asp 2001 2000

5 Ser 2215 12 Ser 2058 2051

6 Glu 2451 13 Glu 2197 2190

14 Gln 2295,9 2297

15 Met 2347,3 -

7 Phe 2512 16 Phe 2434 2436

17 Cys 2443 -2445 -

8 Hyp 2593 18 Hyp 2523,9 2523

19 Trp 2840,3 2979

9 Tyr 3429 20 Tyr 3553,6 3589

10 Lys 5405 21 Lys 5188,6 5181

13

threonine (Pro/Thr) eluted simultaneously with retention times of 1886 s and aspargine + aspartic

acid (Asx) eluted simultaneous with retention times of 2000 s.

3.1.2. Key findings

When performing acid hydrolysis on a sample it always give rise to some uncertainty in whether the

hydrolysis is complete or insufficient. We therefore tested the recovery of Bovine Serum Albumin

(BSA) from two acidic hydrolysis methods (paper 1) and found a recovery of 35.6% (± 1.3%) for

standard hydrolysis and 31.8% (± 1.5%) for fast hydrolysis (data shown as mean ± standard error).

These results were in line with previously reported recoveries of approximately 30% (Fountoulakis

and Lahm, 1998). We also found that there is a high risk of losing material, especially from the lower

Mw fraction, when performing acid hydrolysis. Figure 6 shows the measured concentration of amino

acids after hydrolysis of the < 1 kDa fraction versus the measured concentration of amino acids in the

unhydrolyzed < 1 kDa fraction (i.e., representing the free amino acids). We expected that the

concentration of all amino acids would increase after the hydrolysis, but instead the concentration of

seven of the amino acids was significantly lower after the hydrolysis. Importantly, even though the

hydrolysis affected the amino acid concentrations in the <1 kDa fraction, it did not affect the isotopic

signature (13C and 15N) of the amino acids (Figure 7). Hence, tracing the fate of labeled amino acids is

not compromised, which if further supported by the similar isotopic signature pattern of amino acids

bound in the higher Mw sized fractions (Figure 8).

Figure 6: The content of free amino acids (AA) versus bound amino acids (AA) in the Mw size fraction <1 kDa for

white clover shoot juice using the standard hydrolysis method (n = 3).

Free AA in <1kDa fraction(ng AA/g fresh material)

0 20 40 60 80 100 120 140

Bo

un

d A

A i

n <

1k

Da f

racti

on

(ng

AA

/g f

resh

mate

rial)

0

20

40

60

80

100

120

140Ala

Val

Gly

Leu

Ile

Pro/Thr

Asx

Ser

Glu

Gln

-

Phe

Tyr

Lys

1:1 line

14

Figure 7: The atomic fraction in free amino acids (AA) versus bound amino acids (AA) in the Mw size fraction <1

kDa for (A) 13C in shoot, and (B) 15N in shoot juice of experimentally labeled white clover (n = 3).

Figure 8: Example of the 13C atomic fraction of amino acids in white clover shoot juice for different Mw size

fractions: free amino acids (blue circle), amino acids bound in 1-10 kDa (yellow triangle up), amino acids bound

in 10-100 kDa (green square), and amino acids bound in >100 kDa (orange diamond). For free amino acids both

glutamate and glutaminen were measured (glutamate omitted in this figure), whereas in the hydrolyzed Mw size

fractions >1 kDa glutamate and glutaminen is reported as Glx (n=3). Asterisks indicate significant differences in

the obtained atomic fraction. Double asterisks indicate no normal distribution.

A

13C in free AA in <1kDa fraction

(13

C atomic fraction)

0.00 0.06 0.08 0.10 0.12

13C

in

bo

un

d A

A i

n <

1kD

a f

racti

on

(13C

ato

mic

fra

cti

on

)

0.00

0.06

0.08

0.10

0.12 B

15N in free AA in <1kDa fraction

(15

N atomic fraction)

0.000 0.012 0.014 0.016 0.018

15N

in

bo

un

d A

A i

n <

1kD

a f

racti

on

(15N

ato

mic

fra

cti

on

)

0.000

0.012

0.014

0.016

0.018

Ala

Val

Gly

Leu

Ile

Pro/Thr

Asx

Ser

Glu

Gln

Glx

Phe

Tyr

Lys

1:1 line

0.00

0.02

0.04

0.06

0.08

0.10

Ala*

Val*

Gly*

Leu

Ile**

Pro/Thr*

Asx*

Ser

Gln/Glx*

Phe*

Lys*

Tyr

Shoot free Shoot 1 -10 kDa Shoot 10-100 kDa Shoot >100 kDa

15

3.2. Turnover of higher Mw organic nitrogen (paper 2)

Based on the mineralization of 14C-labelled proteins to 14CO2 (Jan et al., 2009) identified protein

depolymerization as the bottleneck in the plant-soil N cycle. This is in agreement with other studies

identifying depolymerization as the rate limiting step in comparison to the more rapid mineralization

of organic N monomers into inorganic N (Schimel and Bennett, 2004). To test the importance of

organic N molecular size an experiment was set up to follow the turnover of higher Mw (1-10 kDa and

>10 kDa) DON in soils with three pH levels. The hypotheses were that 1) soils differing in pH will differ

in which microbial communities dominate the decomposition of organic matter, and thereby differ in

turnover of higher Mw DON, and 2) part of the added higher Mw DON will be retained in the original

form and the peptide-sized DON will be more strongly retained than protein-sized molecules (Knicker,

2011).

Soils were samped from the Danish Jyndevad long-term field experiments (LTE) on liming and

phosphorus initiated in 1942 (Rubaek, 2008) on a coarse sandy soil (Table 2) used for cultivation of

spring barley. The Jyndevad LTE soil was sampled in August 2015 from the plough layer (5-20 cm) of

the V1 field in the treatments receiving 0, 4, or 12 Mg lime ha-1 (every 6-9 years) and 15.6 kg P ha-1

year-1. At the time of soil sampling contrasting pHCaCl2 levels of 3.6 (low pH, LpH), 5.4 (medium pH,

MpH), and 7.1 (high pH, HpH) were established in the three treatments. Soil was sieved (4 mm) to

remove visible roots and stored at 2°C until the experiment.

Table 2. Basic properties of soils from the Jyndevad LTE on liming and phosphorus fertilization initiated in 1942 (Rubaek, 2008). The experiment is located at Jyndevad Experimental Station, Southern Jutland, Denmark (54o53′N, 09o07′E). The soil is classified as an Orthic Haplohumod (Soil Survey Staff, 1999).

Name Liming pH1 C N Clay Silt Fine sand Coarse sand t ha-1 g/kg soil ------------------ g/kg soil --------------

Low pH 0 3.6 11.7 0.6 Medium pH 4 5.4 10.5 0.6 40 40 170 750 High pH 12 7.1 13.4 0.8

1 pH measured in 0.01 M CaCl2 in a 1:2.5 soil:solution ratio.

The DON solutions used for the first incubation series, were produced as described in paper 1

(Enggrob et al., 2019). Two sized fractions were used: (i) Mw 1-10 kDa, with 14C activity of 8.65 Bq ml-1

and (ii) Mw > 10 kDa (pooled from Mw 10-100 kDa and Mw >100 kDa fractions) with 14C activity of

9.44 Bq ml-1.

16

3.2.1. Incubation experiment

The micro lysimeters were constructed from the insert to a 50 mL centrifugal filter tube (Macrosep®

Advance, Pall Corporation, Ann Arbor, MI, USA) as described in paper 2 (Figure 9). The micro-

lysimeters were added 12 g of field moist soil, which was gently packed by tapping on the insert unit.

The incubation chambers were constructed from 1 L glass jars, where the micro-lysimeter was placed

together with a base trap containing 1 ml NaOH (1 M), for trapping any produced CO2, and a beaker

containing 2 ml water to avoid soil drying.

Figure 9:. Micro-lysimeter setup with the soil packed in an insert unit fitting 50 ml centrifugal tubes, which

allows rapid sampling of soil solution via centrifugation. Micro-lysimeters were constructed using the insert unit

from the 50 ml Macrosep® centrifugal tubes (Pall Coorporation, Ann Arbor, MI, USA) after removal of the

vertical filter-piece. Constructing micro-lysimeters in the insert-unit allowed rapid sampling of soil solution via

centrifugation and the use of a soil quantity great enough to conduct multiple analyses of both soil and soil

solution after treatments with triple-labeled DON. The micro-lysimeters were packed from below of a glass

microfiber filter (Whatman GF/A filter, 25 mm, GE Healthcare Life Sciences), a piece of silk organza cloth, and

another glass microfiber filter. On top, 7 g of purified sea sand (0.1 - 0.315 mm, analytical grade, Merck KGaA,

Darmstadt, Germany) was packed by adding 5 ml of water followed by centrifugation for 5 minutes at 5000g.

The micro-lysimeters were added 12 g of field moist soil, which was gently packed by tapping on the insert unit.

The incubation started with the addition of 2.0 ml DON solution to the micro-lysimeter. As a control

treatment, 2.0 ml water was added instead of the DON solution.

The micro-lysimeters were incubated at room temperature (22°C). Four series of micro-lysimeter

incubations were prepared aiming at final destructive sampling after 1 hour, 1 day, 7 days and 14

days, respectively. The base trap to collect 14CO2 was sampled after 1 hour and 1 day for the 1 hour and

1 day treatments, respectively, and at days 1, 4 and 7 for the 7 days treatment and day 1, 4, 7, and 14

after start of incubation for the 14 days treatment. The resulting mineralization curves for the 14 days

incubation are shown in Figure 10.

Upon termination after 1 hour, 1, 7, and 14 days, the micro-lysimeters were first added 8 ml of water

and centrifuged for 5 minutes at 5000 g, followed by addition of 10 ml of water with repeated

Insert unit setup:

- 25 mm soil layer (12 g fresh weight)- 10 mm sea sand layer (7 g dry weight)- GF/A filter- Disk of silk organza cloth- GF/A filter

17

centrifugation, where after the two solutions were pooled to give one sample of 20 ml of soil solution

washed with water. Then, 10 ml of 1 M KCl was added with subsequent centrifugation, and the

addition of 10 ml KCl and centrifugation was repeated to give a 20 ml pooled sample of soil solution

washed with KCl. The soils were hereafter removed from the micro-lycimeter and frozen before

further analysis. The water and KCl soil solutions were immediately filtrated using 0.45 µm Macrosep

centrifugal tubes (Pall Corporation, New York, USA) with centrifugation for 5 minutes at 5.000 g,

where after two times 250 µl was taken from each sample; one was directly added 4 ml scintillation

cocktail (OptiPhase HiSafe3, PerkinElmer, Waltham, MA, USA), the other was added 250 µl 1 M

HCl and left for 1 hour to allow any dissolved CO2 to escape before addition of 4 ml scintillation

cocktail. The LSC analysis of the KCl soil solution showed no 14C activity and are therefore not shown,

and have not undergoing any other analysis. The water soil solution, from here referred to as the soil

solution, were freeze-dried, dissolved in 500 µl milliQ water (Synergy® System, Millipore,

Molsheim, France) and transferred into individual tin capsuls and analyzed by Flash EA. The results

are shown in Figure 11.

Figure 10: Mineralization of higher Mw labeled organic N to 14CO2 in Jyndevad soils at three pH CaCl2 levels: low

at pH 3.6, medium at pH 5.4, and high at pH 7.1. (A) the 1-10 kDa organic N fraction, and (B) the >10 kDa fraction.

There were no statistical differences among soil pH levels in accumulated 14CO2 after 14 days as shown by ‘ns’

next to the curves (n = 4).

A

0

10

20

30

40

50

ns

B

Time (days)

0 2 4 6 8 10 12 14

Min

era

liza

tio

n t

o C

O2

(ac

cu

mu

late

d 1

4C

O2 o

f a

dd

ed

14C

, %

)

0

10

20

30

40

50

Low pH (3.6)

Medium pH (5.4)

High pH (7.1)

ns

18

The mineralization curves showed a total mineralization of 42-45% and 33-36% for the Mw 1-10 kDa

and the Mw >10 kDa, respectively, but no differences between the soil pH levels. The differences in the

mineralization between the two DON fractions were smaller than we had expected based on the

mineralization curves of protein solutions showed by Jan et al. (2009).

Figure 11: The temporal changes in soil solution 13C and 15N content (in % of added) for (A) the 1-10 kDa, and

(B) the >10 kDa organic N fraction (n = 4).

The Flash EA data from the soil solution revealed an interesting pattern for the relation between 13C

and 15N over time (Figure 11). The percent of added 13C and 15N remaining in soil solution, i.e. still

dissolved, after 1 hour and 1 day, respectively, showed a parallel loss of both 13C and 15N from the soil

solution over time. This correlate with the measured mineralization for both DON fractions. At day 7

and 14, an increase in 15N was observed whereas 13C was only present at low levels in the soil solution.

The interpretation of the data led to the design of a conceptual figure (Figure 12). In this figure, the

A

0 10 20 30 40 50

0

10

20

30

40

50

B

N from organic N in soil solution (% 15

N of added)

0 10 20 30 40 50

C f

rom

org

an

ic N

in

so

il s

olu

tio

n (

%1

3C

of

ad

de

d)

0

10

20

30

40

50 LpH 1 hour

LpH 1 day

LpH 7 days

LpH 14 days

MpH 1 hour

MpH 1 day

MpH 7 days

MpH 14 days

HpH 1 hour

HpH 1 day

HpH 7 days

HpH 14 days

19

parallel loss of 13C and 15N in percent of added from 1 hour to 1 day illustrate the dissipation of the

added dissolved compounds whereas the increase of 15N in percent of added indicate mineralization of

the added compound to inorganic N. These observations were supported by the data from the flash EA

analysis of the soil (data not shown).

Figure 12: Conceptual figure of development in soil solution 13C and 15N content, where first after 1 hour the %

of added remaining in soil solution is the compounds not-sorbed (i.e. still dissolved), second the parallel loss of 13C and 15N show dissipation of the added compounds (still dissolved or in equilibrium with the soil solution),

thirdly the loss of 13C with 15N still present (or even increasing) indicates mineralization of the added compound

to inorganic N.

3.2.2. Three extra sized fractions

Based on the relatively high 14CO2 respiration from the >10 kDa fraction, we wondered if the

bottleneck in organic N mineralization would lie somewhere within this Mw sized fraction. Therefore

the >10 kDa Mw sized fraction was further fractionated into three pools, Mw 10-30 kDa, Mw 30-100

kDa and Mw >100 kDa, and the incubation experiment was repeated for the three different soil pH

levels for two times: 1 hour and 14 d. The incubation and sampling of the Mw sized fractionated DON

(10-30 kDa, 30-100 kDa and >100 kDa) were performed as explained above.

From the 1 hour time series the soil and soil solution were analyzed by Flash EA to investigate the

immediate sorption of the DON solutions to the soil. From the 14 days time series, the base trap to

collect 14CO2 was changed after day 1, 4, 7 and 14, and a mineralization curve for each DON solution

were calculated (Figure 13). From the 14 days incubation experiment, both the soil and soil solution

20

were analyzed by Flash EA as described above. The total 13C and 15N from the Flash EA analysis of the

soil, calculated as percent of added 13C or 15N are listed in Table 3.

Figure 13: Mineralization of higher Mw labeled organic N to 14CO2 after 14 days in Jyndevad soils at three pH

CaCl2 levels: low at pH 3.6, medium at pH 5.4, and high at pH 7.1 of (A) the Mw 10-30 kDa, (B) the Mw 30-100

kDa, and (C) the Mw >100 kDa organic N fractions. Statistical differences are marked by different letter next to

the curves (n = 4).

The relation between the accumulated amount of respired 14CO2, calculated as percent of added, and

the amount of 13C and 15N, also calculated as percent of added, of the Flash EA of the soil solution from

the 1 hour time series, were used to illustrate the negative correlation between the immediate

sorption and the mineralization of each Mw sized fractions. Both the mineralization curve and the

immediate sorption are described in paper 2 and the immediate sorption is displayed in Figure 14.

A

0

10

20

30

40

50

B

Min

era

lizati

on

to

CO

2 (

accu

mu

late

d 1

4C

O2 o

f ad

ded

14C

, %

)

0

10

20

30

40

50

Low pH (3.6)

Medium pH (5.4)

High pH (7.1)

C

Time (days)

0 2 4 6 8 10 12 14

0

10

20

30

40

50

b

a

a

ns

bab

a

21

Table 3. Recovery (% of added) of 13C and 15N in Jyndevad soils after 14 days of incubation. Data is given as

mean ± standard error (n = 4). Statistical differences among organic N fraction within each soil is show with

different letter; no significant differences were found across soils within each organic N fraction.

Fraction 1-10 kDa 10-30 kDa 30-100 kDa >100 kDa

-------------------- 13C recovery (% of added) --------------------

Low pH 25.3 ± 0.8 A 28.6 ± 1.0 B 31.0 ± 1.7 B 44.1 ± 3.1 B Medium pH 33.3 ± 1.0 A 31.7 ± 0.7 A 26.9 ± 1.3 A 44.8 ± 1.6 A High pH 36.9 ± 1.1 A 34.5 ± 1.3 A 33.4 ± 1.1 A 43.6 ± 1.4 B

-------------------- 15N recovery (% of added) ---------------------

Low pH 19.9 ± 0.7 A 18.1 ± 0.6 A 23.5 ± 1.0 B 23.8 ± 2.1 B Medium pH 28.5 ± 0.3 A 23.7 ± 0.5 B 23.0 ± 0.8 C 28.3 ± 1.2 D

High pH 34.2 ± 0.9 A 28.4 ± 1.5 B 32.5 ± 1.8 C 29.6 ± 0.8 A

Figure 14. Correlation between organic N sorption after 1 hour and accumulated 14CO2 after 14 days for (A)

sorption of 13C in the added organic N fractions and (B) sorption of 15N in the added organic N fractions (n = 4).

As described in paper 2, the release of 14CO2 follow a first order kinetic decay model, and statistical

analyses showed that there for some Mw sized fractions was an effect of soil pH. But what especially

comes to mind when looking at the mineralization curve is the clear reduction in accumulated

mineralization from Mw sized fraction 10-30 kDa to the Mw sized 30-100 kDa, indicating that the

bottleneck lies within these two fractions depending on soil pH level. From the data displayed in Table

3 it was concluded that a significant amount of both 13C and 15N remains in the soil after washing with

water, but it is unknown whether the DON are retained in its original form due to sorption to the soil

surface or has been transformed during uptake in the SMB.

A

Sorption of organic N

(13

C retained after 1 hour, %)

0 20 40 60 80 100

Min

era

lizati

on

of

org

an

ic N

to

CO

2

(accu

mu

late

d 1

4C

O2 a

fter

14 d

ays,

%)

0

10

20

30

40

50B

Sorption of organic N

(15

N retained after 1 hour, %)

0 20 40 60 80 100

Low pH

Medium pH

High pH

1-10 kDa

10-30 kDa

30-100 kDa

>100 kDa

22

To investigate to what extent the DON was retained in the soil as the original compounds, the soils

incubated with both the Mw sized 1-10 kDa well below the bottleneck and the Mw sized >100 kDa well

above the bottleneck were selected for further analysis by amino acid CSIA.

3.2.3. The amino acid CSIA of the soil samples

The amino acid CSIA of the two DON solutions was conducted as described in Enggrob et al. (2019).

The procedure for soil samples with complex matrix was developed and optimized during the PhD

project. In the original protocol (Paper 1), after the addition of internal standard, the solution was

immediately transferred to a polypropylene column filled with 1 g Dowex 50WX8 cation exchange

resin. In the optimized protocol, the samples were first freeze dried before being dissolved in 1 ml 0.01

M HCl, and then the sample was transferred to a polypropylene column filled with 2 g Dowex 50WX8

cation exchange resin. This single change to the purification protocol proved to be sufficient to ensure

the analysis of the wanted amino acids.

The second challenge was to find an adequate sample mass to raise analytical results above limit of

detection (LOD), calculated in paper 1. By using too much sample, there is a risk of overload the resin

in the polypropylene column and thereby loosing material. If the concentration in the samples is too

high there is a risk to end up outside the range of the standard curve and the internal standard, and

thereby not be able to calculate the concentration of the amino acid in the samples correctly. Another

risk is to overload the sensitive analytical equipment and thereby contaminating the system, which

again makes it impossible to calculate a valid concentration. However, the sample mass must not be

too small either, due to the fact that the LOD of the N analysis is much lower than that of the C analysis.

The C:N ratio in the derivatizied amino acids were between 6:1 and 14:1. The compromise was to have

an amount of soil containing enough N to raise well above LOD, but not higher than the GC column and

the combustion oven could handle with no to minimum wear and tear. The critical point was risking to

overload and contaminate the IRMS detector, which I avoided by making an analytical dilution before

the detector. This secured that I could use the same derivatized soil sample for both the 13C and the 15N

amino acid analysis. The result obtained from the amino acid CSIA are shown in Figure 15.

3.2.4. Evaluating the amino acid CSIA results

First of all we compared the concentration of amino acids across soils with and without the addition of

organic N. We wanted to make sure that we were monitoring a natural turnover of organic N and not a

system overloading with organic N. We found that there were no significant differences in the

concentration of individual amino acids in soil with or without the addition of organic N.

23

Looking in to the distribution of 13C and 15N within the amino acids after the incubation of 14 d, (Figure

15), we found that across all pH levels and organic N fractions, the lowest recoveries of individual

amino acids (leucine, lysine, phenylalanine) were close to zero, meaning that the added organic N

Figure 15 Bound amino acids remaining in individual amino acids from the peptide-sized (1-10 kDa, a-c) and

protein-sized (>100 kDa, d-f) organic N in the (a,d) low, (b, e) medium, and (c, f) high pH Jyndevad soils.

Significant differences are marked by an asterisk; a double asterisk indicates no 15N data; ‘nn’ indicate no normal

distribution. Amino acids are organized from left on right with increasing steps in their biosynthesis. The amino

acids: asparagine and aspartate (Asx), glutamine and glutamate (Glx), and Proline and Threonine (Pro/Thr) elute

together in the GC-C-IRMS analysis of acid hydrolyzed samples (n = 4).

compounds were not retained in their original form, but had been decomposed. The recovery levels

across individual amino acids were between 0 and 20% for the 15N tracer and between 1 and 30% for

the 13C tracer. The 13C and 15N decoupling, especially in the 1-10 kDa in all soil, further lend support to

microbial decomposition of the added organic N compounds. Thus, in spite of the pronounced sorption

to the soil of the added organic N compounds (Figure 15 d-f), the organic N was not protected against

microbial decomposition.

3.2.5. Confirming the results using different LTE soils

To ensure that our findings could be generalized to other soil types, soil from Askov LTE on animal and

mineral fertilizer were included in the experiment, and the incubations for 1 hours and 14 days were

repeated for the >100 kDa fraction.

The Askov LTE on animal manure and mineral fertilizers was initiated in 1894 (Christensen et al.,

2006) on sandy loam soil used for arable crop rotations (Table 4). Soil was sampled from the plough

layer (5-20 cm) of the treatments designated unfertilized, 1½ NPK, and 1½ AM treatments of the B3

a

Bo

un

d a

min

o a

cid

s r

em

ain

ing

(%

of

ad

de

d)

0

10

20

30

40

13C

15N

b c

d

Ala

As

x

Glx

Se

r

Gly

Pro

/Th

r

Va

l

Le

u

Lys

Ph

e

0

10

20

30

40e

Ala

As

x

Glx

Se

r

Gly

Pro

/Th

r

Va

l

Le

u

Lys

Ph

e

f

Ala

As

x

Glx

Se

r

Gly

Pro

/Th

r

Va

l

Le

u

Lys

Ph

e

*

**

***

* **

*

*

**

* *

*

*

** *

**

*

* * *nn

* **

** **

*

nn

Low pH soil Medium pH soil High pH soil

1-1

0 k

Da

>1

00

kD

a

24

field. Annually, the 1½ NPK and 1½ AM treatments has received on average 150 kg total-N, 30 kg P

and 120 kg K ha-1 in mineral fertilizer and animal manure (slurry since 1974), respectively. All soils

were sieved (4 mm) to remove visible roots and stored at 2°C until the incubation experiment in

October 2015. The sampling and analysis were conducted as described for the Jyndevad LTE, and the

corresponding results are shown in Figure 16.

Table 4. Basic properties of soils from the Askov LTE on animal manure and mineral fertilizers initiated in 1894. The experiment is located at Askov Experimental Station, Southern Jutland, Denmark (55o28′N, 09o07′E). The soil is classified as an Ultic Hapludalf (Soil Survey Staff, 1999).

Name pH1 C N Clay Silt Fine sand Coarse sand g/kg soil ------------------ g/kg soil --------------

Unfertilized 6.6 11.1 0.9 NPK fertilizer 6.2 12.9 1.0 100 120 430 350 Animal Manure 6.4 13.4 1.2

1 pH measured in 0.01 M CaCl2 in a 1:2.5 soil:solution ratio.

Figure 16: Fate of the >100 kDa organic N fraction in Askov soil with three fertilizer treatments: UNF is

unfertilized since 1894, NPK is mineral fertilizers since 1894, and AM is animal manure since 1894. (a)

Mineralization to 14CO2, (b, c) correlation between sorption of organic N after 1 hour and accumulated 14CO2 after

14 days with data from the Askov soils inserted in the findings from the Jyndevad soils, and (d-f) the remaining 15N and 13C in bound amino acids after 14 days of incubation for the (d) unfertilized, (e) mineral fertilized, and (f)

animal manure fertilized soils. Significant differences (panels d, e, f) are marked by an asterisk, ns indicate no

Significant differences (a) and nn indicate no normal distribution (n = 4).

Min

era

liza

tio

n o

f o

rga

nic

N t

o C

O2

(14C

O2 o

f a

dd

ed

14C

, %

)

a

Time (days)

0 2 4 6 8 10 12 14

0

10

20

30

40

50

Unfertilized (UNF)

Mineral fertilizer (NPK)

Animal Manure (AM)

ns

b

Sorption of organic N

(13C retained after 1 hour, %)

0 20 40 60 80 100

UNF

NPK

AM

c

Sorption of organic N

(15N retained after 1 hour, %)

0 20 40 60 80 100

UNF

NPK

AM

d

Ala

Asx

Glx

Ser

Gly

Pro

/Th

r

Val

Leu

Lys

Ph

e

Bo

un

d a

min

o a

cid

s r

em

ain

ing

(%

of

ad

de

d)

0

10

20

30

40

50

13C 15N

e

Ala

Asx

Glx

Ser

Gly

Pro

/Th

r

Val

Leu

Lys

Ph

e

f

Ala

Asx

Glx

Ser

Gly

Pro

/Th

r

Val

Leu

Lys

Ph

e

*

nn

*

*

*

* *

* **

*

* * **

*

*

nn

25

First of all the mineralization of >100 kDa organic N to 14CO2 in Askov soil are in the same range as for

the mineralization of >100 kDa organic N to 14CO2 in Jyndevad soil, but there was no difference across

fertilizer treatments (Figure 16a). Secondly, the sorption of both 13C and 15N matches the negative

correlation found for the incubation of the Jyndevad soil across Mw sized organic N, (Figure 16b and

c). Thirdly, bound amino acids remaining in the soil after 14 days incubation shows the same pattern

in distribution of amino acids as from the 14 days incubation of Jyndevad soil. The decoupling of 13C

and 15N, together with the decrease proportions remaining of amino acids as they increase in

complexity supports the incorporation of organic N by microbial cells (Figure 16 d-f).

3.2.6. The PLFA CSIA of the soil samples

To find out which SMB groups benefits from the incorporation of organic N, incubated soils, both from

Jyndevad LTE and Askov LTE, were analyzed for their content, distribution and 13C enrichment of

PLFA.

The PLFA CSIA analysis itself were conducted by the Stable isotope service lab., Department of Biology,

Lund University, Sweden. The preparation, extraction and derivatization of the PLFA from the soil

samples were performed as described by Petersen and Klug (1994); Petersen et al. (2002) , where 2.5

g freeze-dried soil was used to isolate phospholipids by a Bligh-Dyer single phase extraction followed

by a solid–phase extraction on silicic acid columns and an alkaline transesterification. The SMB

benefitted from the addition of organic N, especially in low pH soil and for all Askov soils. Surprisingly,

we found no substantial increase in the fungal biomarkers in the low pH soil after the addition of

organic N, instead we saw an increase in gram-negative and gram-positive bacteria in low and high pH

soil, and across all Askov soils (Figure 17). This in particular in the Jyndevad LTE low pH soil and all

Askov LTE soils. Gram-positive bacteria and fungi are typically said to contribute to the degradation of

complex compounds due to their ability to facilitate exo-enzymes, whereas gram-negative bacterial

generally decompose lower Mw compounds (Madigan et al., 2018). Looking into the specific activity of

the PLFA Figure 18 for each organic N, all microbial groups were enriched with 13C after 14 days of

incubation. Bacteria dominated the specific activity across all soils from the addition of both the 1-10

kDa and >100 kDa fractions. The addition of > 100 kDa to both Jyndevad LTE and Askov LTE showed

similar patterns with gram-positive bacteria having the highest and fungi the lowest activity. Even

though low pH have been shown to reduce bacterial activity (Rousk and Baath, 2011; Cline and Zak,

2015), this was not the case for the addition of > 100 kDa to the Jyndevad LTE, where both total PLFA

(Figure 17) and the specific activity (Figure 18) showed bacterial dominance across the pH gradient.

26

Figure 17: PLFA biomarkers divided into microbial groups, gram-positive, gram-negative and fungi of (A)

Jyndevad, and (B) Askov LTE soils for controls added water, and soil added 1-10 kDa and >100 kDa organic N

fractions (n = 4).

Figure 18: Specific 13C incorporation gram-positive, gram-negative and fungal PLFAs in Jyndevad soils added (a)

the 1-10 kDa fraction, and (b) the >100 kDa fraction, and (C) in Askov soils added the >100 kDa fraction.

Significant differences are marked by different letter above the bars (n = 4).

B: Askov

UNF NPK AM

Co

ntr

ol

>1

00

kD

a

Co

ntr

ol

>1

00

kD

a

Co

ntr

ol

>1

00

kD

a

Mic

rob

ial b

iom

as

s

(nm

ol P

LF

A g

-1 s

oil)

0

2

4

6

8

10

12

14

A: Jyndevad

Low pH Medium pH High pH

Co

ntr

ol

1-1

0 k

Da

>1

00

kD

a

Co

ntr

ol

1-1

0 k

Da

>1

00

kD

a

Co

ntr

ol

1-1

0 k

Da

>1

00

kD

a

Mic

rob

ial b

iom

as

s

(nm

ol P

LF

A g

-1 s

oil)

0

1

2

3

4

5G+

G-

Fungi

c c

d

ab

d

cc

d

a

b

d

a

b

c

aa

c

c

d

e

abc

e

abc

e

a

ab

c

a

ab

c

ab

b

c

c

d

a

b

e

b

c

e

a

Low pH Medium pH High pH

0.00

0.01

0.02

Gram postive bacteria (G+)

Gram negative bacteria (G-)

Fungi

b

Low pH Medium pH High pH

0.00

0.01

0.02

b a c

a a b

a a b

a b ca b c

a b c

Mic

rbia

l b

iom

ass

ac

tive o

n o

rgan

ic N

(nm

ol

13C

PL

FA

nm

ol-1

C P

LF

A)

c

UNF NPK AM

a b c

a b c a b c

27

3.3. Plant N uptake from higher Mw DON (paper 3)

After investigating the turnover of DON by the SMB alone, we added a plant to the system in a second

series of experiments. We continued using soil from the Jyndevad LTE, table 2, and the Mw sized

fraction > 100 kDa.

We chose maize in four varieties as our test plant: LG 31.218, Alfastar, Atrium and Emblem. The aim of

the study was to investigate the turnover of protein-sized organic N (>100 kDa) and the uptake of this

organic N source in maize grown in soil with three pH levels. The study was based on the hypothesis

that: (i) the turnover of >100 kDa organic N would be greater in the presence of plants due to a more

active microbial community in rhizosphere soil than bulk soil (Blagodatskaya et al., 2014) and plant

exudation of proteolytic exo-enzymes (Godlewski and Adamczyk, 2007), and (ii) higher plant growth

and greater mineralization with increasing soil pH would result in a greater total plant N uptake from

>100 kDa organic N at high soil pH, but a greater proportion of organic N uptake at low pH.

When studying organic N uptake short chase periods in the studies are needed (Nasholm et al., 2009;

Hill and Jones, 2019), but for this study we still needs sufficient time for depolymerization to occur,

assuming that maize is unable to assimilate protein-sized organic N directly. The incubation time of 48

hours was based on the mineralization pattern of the >100 kDa organic N from paper 2.The plant

uptake of organic N was studied through a micro-lysimeter experiment in a similar manner as

described in section 3.2. using CIRO Centrifuge Filters (XPE-45 Maxi-Spin Filters 0.45 PES, Frisinette

ApS, Knebel, Denmark). Each micro-lysimeter was filled with approximately 15 g of field moist soil and

packed by gently tapping on the side of the unit. In addition to the maize receiving the triple-labeled

organic N we had three control treatments: 1) maize with water added (i.e. no organic N), 2) unplanted

soil receiving the triple-labeled organic N (i.e. no plant), 3) unplanted soil with water added (i.e. no

organic N and no plant). All treatments and controls were sampled by end-point sampling of both soil

and plant tissue 48 hours after the addition of triple-labeled organic N (or water for the respective

controls).

Unplanted controls were setup as described in section 3.2.1. These incubations ran for 48h, and the

base trap were changed after 1, 2, 4, 24 and 48 hours, and immediately analyzed by LSC for 14C activity,

the resulting mineralization curves are shown in Figure 19A and B. After end of incubation, the soils

were frozen before further analysis.

Maize seeds were germinated in the dark for 2 days at room temperature before transferring into

micro-lysimeters. After 20 days of growth in the laboratory with a day length of 14h at 24-28°C and

irrigation as needed, the maize plants reached the BBCH growth stage 12-13 and assessed to be ready

for the incubation experiment. Maize grew significantly better in soil at medium and high pH than at

28

Figure 19. Mineralization of protein-sized organic N (>100 kDa) to 14CO2 in Jyndevad soils. (A) Temporal

development of mineralization and (B) accumulated mineralization after 48 hours in soils without maize, and (C)

accumulated mineralization after 48 hours in soils with four maize varieties. The three pH levels are low at

pHCaCl2 3.6, medium at pHCaCl2 5.4, and high at pHCaCl2 7.1. Statistical differences among soil pH levels in

accumulated 14CO2 after 48 hours are indicated by different letters above the bars (n = 4).

the low pH level. Maize plants at the soil low pH level had significantly lower shoot, root and total dry

matter yields than maize in the medium and high soil pH levels (paper 3).

The micro-lysimeters were placed in 1L glass jars, with a beaker containing 4 ml of water and a

scintillation vial containing 1 ml of 0.5 M NaOH. In addition, a hole was drilled in the lid to pull through

the stem of the maize. The incubation were initiated after the addition of 2.00 ml triple-labelled (15N,

14C and 13C) Mw sized > 100 kDa to the micro-lysimeters, or 2.00 ml water for the control treatments;

all done in four replicates. After end incubation the base trap was analyzed for 14C activity, the

resulting mineralization is shown in Figure 19C. The maize was harvested by cutting the stem at soil

level, and then the soil and roots were gently separated and roots washed free of soil. Shoot, root and

soil samples were frozen before further analysis.

The mineralization in unplanted soil followed a first order kinetics with detection of 14CO2 already

after 1 hour across all pH levels (Figure 19A and B). The accumulated mineralization was significantly

higher (P = 0.0259) in the high pH soil than the low pH soil (Figure 19B); with 9.2 ± 0.7 % of added 14C

as 14CO2 at high pH compared to 6.2 ± 0.6 % of added at low pH. The medium pH soil had intermediate

B

Without maize

C

LG Emblem Alfastar Atrium

Min

era

liza

tio

n t

o C

O2 (

ac

cu

mu

late

d 1

4C

O2 o

f a

dd

ed

14C

, %

)

0

2

4

6

8

10

12

14

16

A

Time of incubation (hours)

0 10 20 30 40 50

0

2

4

6

8

10

12Low pH (3.6)

Medium pH (5.4)

High pH (7.1)

b

aba

- a - - a -

- a -

- a -

29

14CO2 evolution with 7.0 ± 0.3 % of added 14C as 14CO2 after 48 hours (Figure 19B). Interestingly, these

differences in mineralization across soil pH levels disappeared in the presence of maize (Figure 19C).

In paper 3, we chose to focus only on one of the four maize variety, but here I will present bulk data for

all four varieties. When plants were present, the mineralization to 14CO2 tended to increase in the low

pH soil and so did the variation among replicate samples across all pH levels and maize varieties. Thus,

there were no significant differences between mineralization across soil pH levels within each maize

variety; neither did we find significant differences in mineralization across maize varieties within each

soil pH levels (Figure 19C).

The bulk analysis of total C and N, and 13C and 15N stable isotope composition was determined by

transferring 5-7 mg shoot or root material to tin capsules before analysis on a PDZ Europa ANCA-GSL

elemental analyzer interfaced to a PDZ Europe 20-20 isotope ratio mass spectrometer (Sercon Ltd.

Cheshire, UK) at the UC Davis Stable Isotope Facility. The corresponding results are shown in Figure

20. The uptake of 15N was significantly (P < 0.0001) greater than the uptake of 13C in all four maize

varieties across all soil pH levels (Figure 20). The total uptake of 15N ranged from 5.8 to 12.4 % of 15N

added with the >100 kDa fraction (Figure 20F) with the 15N equally distributed among the roots and

the shoots (Figure 20B, 20D). The 15N uptake for the LG and Emblem varieties was significantly higher

in the medium and high pH soils than in the low pH soil in line with the mineralization pattern found

for the unplanted soil. However, we found no correlation between 15N uptake and the actual 14CO2

evolution in the planted soils (data not shown). The uptake of 13C was significantly (P < 0.0001) higher

in roots than in shoots 0.9 to 2.6 % of 13C added with the >100 kDa fraction present in the roots and

0.2 to 0.5% of 13C added present in shoots after 48 hours. Although the uptake of 13C in roots tended to

be lower in the low pH soils was there no significant differences across the soil pH levels for the four

maize varieties (Figure 20C). On a whole plant basis, the uptake of 13C was on average 21% (ranging

from 13 to 31%) of the uptake of 15N with no significant differences in the 13C uptake-to-15N uptake-

ratio across soil pH levels or maize varieties. Uptake of organic N is mainly determined using (13C, 15N)

dual-labeled compound with the uptake estimated based on the ratio of bulk 13C and 15N isotope taken

up (Nasholm et al., 1998). This indicate that the uptake of organic N from the depolymerization of

higher Mw DON over the time of 48 hour by maize seedling were on average 21%. The LG maize

variety tended to have greater bulk 13C uptake across the soil pH levels and was therefore chosen for

further CSIA. The amino acid SIP analysis was performed as described above. From the incubation of

LG with and without the addition of Mw sized >100 kDa the concentration of amino acid in soils and in

root tissues after 48 hour incubation are calculated and shown in supporting information for paper 2.

Overall, there were no significant differences in the

30

Figure 20. Uptake (A, C, E) of 13C and (B, D, F) of 15N from protein-sized organic N (>100 kDa) after 48 hours in

maize (A, B) shoots, (C, D) roots, and (E, F) the whole plant. Significant differences in uptake among soils with

different pH are marked by different letters above the bars; ‘nn’ indicates no normal distribution (n = 4).

concentration of individual amino acids, except for Asx in soil at high pH and serine, Pro/Thr and

tyrosine in root at low pH. Indicating that we did not overload the system with organic N added.

Investigating how the added 13C was distributed in the amino acids, calculated as percent of added 13C,

of the soil from both planted and unplanted soil revealed an interesting pattern, Figure 21. In the

presence of maize (variety LG) the general pattern was that significantly higher proportions of

individual amino acids remained in the soil compared to the unplanted control. The exception was

tyrosine, lysine and phenylalanine. For lysine and phenylalanine a similar proportions remained at all

pH levels and at medium and high pH level, respectively. For tyrosine the proportion remained in the

soil where greater in the unplanted control. Before the CSIA the soil were hydrolyzed and we therefore

cannot deduce whether the individual amino acids remaining after 48 hours were bound in the >100

kDa organic N added or had been incorporated in microbial tissue. Assuming an equal degradation of

A Shoot

Reco

very

of

trace

r in

pla

nt

tis

su

e (

%o

f a

dd

ed

)

0

2

4

6

8 Low pH

Medium pH

High pH

C Root

0

2

4

6

8

E Whole plant

LG

Em

ble

m

Alf

asta

r

Atr

ium

0

5

10

15

B

D

F

LG

Em

ble

m

Alf

asta

r

Atr

ium

a b ba b b

a b a

a b b

a b a

a b b a b b

a b b

- a -

- a -

- a -a b a

- a - - a - - a - - a -

- a - - a -- a -

- a -

- a - nn a b abnn

--- bulk 13

C --- --- bulk 15

N ---

31

proteins in the added >100 kDa organic N, we use the individual amino acids with lowest proportions

remaining as an estimate of the proportion of proteins in the >100 kDa organic N remaining intact as

added. This proportion of original organic nitrogen highlighted in Figure 21 by a red line.

Figure 21. Bound amino acids from added >100 kDa organic N remaining after 48 hours without and with maize

in Jyndevad soils at (a) low, (b) medium, and (c) high pH. Significant differences between soils without and with

maize in 13C remaining for individual amino acids are marked by an asterisk above the bars (n = 4). Amino acids

are organized from left on right with increasing steps in their biosynthesis. The amino acids: asparagine and

aspartate (Asx), glutamine and glutamate (Glx), and Proline and Threonine (Pro/Thr) elute together in the GC-C-

IRMS analysis of acid hydrolyzed samples. The red dashed line indicate the lowest proportion of an individual

amino acid remaining in soil without maize.

3.3.1. Key findings

Surprisingly, we found higher proportions of individual amino acids remaining in the soil with maize

than in unplanted soil; this in particular pronounced for the amino acids with fewer than those with

more biosynthetic steps (amino acids to the left in Figure 21). The finding of more amino acids

a: Low pH

0

20

40

60

80

100Without maize

With maize

b: Medium pH

Bo

un

d a

min

o a

cid

re

ma

inin

g i

n s

oil

(%

of

13C

ad

de

d i

n i

nd

ivid

ua

l a

min

o a

cid

s)

0

20

40

60

80

100

c: High pH

Ala

As

x

Glx

Se

r

Gly

Pro

/Th

r

Va

l

Ile

Le

u

Lys

Tyr

Ph

e

0

20

40

60

80

100

*

*

*

* *

*

*

**

*

*

*

*

*

**

*

**

*

*

*

*

*

*

*

**

*

*

32

remaining with maize than in unplanted soil is surprising since turnover is generally consider to be

greater in rhizosphere than bulk soil due to greater microbial activity (Blagodatskaya et al., 2014) and

plant exudation of proteolytic enzymes (Godlewski and Adamczyk, 2007).

The total 15N uptake reached 12.4 % of added and the total 13N uptake reached up to 2.6% of added

after 48 hours across the maize varieties (Figure 20). The presence of individual 13C-labeled amino

acids varied significantly in maize roots of the LG variety at all three soil pH levels (Figure 22).

Figure 22. Presence of 13C-labeled bound amino acids from added >100 kDa organic N in maize roots after 48

hours in Jyndevad soils at (A) low, (B) medium, and (C) high pH. Significant differences between presence among

individual amino acids within each soil pH level are marked by different letters above the bars (n=4). Amino

acids are organized from left on right with increasing steps in their biosynthesis.

The presence ranged from 0 to 1.7% of the 13C added with the >100 kDa organic N across soils; with no

significant effects of soil pH level on the presence of individual 13C-labeled amino acids. The presence

of individual amino acids had a similar pattern at all soil pH levels where glutamine/glutamate,

proline/threonine and leucine had the greatest presence and lysine had the lowest presence

throughout. Importantly, the pattern of amino acid presence in maize roots did not resemble the

A: Low pH

0.0

0.5

1.0

1.5

2.0

B: Medium pH

Pre

se

nc

e o

f la

be

led

bo

un

d a

min

o a

cid

in

ma

ize

ro

ots

(% o

f 1

3C

ad

de

d i

n i

nd

iviu

al

am

ino

ac

ids

)

0.0

0.5

1.0

1.5

2.0

C: High pH

Ala

As

x

Glx

Se

r

Gly

Pro

/Th

r

Va

l

Ile

Le

u

Lys

Tyr

Ph

e

0.0

0.5

1.0

1.5

2.0

b bc

cd

a

aa

b

cd cd

bc

d

de de

ef

b

a

bc

cd

dfef

d

fgg

bcbc

ce

a a

b

cb

defeg

cd

fgg

33

pattern of neither amino acids remaining in the soil nor amino acids lost from the soil. The average

presence of 13C-labeled amino acids was 0.5-0.6 % of added, which correspond to one third of the bulk

13C presence in maize roots; the latter reaching 2.2 % of added in the LG variety.

34

4. General discussion

The aim of this PhD project was to study the role of higher Mw dissolved organic N in the plant-soil N

cycle. Based on three objectives, an extensive experimental and analytical work was performed, which

resulted in three papers that are presented individually in section 3. The following discussion will be

based on the main points of this PhD projects, drawing in results across the three papers.

4.1. The analytical method for amino acid CSIA

Previous studies have typically focused on tracing the turnover of individual amino acids in soil, e.g.,

looking at dissipation kinetic of asparagine (Czaban et al., 2016) or compound specific analysis of di-

and tri-peptides (Jamtgard et al., 2018). But for the study of more complex organic N compounds, such

as large peptides and proteins, better approaches are needed to advance the state of the art. Having a

good analytical method implies that the results can be trusted and repeated, which was the case with

the NAIP derivaterization method (Paper 1) and the optimized soil hydrolysis procedure (Paper 2 and

3). The presently developed methodology for amino acid CSIA of large Mw organic N concerns not only

the GC-C-IRMS analysis itself, but also a detailed and easy-to-follow sample preparation protocol, to

ensure consistency in sample preparation as well as sample analysis.

We used Mw sized fractionation to divide triple labeled DON into size fractions and thereby reduce the

complexity of the organic N compounds. Each fraction was hydrolyzed and analyzed for the amino acid

distribution and enrichment. Results confirmed that hydrolysis is a harsh treatment (Fountoulakis and

Lahm, 1998), and that there is a risk of losing free amino acids, but simultaneously it was shown that

the isotopic signature was not affected by the hydrolysis. Purification procedures were optimized to fit

the soil samples, enabling the GC-C-IRMS analysis of the effect of added organic N on the soil and plant-

soil systems.

I believe that paper 2 and 3 demonstrate the strength of analyzing multiple amino acids. Thus, when

able to create an amino acid profile or fingerprint of the original organic N compound, the fate of the

organic N can be traced through the soil N cycling (paper 2) and soil-plant N cycling (paper 3). This for

example led to the discovery of greater recycling of simple amino acids compared to those with more

complex biosynthesis (Paper 2). Paper 1 showed that there could be an issue with hydrolyzing

samples containing free amino acids. We avoided interference from this issue in the work leading up to

paper 2 by using Mw sized fractions > 1 kDa and by washing the soil with water at the end of

incubation thus removing free amino acids. In the work leading to paper 3, the issue was eliminated by

using the Mw sized fraction >100 kDa, and as shown in studies of the turnover of free amino acids

(Wilkinson et al., 2014; Hill and Jones, 2019), the half-life of these compounds is short. Therefore, the

35

concentration of free amino acids is most likely low and so is the potential bias from loss of free amino

acids upon soil hydrolysis.

CSIA profiling is time-consuming and here we chose the end point sampling for the amino acid CSIA in

different soil management instead of a detailed time series. In the future, more detailed time series of

both soil and plant amino acid CSIA would help the understanding of the uptake dynamic between the

plant and the SMB.

4.2. Mineralization and sorption of organic N in soil with or without

plants

The mineralization of added Mw sized >100 kDa organic N during 48 hour incubation revealed that

there was no lag phase (Figure 19), indicating that a functioning soil microbial population was present

to catabolize higher Mw organic N without preceding growth and most likely also without de novo

enzyme synthesis. The lower mineralization to 14CO2 after 14 days observed for the >100 kDa fraction

than for the 1-10 kDa fraction (Figure 10 and Figure 13) indicated that >100 kDa organic N to a

greater extent was involved in microbial anabolism (Liang et al., 2017). This was also confirmed by the

incorporation of 13C in PLFA. In addition, the low recovery of labeled amino acids after 14 days

incubation (Figure 15 and 16) can hardly be explained by chemical sorption of the originally added

amino acids in peptides and proteins as such processes would result in similar or consistent recoveries

across all amino acids. Thus, C and N from amino acids were incorporated in microbial tissue early in

the degradation; paper 3 shows a large proportion already after 48 hours. This lend support to 14CO2

evolution being at two phases also from protein-sized ON, with the first phase representing the initial

microbial respiration of the added compounds and the second phase representing the subsequent

respiration of organic compounds initially incorporated in the microbial tissue (Boddy et al., 2007;

Boddy et al., 2008; Farrell et al., 2011).

In the plant-soil system studied in paper 3, there was a positive effect from the presence of plants on

14CO2 production from the low pH soil (Figure 19 c). In paper 3 we outlined that this could be due to a

more active SMB, greater overall turnover of added organic N, and root respiration of the added

organic N. The CSIA of amino acids in soil showed a higher microbial amino acid incorporation with

plants than without plants. This strongly indicate a higher microbial C use efficiency of amino acid

derived C, and thus a lower microbial contribution to 14CO2 respiration from the added organic N.

Therefore, root 14CO2 respiration was most likely responsible for the higher mineralization (Nasholm

et al., 2009; Fischer et al., 2010; Warren, 2012; Hildebrandt et al., 2015).

It could be worth repeating the 48 hour incubation with plants to get a more detailed timeline, both for

the respiration of 14CO2 and the incorporation of amino acids in the SMB.

36

4.3. Organic nitrogen in soil with and without plants

In paper 2, the study provides strong evidence for the hypothesis that C and N from labile compounds

persist in soil (Cotrufo et al., 2015), but rather than persisting due to protection of the original

compounds (Schmidt et al., 2011), the C and N persist due to the incorporation via anabolic processes

into microbial tissue. We concluded based on 14 days incubation studies that degradation of large

organic N is rapid and that SMB incorporates especially the simpler amino acids. This picture was

confirmed from the CSIA of amino acids in soil with and without plants after 48 hours as reported in

paper 3. An example is shown for two bound amino acids (alanine and lysine) remaining in soil after

48 hours and 14 days having a high (alanine) and a low (lysine) percent of added remaining in the soil

(Figure 23). Alanine is a simple amino acid with few biosynthetic steps (Kirchman et al., 1986), and

was one of the amino acids with highest amount remaining. Lysine is a complex amino acid with

several biosynthetic steps (Kirchman et al., 1986), and was one of the amino acids with lowest amount

remaining. A curve fitting of these data would have a striking resembling to that of the mineralization

curves, although of course the dissipation curves show decreases in amino acid remaining and

mineralization curves show increases in CO2 respiration. The tendency of the more complex amino

acids to decrease more rapidly than the simpler amino acids was surprising and diverges from the

expected direct anabolic microbial use of amino acids with complex biosynthesis saving the

Figure 23: The rapid degradation of the added >100 kDa organic N is illustrated by the presence of bound 13C-

labeled alanine and lysine from the 48 hours incubation (paper 3) and the 14 days incubation (paper 2) in

Jyndevad soil at low pH. The differences in the level of 13C remaining of alanine and lysine illustrate the higher

use of complex amino acids (lysine) for energy and the higher incorporation of simple amino acids (alanine) as

microbial building blocks.

Incubation time (days)

0 2 4 6 8 10 12 14

Bo

un

d a

min

o a

cid

s r

em

ain

ing

(%

of

ad

de

d 1

3C

)

0

20

40

60

80

100

Lysine in low pH soil without plant

Alanine in low pH soil without plant

37

microorganisms most energy (Kirchman et al., 1986). The majority of the labeled amino acids

recovered from soil after 48 hours were most likely in microbial tissue in the unplanted soil; with a

good proportion of C-skeletons entering biosynthesis of other compounds as less than 10% of C was

respired as 14CO2 in unplanted soil. Interestingly, simpler amino acids, such as alanine,

asparagine/aspartate, glutamine/glutamate, and glycine, are typically among the most abundant

constituents of the peptidoglycan layers of bacterial cell walls (Simelyte et al., 2003; Vollmer et al.,

2008; Schneewind and Missiakas, 2012). Hence, these findings support substantial incorporation of

organic N by microbial cells.

4.4. Plant N uptake

The variation in the maize root presence of individual amino acids from the added >100 kDa organic N

showed that organic N uptake contributed to maize N uptake (paper 3). We estimated that 20-30%

was taken up in organic form based on the bulk 13C and 15N uptake. Since, plants most likely

contributed to the 14CO2 respiration (see section 4.1.), then the 20-30% N uptake in organic form is

probably underestimated due to post-uptake metabolism (Nasholm et al., 2009; Warren, 2012).

Supporting this, is the soil solution data from experiment 2, where the temporal development in 13C

and 15N presence show a parallel loss of 13C and 15N from 1 hour to 1 days with no indication of 15N-

inorgaic N in this period (Figure 10b). This point to that for at least the first 24 h, the disappearance of

13C and 15N are correlated, meaning that the N available was in organic form and that the release of

DIN do not occur until after 24 hours. Hence supporting, a higher N uptake in organic form than the

20-30% estimate, which contradicting the recent finding by Hill and Jones (2019) showing N uptake

being dominated by inorganic N forms when adding alanine to a plant-soil system.

With the present data we cannot determine at what rate the >100 kDa was depolymerized and hence

made bio-available, and therefore the uptake rate of organic N by plants cannot be determined either.

But, the post-uptake fate of individual amino acids may be indicated by the specific enrichment of

amino acids in the maize roots (paper 3), where a greater specific enrichment indicate a greater

recycling of amino acid C-skeletons. We speculate that the post-uptake fate of amino acids is a balance

between abundance of the amino acid in the plant tissue and the energy gain when using the amino

acid in catabolism (Hildebrandt et al., 2015). Again, future studies with more detailed time scales

would be interesting to determine the plant uptake rate of organic N compounds and their fate in root

after uptake.

38

5. Conclusion

In this PhD project I examined the turnover of higher Mw dissolved organic N in soils from two LTE,

Jyndevad and Askov. The studies were based on among other the setup, testing and optimization of a

CSIA method for labeled amino acid from complex organic N compounds. The LTE soils, with and

without growing plants, were incubated in micro-lysimeters in 1 L glass jars at varies times, from 1

hour to 14 days. The experiments were setup to test four hypothesis:

(i) In cultivated soil the pool of Higher Mw DON represents the bottleneck in the production of plant

available N from soil organic N. This hypothesis was rejected in paper 2, where there was a rapid

turnover of large molecular size organic N compounds. It was concluded that large organic N primarily

contributes to SOM formation via build-up of microbial tissue, where incorporation of C and N in the

short-chained peptides of bacterial cell walls potentially results in longer-term storage of plant-

derived C. The study, thus, provides strong evidence for the hypothesis that C and N from labile

compounds persist in soil via anabolic incorporation into microbial tissue.

(ii) Differences in soil pH will affect which microbial communities dominate the decomposition of DON; at

low pH fungi is expected to dominate and at increasing pH the dominating microbial communities will

shift towards bacteria. However, I found that bacteria dominated the decomposition of both 1-10 kDa

and >100 kDa across all LTE soils. Turnover of the largest organic N (>100 kDa) was dominated by

gram-positive bacteria, and we suggested that this could be coupled to direct uptake of organic N

larger than the presently acknowledged assimilation limit of 0.6 kDa.

(iii) The competition between plants and the microbial communities for plant available DON will increase

the turnover of high Mw DON compared to a soil without plants. The findings in paper 3 showed

interestingly that the concentration of individual amino acids was higher in soils with maize growing

than in soils with no maize growing. The amino acids found to have the highest concentration was

consistent with the amino acids most abundant constituents of the peptidoglycan layers of bacterial

cell walls, alanine, Asx, Glx and glycine. The interpretation of this was that with a plant growing in the

soil, there is a second source of C for the SMB and therefore more amino acids were taken up by the

SMB to be used as building blocks in creating new cell growth, instead of undergoing deamination and

being used for energy.

(iv) At low soil pH, organic N turnover is expected to be slower and hence there will be a greater chance

of direct organic N plant uptake as indicated by 13C presence in roots, whereas at higher pH

mineralization will be greater and so will dissolved inorganic 15N (DI15N). The plant uptake of 15N from

the > 100 kDa organic N after 48 hours was increasing with soil pH reaching 12% of the added 15N. The

maize uptake of organic N, confirmed by the presence of 13C-labeled amino acid in the maize roots, was

39

estimated based on the ratio between the net-uptake of 13C-to-15N to be 20-30% of the total 15N uptake

with no significantly different across soil pH.

Overall, I conclude based on my finding during this PhD project that depolymerization does not pose a

bottleneck to the turnover of dissolved organic N into bioavailable N. The decomposition of higher Mw

organic N was dominated by the exo-enzymatic activity of gram-positive bacteria, which would

potentially allow plants to assimilate degradation metabolites, such as amino acids or short peptides.

The latter was confirmed by the uptake of organic N in young maize plants. Hence, the study showed

that large organic N can make a significant contribution to plant and microbial N nutrition.

40

6. References

Berg, J.M., Tymoczko, J.L., Stryer, L., 2006. Biochemistry. Sara Tenney, W. H. Freeman and Company, New York.

Blagodatskaya, E., Blagodatsky, S., Anderson, T.H., Kuzyakov, Y., 2014. Microbial Growth and Carbon Use Efficiency in the Rhizosphere and Root-Free Soil. Plos One 9, 9.

Blagodatskaya, E., Kuzyakov, Y., 2013. Active microorganisms in soil: Critical review of estimation criteria and approaches. Soil Biology & Biochemistry 67, 192-211.

Boddy, E., Hill, P.W., Farrar, J., Jones, D.L., 2007. Fast turnover of low molecular weight components of the dissolved organic carbon pool of temperate grassland field soils. Soil Biology & Biochemistry 39, 827-835.

Boddy, E., Roberts, P., Hill, P.W., Farrar, J., Jones, D.L., 2008. Turnover of low molecular weight dissolved organic C (DOC) and microbial C exhibit different temperature sensitivities in Arctic tundra soils. Soil Biology & Biochemistry 40, 1557-1566.

Boschker, H.T.S., Vasquez-Cardenas, D., Bolhuis, H., Moerdijk-Poortvliet, T.W.C., Moodley, L., 2014. Chemoautotrophic Carbon Fixation Rates and Active Bacterial Communities in Intertidal Marine Sediments. Plos One 9, 12.

Burns, R.G., DeForest, J.L., Marxsen, J., Sinsabaugh, R.L., Stromberger, M.E., Wallenstein, M.D., Weintraub, M.N., Zoppini, A., 2013. Soil enzymes in a changing environment: Current knowledge and future directions. Soil Biology & Biochemistry 58, 216-234.

Christensen, B.T., Petersen, J., Trentemoller, U.M., 2006. The Askov Long-Term Experiments on Animal Manure and Mineral Fertilizers: The Lermarken site 1894-2004., DIAS report Plant productions.

Cline, L.C., Zak, D.R., 2015. Soil microbial communities are shaped by plant-driven changes in resource availability during secondary succession. Ecology 96, 3374-3385.

Corr, L.T., Berstan, R., Evershed, R.P., 2007. Optimisation of derivatisation procedures for the determination of delta C-13 values of amino acids by gas chromatography/combustion/isotope ratio mass spectrometry. Rapid Communications in Mass Spectrometry 21, 3759-3771.

Cotrufo, M.F., Soong, J.L., Horton, A.J., Campbell, E.E., Haddix, M.L., Wall, D.H., Parton, A.J., 2015. Formation of soil organic matter via biochemical and physical pathways of litter mass loss. Nature Geoscience 8, 776-+.

Czaban, W., Rasmussen, J., Nicolaisen, M., Fomsgaard, I.S., 2016. Dissipation kinetics of asparagine in soil measured by compound-specific analysis with metabolite tracking. Biology and Fertility of Soils 52, 911-916.

Dumont, M.G., Murrell, J.C., 2005. Stable isotope probing — linking microbial identity to function. Nature Reviews Microbiology 3, 499.

Enggrob, K.L., Larsen, T., Larsen, M., Elsgaard, L., Rasmussen, J., 2019. The influence of hydrolysis and derivatization on the determination of amino acid contentand isotopic ratios in dual‐labeled (13C,15N) white clover. Rapid Commun Mass Spectrom 33, 21-30.

Farrell, M., Hill, P.W., Wanniarachchi, S.D., Farrar, J., Bardgett, R.D., Jones, D.L., 2011. Rapid peptide metabolism: A major component of soil nitrogen cycling? Global Biogeochemical Cycles 25, 11.

Fierer, N., Schimel, J.P., Holden, P.A., 2003. Variations in microbial community composition through two soil depth profiles. Soil Biology & Biochemistry 35, 167-176.

41

Fischer, H., Eckhardt, K.U., Meyer, A., Neumann, G., Leinweber, P., Fischer, K., Kuzyakov, Y., 2010. Rhizodeposition of maize: Short-term carbon budget and composition. J. Plant Nutr. Soil Sci. 173, 67-79.

Fountoulakis, M., Lahm, H.W., 1998. Hydrolysis and amino acid composition analysis of proteins. Journal of Chromatography a 826, 109-134.

Frostegard, A., Baath, E., Tunlid, A., 1993. SHIFTS IN THE STRUCTURE OF SOIL MICROBIAL COMMUNITIES IN LIMED FORESTS AS REVEALED BY PHOSPHOLIPID FATTY-ACID ANALYSIS. Soil Biology & Biochemistry 25, 723-730.

Frostegard, A., Tunlid, A., Baath, E., 2011. Use and misuse of PLFA measurements in soils. Soil Biology & Biochemistry 43, 1621-1625.

Ge, T.D., Song, S.W., Roberts, P., Jones, D.L., Huang, D.F., Iwasaki, K., 2009. Amino acids as a nitrogen source for tomato seedlings: The use of dual-labeled (C-13, N-15) glycine to test for direct uptake by tomato seedlings. Environ. Exp. Bot. 66, 357-361.

Godlewski, M., Adamczyk, B., 2007. The ability of plants to secrete proteases by roots. Plant Physiology and Biochemistry 45, 657-664.

Hildebrandt, T.M., Nesi, A.N., Araujo, W.L., Braun, H.P., 2015. Amino Acid Catabolism in Plants. Molecular Plant 8, 1563-1579.

Hill, P.W., Jones, D.L., 2019. Plant–microbe competition: does injection of isotopes of C and N into the rhizosphere effectively characterise plant use of soil N? New Phytologist 221, 796-806.

Hodge, A., Robinson, D., Fitter, A., 2000. Are microorganisms more effective than plants at competing for nitrogen? Trends in plant science 5, 304-308.

Jamtgard, S., Nasholm, T., Huss-Danell, K., 2010. Nitrogen compounds in soil solutions of agricultural land. Soil Biology & Biochemistry 42, 2325-2330.

Jamtgard, S., Robinson, N., Moritz, T., Colgrave, M.L., Schmidt, S., 2018. Optimising methods for the recovery and quantification of di- and tripeptides in soil. Soil Research 56, 404-412.

Jan, M.T., Roberts, P., Tonheim, S.K., Jones, D.L., 2009. Protein breakdown represents a major bottleneck in nitrogen cycling in grassland soils. Soil Biology & Biochemistry 41, 2272-2282.

Jones, D.L., 1999. Amino acid biodegradation and its potential effects on organic nitrogen capture by plants. Soil Biology & Biochemistry 31, 613-622.

Jones, D.L., Healey, J.R., Willett, V.B., Farrar, J.F., Hodge, A., 2005. Dissolved organic nitrogen uptake by plants - an important N uptake pathway? Soil Biology & Biochemistry 37, 413-423.

Jones, D.L., Kielland, K., 2002. Soil amino acid turnover dominates the nitrogen flux in permafrost-dominated taiga forest soils. Soil Biology & Biochemistry 34, 209-219.

Jones, D.L., Murphy, D.V., 2007. Microbial response time to sugar and amino acid additions to soil. Soil Biology & Biochemistry 39, 2178-2182.

Jones, D.L., Willett, V.B., Stockdale, E.A., Macdonald, A.J., Murphy, D.V., 2012. Molecular Weight of Dissolved Organic Carbon, Nitrogen, and Phenolics in Grassland Soils. Soil Science Society of America Journal 76, 142-150.

42

Kemmitt, S.J., Lanyon, C.V., Waite, I.S., Wen, Q., Addiscott, T.M., Bird, N.R.A., O'Donnell, A.G., Brookes, P.C., 2008. Mineralization of native soil organic matter is not regulated by the size, activity or composition of the soil microbial biomass - a new perspective. Soil Biology & Biochemistry 40, 61-73.

Kirchman, D.L., Newell, S.Y., Hodson, R.E., 1986. INCORPORATION VERSUS BIOSYNTHESIS OF LEUCINE - IMPLICATIONS FOR MEASURING RATES OF PROTEIN-SYNTHESIS AND BIOMASS PRODUCTION BY BACTERIA IN MARINE SYSTEMS. Marine Ecology Progress Series 32, 47-59.

Knicker, H., 2011. Soil organic N - An under-rated player for C sequestration in soils? Soil Biology & Biochemistry 43, 1118-1129.

Knief, C., Altendorf, K., Lipski, A., 2003. Linking autotrophic activity in environmental samples with specific bacterial taxa by detection of C-13-labelled fatty acids. Environ. Microbiol. 5, 1155-1167.

Kusliene, G., Rasmussen, J., Kuzyakov, Y., Eriksen, J., 2014. Medium-term response of microbial community to rhizodeposits of White clover and ryegrass and tracing of active processes induced by C-13 and N-15 labelled exudates. Soil Biology & Biochemistry 76, 22-33.

Larsen, T., Ventura, M., Andersen, N., O'Brien, D.M., Piatkowski, U., McCarthy, M.D., 2013. Tracing Carbon Sources through Aquatic and Terrestrial Food Webs Using Amino Acid Stable Isotope Fingerprinting. Plos One 8, e73441-e73441.

Liang, C., Schimel, J.P., Jastrow, J.D., 2017. The importance of anabolism in microbial control over soil carbon storage. Nature Microbiology 2.

Liebig, J., 1842. Chemistry in its application to Agriculture and Physiology. John Owen, Cambridge.

Madigan, M.T., Bender, K.S., Buckley, D.H., Sattley, W.M., Stahl, D.A., 2018. Brock Biology of Microorganisms. Pearson Education Limited, London, UK.

Nasholm, T., Ekblad, A., Nordin, A., Giesler, R., Hogberg, M., Hogberg, P., 1998. Boreal forest plants take up organic nitrogen. Nature 392, 914-916.

Nasholm, T., Huss-Danell, K., Hogberg, P., 2001. Uptake of glycine by field grown wheat. New Phytologist 150, 59-63.

Nasholm, T., Kielland, K., Ganeteg, U., 2009. Uptake of organic nitrogen by plants. New Phytologist 182, 31-48.

Owen, A.G., Jones, D.L., 2001. Competition for amino acids between wheat roots and rhizosphere microorganisms and the role of amino acids in plant N acquisition. Soil Biology & Biochemistry 33, 651-657.

Petersen, S.O., Frohne, P.S., Kennedy, A.C., 2002. Dynamics of a soil microbial community under spring wheat. Soil Science Society of America Journal 66, 826-833.

Petersen, S.O., Klug, M.J., 1994. EFFECTS OF SIEVING, STORAGE, AND INCUBATION-TEMPERATURE ON THE PHOSPHOLIPID FATTY-ACID PROFILE OF A SOIL MICROBIAL COMMUNITY. Appl. Environ. Microbiol. 60, 2421-2430.

Rasmussen, J., Gjettermann, B., Eriksen, J., Jensen, E.S., Hogh-Jensen, H., 2008. Fate of (15)N and (14)C from labelled plant material: Recovery in perennial ryegrass-clover mixtures and in pore water of the sward. Soil Biology & Biochemistry 40, 3031-3039.

Rousk, J., Baath, E., 2011. Growth of saprotrophic fungi and bacteria in soil. FEMS microbiology ecology 78, 17-30.

43

Rubaek, G.H., 2008. Long-term effects of liming and phosphorus fertilisation on soil properties. In: Christensen, B.T., Petersen, J., Schact, M. (Eds.), Long-term field experiments - a unique research platform: Proceedings of NJF Seminar 407. Aarhus University, Denmark, Askov Experimental Station and Sandbjerg Estate, Denmark.

Schimel, J.P., Bennett, J., 2004. Nitrogen mineralization: Challenges of a changing paradigm. Ecology 85, 591-602.

Schmidt, M.W.I., Torn, M.S., Abiven, S., Dittmar, T., Guggenberger, G., Janssens, I.A., Kleber, M., Kogel-Knabner, I., Lehmann, J., Manning, D.A.C., Nannipieri, P., Rasse, D.P., Weiner, S., Trumbore, S.E., 2011. Persistence of soil organic matter as an ecosystem property. Nature 478, 49-56.

Schneewind, O., Missiakas, D.M., 2012. Protein secretion and surface display in Gram-positive bacteria. Philosophical Transactions of the Royal Society B-Biological Sciences 367, 1123-1139.

Simelyte, E., Rimpilainen, M., Zhang, X., Toivanen, P., 2003. Role of peptidoglycan subtypes in the pathogenesis of bacterial cell wall arthritis. Annals of the Rheumatic Diseases 62, 976-982.

Sinsabaugh, R.L., Lauber, C.L., Weintraub, M.N., Ahmed, B., Allison, S.D., Crenshaw, C., Contosta, A.R., Cusack, D., Frey, S., Gallo, M.E., Gartner, T.B., Hobbie, S.E., Holland, K., Keeler, B.L., Powers, J.S., Stursova, M., Takacs-Vesbach, C., Waldrop, M.P., Wallenstein, M.D., Zak, D.R., Zeglin, L.H., 2008. Stoichiometry of soil enzyme activity at global scale. Ecology Letters 11, 1252-1264.

Soil Survey Staff, 1999. Soil Taxonomy: A Basic System of Soil Classification for Making and Interpreting Soil Surveys. Natural Resources Conservation Service, United States Department of Agriculture, Washington D.C., USA.

Stromberger, M.E., Keith, A.M., Schmidt, O., 2012. Distinct microbial and faunal communities and translocated carbon in Lumbricus terrestris drilospheres. Soil Biology & Biochemistry 46, 155-162.

Vollmer, W., Blanot, D., de Pedro, M.A., 2008. Peptidoglycan structure and architecture. Fems Microbiology Reviews 32, 149-167.

Vranova, V., Rejsek, K., Formanek, P., 2013. Proteolytic activity in soil: A review. Applied Soil Ecology 70, 23-32.

Waksman, S.A., 1932. Principles of Soil Microbiology. The Williams & Wilkins Company, Baltimore.

Warren, C.R., 2012. Post-uptake metabolism affects quantification of amino acid uptake. New Phytologist 193, 522-531.

Warren, C.R., 2014. Organic N molecules in the soil solution: what is known, what is unknown and the path forwards. Plant and Soil 375, 1-19.

Wichern, F., Eberhardt, E., Mayer, J., Joergensen, R.G., Muller, T., 2008. Nitrogen rhizodeposition in agricultural crops: Methods, estimates and future prospects. Soil Biology & Biochemistry 40, 30-48.

Wilkinson, A., Hill, P.W., Farrar, J.F., Jones, D.L., Bardgett, R.D., 2014. Rapid microbial uptake and mineralization of amino acids and peptides along a grassland productivity gradient. Soil Biology & Biochemistry 72, 75-83.

Yarnes, C.T., Herszage, J., 2017. The relative influence of derivatization and normalization procedures on the compound-specific stable isotope analysis of nitrogen in amino acids. Rapid Communications in Mass Spectrometry 31, 693-704.

44

7. Appendices

45

Paper 1

The influence of hydrolysis and derivatization on the determination of amino acid content and isotopic

ratios in dual-labeled (13C, 15N) white clover

Kirsten Lønne Enggrob, Thomas Larsen, Mogens Larsen, Lars Elsgaard, Jim Rasmussen

Puplished in Rapid Comunication in Mass Spectrometry. DOI: 10.1002/rcm.8300

Received: 9 July 2018 Revised: 26 September 2018 Accepted: 27 September 2018

DOI: 10.1002/rcm.8300

R E S E A R CH AR T I C L E

The influence of hydrolysis and derivatization on thedetermination of amino acid content and isotopic ratios indual‐labeled (13C, 15N) white clover

Kirsten Lønne Enggrob1 | Thomas Larsen2,3 | Mogens Larsen4 | Lars Elsgaard1 |

Jim Rasmussen1

1Department of Agroecology, Aarhus

University, Foulum, Denmark

2Max Planck Institute for the Science of

Human History, Kahlaische Str. 10, 07745

Jena, Germany

3Leibniz‐Laboratory for Radiometric Dating

and Stable Isotope Research, Christian‐Albrechts Universität zu Kiel, Kiel, Germany

4Department of Animal Science, Aarhus

University, Foulum, Denmark

Correspondence

J. Rasmussen, Department of Agroecology,

Aarhus University, Foulum, Denmark.

Email: [email protected]

Funding information

Teknologi og Produktion, Det Frie

Forskningsråd, Grant/Award Number: 1335‐00760B; The Independent Research Fund

Denmark – Technology and Production

Rapid Commun Mass Spectrom. 2019;33:21–30.

Rationale: The cycling of peptide‐ and protein‐bound amino acids (AAs) is

important for studying the rate‐limiting steps in soil nitrogen (N) turnover. A strong

tool is stable C and N isotopes used in combination with compound‐specific isotope

analysis (CSIA), where a prerequisite for analysis is appropriate methods for peptide

and protein hydrolysis and appropriate methods for derivatization of AAs for analysis

by gas chromatography (GC).

Methods: We examined the efficiency of a standard acidic hydrolysis (6M HCl, 20 h

at 110°C) and a fast acidic hydrolysis (6M HCl, 70min at 150°C) on the recovery of

AAs from a protein standard (bovine serum albumin). The best methods were used

on dual‐labeled (13C and 15N) clover shoot and root juice, divided into four molecular

weight (Mw) size fractions. We used NAIP (N‐acetyl isopropyl esterification)

derivatization for GC/combustion‐isotope ratio mass spectrometry (C‐IRMS) analysis

of AA standards.

Results: The NAIP derivatization gave very low limits of detection (LODs) (< 2 pmol)

and limits of quantification (LOQs) ranging from 0.55 to 4.89 pmol. Comparing the

concentrations of individual AAs in hydrolyzed versus unhydrolyzed clover juice

samples of the low Mw size fraction (<1 kDa) showed a significant decline in

concentration (p <0.03) for seven AAs after hydrolysis. Despite the decline in AA

concentration, we found a linear connection between the obtained atomic fraction

(13C/total carbon and 15N/total nitrogen) for individual AAs of hydrolyzed versus

unhydrolyzed samples.

Conclusions: The methodology distinguished differences in atomic fractions across

AAs, in individual AAs in Mw size fractions, and between shoot and root samples of

experimentally labeled white clover. Specifically, the method separated L‐glutamate

(Glu) and glutamine (Gln). Thus, for a broader use in plant and soil ecology, we present

an optimized methodology for GC/C‐IRMS analysis of AAs from organic nitrogen

samples enriched with 13C and 15N – AA stable isotope probing (SIP).

1 | INTRODUCTION

Nitrogen (N) is a required nutrient for all lifeforms and a building block

for the backbone of amino acids (AAs). In studies of plant–soil

wileyonlinelibrary

interactions, AAs have gained considerable interest as plant nutrients,1

as components of root exudates,2 and as key compounds for under-

standing key fluxes through soil organic N pools.3,4 A strong tool to

study these issues is stable isotope labeling, where isotopically

© 2018 John Wiley & Sons, Ltd..com/journal/rcm 21

22 ENGGROB ET AL.

enriched AAs are used to trace the fate of single or a few AAs. Such

studies have, for example, documented direct acquisition of asparagine

by white clover,5,6 direct uptake of glycine in a number of crop species,7

and direct uptake of a mixture of glycine, valine, tyrosine and lysine by

plantain.8 However, it is generally acknowledged that AAs and small

peptides (<1 kDa) have a rapid turnover in soil3,9 and this step therefore

does not constitute the rate‐limiting step of the turnover of organic N

in plant–soil N cycling. However, progress in characterizing the cycling

and role of more complex organic N molecules, such as AAs bound in

larger peptides and proteins (>1 kDa), has been hampered by two

methodological issues. One is the need for testing the optimal

hydrolysis of peptides and proteins, and the other is separating as many

isotopically labeled AAs as possible.

To analyze proteinogenic AAs in plant and soil samples, it is

necessary to perform several analytical steps. The first step is breaking

the peptide bonds between individual AAs, which is commonly done

by acidic hydrolysis.10,11 The second step is purifying the hydrolysates

and isolating the AA fraction using a cation‐exchange column. For gas

chromatographic separation, the third and final step is making the AAs

more volatile by derivatizing them.

Several protocols have been used for acid hydrolysis, which

basically differ in the type and strength of acids, reaction temperature

and incubation time.12,13 Samples can be incubated in heating blocks,

ovens or microwave ovens. Joergensen et al14 reported similar results

when comparing microwave oven incubation for between 10 and

30min and oven incubation for 20 h. The benefit of the microwave

oven is the reduction in incubation time. The hydrolysis of protein

bonds of the hydrophobic AAs, valine, isoleucine and leucine, may

require an extended hydrolysis time up to 72 h.15 However, hydrolysis

is a harsh treatment that may also destroy AAs. A typical result is

deamination of asparagine and glutamine into aspartate and glutamate,

respectively, and the destruction of tryptophan and cysteine.15 In

addition, residual oxygen left in the hydrolysis reaction vessel may

induce thermal breakdown, resulting in reduced recovery of the

hydroxyl‐ and C‐heavy AAs, such as serine, threonine, tyrosine and

methionine.12 Thus, it is challenging to balance the release of strongly

bound AAs against the loss of labile AAs. We here selected two

methods for investigation, differing in temperature and hydrolysis

length, reported in recent literature.13,16

Theanalysisof labeledAAs isperformedbycompound‐specificstable

isotope analysis (CSIA), using either gas chromatography/combustion‐

FIGURE 1 Step‐by‐step structural information of the two amino acid derisopropyl esterification (B)

isotope ratio mass spectrometry (GC/C‐IRMS) or liquid chromatography

(LC) coupled to either IRMS or to time‐of‐flight mass spectrometry

(TOF‐MS). GC/C‐IRMS analysis requires the derivatization of AAs to

make them volatile, but the derivatization step introduces extra carbon

atoms to the AAs and thereby induces a potential error in measurement

of δ13C values,which have to be adjusted by correction factors calculated

for each individual AA.17,18WehereusedGC/C‐IRMS, since it is themost

widely usedmethod to rapidly separate multiple AAs.

In selecting the most appropriate derivatization method for

GC/C‐IRMS, several issues must be taken into account.19,20 These

include the number of C atoms introduced, the stability of the

derivative, the ease of handling and use of hazardous reagents, the

effect of the derivative on the durability of the combustion oven,

and the type of study in terms of natural isotope abundance or

experimental labeling. Corr et al19 compared seven derivatization

methods, two of which they detailed in a subsequent study,21

namely N‐acetyl methyl esterification (NACME) and N‐acetyl

isopropyl esterification (NAIP), as depicted in Figure 1. Corr et al21

recommended the NACME method as it only introduces the small

number of additional C atoms and had the smallest errors associated

with δ13C value determinations. Yet, a disadvantage of NACME

compared with NAIP derivatization22 is that it is harder to obtain a

good GC baseline separation.13,23 The NAIP method has the

disadvantage that it introduces five C atoms (Figure 1) and is thus

associated with a potentially greater error on the measured δ13C

values; this may be of particular concern in natural abundance

studies. However, when working with experimental labeling the main

concern is not related to (relatively small) errors on the measured

δ13C values, but rather to kinetic isotope effects,18 arising from the

faster reaction of lighter than heavier atoms. Thus, for experimental

isotope labeling studies, the most important criterion is choosing a

method with superior baseline separation. For this reason, we used

the NAIP derivatization method.

The aim of this study was to refine and test the GC/C‐IRMS

analysis of labeled AAs, both free and bound in peptides and proteins,

separating as many AAs as possible to be used in experimental labeling

studies. We compared the recovery of AAs from a standard protein

using two common acid hydrolysis protocols: standard hydrolysis in

6N HCl for 20 h at 110°C12 and fast hydrolysis in 6N HCl for

70min at 150°C.13 Finally, we then used the best methods for acid

hydrolysis on dual‐labeled (13C and 15N) clover shoot and root juice,

ivatization methods, i.e. N‐acetyl methyl esterification (A) and N‐acetyl

ENGGROB ET AL. 23

divided into four molecular weight (Mw) size fractions, to test the

power of the methods in separating the AA and then measuring the

isotopic enrichment of free and bound AAs in the clover.

2 | EXPERIMENTAL

2.1 | Amino acid preparation, derivatization andanalysis

2.1.1 | Standards and reagent

Twenty‐one individual unlabeled amino acid standards were purchased,

namely L‐alanine (Ala), L‐valine (Val), glycine (Gly), L‐leucine (Leu),

L‐isoleucine (Ile), L‐proline (Pro), L‐threonine (Thr), aspargine (Asn),

L‐aspartate (Asp), L‐serine (Ser), L‐glutamate (Glu), glutamine (Gln),

methionine (Met), L‐phenylalanine (Phe), L‐cysteine (Cys), hydroxy‐L‐

proline (Hyp), tryptophan (Trp), L‐tyrosine (Tyr), L‐lysine (Lys) and

norvaline (Avl) from Sigma‐Aldrich (St Louis, MO, USA) and D‐(−)‐

norleucine (Nle) from Alfa Aesar (Thermo Fisher Scientific, Heysham,

UK). D‐(−)‐Norleucine and norvaline were used as internal standards

(ISTDs). All solvents used were of HPLC grade and purchased from

VWR International (Herlev, Denmark).

2.1.2 | Acid hydrolysis

In order to release AAs from protein and peptide bonds, the samples

were exposed to a standard acidic hydrolysis16 and a fast acidic

hydrolysis.13 The two methods were tested (with three replicates

of the entire procedure) on 10 μL of a bovine serum albumin (BSA)

protein standard (200mgmL−1) in 16 × 100 soda‐lime disposable cul-

ture tubes (Duran Group, Mainz, Germany). BSA was chosen as a

protein standard to test the influences of the hydrolysis because it

is a relatively small, water‐soluble, standardized protein, and there-

fore directly comparable with the easily degradable DON (dissolved

organic nitrogen) compounds found in the juice fraction of the white

clover. For both methods, 1 mL 6M HCl was added to the sample

under an N2 atmosphere to eliminate oxygen and the culture tubes

were sealed to prevent oxygen penetration. In the standard

hydrolysis procedure, the samples were heated to 110°C for 20 h

using aluminum blocks, while, in the fast hydrolysis, samples were

heated to 150°C for 70min using aluminum blocks. After hydrolysis,

the AAs were purified and derivatized before GC/C‐IRMS analysis as

described below.

2.1.3 | Sample purification

Prior to derivatization, the hydrolyzed samples were purified, using a

modified version of the method of Amelung and Zhang.10 Solids and

lipophilic compounds were removed from the hydrolysate by adding

2mL n‐hexane/dichloromethane (DCM) (6:5, v/v), vortexing for 30 s

and centrifugation at 1000 g for 2min. The aqueous phase was then

transferred through a Pasteur pipette lined with glass wool into a

new glass tube, followed by 2 × 0.5mL 0.1M HCl to rinse the

pipette. An ISTD was then added to the aqueous sample, which

was diluted to 8mL with MilliQ water (Synergy® system, Millipore,

Molsheim, France) and transferred to a polypropylene column (10mL)

containing 1 g Dowex 50W‐X8 cation‐exchange resin (analytical

grade, 100–200 mesh, hydrogen form; Bio‐Rad Laboratory Inc.,

Hercules, CA, USA) and rinsed with 3 × 3mL 0.1M oxalic acid,

3 × 3mL 0.01M HCl and 3 × 3mL MilliQ water. Finally, the AAs

were eluted with 3 × 1mL 2.5M ammonium hydroxide solution

(Merck, Darmstadt, Germany). To avoid heating of the samples prior

to derivatization, the ammonium hydroxide solution was evaporated

by freeze centrifugation.

2.1.4 | Derivatization

The NAIP derivatization method was modified from Corr et al and

Larsen et al21,22 and tested on individual AA standards to assess and

optimize the stabilization of the derivatives. First, solutions of standard

AAs were dried as described above by freeze centrifugation. Next, the

dried AAs were added to 0.70mL acidified isopropanol, prepared by

transferring 1:4 cooled acetyl chloride to cooled isopropanol. The

reaction vessels were then flushed with N2, closed and placed in a

heating block at 75°C for 60min, followed by cooling to room

temperature (20°C) in aluminum blocks. After cooling of the sample, the

excess solvent was evaporated under a gentle N2 flow at 60°C. Any

residual acetate left over from the propylation was removed by the

addition and evaporation of 0.5mL DCM to ensure proper acetylation.

The derivatized AAs were acetylated by adding 0.75mL of an acetylation

mixture containing acetic anhydride, triethylamine and acetone (1:2:5,

v/v/v, 10min, 60°C). Excess solvent was evaporated under a gentle N2

flow for 8min at 60°C. Salts and precipitates were removed by adding

ethyl acetate and a saturated NaCl solution (1:1, v/v) and centrifuging

for 2min at 1500 g. After centrifugation, the organic phase was

transferred to a new test tube and gently evaporated under N2 flow at

room temperature. Residual water was removed by the addition and

evaporation of 0.5mL DCM. Finally, the derivatized AAs were dissolved

in ethyl acetate and transferred to GC glass vials.

2.1.5 | GC/C‐IRMS analysis

GC analyses were performed on a Trace GC Ultra gas chromatograph

interfaced to a TriPlus autosampler (both from Thermo Scientific,

Hvidovre, Denmark). The derivatives were separated with a VF‐23m

capillary column (60m × 0.25mm i.d. × 0.25 μm film thickness; Agilent

Technologies, Amstelveen, The Netherlands). The inlet was operated

with a temperature of 250°C in splitless mode, with a helium column

flow rate of 1.4mLmin−1. The temperature gradient was as follows:

initial oven temperature set to 90°C for 1min, raised to 120°C at

15°C·min−1, then raised to 250°C at 3°C·min−1 and held at 250°C for

5–45min, depending on the AA retention time. The gas chromatograph

was coupled via a combustion reactor (GC IsoLink, Thermo Scientific),

oxidation at 1000°C, to a Delta V Plus isotope ratio mass spectrometer

(Thermo Scientific). All mass spectrometer related parameters were

controlled by Isodat version 3.0 software (Thermo Scientific). All δ13C

values are reported relative to the Vienna PeeDee Belemnite (VPDB)

international isotope standard. All δ15N values are reported relative to

atmospheric N2. A standard curve based on analyses of Asn with

24 ENGGROB ET AL.

increasing percentages of dual‐labeled Asn (13C4,15N2) showed a

strong linearity of all δ13C and δ15N values, with increasing amounts

of dual‐labeled Asn (13C4,15N2), with a coefficient of determination of

R2 = 0.984 and R2 = 0.982, respectively. The AAs were identified by

the retention time of standards and the concentration calculated

relative to individual standard curves.

2.1.6 | Correction factor and kinetic isotope effect(KIE)

The obtained δ13C values were converted into atomic fractions (AFs)

using the following equation:17

AF ¼VPDB 10−3δ13Cþ 1

� �

1þ VPDB 10−3δ13Cþ 1� � (1)

To adjust for the carbon added during the derivatization a

correction factor, AFL, was calculated for each AA according to:17

AFL ¼ yþ xð ÞAFD − xAFUy

(2)

where y is the number of added C atoms, x is the number of C atoms

in the AA, AFD is the δ13C atomic fraction of the derivatizied AA, and

AFU is the δ13C atomic fraction of the underivatizied AA.

The AF of the enriched AA can now be calculated as:

AFAA ¼ xþ yð ÞAFAAD − yAFLx

(3)

During derivatization, isotopic fractionation occurs when bonds

involving heavier isotopes are broken, thereby creating a kinetic

isotopic effect (KIE).18 The KIE can be calculated according to the

following equation:

KIE ¼ − 1þ AFL xþ yð Þ1000z

� �(4)

where z is the number of functional groups available for acetylation.

2.2 | Determination of limit of detection (LOD) andlimit of quantification (LOQ)

Typically, the LOD and LOQ are calculated by multiplying the signal‐

to‐noise ratio by 3 for the LOD, and by 10 for the LOQ. This method,

unfortunately, does not take account of the stability of the signal or

background noise of the analysis. We applied the more stringent

three‐step method suggested by Harris24 and first established a

calibration curve for the AAs of interest, spanning from 0.75 to

7.5mM, carried out in triplicate. Secondly, eight standard mixtures,

containing all the AAs of interest, were analyzed at an equal low

concentration of 0.75mM to reflect the stability of the analyses.

Thirdly, eight blind samples were analyzed to reflect the stability of

the background noise. The lowest detectable signal (ydl), LOD and

LOQ were determined using the following equations:24

ydl ¼ yblank þ 3·s (5)

LOD ¼ 3·sm

(6)

LOQ ¼ 10·sm

(7)

where yblank is the average of the signal obtained from blind

samples, s is the standard deviation from the analysis of standards

in mixtures, and m is the slope of the standard curve determined

for individual AAs.

2.3 | Test of analytical procedure on AAs in whiteclover

2.3.1 | Cultivating and labeling of white clover

White clover was sown in pots containing sand and irrigated once a

week for the first 8weeks, twice a week for the next 5weeks and

three times a week for the last 6weeks. The water used for irrigation

contained 3 at% 15N‐PK fertilizer. From week 8 and onwards, the

clover was labeled with 13CO2 prior to irrigation as previously

described.25 Briefly, 5mL of a saturated solution of sodium

bicarbonate (13C 99at%) dissolved in 1M NaOH was placed in a beaker

in each pot of clover. The pot was then covered with a transparent

plastic bag and 5mL 2M HCl was added to the 13C‐bicarbonate

solution, hence releasing 13CO2. After 2 h the labeling was stopped by

removing the plastic bags.

2.3.2 | Harvest and preparation of juice samples

White clover was harvested after 19weeks of growth and divided into

shoot and root biomass by cutting the shoots at the sand surface (i.e.

everything containing chlorophyll was considered a part of the

shoots). The shoots were rinsed with water and dabbed dry with a

towel. The roots were carefully separated from the sand and rinsed

in the same manner.

Juice from shoots and roots was extracted by screw pressing.26

The resulting juice was centrifuged at 10.000 g for 30min followed

by filtration through a 0.45 μm syringe CA filter (VWR International,

Søborg, Denmark).

Prior to determination of AAs, the juice was subjected to

molecular weight (Mw) size fractionation to reduce interference from

larger molecules. Molecular weight size fractionation was performed

by a modified ultrafiltration method27 using 20‐mL centrifugal filter

tubes (Macrosep® Advance, Pall Corporation, Ann Arbor, MI, USA)

equipped with permeable membranes of pore sizes 1, 10, and

100 kDa. First, each centrifugal tube was washed with 0.1 N HCl

and rinsed with MilliQ water three times before use. Then, the juice

was added to the 100 kDa filter tubes and centrifuged at 5000 g for

up to 180min. After centrifugation of the juice sample, the filter was

washed twice by adding 5.0mL MilliQ water to the filter and

centrifuged at 5000 g for up to 60min. The filtrate that had passed

through the 100 kDa filter was collected and the process was

repeated for the 10 kDa and 1 kDa filters. The residues that had

not passed through the filters were washed out by shaking the filter

ENGGROB ET AL. 25

three times in 5.0mL MilliQ water. The fractionation procedure

eventually resulted in four size fractions of juice from both root

and shoot biomass, i.e. >100 kDa, 10–100 kDa, 1–10 kDa

and <1 kDa. All fractionated samples were stored frozen until

derivatization and analysis as described above.

2.4 | Data analysis and statistics

The influence of the hydrolysis and the Mw size distribution on the

atomic fraction was tested with a linear mixed‐effects model using

the statistical analysis program R (version 3.3.1; R Core Team,

2016).28 For statistical analysis, the datasets were divided into sub-

sets; each subset was tested for normal distribution by the Shapiro–

Wilk normality test and for homogeneity of variances by the Bartlett

test. Models describing each subset were tested by either an analysis

of variance (ANOVA) test or a pairwise test.

TABLE 1 Obtained retention times (s) of AAs after NAIP derivati-zation (ISTD= internal standard)

NAIP AAIndividualstandards (s)

Mixedstandards (s)

1 Ala 1232.3 1230

2 Val 1360.2 1360

3 AvlISTD 1438.5 1444

3 | RESULTS

3.1 | Efficiency of the hydrolysis

BSA standard solutions, subjected to the two acid hydrolysis methods

and GC/C‐IRMS analysis, revealed between 21 and 23 peaks, where

12 could be assigned to a specific AA or group of AAs, including one

peak from the ISTD. The total area ± SE of both identified and

unidentified AAs was significantly higher (P <0.001) with the standard

acid hydrolysis (1083 ± 104 Vs) than with the fast acid hydrolysis (636

Vs ± 55 Vs). The areas of identified AAs were 685 ± 46 Vs and

502 ± 39 Vs for the standard and fast hydrolysis, respectively. We

then calculated that the total recoveries, based on the identified

AAs, were 35.6% (±1.3%) and 31.8% (±1.5%) of the total mass of the

BSA standard for the standard and fast hydrolysis, respectively. The

recovery was significantly higher (P <0.04) with the standard acid

hydrolysis for 7 of the 11 identified AAs, whereas the fast acid

hydrolysis was comparable with the standard acid hydrolysis for the

remaining 4 identified AAs (Figure 2).

FIGURE 2 Recovery of 11 identified amino acids or groups of aminoacids from bovine serum albumin (BSA) standard solutions afterstandard acid hydrolysis (“black bars”) and fast acid hydrolysis (“greybars”). Data are mean ± standard error (n = 3); Asterisks indicatesignificant differences in the recovery among the hydrolysis methods

3.2 | Efficiency of the NAIP derivatization methodfor separating AAs

Individual AA standard solutions were derivatized to optimize the

gradient of the GC column for maximum separation of the AA.

We were able to obtain separation and stable retention time for

all 21 AAs in single standards with the NAIP method (Table 1).

Ala eluted first with a retention time of 1232 s and Lys eluted last

with a retention time of 5181 s (Figure S1, supporting information).

However, in mixed standards, Met and Cys disappeared, whereas

Pro + Thr (Pro/Thr) eluted simultaneously with retention times of

1886 s and Asn + Asp (Asx) eluted simultaneously with retention

times of 2000 s. In addition, the ISTD Avl eluted simultaneously

with Gly; we, therefore, decided to use Nle as the ISTD for the

sample analysis.

3.3 | LODs and LOQs for the C‐IRMS analysis

Using the three‐step methodology suggested by Harris,24 we deter-

mined the LODs and LOQs for 15 AAs or co‐eluting AAs using

NAIP derivatization for C‐IRMS analysis (Table 2). The LODs

ranged from 0.17 pmol for Leu to 1.47 pmol for Lys. The LOQs

ranged from 0.55 pmol for Ser to 4.89 pmol for Lys. The LOD

was below 1 pmol for seven AAs and below 2 pmol for another

four AAs.

4 Gly 1443.8 1444

5 Leu 1465.3 1464

6 Ile 1477.6 1478

7 NleISTD 1556.2 1554

8 Pro 1874 1886

9 Thr 1883 1886

10 Asn 1996.8 2000

11 Asp 2001 2000

12 Ser 2058 2051

13 Glu 2197 2190

14 Gln 2295.9 2297

15 Met 2347.3 ‐

16 Phe 2434 2436

17 Cys 2443–2445 ‐

18 Hyp 2523.9 2523

19 Trp 2840.3 2979

20 Tyr 3553.6 3589

21 Lys 5188.6 5181

TABLE 2 The standard equation, the LOD and LOQ, lowest detectable signal (ydl) and13C AF stability for 15 AAs or co‐eluting AAs after

NAIP derivatization of non‐enriched standard mixtures for C‐IRMS analysis. (ISTD= internal standard). AF stability of 13C is listed as the lowestconcentration with stable AF (p >0.05)

AAMolar mass(g/mol) Slope Intercept R2

ydl(signal intensity)

LOD C(pmol)

LOQ C(pmol)

AF stability(pmol)

Ala 89.09 7.714 3.284 0.9527 2.117 0.23 0.77 0.74

Val 117.15 2.949 1.039 0.9599 1.213 0.19 0.63 1.23

Gly 75.07 5.512 3.122 0.9489 2.100 0.25 0.85 0.74

Leu 131.17 9.443 3.733 0.9697 2.544 0.17 0.58 2.55

Ile 131.17 3.957 ‐0.473 0.9239 1.828 0.29 0.96 2.48

Nle ISTD 131.17 9.653 3.898 0.9691 3.947 0.34 1.12 0.75

Pro/Thr 115.13 6.224 9.858 0.9459 6.900 0.92 3.07 1.51

Asx 132.12 5.558 5.792 0.9691 3.138 0.28 0.92 2.52

Ser 105.09 7.081 −1.144 0.8589 3.013 0.16 0.55 0.74

Glu 147.13 4.464 1.485 0.9439 5.734 0.76 2.53 2.5

Gln 146.14 14.801 7.417 0.9731 8.358 0.35 1.18 0.75

Phe 165.19 7.813 6.359 0.9574 6.631 0.36 1.19 2.51

Hyp 131.13 4.662 7.828 0.8245 5.641 0.64 2.12 0.74

Tyr 181.19 5.231 7.937 0.9194 4.324 0.60 1.98 1.24

Lys 182.65 9.195 0.167 0.9081 18.88 1.47 4.89 2.5

26 ENGGROB ET AL.

3.4 | AAs in white clover samples

3.4.1 | Low Mw size fraction

The low Mw size fraction (<1 kDa) was analyzed both for free AAs and

for bound AAs after hydrolysis in order to release the AAs in small

peptides. We found that the unhydrolyzed samples contained 11 single

eluting AAs (Ala, Val, Gly, Leu, Ile, Ser, Glu, Gln, Phe, Tyr and Lys),

and two pairs of co‐eluting AAs: Pro/Thr, and Asx. The hydrolyzed

sample contained the same eluting AAs, except for Glu and Tyr, that

were completely lost. Furthermore, the amount of seven AAs (Asx,

Gln, Ser, Pro/Thr, Ala and Val) was significantly lower (P <0.03) after

acid hydrolysis, whereas four AAs (Phe, Leu, Ile and Gly) were

unaffected (P >0.09) by the hydrolysis (Figure 3).

FIGURE 3 The content of free amino acids (AAs) versus bound AAs in thusing the standard hydrolysis method to retrieve bound AAs

3.4.2 | High Mw size fraction

The three larger Mw size fractions (>1 kDa) were all subject to

hydrolysis prior to derivatization, in order to release all the

peptide‐ and protein‐bound AAs (Table S1, supporting information).

For the 1–10 kDa size fraction we detected, in both shoot and

root juice, six single eluting AAs (Ala, Val, Gly, Leu, Ile and Ser)

and one pair of co‐eluting AAs: Pro/Thr; with an additional co‐

eluting pair, Asx, found in the shoot juice. For the 10–100 kDa size

fraction, both shoot and root juice gave four single eluting AAs

(Val, Leu, Ile and Phe). For the >100 kDa size fraction, both shoot

and root juice gave 10 single eluting AAs (Ala, Val, Gly, Leu, Ile,

Ser, Gln, Phe, Tyr and Lys), and two pairs of co‐eluting AAs:

Pro/Thr, and Asx.

e <1 kDa Mw size fraction for (A) root juice and (B) shoot juice (n = 3)

ENGGROB ET AL. 27

3.5 | Enrichment of AAs in clover samples

The correction factor and KIE were calculated based on obtained

δ13C data from analysis of underivatizied AA standards on a Flash

elemental analyzer (Thermo Scientific) and GC/C‐IRMS analysis of

derivatized AA standards; the KIE values were 0.13 ± 0.06‰

(Table 3). The correction factor reflects the contribution of 13C from

the added C atoms during derivatization. The correction factor must

therefore be significantly lower than the 13C atomic fraction of the

samples for the results to be reliable. In the present study, the 13C

atomic fraction for AAs was five to ten times higher than the

correction factor, substantiating that, for 13C‐enriched samples, the

number of C atoms added with the NAIP derivatization is not an

issue of concern.

In the low Mw size fraction (<1 kDa) we found a linear relation-

ship, with a slope close to 1, between the 13C and 15N atomic fractions

for individual AAs as both free and bound for both shoot and root

juice (Figures 4A–4D).

The statistical comparison of the 13C atomic fractions in shoot and

root from unhydrolyzed versus hydrolyzed samples showed that there

was no significant difference between the atomic fraction obtained

from free and bound AAs in the <1 kDa fraction for both shoot

(p = 0.73) and root (p = 0.84). For the 15N atomic fraction in shoot

and root juice from hydrolyzed versus unhydrolyzed samples there

was no significant difference for Ile (p = 0.82) and Phe (p = 0.19) in

shoot juice and for Val (p = 0.15), Gly (p = 0.08), Leu (p = 0.27), Ser

(p = 0.41), Phe (p = 0.31) and Lys (p = 0.16) in root juice. For the three

larger Mw size fractions (>1 kDa) there were significant differences

between the atomic fraction obtained across Mw size fractions for

both 15N and 13C, except for Leu and Ser for 13C in shoot juice, and

TABLE 3 Correction factors and KIEs for individual AAs for theIRMS determination of 13C

AA Correction factor KIE [‰]

Ala 0.0106229 0.08

Val 0.0105033 0.11

Gly 0.0106524 0.07

Leu 0.0104997 0.12

Ile 0.0104222 0.11

Nle 0.0105710 0.12

Pro 0.0104943 0.10

Thr 0.0106344 0.13

Asn 0.0106472 0.23

Asp 0.0106316 0.13

Ser 0.0105209 0.12

Glu 0.0109726 0.14

Gln 0.0105238 0.00

Phe 0.0103794 0.15

Hyp 0.0105120 0.14

Tyr 0.0102663 0.14

Lys 0.0105582 0.30

Average 0.0105537 0.13

SE 0.0001444 0.03

Gln, Tyr and Lys for 13C in root juice (Figure 5 and Table S2,

supporting information). The 13C atomic fraction of AAs in shoot juice

generally had a higher 13C enrichment than root juice compared across

Mw size fractions (Figure 6).

4 | DISCUSSION

4.1 | Influence of hydrolysis and derivatization onrecovery of AAs

The recoveries of BSA from the two acidic hydrolysis methods, i.e.

35.6% (±1.3%) for standard hydrolysis and 31.8% (±1.5%) for

fast hydrolysis, are in line with previously reported recoveries of

approximately 30%.12 Based on the total recovery and the recovery of

individual AAs in BSA standard solutions, the standard acid hydrolysis

was chosen for the further work.

The NAIP derivatization method could separate and identify 13

proteinogenic AAs, one synthetic AA (Nle) used as the ISTD, and

two pairs of co‐eluting AAs. Compared with other methods,19,22 we

were uniquely able to separate Gln and Glu in a mixed standard and

we were also able to obtain signals from Hyp and Trp and to separate

Pro and Thr. Both Corr et al19 and Larsen et al16 showed co‐elution of

these two AAs with the NACME derivatization method using VF‐

23ms GC and TG‐200MS GC columns, respectively. Wang et al29

recently separated NACME‐derivatized Pro and Thr using a VF‐

35ms GC column, but further studies are required to test whether this

GC column is suited for AAs enriched in heavy isotopes.

We obtained markedly lower LODs and LOQs than previously

reported for the GC/C‐IRMS analysis of AAs. Walsh et al20 reported

a LOQ of 10–50 pmol and Sessions30 stated that the instrument

sensitivity for GC/C‐IRMS typically lies between 0.1 and 10 nmol.

Our LOQs ranged from 0.55 to 4.89 pmol with the lowest concentrations

with stable AFs ranging from 0.74 to 2.55 pmol of the non‐enriched

standards. The low LODs and LOQ for C‐IRMS were achieved in part

because at least five C atoms were added during NAIP derivatization

(Figure 1). As demonstrated in this study, the dilution of the13C‐labeled atoms is a relatively low source of error compared with

analyzing peaks next to or below the LOD and LOQ. The improvement

in LODs and LOQs for C analysis by using NAIP does not apply to the

N analysis since there is no addition of N atoms during the

derivatization. The low LODs and LOQs of the NAIP derivatization

are particularly important when analyzing labeled natural samples with

low AA concentrations.

4.2 | Content and enrichment of AAs in Mw sizefractionated clover shoot and root juice

As stated in the literature,12,15 a decline in the concentration of Asn,

Gln, Ser, Thr and Tyr is expected due to hydrolysis, along with an

increase in Val, Ile and Leu. Yet, we were surprised to find that the

AA concentration from the <1 kDa Mw size fraction was generally

lower after hydrolysis than the concentrations of free AAs (i.e. in the

unhydrolyzed samples). This indicated that the hydrolysis affects the

composition of AAs much more than previously recognized.11 This

FIGURE 4 The atomic fraction in free amino acids (AAs) versus bound AAs in the <1 kDa Mw size fraction for (A) 13C in shoot juice, (B) 15N inshoot juice, (C) 13C in root juice, and (D) 15N in root juice of experimentally labeled white clover (n = 3)

FIGURE 5 Example of the 13C atomicfraction of amino acids (AAs) in white clovershoot juice for different Mw size fractions:free AAs (black circles), AAs bound in 1–10 kDa fraction (dark grey triangles), AAsbound in 10–100 kDa fraction (light greysquares), and AAs bound in >100 kDa fraction(white diamonds). For free AAs both Glu andGln were measured (Glu omitted in thisfigure), whereas in the hydrolyzed Mw sizefractions >1 kDa, Glu/Gln is reported as Glx.Table S2 (supporting information) gives 13Cand 15N atomic fractions for AAs in all Mwsize fractions from white clover shoot androot juice (n = 3). Asterisks indicate significantdifferences in the obtained AFs. Doubleasterisks indicate no normal distribution

28 ENGGROB ET AL.

FIGURE 6 Comparison of the 13C atomic fraction in root juiceversus shoot juice in the Mw size fractions >1 kDa (n = 3)

ENGGROB ET AL. 29

highlights the challenge of calculating the total amount of AAs, peptides

and proteins in natural samples and the need to adjust hydrolysis

procedures to the actual samples under investigation.10,12,14

Importantly, the atomic fraction of 13C proved to be independent of

the concentration and did not show the same tendency to decline

following the hydrolysis; meaning that even though AAs were lost

during the hydrolysis, the 13C enrichment was not affected.

Due to the influences of the hydrolysis and the fact that there is

no significant difference between the 13C atomic fraction from the

hydrolyzed and unhydrolyzed <1 kDa Mw size fraction, the >1 kDa

Mw size fraction was not included in the comparison of 13C atomic

fraction across Mw size fraction in Figure 5.

The atomic fraction of 13C in shoot and root showed similar

distributions, with a tendency to a higher relative abundance of 13C

in the shoots than in the roots. This is in agreement with the fact that13C was induced into the white clover through atmospheric labeling.

The atomic fraction of 15N in shoot showed a similar distribution

in all Mw size fractions except for Tyr and Lys, but the 15N atomic

fraction in roots was higher in the low than high Mw size fractions.

This aligns with the way of labeling, since 15N was added to the white

clover through irrigation.

5 | CONCLUSIONS

We investigated two hydrolysis methods for the GC/C‐IRMS analysis

of AAs. We found that the standard hydrolysis procedure had

significantly higher recovery of AAs from BSA standards than fast

hydrolysis. Using mixtures of AAs, we found that NAIP derivatization

resulted in a very low LODs and LOQs, due to the addition of extra

C atoms, with the latter ranging from 0.55 to 4.89 pmol for the 15

AAs determined. We tested the methodology (i.e. standard hydrolysis

and NAIP derivatization) on Mw size fractionated organic N in juice

from dual‐labeled white clover shoot and root and documented the

ability to distinguish differences in atomic fractions across AAs, in

individual AAs in four Mw size fractions, and between shoot and root

samples. Of particular interest, we showed that hydrolysis of the

smallest fraction (<1 kDa), to release AAs bound in small peptides,

caused a notable loss of AAs, actually resulting in lower concentrations

of most AAs in the hydrolyzed sample than of the free AAs in the

unhydrolyzed sample. The present study highlighted the importance

of determining the recovery efficiency of the applied hydrolysis

method and the great potential for using NAIP derivatization for

GC/C‐IRMS analysis of enriched (13C and 15N) samples. In contrast

to previous studies, the hydrolysis and NAIP procedures presently

described were specifically able to separate Gln and Glu.

ACKNOWLEDGEMENT

The studywas financially supported by The Independent Research Fund

Denmark – Technology and Production (Project No. 1335‐00760B).

ORCID

Kirsten Lønne Enggrob http://orcid.org/0000-0003-1667-1610

REFERENCES

1. Nasholm T, Kielland K, Ganeteg U. Uptake of organic nitrogen byplants. New Phytol. 2009;182(1):31‐48. https://doi.org/10.1111/j.1469‐8137.2008.02751.x

2. Lesuffleur F, Paynel F, Bataille M, Le Deunff E, Cliquet J. Root aminoacid exudation: measurement of high efflux rates of glycine and serinefrom six different plant species. Plant and Soil. 2007;294(1–2):235‐246.https://doi.org/10.1007/s11104‐007‐9249‐x

3. JanMT, Roberts P, Tonheim SK, Jones DL. Protein breakdown representsa major bottleneck in nitrogen cycling in grassland soils. Soil Biol Biochem.2009;41(11):2272‐2282. https://doi.org/10.1016/j.soilbio.2009.08.013

4. Jones DL, Kielland K. Amino acid, peptide and protein mineralizationdynamics in a taiga forest soil. Soil Biol Biochem. 2012;55:60‐69.https://doi.org/10.1016/j.soilbio.2012.06.005

5. Czaban W, Jamtgard S, Nasholm T, Rasmussen J, Nicolaisen M,Fomsgaard IS. Direct acquisition of organic N by white clover even inthe presence of inorganic N. Plant and Soil. 2016;407(1–2):91‐107.https://doi.org/10.1007/s11104‐016‐2896‐z

6. Czaban W, Rasmussen J, Laursen BB, et al. Multiple effects ofsecondary metabolites on amino acid cycling in white clover rhizo-sphere. Soil Biol Biochem. 2018;123:54‐63. https://doi.org/10.1016/j.soilbio.2018.04.012

7. Nasholm T, Huss‐Danell K, Hogberg P. Uptake of organic nitrogen inthe field by four agriculturally important plant species. Ecology.2000;81(4):1155‐1161. https://doi.org/10.1890/0012‐9658(2000)081[1155:UOONIT]2.0.CO;2

8. Sauheitl L, Glaser B, Weigelt A. Advantages of compound‐specificstable isotope measurements over bulk measurements in studies onplant uptake of intact amino acids. Rapid Commun Mass Spectrom.2009;23(20):3333‐3342. https://doi.org/10.1002/rcm.4255

9. Hill PW, Farrell M, Jones DL. Bigger may be better in soil N cycling:Does rapid acquisition of small l‐peptides by soil microbes dominatefluxes of protein‐derived N in soil? Soil Biol Biochem.2012;48:106‐112. https://doi.org/10.1016/j.soilbio.2012.01.023

10. Amelung W, Zhang X. Determination of amino acid enantiomers insoils. Soil Biol Biochem. 2001;33(4):553‐562. https://doi‐org.ez.statsbiblioteket.dk:12048/10.1016/S0038‐0717(00)00195‐4

11. Weiss M, Manneberg M, Juranville J, Lahm H, Fountoulakis M. Effectof the hydrolysis method on the determination of the amino acid com-position of proteins. J Chromatogr A. 1998;795(2):263‐275. https://doi.org/10.1016/S0021‐9673(97)00983‐7

30 ENGGROB ET AL.

12. Fountoulakis M, Lahm H. Hydrolysis and amino acid composition anal-ysis of proteins. J Chromatogr A. 1998;826(2):109‐134. https://doi.org/10.1016/S0021‐9673(98)00721‐3

13. Yarnes CT, Herszage J. The relative influence of derivatization and nor-malization procedures on the compound‐specific stable isotopeanalysis of nitrogen in amino acids. Rapid Commun Mass Spectrom.2017;31(8):693‐704. https://doi.org/10.1002/rcm.7832

14. Joergensen L, Thestrup H. Determination of Amino‐Acids in Biomassand Protein Samples by Microwave Hydrolysis and Ion‐ExchangeChromatography. J Chromatogr A. 1995;706(1–2):421‐428. https://doi.org/10.1016/0021‐9673(94)01107‐P

15. Roberts P, Jones DL. Critical evaluation of methods for determiningtotal protein in soil solution. Soil Biol Biochem. 2008;40(6):1485‐1495.https://doi.org/10.1016/j.soilbio.2008.01.001

16. Larsen T, Bach LT, Salvatteci R, et al. Assessing the potential ofamino acid C−13 patterns as a carbon source tracer in marine sedi-ments: effects of algal growth conditions and sedimentary diagenesis.Biogeosciences. 2015;12(16):4979‐4992. https://doi.org/10.5194/bg‐12‐4979‐2015

17. Tetens V, Kristensen N, Calder A. Measurement of C‐13 Enrichmentof Plasma Lactate by Gas Chromatography/isotope Ratio Mass‐Spectrometry. Anal Chem. 1995;67(5):858‐862. https://doi.org/10.1021/ac00101a011

18. Docherty G, Jones V, Evershed R. Practical and theoretical consider-ations in the gas chromatography/combustion/isotope ratio massspectrometry delta C‐13 analysis of small polyfunctional compounds.Rapid Commun Mass Spectrom. 2001;15(9):730‐738. https://doi.org/10.1002/rcm.270

19. Corr LT, Berstan R, Evershed RP. Optimisation of derivatisation proce-dures for the determination of delta C‐13 values of amino acids by gaschromatography/combustion/isotope ratio mass spectrometry. RapidCommun Mass Spectrom. 2007;21(23):L3759‐L3771. https://doi.org/10.1002/rcm.3252

20. Walsh RG, He S, Yarnes CT. Compound‐specific delta C‐13 and deltaN‐15 analysis of amino acids: a rapid, chloroformate‐basedmethod for ecological studies. Rapid Commun Mass Spectrom.2014;28(1):96‐108. https://doi.org/10.1002/rcm.6761

21. Corr LT, Berstan R, Evershed RP. Development of N‐acetyl methylester derivatives for the determination of delta(13) C values of aminoacids using gas chromatography‐combustion‐isotope ratio mass spec-trometry. Anal Chem. 2007;79(23):9082‐9090. https://doi.org/10.1021/ac071223b

22. LarsenT, Ventura M, Andersen N, O'Brien DM, Piatkowski U, McCarthyMD. Tracing Carbon Sources through Aquatic and Terrestrial FoodWebs Using Amino Acid Stable Isotope Fingerprinting. PLoS One.2013;8(9):e73441. https://doi.org/10.1371/journal.pone.0073441

23. Sabadel AJM, Woodward EMS, Van Hale R, Frew RD. Compound‐spe-cific isotope analysis of amino acids: A tool to unravel complexsymbiotic trophic relationships. Food Webs. 2016;6:9‐18. https://doi.org/10.1016/j.fooweb.2015.12.003

24. Harris DC (Ed). Quality Assurance and Calibration Methods. In: Quanti-tative Chemical Analysis. 7th ed. New York: W. H. Freeman andCompany; 2007:78‐95.

25. Rasmussen J, Gjettermann B, Eriksen J, Jensen ES, Hogh‐Jensen H.Fate of (15) N and (14) C from labelled plant material: Recovery inperennial ryegrass‐clover mixtures and in pore water of the sward. SoilBiol Biochem. 2008;40(12):3031‐3039. https://doi.org/10.1016/j.soilbio.2008.08.025

26. Colas D, Doumeng C, Pontalier PY, Rigal L. Green crop fractionation bytwin‐screw extrusion: Influence of the screw profile on alfalfa(Medicago sativa) dehydration and protein extraction. Chem Eng Pro-cess. 2013;72:1‐9. https://doi.org/10.1016/j.cep.2013.05.017

27. PourretO,Dia A, DavrancheM, GruauG,HeninO, AngeeM.Organo‐colloi-dal control on major‐ and trace‐element partitioning in shallowgroundwaters: Confronting ultrafiltration and modelling. Appl Geochem.2007;22(8):1568‐1582. https://doi.org/10.1016/j.apgeochem.2007.03.022

28. R Core Team. R: A language and environment for statistical computing.Vienna, Austria: R Foundation for Statistical Computing; 2016 Avail-able at https://www.R‐project.org/.

29. Wang YV, Wan AHL, Lock E, Andersen N, Winter‐Schuh C, Larsen T.Know your fish: A novel compound‐specific isotope approach for trac-ing wild and farmed salmon. Food Chem. 2018;256:380‐389. https://doi.org/10.1016/j.foodchem.2018.02.095

30. Sessions AL. Isotope‐ratio detection for gas chromatography. J Sep Sci.2006;29(12):1946‐1961. https://doi.org/10.1002/jssc.200600002

SUPPORTING INFORMATION

Additional supporting information may be found online in the

Supporting Information section at the end of the article.

How to cite this article: Enggrob KL, Larsen T, Larsen M,

Elsgaard L, Rasmussen J. The influence of hydrolysis and

derivatization on the determination of amino acid content

and isotopic ratios in dual‐labeled (13C, 15N) white clover.

Rapid Commun Mass Spectrom. 2019;33:21–30. https://doi.

org/10.1002/rcm.8300

56

Paper 2

Molecular size doesn't matter for turning over large organic N in soil

Kirsten Lønne Enggrob, Thomas Larsen, Jim Rasmussen

Prepared for sumbission to Nature

1  

Original paper Date of preparation: 30-01-2019 1 

Pages: 18 Figures: 5 Extended Data Figures: 4 Extended Data Tables: 5 2 

Molecular size doesn’t matter for turning over large organic N in soil 4 

Kirsten Lønne Enggrob,1 Thomas Larsen,2 and Jim Rasmussen1* 6 

1Department of Agroecology, Faculty of Science and Technology, Aarhus University, Denmark. 8 

2Department of Archaeology, Max Planck Institute for the Science of Human History, Jena, 9 

Germany. 10 

*Corresponding author: Post Box 50, 8830 Tjele, Denmark, [email protected] 11 

12 

13 

Summary 14 

Global nitrogen use efficiency urgently needs to increase to reduce environmental emissions. This 15 

necessitates better predictions of the cycling of organic nitrogen from plants into stable soil organic 16 

carbon and nitrogen pools. We investigated the fate of peptide-size and protein-size fractions in 17 

soils from two long-term field experiments markedly differing in condition for microorganisms. 18 

Contrary to the present paradigm, we found for both soils that exo-enzymatic depolymerization was 19 

not per se the rate-limiting step in the turnover of these compounds neither was protection via 20 

strong sorption to the soil mineral phase. Instead, we found strong evidence that gram-positive 21 

bacteria are the key actors in the decomposition of protein-sized nitrogen compounds and that large 22 

organic nitrogen compounds contribute directly to bacterial tissue build-up. We conclude that when 23 

large organic nitrogen compounds are dissolved, turnover occurs rapidly, irrespective of molecular 24 

2  

size, and that the bacterial incorporation of these rapid cycling compounds makes an important 25 

contribution to soil organic matter formation. 26 

27 

Main text 28 

Turnover and stabilization of soil carbon (C) and nitrogen (N) are critical processes for enhancing N 29 

use efficiency in cultivated soils and mitigating increasing atmospheric loads of greenhouse gases 30 

through soil C sequestration1. Carbon plays a pivotal role for N-cycling in soils because one-third of 31 

stored C is bound in compounds containing N2. Thus, to enhance soil C storage it is vital to improve 32 

our understanding the fate of organic N compounds. The present view on soil organic matter (SOM) 33 

formation3,4 implies that microbial turnover is key to C and N stabilization5,6, which requires insight 34 

in the short-term cycling of labile organic compounds. During litter decomposition, small plant-35 

derived compounds, like amino acids and sugars, turn over within hours or days7,8, which may 36 

contribute to the build-up of microbial biomass9 and eventual SOM stabilization in the microbial 37 

necromass10,11. However, the turnover rates of the larger (>1 kDa) soluble compounds derived from 38 

plants and their role in the formation of SOM is only scarcely examined. 39 

40 

Soil organic matter is composed of progressively decomposing organic compounds in a continuum 41 

of size classes4. In the soil continuum model, Lehmann and Kleber4 makes a generally accepted 42 

distinction between small biopolymers (<0.6 kDa) that can be directly assimilated by 43 

microorganisms and larger compounds (>0.6 kDa) requiring extracellular depolymerization prior to 44 

microbial assimilation. The majority of organic N compounds enter the soil in the form of proteins 45 

inevitably larger than 0.6 kDa12 and the depolymerization of proteins to peptides and amino acids 46 

are considered the bottleneck in the turnover of organic N13. The slower turnover of proteins14,15 47 

than of small peptides and amino acids is considered to be a result of strong retention of proteins in 48 

3  

the soil mineral phase15,16 or the need for production of complex energetically demanding exo-49 

enzymes, which represents a risky investment for the microorganisms in the soil ecosystem17. It has 50 

also been suggested that peptides are strongly bound to the soil mineral phase2, but it is not clear 51 

whether these peptides are bound in the original form or in microbial necromass11. Therefore, to 52 

determine the mechanisms controlling large molecular weight (Mw) organic N cycling in soil, we 53 

studied the short-term fate of non-structural organic N compounds in four molecular size classes 54 

above the 0.6 kDa threshold for assimilation in soil. We triple-labeled (14C, 13C, 15N) white clover 55 

during the growth phase and retrieved non-structural organic N compounds from shoots via screw 56 

pressing and molecular size fractionation of the resulting plant juice18. The decomposition of these 57 

organic N fractions was determined in topsoil from long-term field experiments (LTE) with 58 

contrasting pH and fertilizer management; specifically the Jyndevad LTE on liming and phosphorus 59 

fertilization initiated in 194219 and the Askov LTE on animal manure and mineral fertilizer initiated 60 

in 189420. We determined organic N sorption and mineralization, investigated to which extent the 61 

organic N compounds were retained in their original forms or had been metabolized using amino 62 

acid (AA) stable isotope probing (SIP)18, and we identified active microbial groups metabolizing 63 

the labeled organic N using phospholipid acid stable isotope probing (PLFA-SIP)21. 64 

65 

Larger organic N size reduces mineralization 66 

The release of CO2 from organic N mineralization followed first-order kinetics without a lag-phase 67 

across all Mw size fractions (1-10, 10-30, 30-100, >100 kDa) and soil pH levels (low: pHCaCl2 3.6; 68 

medium: pHCaCl2 5.4; high: pHCaCl2 7.1) (Fig. 1a-d). This shows that the microbial community was 69 

immediately capable of decomposing the added organic N compounds across all pH levels, 70 

although proteolysis is believed to be under strong pH control22. However, the accumulated CO2 71 

released decreased with increasing molecular size. In the 1-10 kDa fraction, 40-45% of added 14C 72 

4  

was respired as 14CO2 after 14 days at all soil pH levels (Fig. 1a). Likewise the 10-30 kDa fraction 73 

in the high pH soil resembled the mineralization patterns often found for small organic N 74 

compounds (< 1kDa)8,23. Respiration from the low pH soil decreased with increasing molecular size 75 

compared to the medium and high pH soils (Fig. 1b-d). Based on the C/N ratios (Extended Dsata 76 

Table 1), the 1-10, 10-30 and 30-100 kDa fractions most likely contained C-compounds other than 77 

organic N, which may have contributed to 14CO2 production. Nevertheless, the level of respiration 78 

halved from the 1-10 kDa to both the 30-100 kDa and >100kDa fractions. Classically this would be 79 

interpreted as a bottleneck in the decomposition of organic N above the Mw of 30 kDa14,15 with a 80 

more pronounced bottleneck under less favorable conditions (i.e. in the 10-30 kDa range at low pH) 81 

and less pronounced under more favorable conditions (i.e. in the 30-100 kDa range at high pH). 82 

83 

Three main mechanisms can explain the reduced respiration with increasing organic N molecular 84 

sizes: (i) increased protection with higher molecular size making organic N inaccessible for 85 

microorganisms15,16, (ii) an increase in ex vivo modification of the original compounds6 or (iii) a 86 

higher microbial C use efficiency with increasing molecular sizes (i.e. a shift from catabolism to 87 

anabolism)6. We first examined relationships between respiration and organic N sorption 88 

determined as removal of labeled (13C, 15N) organic N compounds from soil solution after one hour. 89 

We found a negative correlation between respiration and sorption of organic N (Fig. 1e-f) indicating 90 

that microbial decomposition is controlled the accessibility of organic N24,25. In addition, we found 91 

a significantly larger recovery after 14 days of added 13C in soil for the >100 kDa fraction compared 92 

to the 1-10 kDa fraction (Extended Data Table 2). In order to elucidate to what extent the added 93 

organic N compounds were retained in their original form, we determined the isotopic signatures of 94 

soil-bound amino acids for the 1-10 kDa fraction where a bottleneck was not observed and the >100 95 

kDa fraction where the bottleneck was most pronounced and present in all studied soils. 96 

5  

97 

Organic N not retained in original form 98 

Across all pH levels and organic N fractions, the lowest recoveries of individual amino acids 99 

(generally leucine, lysine, phenylalanine) were close to zero (Fig. 2). In general, when organic N in 100 

form of proteins or peptides is decomposed it occur via proteases breaking peptide bonds either on 101 

terminal amino acids releasing single amino acids or on internal peptide bonds releasing peptide 102 

fragments of various length. Assuming equal degradation of the added organic N compounds the 103 

low recoveries means that the organic N compounds were decomposed rather than retained in their 104 

original form. The recovery levels across individual amino acids were 0-20% for the 15N tracer and 105 

1-30% for the 13C tracer with significant decoupling of the remaining 13C and 15N in individual 106 

amino acids (Fig. 2). The 13C and 15N decoupling further supports the microbial decomposition of 107 

the added organic N compounds, as deaminating amino-groups is often the initial step in during 108 

microbial metabolism26. It is unlikely that our low recovery of the dual-labeled amino acids is 109 

associated with the hydrolysis procedure27 as the isotopic ratios of amino acids are unaffected by 110 

amino acid decomposition during hydrolysis18. Further, it is highly unlikely that chemical sorption 111 

of the originally added amino acids in peptides and proteins explains the low recovery as such 112 

processes would result in similar or consistent recoveries across all amino acids. Thus, our amino 113 

acid 13C and 15N is representative of the isotopic signature prior to extraction and, despite the 114 

pronounced sorption to the soil of the added organic N compounds (Fig. 1e-f), the organic N was 115 

not protected against microbial decomposition. Hence, any sorption of large organic N must be an 116 

equilibrium between the soil matrix and soil solution allowing decomposition and the lower 117 

mineralization to 14CO2 observed for the >100 kDa fraction than for the 1-10 kDa fraction. This 118 

together with the similar levels of amino acids remaining implies that the >100 kDa organic N is 119 

involved in microbial anabolism to a greater extent than the 1-10 kDa organic N6. 120 

6  

121 

To confirm that the almost complete degradation of >100 kDa organic N was not due soil specific 122 

conditions, we validated our findings for the protein-sized fraction (>100 kDa) using soils from the 123 

Askov LTE. Soil from Askov is more clayey than the Jyndevad soils20 (Extended Data Table 3 & 124 

4), and the long-term experimental treatments have resulted in different microbial communities28,29 125 

and fertility levels30. The degradation patterns of the >100 kDa fraction across the fertility levels in 126 

the Askov soils resembled those observed at the different soil pH levels in Jyndevad soils. The 127 

mineralization to 14CO2 reached 15-20% after 14 days (Fig. 3a) and had a negative relationship to 128 

sorbed organic N (Fig. 3b-c). The added organic N was degraded with less than 25% of the 13C and 129 

15N in bound amino acids remaining after two weeks at all three fertility levels (Fig. 3d-f), thereby 130 

corroborating the equilibrium between sorbed and dissolved organic N as aforementioned. This is 131 

remarkable especially for the soil unfertilized since 1894 and known to be unsaturated with organic 132 

matter31 and thus exhibiting strong potential for sorption32. 133 

134 

The patterns of amino acid decomposition were surprising. Amino acids with more complex 135 

biosynthetic pathways (e.g. leucine, lysine, phenylalanine) were decomposed at a greater rate than 136 

the simpler amino acids (alanine, asparagine/aspartate, glutamine/glutamate) (Fig. 2d-i and 3d-f), 137 

which diverges from the expected direct anabolic microbial use of amino acids to reduce energy for 138 

de novo synthesis33. The results can neither be explained by different additions of these specific 139 

amino acids in the 1-10 and >100 kDa fractions (Extended Data Fig. 1) nor by changes in soil 140 

amino acid composition upon organic N addition (Extended Data Fig. 2). It is also notable that more 141 

amino acids remained in the 1-10 kDa than >100 kDa fractions at Jyndevad medium and high pH 142 

soils underlining that greater molecular size does not protect against degradation. Moreover, 13C 143 

and 15N decoupling were highest for the 1-10 kDa fraction with more 13C than 15N remaining at all 144 

7  

Jyndevad pH levels. The most likely explanation for this decoupling is a higher rate of deaminating 145 

amino-groups during microbial metabolism26 than catabolizing amino acid carbon skeletons. The C 146 

and N decoupling were highest for amino acids associated with aminotransferases, thus supporting 147 

that bacterial cells incorporated intact amino acid from the soil medium. In other words, simpler 148 

amino acids such as alanine and asparagine were incorporated intact into microbial tissue to a 149 

greater extent than more complex amino acids such as lysine and phenylalanine. Interestingly, 150 

simpler amino acids, alanine, asparagine/aspartate, glutamine/glutamate, and glycine, are typically 151 

among the most abundant constituents of the peptidoglycan layers of bacterial cell walls34-36. Hence, 152 

our findings support substantial incorporation of organic N by microbial cells. 153 

154 

Bacteria are doing the work 155 

To elucidate microbial groups active in organic N turnover, we determined the incorporation of 13C 156 

in PLFA. All microbial groups were 13C enriched after 14 days of the incubation, but bacteria 157 

dominated the specific incorporation of 13C from both the 1-10 kDa and >100 kDa fractions across 158 

all soils (Fig. 4c-e). Both bacteria and fungi have the capacity to facilitate exo-enzymatic 159 

depolymerization with gram-positive bacteria and fungi typically contributing to the degradation of 160 

complex compounds and gram-negative generally decomposing lower Mw compounds37. The 161 

specific incorporation of 13C in microbial PLFA from the protein-sized organic N compounds 162 

showed a surprisingly similar pattern across all Jyndevad pH levels (Fig. 4d) and Askov fertilizer 163 

treatments (Fig. 4e). In all soils, gram-positive bacteria had a significantly higher specific 13C 164 

incorporation from the >100 kDa fraction than gram-negative bacteria, and subsequently higher 13C 165 

incorporation than fungi. Low soil pH is generally considered to reduce bacterial activity, thus 166 

enhancing the relative importance of fungal activity38. Although fungi and gram-negative bacteria 167 

activity (Fig. 4) and biomass (Extended Data Fig. 3) responded to organic N addition, the higher 168 

8  

activity of gram-positive bacteria on protein-sized organic N compounds show that this microbial 169 

group must both have the proteolytic ability and uptake mechanisms to outcompete other microbial 170 

groups for protein-derived organic N. The production of extracellular enzymes is expected to be 171 

greater in gram-positive than in gram-negative bacteria where enzymes to a larger extent 172 

accumulate in the periplasmic space rather than being exuded39. Furthermore, gram-positive 173 

bacterial species can directly assimilate organic N well above the 0.6 kDa threshold 40,41. This 174 

suggests that the gram-positive bacterial group outcompeted the gram-negative and fungal groups in 175 

a two-step process involving depolymerization of protein-sized N and subsequent direct 176 

assimilation of the released peptide helixes (Fig. 5a). Gram-negative bacteria needing organic N 177 

smaller than the 0.6 kDa threshold for assimilation would then scavenge on any further 178 

depolymerization of peptide N. A similar mechanism could apply to the 1-10 kDa fraction (Fig. 5b), 179 

where the specific 13C incorporation showed an activity equivalent with gram-positive and gram-180 

negative bacteria and a greater fungal activity than found with the >100 kDa fraction (Fig. 4c). As 181 

mentioned previously, the 1-10 kDa fraction most likely contained labeled C-compounds other than 182 

organic N, which could also have contributed to the 13C incorporation in microbial biomarkers; this 183 

may explain the 13C incorporation in the fungal biomarker for this organic N fraction. In conclusion, 184 

the results suggest that direct assimilation above the 0.6 kDa threshold may be more prevalent for 185 

gram-positive bacteria than previously thought, and importantly the PLFA-SIP results identify a 186 

microbial group active on the added organic N having the toolbox to rapidly turnover protein-sized 187 

N. 188 

189 

Discussion 190 

The rate-limiting step in soil organic N turnover is generally thought to be the depolymerization of 191 

higher Mw organic N into lower Mw compounds that can be directly assimilated by 192 

9  

microorganisms and plants4,6,13. The rate of depolymerization is affected by sorption of the 193 

substrate4,16 or exo-enzymes to the soil mineral phase42. The latter causing hindrance of exo-enzyme 194 

activity due to the spatial separation from the substrate3 or by blockage of the enzyme reactive 195 

sites15. In line with this, we found that increasing molecular size of added organic N reduced the 196 

mineralization; a reduction in mineralization strongly correlated to sorption of the higher organic N 197 

sizes. However, in all treatments, we saw a rapid degradation of the added organic N compounds 198 

irrespective of molecular size (Fig. 2 and 3d-f). Such rapid dissipation of protein-sized organic N 199 

could be expected in fertile agricultural soils13, but we were surprised to find that even in the low 200 

productivity soil (low pH or unfertilized since 1894) with lower microbial biomass (Fig. 4a-b), the 201 

remaining added organic N was at a similarly low level as in the more fertile soils. This shows that 202 

the ability of the existing microbial biomass for depolymerization of large size organic N was 203 

sufficient in these soils and that the high sorption of >100 kDa organic N in all soils did not prevent 204 

an almost complete turnover of the added compounds. Thus, we demonstrate that depolymerization 205 

of proteins is not per se the rate-limiting step in large size organic N turnover, neither is sorption of 206 

protein-sized organic N to the soil mineral phase. In the latter, we show that sorbed organic N is in 207 

equilibrium with the soil solution where dissolved organic N compounds are rapidly turned over; 208 

i.e. sorption is not protecting the compounds. Instead, part of the plant-derived protein must – at 209 

least short-term – be physically protected in cell structures, which needs to be degraded before 210 

proteins can be turned over. Hence, in contrast to the presently viewed importance of proteolytic 211 

enzymes43,44, other enzyme classes may play the key role in the eventual turnover of bound organic 212 

N. In the long term, parts of the structurally bound organic N may be physically protected in 213 

particulate organic matter predominately of plant origin32. 214 

215 

10  

To study stabilization of plant-derived C, Liang et al.6 suggests to differentiate between in vivo 216 

turnover and ex vivo modification, i.e. biotic, abiontic, and abiotic processes taking place inside or 217 

outside a microbial cell. The rapid and almost complete dissipation of all large size organic N 218 

fractions found in the present study and the incorporation of organic N derived 13C into bacterial 219 

PLFAs point to in vivo turnover as the dominant route of decomposition. The turnover of >100 kDa 220 

organic N was dominated by gram-positive bacteria, which have the exo-enzymatic tools to 221 

depolymerize these large compounds. Thus, the potential for ex vivo modification should potentially 222 

be higher for the >100 kDa organic N as would retention of modified compounds to the soil mineral 223 

phase. However, our data do not support this. Therefore, large organic N primarily contributes to 224 

SOM formation via build-up of microbial tissue, where incorporation of C and N in the short-225 

chained peptides of bacterial cell walls potentially results in longer-term storage of plant-derived C. 226 

Our study provides strong evidence for the hypothesis that C and N from labile compounds persist 227 

in soil5, but rather than persisting due to protection of the original compounds3, the C and N persist 228 

due to the incorporation via anabolic processes into microbial tissue. Furthermore, the rapid 229 

turnover of large molecular size organic N compounds in our study suggests that it will be 230 

beneficial to make a distinction between organic N contained inside the cell (non-structural5) and 231 

within cell structures when predicting the release of plant-available N from plant residues. 232 

Additionally, non-structural N inside microbial cells should be considered as a temporary pool that 233 

is highly prone to rapid decomposition. Finally, by identifying gram-positive bacteria as the 234 

dominating organic N decomposers, our study suggests that exo-enzymatic decomposition would 235 

potentially allow plants to assimilate degradation metabolites, such as amino acids or short peptides, 236 

in the same manner that gram-negative bacteria in the present study assimilated degradation 237 

metabolites from the protein-sized organic N. 238 

239 

11  

Materials and methods 240 

Soils came from the Jyndevad and Askov long-term field experiments (LTE) in Denmark. The 241 

Jyndevad LTE on liming and phosphorus was initiated in 194219 on a coarse sandy soil (Extended 242 

Data Table 3) cultivated with spring barley for at least 30 years. Soil was sampled in August 2015 243 

from the plough layer (5-20 cm) of the V1 field in the treatments receiving 0, 4 and 12 Mg lime ha-1 244 

every 6-9 years and yearly doses of 15.6 kg P ha-1 year-1. At the time of soil sampling contrasting 245 

pHCaCl2 levels of 3.6 (low pH), 5.4 (medium pH), and 7.1 (high pH) were established in the three 246 

treatments. The Askov LTE on animal manure and mineral fertilizers was initiated in 189420 on a 247 

sandy loam soil in an arable crop rotation (Extended Data Table 4). Soil was sampled in October 248 

2015 from the plough layer (5-20 cm) of the treatments designated unfertilized, 1½ mineral 249 

fertilizer (NPK), and 1½ animal manure (AM) treatments of the B3 field. Annually, the 1½ NPK 250 

and 1½ AM treatments have received on average 150 kg total-N, 30 kg P and 120 kg K ha-1 in 251 

mineral fertilizer and animal manure (slurry since 1974), respectively. All soils were sieved (4 mm) 252 

to remove visible roots and stored at 2 ºC until the incubation experiment in September 2015 for 253 

Jyndevad soils and October 20145 for Askov soils. 254 

255 

Organic N fractions were produced from greenhouse grown triple-labeled (14C, 13C, 15N)45 white 256 

clover shoots using a screw press and subsequent Mw size fraction of the juice18 into the fractions: 257 

1-10, 10-30, 30-100, and >100 kDa. The organic N fractions were characterized for total C and N, 258 

bulk isotopic and amino acid specific composition as described in Enggrob et al.18 (Extended Data 259 

Table 1, Extended Data Fig. 1). The organic N fractions were incubated in packed micro-lysimeters 260 

holding 12 g field moist soil. The micro-lysimeters were constructed from inserts in 50 mL 261 

centrifuge tubes (Extended Data Fig. 4) to allow rapid recovery of soil solution by centrifugation 262 

upon termination of incubation. Incubation time was one-hour and 14 days at room temperature (22 263 

12  

°C) and all soil amendments were made in four replicates. The one hour incubation allowed the 264 

determination organic N sorption. The 14 days incubation was chosen for the mineralization 265 

response because at that time the 14CO2 production from the fastest mineralizing organic N fraction 266 

started to level off; based on the assumption that if sorption controls mineralization this would be 267 

the time when labeled organic N had been depleted from soil solution. The organic N fractions were 268 

added in low amounts in 2 mL water (100-190 µg C g-1 soil and 9-40 µg N g-1 soil), and sufficiently 269 

low to have a minor or no influence on the concentration of extractable amino acids in soil 270 

(Extended Data Fig. 2). The micro-lysimeters were incubated in the dark in 1 L glass jars at room 271 

temperature (22ºC) with a beaker holding 1 mL of 0.5 M NaOH to trap 14CO2. CO2 traps were 272 

replaced after 1, 4, 7, and 14 days. Liquid scintillation cocktail (OptiPhase HiSafe3, PerkinElmer, 273 

Waltham, MA, USA) was added to the trap solution and 14C-activity counted on a Tri-Carb® 274 

2910TR Liquid Scintillation Analyser (PerkinElmer, Waltham, MA, USA). All four organic N 275 

fractions were incubated in the Jyndevad soils, whereas the >100 kDa fraction was incubated in the 276 

Askov soils for comparison of results across two soil types. Control treatments with addition of 2 277 

mL water instead of organic N were run for all soils and sampling times. 278 

279 

Upon terminating the incubation, the micro-lysimeters were first added 8 mL of water and 280 

centrifuged for 5 minutes at 5000 g followed by addition of 10 mL of water and repeated 281 

centrifugation. The two solutions were pooled to give one sample of 20 mL containing the soluble 282 

N fractions. After this, the soil in the micro-lysimeters was washed in a similar manner with two 283 

times 10 mL 1 M KCl and the KCl solutions were pooled. Both water and KCl solutions were 284 

immediately filtered through 0.45 µm Macrosep centrifuge filters (Pall Corporation, New York, 285 

USA) and the filtrates were sampled for 14C-analysis (see above). The remaining liquid samples 286 

were stored frozen until further analysis of total C and N content and 13C and 15N isotope 287 

13  

composition. After the final KCl wash the soil was immediately recovered from the micro-288 

lysimeters and stored frozen until further analysis. 289 

290 

Soil solution samples were freeze-dried, re-dissolved in 1 mL MilliQ water (Synergy® System, 291 

Millipore, Molsheim, France), transferred to tin capsules and freeze-dried before analysis of total C 292 

and N content and 13C and 15N stable isotope composition. Analyses were performed on a Flash 293 

Elemental Analyser (Thermo Scientific, Hvidovre, Denmark) coupled via a TCD to an isotope ratio 294 

mass spectrometer (Delta V Plus IRMS, Thermo Scientific, Hvidovre, Denmark). Mass 295 

spectrometer related parameters were controlled by the Isodat software version 3.0 (Thermo 296 

Scientific, Hvidovre, Denmark). All δ13C values are reported relative to the Vienna PeeDee 297 

Belemnite (VPDB) international isotope standard. All δ15N values are reported relative to the δ15N 298 

values of atmospheric N2. 299 

300 

Soil samples were freeze-dried and homogenized by ball milling to allow representative sub-301 

sampling. Analysis of total C and N and 13C and 15N composition was carried out as described 302 

above after weighing 25-30 mg soil samples into tin capsules. Jyndevad soils with added 1-10 and > 303 

100 kDa organic N fractions and Askov soils with added >100 kDa fraction (all incubated for 14 304 

days) underwent compound-specific isotope analysis aimed at determining organic N bound in 305 

amino acids (AA-SIP) and biomarkers for active microbial biomass (PLFA-SIP). For AA-SIP, 800 306 

mg freeze-dried soil was weighed into 16x100 soda-lime disposable test tubes (Duran Group, 307 

Mainz, Germany) added 2 mL 6 M HCl and hydrolyzed for 20 hours at 110°C. To remove solids 308 

and lipophilic compounds 4 mL n-hexane/dichloromethane (6:5, v/v) was added and vortexed for 309 

30 s, upon centrifugation (1600 rpm at 2 min). After mixing and centrifugation, the aquatic phase 310 

was transferred through a Pasteur pipette lined with glass wool, to remove visible floating particles 311 

14  

from the aquatic phase, followed by washing the Pasteur pipette lined with glass wool by 2 x 0.5 312 

mL 0.1 M HCl into new test tubes. The remaining sample preparation (purification and 313 

derivatization of amino acids) along with the GC/C-IRMS analysis was done as described in 314 

Enggrob et al.18, except for an extra freeze-drying step added during purification after the addition 315 

of the internal standard (300µl 2.5 M norLeucine) and before the filtration of the samples on resin 316 

columns. The amino acids: asparagine and aspartate (Asx), glutamine and glutamate (Glx), and 317 

Proline and Threonine (Pro/Thr) elute together in the GC-C-IRMS analysis of acid hydrolyzed 318 

samples. For PLFA-SIP, 2.5 g freeze-dried soil was used to isolate phospholipids by a Bligh-Dyer 319 

single phase extraction followed by a solid–phase extraction on silicic acid columns and an alkaline 320 

transesterification46,47. The PLFA’s were analyzed for isotopic composition by a GC-C-IRMS at the 321 

Stable isotope service lab., Department of Biology, Lund University, Sweden. Individual PLFA’s 322 

were assigned to specific microbial groups48-50 (Extended Data Table 5); both specified and 323 

unspecified PLFAs were used for estimating active microbial biomass. 324 

325 

Statistical analyses 326 

The influence of the soil pH levels and comparisons across 13C and 15N content, amino acid 327 

contents and groupings of soil microorganisms were tested with a linear mixed-effects model using 328 

the statistical analysis program R version 3.3.1 (R Core Team, 2016, URL https://www.R-329 

project.org/). For statistical analysis, the datasets were tested for normal distribution by the Shapiro-330 

Wilk test of normality. Linear models describing each subset were tested by pairwise comparisons. 331 

Correlations between sorption and mineralization were conducted using the Spearman’s rank 332 

correlation coefficient or Spearman’s rho. 333 

334 

References 335 

15  

1  Paustian, K. et al. Climate‐smart soils. Nature 532, 49‐57 (2016). 336 2  Knicker, H. Soil organic N ‐ An under‐rated player for C sequestration in soils? Soil Biol. Biochem. 43, 337 

1118‐1129 (2011). 338 3  Schmidt, M. W. I. et al. Persistence of soil organic matter as an ecosystem property. Nature 478, 49‐339 

56 (2011). 340 4  Lehmann, J. & Kleber, M. The contentious nature of soil organic matter. Nature 528, 60‐68, 341 

doi:10.1038/nature16069 (2015). 342 5  Cotrufo, M. F. et al. Formation of soil organic matter via biochemical and physical pathways of litter 343 

mass loss. Nat. Geosci. 8, 776‐779 (2015). 344 6  Liang, C., Schimel, J. P. & Jastrow, J. D. The importance of anabolism in microbial control over soil 345 

carbon storage. Nat. Microbiol. 2, Article no. 17105 (2017). 346 7  Czaban, W., Rasmussen, J., Nicolaisen, M. & Fomsgaard, I. S. Dissipation kinetics of asparagine in 347 

soil measured by compound‐specific analysis with metabolite tracking. Biol. Fertil. Soils 52, 911‐916 348 (2016). 349 

8  Farrell, M. et al. Rapid peptide metabolism: A major component of soil nitrogen cycling? Global 350 Biogeochem. Cy. 25 (2011). 351 

9  Apostel, C., Dippold, M. A., Bore, E. & Kuzyakov, Y. Sorption of Alanine changes microbial 352 metabolism in addition to availability. Geoderma 292, 128‐134 (2017). 353 

10  Khan, K. S., Mack, R., Castillo, X., Kaiser, M. & Joergensen, R. G. Microbial biomass, fungal and 354 bacterial residues, and their relationships to the soil organic matter C/N/P/S ratios. Geoderma 271, 355 115‐123 (2016). 356 

11  Miltner, A., Bombach, P., Schmidt‐Brücken, B. & Kästner, M. SOM genesis: microbial biomass as a 357 significant source. Biogeochem., 1‐15 (2012). 358 

12  Nannipieri, P. & Eldor, P. The chemical and functional characterization of soil N and its biotic 359 components. Soil Biol. Biochem. 41, 2357‐2369 (2009). 360 

13  Schimel, J. P. & Bennett, J. Nitrogen mineralization: Challenges of a changing paradigm. Ecology 85, 361 591‐602 (2004). 362 

14  Jan, M. T., Roberts, P., Tonheim, S. K. & Jones, D. L. Protein breakdown represents a major 363 bottleneck in nitrogen cycling in grassland soils. Soil Biol. Biochem. 41, 2272‐2282 (2009). 364 

15  Jones, D. L. & Kielland, K. Amino acid, peptide and protein mineralization dynamics in a taiga forest 365 soil. Soil Biol. Biochem. 55, 60‐69 (2012). 366 

16  Mariano, E., Jones, D. L., Hill, P. W. & Trivelin, P. C. O. Mineralisation and sorption of dissolved 367 organic nitrogen compounds in litter and soil from sugarcane fields. Soil Biol. Biochem. 103, 522‐368 532 (2016). 369 

17  Bosatta, E. & Ågren, G. I. Soil organic matter quality interpreted thermodynamically. Soil Biol. 370 Biochem. 31, 1889‐1891 (1999). 371 

18  Enggrob, K. L., Larsen, T., Larsen, M., Elsgaard, L. & Rasmussen, J. The influence of hydrolysis and 372 derivatization on the determination of amino acid content and isotopic ratios in dual labeled (13C, 373 15N) white clover. Rapid Commun. Mass Spectrom. 33, 21‐30 (2019). 374 

19  Rubaek, G. H. Effects of liming an P fertilisation. in Long continued Agriculatural Soil experiments: A 375 Nordic research platform – a catalogue, eds Petersen, J. et al., DJF Plant Science 16, 12‐14, Aarhus 376 University, Denmark (2008). 377 

20  Christensen, B. T., Petersen, J. & Trentemoller, U. M. The Askov long‐term experiments on animal 378 manure and mineral fertilizers: The Lermarken site 1894‐2004. Report No. 121, (2006). 379 

21  Dumont, M. G. & Murrell, J. C. Stable isotope probing ‐ linking microbial identity to function. Nat. 380 Rev. Microbiol. 3, 499‐504 (2005). 381 

22  Nasholm, T., Kielland, K. & Ganeteg, U. Uptake of organic nitrogen by plants. New Phytol. 182, 31‐382 48 (2009). 383 

16  

23  Hill, P. W., Farrell, M. & Jones, D. L. Bigger may be better in soil N cycling: Does rapid acquisition of 384 small l‐peptides by soil microbes dominate fluxes of protein‐derived N in soil? Soil Biol. Biochem. 385 48, 106‐112 (2012). 386 

24  Jones, D. L. & Hodge, A. Biodegradation kinetics and sorption reactions of three differently charged 387 amino acids in soil and their effects on plant organic nitrogen availability. Soil Biol. Biochem. 31, 388 1331‐1342 (1999). 389 

25  Gonod, L. V., Jones, D. L. & Chenu, C. Sorption regulates the fate of the amino acids lysine and 390 leucine in soil aggregates. Eur. J. Soil Sci. 57, 320‐329 (2006). 391 

26  Dippold, M. A. & Kuzyakov, Y. Biogeochemical transformations of amino acids in soil assessed by 392 position‐specific labelling. Plant Soil 373, 385‐401 (2013). 393 

27  Kogel‐Knabner, I. & Rumpel, C. Advances in molecular approaches for understanding soil organic 394 matter composition, origin, and turnover: A historical overview. Advances in Agronomy, 149, 1‐48 395 (2018). 396 

28  Tamez‐Hidalgo, P., Christensen, B. T., Lever, M. A., Elsgaard, L. & Lomstein, B. A. Endospores, 397 prokaryotes, and microbial indicators in arable soils from three long‐term experiments. Biol. Fertil. 398 Soils 52, 101‐112 (2016). 399 

29  Hemkemeyer, M., Christensen, B. T., Martens, R. & Tebbe, C. C. Soil particle size fractions harbour 400 distinct microbial communities and differ in potential for microbial mineralisation of organic 401 pollutants. Soil Biol. Biochem. 90, 255‐265 (2015). 402 

30  Suarez‐Tapia, A., Thomsen, I. K., Rasmussen, J. & Christensen, B. T. Residual N effect of long‐term 403 applications of cattle slurry using winter wheat as test crop. Field Crop. Res. 221, 257‐264 (2018). 404 

31  Jensen, J. L., Schjonning, P., Christensen, B. T. & Munkholm, L. J. Suboptimal fertilisation 405 compromises soil physical properties of a hard‐setting sandy loam. Soil Res. 55, 332‐340 (2017). 406 

32  Castellano, M. J., Mueller, K. E., Olk, D. C., Sawyer, J. E. & Six, J. Integrating plant litter quality, soil 407 organic matter stabilization, and the carbon saturation concept. Global Change Biol. 21, 3200‐3209 408 (2015). 409 

33  Kirchman, D. L., Newell, S. Y. & Hodson, R. E. Incorporation versus biosynthesis of leucine – 410 implications for measuring rates of protein‐synthesis and biomass production by bacteria in marine 411 systems. Mar. Ecol. Prog. Ser. 32, 47‐59 (1986). 412 

34  Simelyte, E., Rimpilainen, M., Zhang, X. & Toivanen, P. Role of peptidoglycan subtypes in the 413 pathogenesis of bacterial cell wall arthritis. Ann. Rheum. Dis. 62, 976‐982 (2003). 414 

35  Schneewind, O. & Missiakas, D. M. Protein secretion and surface display in Gram‐positive bacteria. 415 Philos. Trans. R. Soc. Lond. B Biol. Sci. 367, 1123‐1139 (2012). 416 

36  Vollmer, W., Blanot, D. & de Pedro, M. A. Peptidoglycan structure and architecture. FEMS 417 Microbiol. Rev. 32, 149‐167 (2008). 418 

37  Madigan, M. T., Bender, K. S., Buckley, D. H., Sattley, W. M. & Stahl, D. A. Brock Biology of 419 Microorganisms. 15th edition, Pearson Education Limited, London, UK (2018). 420 

38  Rousk, J. et al. Soil bacterial and fungal communities across a pH gradient in an arable soil. ISME J. 421 4, 1340‐1351 (2010). 422 

39  Burns, R. G. Enzyme activity in soil: some theoretical and practical considerations. in Soil Enzymes, 423 ed R.G. Burns, Academic Press Inc., London, UK (1978). 424 

40  Detmers, F. J. M., Kunji, E. R. S., Lanfermeijer, F. C., Poolman, B. & Konings, W. N. Kinetics and 425 specificity of peptide uptake by the oligopeptide transport system of Lactococcus lactis. Biochem. 426 37, 16671‐16679 (1998). 427 

41  Scherrer, R., Beaman, T. C. & Gerhardt, P. Macromolecular sieving by dormant spore of bacillus‐428 cereus. J. Bacteriol. 108, 868‐& (1971). 429 

42  Nannipieri, P., Trasar‐Cepeda, C. & Dick, R. P. Soil enzyme activity: a brief history and biochemistry 430 as a basis for appropriate interpretations and meta‐analysis. Biol. Fertil. Soils 54, 11‐19 (2018). 431 

17  

43  Simpson, J., Warren, C. & Adams, P. Potential protease activity and organic nitrogen concentration 432 are rapid tests and accurate indicators of N‐availability in Tasmanian Eucalyptus nitens plantations. 433 Soil Biol. Biochem. 115, 152‐160 (2017). 434 

44  Lipson, D. & Nasholm, T. The unexpected versatility of plants: organic nitrogen use and availability 435 in terrestrial ecosystems. Oecologia 128, 305‐316 (2001). 436 

45  Rasmussen, J., Gjettermann, B., Eriksen, J., Jensen, E. S. & Hogh‐Jensen, H. Fate of 15N and 14C from 437 labelled plant material: Recovery in perennial ryegrass‐clover mixtures and in pore water of the 438 sward. Soil Biol. Biochem. 40, 3031‐3039 (2008). 439 

46  Petersen, S. O. & Klug, M. J. Effects of sieving, storage, and incubation‐temperature on the 440 phospholipid fatty‐acid profile of a soil microbial community. Appl. Environ. Microbiol. 60, 2421‐441 2430 (1994). 442 

47  Petersen, S. O., Frohne, P. S. & Kennedy, A. C. Dynamics of a soil microbial community under spring 443 wheat. Soil Sci. Soc. Am. J. 66, 826‐833 (2002). 444 

48  Fierer, N., Schimel, J. P. & Holden, P. A. Variations in microbial community composition through two 445 soil depth profiles. Soil Biol. Biochem. 35, 167‐176 (2003). 446 

49  Stromberger, M. E., Keith, A. M. & Schmidt, O. Distinct microbial and faunal communities and 447 translocated carbon in Lumbricus terrestris drilospheres. Soil Biol. Biochem. 46, 155‐162 (2012). 448 

50  Frostegard, A., Baath, E. & Tunlid, A. Shifts in the structure of soil microbial communities in limed 449 forests as revealed by phospholipid fatty‐acid analysis. Soil Biol. Biochem. 25, 723‐730 (1993). 450 

451 

Extended data 452 

Extended method description and data are available online. 453 

454 

Acknowledgement 455 

We thank L. Elsgaard, G.H. Rubæk, L. Peixoto, Z. Liang and J.E. Olesen for discussions. The study 456 

was financially supported by The Independent Research Fund Denmark – Technology and 457 

Production (Project no. 1335-00760B). 458 

459 

Author contributions 460 

K.L.E. and J.R. designed and executed the experiment and analysis; K.L.E. performed statistical 461 

analysis; K.L.E., T.L. and J.R. made AA-SIP interpretations; K.L.E. and J.R. drafted the 462 

manuscript; all authors revised the manuscript and approved the final version. 463 

464 

465 

18  

Author Information 466 

Reprints and permissions information is available at www.nature.com/reprints. The authors declare 467 

no competing interests. Correspondence and requests for materials should be addressed to J.R. 468 

([email protected]). 469 

 470 

Figure 1. 

Fig. 1. Mineralization and sorption of large organic N. Mineralization of labeled organic N to 14CO2 in Jyndevad soils at three pHCaCl2 levels: low at pH 3.6, medium at pH 5.4, and high at pH 7.1. Panels show (a) the 1-10 kDa organic N fraction, (b) the 10-30 kDa fraction, (c) the 30-100 kDa fraction, and (d) the >100 kDa fraction. Statistical differences among soil pH levels in accumulated 14CO2 after 14 days are indicated by different letters next to the curves. Correlation between organic N sorption after 1 hour and accumulated 14CO2 after 14 days with (e) sorption of 13C in added organic N fractions (Rs = -0.90, P<0.001; for 15N Rs = -0.84, P <0.001) and (f) sorption of 15N in added organic N fractions (Rs = -0.84, P <0.001).

a

0

10

20

30

40

50b

c

Time (days)0 2 4 6 8 10 12 14

Min

eral

iza

tio

n o

f o

rgan

ic N

to

CO

2

(cu

mu

lati

ve 1

4 CO

2 p

rod

uct

ion

ad

ded

14C

, %

)

0

10

20

30

40

50

Low pH (3.6)Medium pH (5.4)High pH (7.1)

d

0 2 4 6 8 10 12 14

ns b

a

a

ns

baba

e

Sorption of organic N(13C retained after 1 hour, %)

0 20 40 60 80 1000

10

20

30

40

50f

Sorption of organic N(15N retained after 1 hour, %)

0 20 40 60 80 100

Low pHMedium pHHigh pH1-10 kDa10-30 kDa30-100 kDa>100 kDa

Figure 2.

  

Fig. 2. Organic N derived amino acids remaining in soil. Amino acids remaining in % of added bound amino acids from the peptide-sized (1-10 kDa, a-c) and protein-sized (>100 kDa, d-f) organic N in the low (a,d), medium (b,e) and high pH Jyndevad soils (c,f). Significant differences are marked by an asterisk; a double asterisk indicates no 15N data; ‘nn’ indicates non-normal distribution. Amino acids are organized from left on right with increasing number of steps in their biosynthesis.

a

Bo

un

d a

min

o a

cid

s re

mai

nin

g (

% o

f ad

ded

)

0

10

20

30

4013C 15N

b c

d

Ala

Asx Glx

Se

r

Gly

Pro

/Th

r

Val

Leu Lys

Ph

e

0

10

20

30

40e

Ala

As

x

Glx

Ser

Gly

Pro

/Th

r

Va

l

Leu Lys

Ph

e

f

Ala

As

x

Glx

Ser

Gly

Pro

/Th

r

Val

Leu Lys

Ph

e

*

**

* *** **

*

*

**

* **

*

** *

**

*

* * *nn

* **

** **

*

nn

Low pH soil Medium pH soil High pH soil

1-1

0 k

Da

>100 kD

a

Figure 3.

 

Fig. 3. Mineralization, sorption and amino acids remaining in soils with different fertility. Fate of the >100 kDa organic N fraction in Askov soil with three fertilizer treatments: UNF is unfertilized since 1894, NPK is mineral fertilizers since 1894, and AM is animal manure since 1894. Mineralization to 14CO2 (a), correlation between sorption of organic N after 1 hour and accumulated 14CO2 after 14 days with results from the Askov soils inserted in the Jyndevad results (grey) (b, c), and remaining 15N and 13C in bound amino acids after 14 days of incubation for the UNF (d), NPK (e), and AM (f) soils. Significant differences are marked by an asterisk; ‘nn’ indicates non-normal distribution.

Min

era

liza

tio

n o

f o

rga

nic

N t

o C

O2

(14 C

O2

of

add

ed

14 C

, %

)

a

Time (days)

0 2 4 6 8 10 12 140

10

20

30

40

50

Unfertilized (UNF)Mineral fertilizer (NPK)Animal Manure (AM)

ns

b

Sorption of organic N(13C retained after 1 hour, %)

0 20 40 60 80 100

UNFNPKAM

c

Sorption of organic N(15N retained after 1 hour, %)

0 20 40 60 80 100

UNFNPKAM

d

Ala

As

x

Glx

Se

r

Gly

Pro

/Th

r

Va

l

Leu Lys

Ph

e

Bo

un

d a

min

o a

cid

s re

mai

nin

g

(% o

f ad

ded

)

0

10

20

30

40

5013C 15N

e

Ala

As

x

Glx

Se

r

Gly

Pro

/Th

r

Va

l

Leu Lys

Ph

e

f

Ala

As

x

Glx

Se

r

Gly

Pro

/Th

r

Va

l

Leu Lys

Ph

e

*

nn

**

*

* *

* **

*

* * * *

*

*

nn

Figure 4.

Fig. 4. Microbial biomass and specific activity of 13C from added organic N fractions in the Jyndevad and Askov soils. Total PLFA after 14 days of incubation in control added water, the 1-10 kDa fraction, and the >100 kDa fraction for the low, medium and high pH Jyndevad soils (a), Total PLFA after 14 days of incubation in control added water and the >100 kDa fraction for the unfertilized, mineral fertilized, and animal manure fertilized Askov soil (b), Specific 13C incorporation in gram-positive, gram-negative and fungal PLFAs in Jyndevad soils added the 1-10 kDa fraction (c), Specific 13C incorporation in gram-positive, gram-negative and fungal PLFAs in Jyndevad soils added the >100 kDa fraction (d), and Specific 13C incorporation in gram-positive, gram-negative and fungal PLFAs in Askov soils added the >100 kDa fraction (e). Significantly differences are marked by different letter above the bars.

a

Low pH Medium pH High pH

Mic

rob

ial b

iom

as

s(n

mo

l PL

FA

g-1

so

il)

0

5

10

15

20

25 Control1-10 kDa>100 kDa

c

Low pH Medium pH High pH0.00

0.01

0.02

Gram postive bacteria (G+) Gram negative bacteria (G-) Fungi

d

Low pH Medium pH High pH0.00

0.01

0.02

c a b - ns - - ns -

b a c

a a ba a b

a b ca b c

a b c

Mic

rbia

l b

iom

ass

acti

ve o

n o

rgan

ic N

(nm

ol

13 C

PL

FA

nm

ol-1

C P

LF

A)

b

UNF NPK AM0

10

20

30

40

50

b

e

UNF NPK AM

a b c

a b c a b c

ab a

b a

Figure 5.

Fig. 5. The suggested microbial effects on large organic N turnover in soil. (a) The suggested route of protein-size (>100 kDa) organic N turnover via gram-positive (G+) exo-enzymatic depolymerization to peptides directly assimilated by gram-positive bacteria, but also benefitting gram-negative (G-) bacteria. (b) The suggested routes for peptide-sized (1-10 kDa) organic N turnover directly through gram-positive (G+) bacteria and via exo-enzymatic depolymerization for both gram-positive (G+) and gram-negative (G-) bacteria.

LowerMw DON

1-100 kDaDON

G+Bacteria

G-Bacteria

a

NH4+

>100 kDaDON

Soil microbial

community

LowerMw DON

1-10 kDaDON

Soil microbial

community

G+Bacteria

G-Bacteria

b

NH4+

Mobilizing

Immobilizing

Exo-enzymatic

Extended data for Enggrob et al.

Extended Data Fig. 1. Concentration, 13C and 15N atom fractions of individual amino acids bound in the 1-10 kDa and >100 kDa fractions added to the soils. Mean ± standard error (n = 4).

Am

ino

aci

d c

on

cen

trat

ion

(mg

ml-

1)

0.0

0.1

0.2

0.3

0.4

0.5

0.61-10 kDa>100 kDa

Am

ino

aci

d 13

C a

tom

fra

ctio

n

0.00

0.05

0.10

0.15

Ala

Asx Glx

Se

r

Gly

Pro

/Th

r

Val

Leu Lys

Ph

e

Am

ino

aci

d 15

N a

tom

fra

ctio

n

0.00

0.01

0.02

Extended Data Fig. 2. Concentrations of amino acids in control and organic N treated soils. Mean ± standard error (n = 4).

A: Low pH

Co

nce

ntr

atio

n o

f am

ino

aci

ds

in s

oil

(m

g/g

)

0.0

0.1

0.2

0.3

0.4Control1-10 kDa>100 kDa

B: Medium pH

0.0

0.1

0.2

0.3

0.4

C: High pH

Ala

Asx Glx

Se

r

Gly

Pro

/Th

r

Val

Leu Lys

Ph

e

0.0

0.1

0.2

0.3

0.4

D: UnfertilizedControl>100 kDa

E: NPK

F: Animal manure

Ala

Asx Glx

Se

r

Gly

Pro

/Th

r

Val

Leu Lys

Ph

e

Jyndevad Askov

Extended Data Fig. 3. Microbial community structure of Jyndevad (A) and Askov (B) soils in control added water, and soil added 1-10 and >100 kDa organic N fractions. Letters above the bars show significant differences among microbial groups within each soil (n = 4).

B: Askov

UNF NPK AM

Co

ntro

l

>10

0 kD

a

Co

ntro

l

>10

0 kD

a

Co

ntro

l

>10

0 kD

a

Mic

rob

ial b

iom

as

s(n

mo

l PL

FA

g-1

so

il)

0

2

4

6

8

10

12

14

A: Jyndevad

Low pH Medium pH High pH

Con

trol

1-1

0 k

Da

>1

00 k

Da

Con

trol

1-1

0 k

Da

>1

00 k

Da

Con

trol

1-1

0 k

Da

>1

00 k

Da

Mic

rob

ial b

iom

as

s(n

mo

l PL

FA

g-1

so

il)

0

1

2

3

4

5G+ G- Fungi

c c

d

ab

d

cc

d

a

b

d

a

b

c

aa

c

c

d

e

abc

e

abc

e

a

ab

c

a

ab

c

ab

b

c

c

d

a

b

e

b

c

e

Extended Data Table 1. Composition of the four organic N Mw size fraction used in the experiment; 13C and 15N enrichment expressed as atom fraction (AF) of the isotope. Data is given as mean ± standard error (n = 4).

Fraction C quantity N quantity C/N ratio 14C-activity AF 13C AF 15N [mg ml-1] [mg ml-1] [Bq ml-1]

1-10 kDa 0.87 ±0.02 0.11 ±0.002 8.1 ±0.2 8.65 ±0.12 0.081 ±8.4E-5 0.015 ±1.3E-5 10-30 kDa 0.97 ±0.02 0.16 ±0.003 6.1 ±0.1 10.2 ±0.09 0.081 ±7.8E-5 0.015 ±0.1E-5 30-100 kDa 0.48 ±0.13 0.04 ±0.009 10.7 ±3.7 1.8 ±0.07 0.047 ±6.0E-4 0.013 ±8.5E-4 >100 kDa 0.90 ±0.01 0.20 ±0.001 4.6 ±0.1 9.78 ±0.15 0.082 ±3.7E-5 0.014 ±0.7E-5

Extended Data Table 2. Recovery (% of added) of 13C and 15N in Jyndevad soils after 14 days of incubation. Data is given as mean ± standard error (n = 4). Statistical differences among organic N fraction within each soil is show with different letter; no significant differences were found across soils within each organic N fraction.

Fraction 1-10 kDa 10-30 kDa 30-100 kDa >100 kDa

-------------------- 13C recovery (% of added) --------------------

Low pH 25.3 ± 0.8 A 28.6 ± 1.0 B 31.0 ± 1.7 B 44.1 ± 3.1 B Medium pH 33.3 ± 1.0 A 31.7 ± 0.7 A 26.9 ± 1.3 A 44.8 ± 1.6 A High pH 36.9 ± 1.1 A 34.5 ± 1.3 A 33.4 ± 1.1 A 43.6 ± 1.4 B

-------------------- 15N recovery (% of added) ---------------------

Low pH 19.9 ± 0.7 A 18.1 ± 0.6 A 23.5 ± 1.0 B 23.8 ± 2.1 B Medium pH 28.5 ± 0.3 A 23.7 ± 0.5 B 23.0 ± 0.8 C 28.3 ± 1.2 D

High pH 34.2 ± 0.9 A 28.4 ± 1.5 B 32.5 ± 1.8 C 29.6 ± 0.8 A

Extended Data Table 3. Basic properties of soils from the Jyndevad LTE on liming and phosphorus fertilization initiated in 1942. The experiment is located at Jyndevad Experimental Station, Southern Jutland, Denmark (54o53′N, 09o07′E). The soil is classified as an Orthic Haplohumod a.

Name Liming pH b C N Clay Silt Fine sand Coarse sand t ha-1 g/kg soil ------------------ g/kg soil --------------

Low pH 0 3.6 11.7 0.6 Medium pH 4 5.4 10.5 0.6 40 40 170 750 High pH 12 7.1 13.4 0.8

a Soil Survey Staff. Soil Taxonomy: A Basic System of Soil Classification for Making and Interpreting Soil Surveys. 2nd edition, Natural Resources Conservation Service, United States Department of Agriculture (1999). b pH measured in 0.01 M CaCl2 in a 1:2.5 soil:solution ratio.

Extended Data Table 4. Basic properties of soils from the Askov LTE on animal manure and mineral fertilizers initiated in 1894. The experiment is located at Askov Experimental Station, Southern Jutland, Denmark (55o28′N, 09o07′E). The soil is classified as an Ultic Hapludalf a.

Name pH b C N Clay Silt Fine sand Coarse sand g/kg soil ------------------ g/kg soil --------------

Unfertilized 6.6 11.1 0.9 NPK fertilizer 6.2 12.9 1.0 100 120 430 350 Animal Manure 6.4 13.4 1.2

a Soil Survey Staff. Soil Taxonomy: A Basic System of Soil Classification for Making and Interpreting Soil Surveys. 2nd edition, Natural Resources Conservation Service, United States Department of Agriculture (1999). b pH measured in 0.01 M CaCl2 in a 1:2.5 soil:solution ratio.

Extended Data Fig. 4. Micro-lysimeter setup with the soil packed in an insert unit fitting 50 ml centrifugal tubes, which allows rapid sampling of soil solution via centrifugation. Micro-lysimeters were constructed using the insert unit from the 50 ml Macrosep® centrifugal tubes (Pall Coorporation, Ann Arbor, MI, USA) after removal of the vertical filter-piece. Constructing micro-lysimeters in the insert-unit allowed rapid sampling of soil solution via centrifugation and the use of a soil quantity great enough to conduct multiple analyses of both soil and soil solution after treatments with triple-labeled DON. The micro-lysimeters were packed from below of a glass microfiber filter (Whatman GF/A filter, 25 mm, GE Healthcare Life Sciences), a piece of silk organza cloth, and another glass microfiber filter. On top, 7 g of purified sea sand (0.1 - 0.315 mm, analytical grade, Merck KGaA, Darmstadt, Germany) was packed by adding 5 ml of water followed by centrifugation for 5 minutes at 5000g. The micro-lysimeters were added 12 g of field moist soil, which was gently packed by tapping on the insert unit.

Insert unit setup:

‐ 25 mm soil layer (12 g fresh weight)‐ 10 mm sea sand layer (7 g dry weight)‐ GF/A filter‐ Disk of silk organza cloth‐ GF/A filter

Extended Data Table 5. Overview of individual PLFA’s used as specific for gram-positive bacteria, gram-negative bacteria, and fungi, and individual PLFA’s not specified for microbial groups.

Group Name References

Gram positive bacteria a15:0, i15:0, i16:0, i17:0 48

Gram negative bacteria 16:1w7c, 18:1w9c 49

Fungi 18:2w6,9 50

Unspecified

14:0, 15:0, 16:0, 17:0, 18:0, 19:0, 18:1w9t

86

Paper 3

Newly depolymerized large organic N contributes directly to maize amino acid uptake

Kirsten Lønne Enggrob, Charlotte Marie Jakobsen, Ingeborg Frøsig Pedersen, Jim Rasmussen

Prepared for submission to New Phytologist

1  

Title: Newly depolymerized large organic N contributes directly to amino acid uptake in young 1 

maize plants 2 

Kirsten Lønne Enggrob,1 Charlotte Marie Jakobsen,1 Ingeborg Frøsig Pedersen,1 and Jim 4 

Rasmussen1* 5 

1Department of Agroecology, Faculty of Science and Technology, Aarhus University, Denmark. 7 

*Corresponding author: Post Box 50, 8830 Tjele, Denmark, [email protected]

Total word count for the main body of the text: 5777 10 

Word count for Introduction: 1086 11 

Word count for Material and Methods: 2048 12 

Word count for Results: 847 13 

Word count for Discussion: 1796 14 

Word count for Acknowledgement: 31 15 

Number of Figures: 5 16 

Number of Tables: 3 17 

Number of Supporting Material: 1 18 

19 

20 

Summary 21 

The contribution of large molecular size organic nitrogen (N) to microbial and plant carbon 22 

(C) and N nutrition is unclear. 23 

Soils with and without maize at three pH levels were added (14C, 13C, 15N) triple-labeled 24 

>100 kDa organic N. Soil and maize sampled 48 hours after addition of organic N were 25 

analyzed by bulk and compound specific isotope analysis to study plant and microbial 13C 26 

and 15N uptake. 27 

Mineralization of >100 kDa organic N increased with soil pH in soil without maize, but no 28 

effect of soil pH was seen for soil with maize. The >100 kDa organic N disappeared rapidly 29 

in soils with and without maize, but surprisingly more >100 kDa organic N derived amino 30 

2  

acids remained in soil with than without maize – most likely in the microbial biomass. Total 31 

15N uptake in maize increased with higher soil pH and the organic N uptake estimated to 32 

account for 20-30% of the total 15N uptake across the soil pH gradient. Organic N uptake 33 

was confirmed by presence of 13C-labeled amino acids in the maize roots. 34 

The study shows that when bio-available N is derived from large molecular size organic N 35 

then the importance of plant organic N uptake increases, and that rhizosphere 36 

microorganisms increase anabolic utilization of organic N compared to bulk soil 37 

microorganisms. 38 

39 

Keywords: amino acid, large molecular size organic N, maize, organic N uptake, soil pH, stable 40 

isotope probing. 41 

42 

1. Introduction 43 

Insufficient nitrogen (N) supply limits crop production and N fertilizer are supplied to meet this 44 

demand, although surplus N inputs in the agricultural sector has adverse environmental effects. We 45 

thus need to enhance the use efficiency of N cycling in agricultural systems (Lassaletta et al., 2014). 46 

Nitrogen entering or cycling in soil bound in organic form poses a great challenge as it is hard to 47 

achieve synchrony of plant-available N and plant N demands. Most N bound in organic form enters 48 

soil as amino acids bound in proteinaceous material (Jan et al., 2009), and organic N in soil organic 49 

matter is dominated by amino acids bound in peptides or proteins (polypeptides) (Knicker, 2011). In 50 

order to become plant available this organic N needs to be depolymerized to smaller organic N 51 

forms like small peptides and amino acids or further mineralized to inorganic N forms (Fig. 1) 52 

(Schimel & Bennett, 2004; Jones et al., 2005a). This turnover of organic N sources is both plant and 53 

microbial driven. We know that small peptides and amino acids in soil solution turnover rapidly 54 

(within minutes to hours) (Jones et al., 2005a; Czaban et al., 2016b; Hill & Jones, 2019), whereas 55 

larger organic N like proteins are decomposed a slower rates (Jan et al., 2009), although when 56 

dissolved in soil solution at a rate of days rather than weeks (Enggrob et al. subm.). The proteolytic 57 

activity for large organic N depolymerization is known to be affected by among other soil pH 58 

(Godlewski & Adamczyk, 2007; Sinsabaugh et al., 2008; Vranova et al., 2013), and the presence of 59 

plants (Godlewski & Adamczyk, 2007) where rhizosphere soil have a more active microbial 60 

community than bulk soil (Clarholm, 1985; Blagodatskaya et al., 2014). Thus, we expect a higher 61 

turnover of organic N in the rhizosphere than in bulk soil. Furthermore, soil pH generally affect the 62 

3  

active microbial communities with low soil pH favoring fungi and higher soil pH favoring bacteria 63 

(Rousk & Baath, 2011), although we recently showed that gram-positive bacteria mainly benefit 64 

from dissolved organic N turnover irrespective of soil pH (Enggrob et al. subm.). However, we lack 65 

knowledge of the short-term contribution of N bound in large organic N forms to plant N nutrition. 66 

67 

In fertile soil crop N uptake is viewed as dominated by the inorganic N forms (Schimel & Bennett, 68 

2004; Hill & Jones, 2019), although most plants have the capacity for organic N uptake (Chapin et 69 

al., 1993; Kielland et al., 2006; Näsholm et al., 2009; Paungfoo-Lonhienne et al., 2012). Our main 70 

knowledge of plant organic N uptake comes from studies with addition of individual amino acids 71 

e.g. glycine (Näsholm et al., 1998), alanine (Hill & Jones, 2019), asparagine (Czaban et al., 2016a; 72 

Czaban et al., 2018), glutamate (Jones et al., 2013), mixtures of amino acids (Forsum et al., 2008; 73 

Jämtgard et al., 2008; Sauheitl et al., 2009b) or short peptides (Paungfoo-Lonhienne et al., 2009; 74 

Hill et al., 2011a; Soper et al., 2011). These studies confirms direct plant uptake of organic N, but 75 

also point to that the microbial competition for these small organic N compounds may reduce the 76 

importance of organic N uptake for crop N nutrition (Jones et al., 2013; Hill & Jones, 2019). 77 

However, studies with the addition of a single pulse of one or a few amino acids may not reflect 78 

conditions in soil (Hill & Jones, 2019), where amino acids and peptides released via 79 

depolymerization of proteins most likely would be present at lower and more constant 80 

concentrations. Furthermore, release of N from protein-sized organic N would give a broader 81 

profile of small organic N compounds (i.e. more individual amino acids and small peptides) than 82 

previously studied. Thus, there is a lack of studies on the release of N from protein-sized organic N 83 

and the subsequent contribution to plants N nutrition. 84 

85 

Uptake of organic N is mainly determined using (13C, 15N) dual-labeled compounds with the uptake 86 

estimated based on the ratio of bulk 13C and 15N isotope uptake (Näsholm et al., 1998), and the 87 

intact uptake of the added compound confirmed by compound specific isotope analysis (Näsholm et 88 

al., 2001; Sauheitl et al., 2009a; Czaban et al., 2016a). Although these methods have mainly been 89 

used for individual amino acids they ought to be applicable in studies of larger dual-labeled organic 90 

N, even though it may pose several challenges: Firstly, since the large organic N most likely needs 91 

to undergo depolymerization before being plant available (Fig. 1), the short chase periods normally 92 

recommended for the study of intact amino acid uptake (Näsholm et al., 2009; Hill & Jones, 2019) 93 

cannot be used. Sufficient time must be allowed for the depolymerization to small peptides or 94 

4  

individual amino acids. Secondly, upon depolymerization we know presently little of the strength of 95 

microbial competition towards released peptides, although based on studies of small peptides and 96 

individual amino acids we expect microorganisms to better than plants at taking up small organic N 97 

(Jones et al., 2013; Hill & Jones, 2019). Recently, we showed that in unplanted soil increasing the 98 

molecular size of organic N changed the microbial catabolism-anabolism balance towards greater 99 

anabolism (Enggrob et al. subm.), but we have a lack of knowledge of this microbial balance in the 100 

presence of plants. Thirdly, given that the protein-sized organic N most likely needs 101 

depolymerization prior to plant uptake, then the estimation of intact organic N uptake cannot be 102 

directly based on the quantity of labeled compound added as normally done with bulk or 103 

compound-specific methods. Instead, it must be related to the release of bio-available organic N, 104 

which is difficult to estimate non-destructively. Finally, the ‘normal’ potential biases like C-tracer 105 

uptake via bicarbonate (Rasmussen & Kuzyakov, 2009; Rasmussen et al., 2010) or keto acids (Hill 106 

& Jones, 2019) also needs to be taken into account. 107 

108 

The aim of the present study was to investigate the turnover of protein-sized organic N (>100 kDa) 109 

and determine the contribution of this organic N source to the N nutrition in young maize plants 110 

grown in soil with a pH gradient. We hypothesized that (i) the turnover of >100 kDa organic N 111 

would be greater in the presence of plants due to higher microbial activity and more plant and 112 

microbially derived exo-enzymes in the rhizosphere, and (ii) higher plant growth and greater 113 

mineralization with increasing soil pH would result in a greater total 15N uptake from >100 kDa 114 

organic N at high soil pH, but a greater proportion of the total 15N uptake being in organic form at 115 

low pH due to lower mineralization. 116 

117 

2. Material and Methods 118 

An experiment was carried out with maize in 20 ml ”pots” (hence fort termed micro-lysimeters) 119 

where the mineralization and bulk uptake of C- and N-tracer from triple-labeled >100 kDa organic 120 

N was investigated in soil with three pH levels established through long-term liming. 121 

122 

2.1 Soils 123 

Soils came from the Jyndevad long-term field experiment (LTE) on liming and phosphorus initiated 124 

in 1942 (Rubaek, 2008) located at St. Jyndevad Experimental Station, Southern Jutland, Denmark 125 

(54˚53’N, 09˚07’E). The soil is coarse sandy with 40 g kg-1 clay, 40 g kg-1 silt, 170 g kg-1 fine sand, 126 

5  

and 750 g kg-1 coarse sand in the plough layer and classified as an Ultic Hapludalf (Soil Survey 127 

Staff, 1999), which has been used for spring barley cultivation for at least 30 years prior to 128 

sampling. Soil for the experiment was sampled in August 2015 from the plough layer (5-20 cm) of 129 

the V1 field in the treatments receiving 15.6 kg P ha-1 year-1. Soil was taken from plots receiving 0, 130 

4, or 12 Mg lime ha-1, which had contrasting pHCaCl2 levels (Table 1); with pH measured in 0.01 M 131 

CaCl2 in a 1:2.5 soil:solution ratio. Soil was sieved (4 mm) to remove visible roots and stored at 2ºC 132 

until the experiment. 133 

134 

2.2. Triple-labeled >100 kDa organic N 135 

The protein-sized organic N (>100 kDa) was produced from greenhouse grown triple-labeled (14C, 136 

13C, 15N) white clover shoots (Enggrob et al., 2019). Briefly, white clover grown in sterile sand 137 

received a standard nutrient solution supplemented with 15N-labeled (15NH4)2SO4 (98 at%) and 138 

clover was C-labeled via 14CO2 and 13CO2 as described by Rasmussen et al. (2008). Upon harvest, 139 

clover shoots were passed through a screw press, (Colas et al., 2013) to obtain a juice, which 140 

following was filtrated through a 0.45µm filter and Mw size fractionized using 20 ml centrifugal 141 

filter tubes (Macrosep® Advance, Pall Corporation, Ann Arbor, MI, USA) with a pore size of 100 142 

kDa. The residue remaining on the filter after three times rising with 5 ml MilliQ water was used 143 

for the experiment; thus, the >100 kDa organic N was in the size >100 kDa and <0.45 µm (hereafter 144 

referred as “>100 kDa organic N”). The >100 kDa organic N was characterized for total C and N, 145 

bulk isotopic and amino acid specific composition as described in Enggrob et al. (2019) (Table 2, 146 

Fig. S1). 147 

148 

2.3. Micro-lysimeter experiment 149 

The micro-lysimeters were constructed in the 20 ml inset of 50 ml centrifugal tubes (Maxi-Spin 150 

Filter tubes XPE-45, Ciro, Deerfield Beach, Florida, USA) to allow free drainage from the soil 151 

reducing the risk of anaerobic conditions. Each inset was filled with 15 g of field moist soil and 152 

gently packed by tapping on the side of the insert unit. In addition to the maize receiving the triple-153 

labeled >100 kDa organic N we had the following control treatments: (i) maize with water added 154 

(i.e. no organic N), (ii) soil without maize receiving the >100 kDa organic N, and (iii) soil without 155 

maize added water. In all treatments and controls end-point sampling of soil and plant tissue was 156 

done 48 hours after addition of the >100 kDa organic N (or water for the respective controls). The 157 

48 hours was based on the mineralization pattern of the >100 kDa organic N in the same soils in a 158 

6  

previous study (Enggrob et al. subm.) recognizing the need for short chase periods in organic N 159 

uptake studies (Näsholm et al., 2009; Hill & Jones, 2019) and allowing sufficient time for 160 

depolymerization to occur assuming that maize is unable to directly take up protein-size organic N. 161 

Thus, the duration of the treatments with >100 kDa organic N was a balance between a short chase 162 

period known to be important for detection of intact amino acids in plant roots and allowing time 163 

for depolymerization of the >100 kDa organic N. 164 

165 

2.3.1. Control soil without maize – >100 kDa organic N incubation 166 

Soil without plants was incubated with the triple-labeled >100 kDa organic N or with water for 48 167 

hours in a glass jar setup (Enggrob et al. subm.). Briefly, micro-lysimeters were placed in a 1 L 168 

glass jar together with a base trap containing 1 ml 1 M NaOH, and a beaker containing 2 ml water 169 

to insure humidity. Prior to incubation the soils were allowed to temperature adjust for a day. The 170 

incubation started by addition of 2.0 ml the >100 kDa organic N or in the control 2.0 ml of MilliQ 171 

water. The C and N quantity added in the >100 kDa organic N corresponded to 117 µg C g-1 soil 172 

and 22 µg N g-1 soil. 173 

174 

Mineralization to 14CO2 of the >100 kDa organic N was measured by exchanging the base trap after 175 

1, 2, 4, 24 and 48h. Immediately after removal of the base trap, 4 ml of liquid scintillation cocktail 176 

(OptiPhase HiSafe3, PerkinElmer, Waltham, MA, USA) was added to the base trap and it was 177 

counted for 14C-activity on a Tri-Carb® 2910TR Liquid Scintillation Analyzer (PerkinElmer, 178 

Waltham, MA, USA). After 48 hours, the incubation was terminated and the soil was immediately 179 

frozen. The soil was freeze dried and ground for two minutes to fine powder in 2 ml Eppendorf 180 

tubes prior to analysis. 181 

182 

2.3.2. Maize treatment and control with plant 183 

Maize seeds (variety LG31.218, Limagrain A/S, Horsens, Denmark) were germinated in the dark 184 

for two days at room temperature before transplanting to the micro-lysimeters. The maize were 185 

grown in the laboratory for 20 days (14 hour day length, 24-28 ˚C, irrigated when needed) reaching 186 

the BBCH growth stage 12-13 (Meier, 2001) prior to the addition of the >100 kDa organic N. At 187 

the time of addition of the >100 kDa organic N maize roots had occupied the whole soil volume 188 

effectively making all soil ‘rhizosphere soil’ (Jones et al., 2005b; Rasmussen et al., 2010). 189 

190 

7  

Before the >100 kDa organic N treatment, maize plants were placed in glass jars with a setup as 191 

described above, with the exception that the lid of the jar had a hole where maize shoot was gently 192 

pulled through. The shoot was then sealed from the root and soil inside the glass jar by placing an 193 

inert plastic material (UNIGUM Sanitary putty, Unipak A/S, Galten, Denmark) around the stem 194 

making the hole airtight. Again, the incubation started by addition of 2.0 ml the >100 kDa organic 195 

N or in the control 2.0 ml of MilliQ water, this time with a syringe through the hole in the lid. The 196 

mineralization of the >100 kDa organic N was determined as an end-point sampling after 48 hours 197 

of incubation, where the base trap was removed and analyzed for 14C-activity as described above. 198 

Upon termination, the shoots were cut at the soils surface and the roots were gently shaken free of 199 

the soil and firstly rinsed with demineralized water and secondly rinsed with a 1M KCl solution, to 200 

remove any DON solution sorbed to the root surface. The shoot, root and soils were then 201 

immediately frozen, freeze dried and ground to fine powder as described above. 202 

203 

2.4. Analysis 204 

2.4.1. Bulk 13C and 15N analysis 205 

The total C and N, and 13C and 15N stable isotope composition was determined by transferring 5-7 206 

mg shoot or root material to tin capsules before analysis on a PDZ Europa ANCA-GSL elemental 207 

analyzer interfaced to a PDZ Europe 20-20 isotope ratio mass spectrometer (Sercon Ltd. Cheshire, 208 

UK) at the UC Davis Stable Isotope Facility. 209 

210 

2.4.2. GC-C-IRMS analysis of bound amino acids 211 

For the GC-C-IRMS analysis of bound amino acids approximately 800 mg ground soil and 212 

approximately 70 mg ground root sample were weighted into separate 16x100 soda-lime disposable 213 

test tubes (Duran Group, Mainz, Germany).The samples were hydrolyzed as followed: 2 ml 6 M 214 

HCl was added to each sample and heated to 110 °C for 20 hours. To remove solids and lipophilic 215 

compounds 4 ml n-hexane/dichloromethane (6:5, v/v) was added to the soil samples and 2 ml to the 216 

root samples. After mixing and centrifugation, the aquatic phase was transferred through a Pasteur 217 

pipette lined with glass wool followed by 2 x 0.5 mL 0.1 M HCl into new test tubes and the internal 218 

standard was added. After freeze drying and resolving in 1 ml 0.01 M HCl, the sample was 219 

transferred to a polypropylene column containing 2 g Dowex 50w x 8 cation exchange resin for 220 

separation of compounds containing N and compounds not containing N. After eluting the amino 221 

8  

acid with 2.5 M ammonium hydroxide solution, the sample was freeze-dried and derivatizied 222 

according to Enggrob et al. (2019). 223 

224 

The GC-IRMS analyses were performed as described by Enggrob et al. (2019), briefly: A VF-23m 225 

capillary column (60 m x 0.25 mm inner diameter x 0.25 µm film thickness, Agilent Technologies, 226 

Amstelveen, Netherland) fittet in a Trace GC Ultra mounted with a TriPlus autosampler (both from 227 

Thermo Scientific, Hvidovre, DK), were used to separate the derivate. The inlet was operating at 228 

250°C in splitless mode, with a Helium column flow of 1.4 mL min-1. The gas chromatograph was 229 

coupled via a combustion reactor (GC IsoLink, Thermo Scientific, Hvidovre, DK), oxidation at 230 

1000°C, to an isotope ratio mass spectrometer (Delta V Plus IRMS, Thermo Scientific, Hvidovre, 231 

DK). All MS related parameters were controlled by the Isodat software version 3.0 (Thermo 232 

Scientific, Hvidovre, DK). All δ13C values were reported relative to the Vienna PeeDee Belemnite 233 

(VPDB) international isotope standard. All δ15N values were reported relative to the δ15N values of 234 

atmospheric N2. A standard curve based on analyses of Asparagine (Asn) with an increasing 235 

percentages of fully-labeled Asn (13C-4, 15N-2), showed a strong linearity of all δ13C and δ15N 236 

values, with increasing amounts of fully-labeled Asn (13C-4, 15N-2), with coefficient of 237 

determination of R2=0.984 and R2=0.982 for 13C and 15N, respectively. The AAs were identified by 238 

the retention time of standards and the concentration calculated relative to individual standard 239 

curves. 240 

241 

2.5. Calculations and statistical analysis 242 

Mineralization of the >100 kDa organic N to 14CO2 was calculated as percent of added, where the 243 

measured 14C-activity in the base traps were background corrected based on the respective controls 244 

added water, and then divided by the 14C-activity added initially. Controls were soil added water for 245 

the treatment without maize, and maize and soil added water for the treatments with plants. Total 246 

uptake of 13C and 15N in maize shoot and root was calculated from the excess quantify of tracer 247 

using maize receiving water as natural abundance backgrounds, and expressed as percent of added 248 

13C and 15N with the >100 kDa organic N. 249 

250 

The concentration of amino acid in hydrolyzed soils and roots was calculated based on individual 251 

standard curves for each amino acid together with internal standard present in each sample. The 252 

13C-labeled amino acids remaining in soil 48 hours after addition of the 100 kDa organic N was 253 

9  

calculated as the concentration of each individual amino acid in soil multiplied by the 13C at% 254 

excess of the particular amino acid using the respective controls receiving water as natural 255 

abundance backgrounds, and expressed as percent of 13C added in each individual amino acid in the 256 

>100 kDa organic N. The 15N uptake occurring in organic form was estimated as the ratio between 257 

total 13C and total 15N uptake in the whole plant. The presence of 13C-labeled amino acids in roots 258 

after 48 hours was calculated as the concentration of each individual amino acid in root multiplied 259 

by the 13C at% excess of the particular amino acid using control maize added water as natural 260 

abundance backgrounds, and expresses as percent of 13C added in each individual amino acid in the 261 

>100 kDa organic N. The specific enrichment of individual amino acids in root was calculated as 262 

the ratio between the total 13C amount and the total C amount in each individual amino acid. 263 

264 

The influence of the soil pH levels and the presence of maize on mineralization to 14CO2, bulk 13C 265 

and 15N in maize shoot and root, total amino acids content and 13C-labeled amino acid content in 266 

soil and maize roots were tested with a linear mixed-effects model using the statistical analysis 267 

program R version 3.5.1 using R-package lme4 (RCoreTeam, 2018). For statistical analysis, the 268 

datasets were divided into subsets, each subset were tested for normal distribution by the Shapiro-269 

Wilk normality test (Royston, 1982). For each subset, a two-sample t-test comparing the least-mean 270 

squares was conducted using R-package emmeans. Significance was declared at P ≤ 0.05. 271 

272 

3. Results 273 

Maize grew better in soil at medium and high pH than at the low pH level. Maize seedlings at the 274 

soil low pH level had significantly lower shoot, root and total dry matter yields than maize in the 275 

medium and high soil pH levels (Table 3). In presence of maize, the addition of the >100 kDa 276 

organic N did not significantly affect the concentrations of individual amino acids in hydrolyzed 277 

soil irrespective of pH level (Fig. S2) nor was the concentration of individual amino acids in the 278 

root samples generally affected (Fig. S2). 279 

280 

3.1. Mineralization of >100 kDa organic N with and without maize 281 

The mineralization of the >100 kDa organic N in soil without followed first order kinetics with 282 

detection of 14CO2 already after 1 hour across the pH gradient (Fig. 2A). The accumulated 283 

mineralization was significantly higher (P = 0.0259) in the high pH soil than the low pH soil (Fig. 284 

2B). Mineralization in the high pH soil was 9.2 ± 0.7 % of added 14C as 14CO2 compared to 6.2 ± 285 

10  

0.6 % of added in the low pH soil, whereas the medium pH soil had intermediate 14CO2 evolution 286 

with 7.0 ± 0.3 % of added after 48 hours (Fig. 2B). Interestingly, these differences in mineralization 287 

across soil pH levels disappeared in the presence of maize, where there were no significant 288 

differences between mineralization across the pH gradient (Fig. 2B). 289 

290 

3.2. 13C-labeled amino acids remaining in soil with and without maize 291 

After 48 hours 6 to 50% of the bound amino acids added in the >100 kDa organic N remained with 292 

intact C-skeletons in the soils without maize (Fig. 3). There was a considerable variation in the 293 

proportion remaining among individual amino acids across the pH gradient with 13-50%, 16-50% 294 

and 6-38% of individual amino acids remaining in the soils with low, medium and high pH levels, 295 

respectively. Surprisingly, in the presence of maize the general pattern was that significantly higher 296 

proportions of individual amino acids remained in the soil compared to the soil without maize (Fig. 297 

3); except lysine where similar proportions remained, and tyrosine at all pH levels and 298 

phenylalanine at the low pH level where greater proportions remained in the soil without maize. 299 

The lowest proportions of individual amino acid remaining was 19, 19 and 18 % of the added at the 300 

low, medium and high soil pH levels, respectively, with the highest proportions remaining reaching 301 

as much as 78% of added (alanine in the medium pH soil). In line with the controls without maize, 302 

there was also in the presence of maize a considerable variation in the proportion remaining among 303 

individual amino acids with 19-74%, 19-78% and 18-66% of the added 13C in individual amino 304 

acids remaining in the low, medium and high pH level soils, respectively. 305 

306 

3.3. Bulk 13C and 15N uptake in maize 307 

The uptake of 15N was significantly (P < 0.001) greater than the uptake of 13C across the soil pH 308 

gradient (Fig. 4). The total uptake of 15N ranged from 6.5 to 12.0 % of 15N added the >100 kDa 309 

fraction (Fig. 4c) with the 15N equally distributed among the roots and the shoots (Fig. 4a,b). The 310 

15N uptake was significantly higher in the medium and high pH soils than in the low pH soil in line 311 

with the mineralization pattern found for the soils without maize. However, we found no correlation 312 

between 15N uptake and the actual 14CO2 evolution in the soils with maize. The uptake of 13C 48 313 

hours after addition of the >100 kDa organic N was significantly (P < 0.001) higher in roots (1.4 to 314 

2.2 % of 13C added) than in the shoots (0.4 to 0.5% of 13C added). Although the uptake of 13C in 315 

roots tended to be lower in the low pH soils there was no significant differences across the soil pH 316 

gradient (Fig. 4b). On a whole plant basis, the 13C-to-15N uptake-ratio was 28 ± 5%, 20 ± 1% and 22 317 

11  

± 4% in the low, medium and high pH soil, respectively with no significant differences in the 13C-318 

to-15N uptake-ratio across the soil pH gradient. 319 

320 

3.4. Presence of 13C-labeled amino acids in maize 321 

The presence of individual 13C-labeled amino acids varied significantly in maize roots at all three 322 

soil pH levels (Fig. 5). The presence ranged from 0 to 1.7% of the 13C added with the >100 kDa 323 

organic N across soils; with no significant effects of soil pH level on the presence of individual 13C-324 

labeled amino acids. The presence of individual amino acids had a similar pattern across the soil pH 325 

gradient where glutamine/glutamate, proline/threonine and leucine had the greatest presence 326 

reaching close to 2% of added, and lysine had the lowest presence throughout. The pattern of amino 327 

acid presence in maize roots did not resemble the pattern of neither amino acids remaining in the 328 

soil nor amino acids lost from the soil. The average presence of 13C-labeled amino acids in maize 329 

roots was 0.5-0.6% of the added. Thus, the 13C in the root amino acids corresponded to one third of 330 

the bulk 13C recovered. 331 

332 

4. Discussion 333 

334 

4.1. Mineralization with and without maize 335 

Mineralization of the >100 kDa organic N in the soils without maize started immediately across the 336 

soil pH gradient, confirming that the microbial toolbox for depolymerization of large organic N was 337 

ready and available irrespective of soil pH level (Enggrob et al. subm.). Individual amino acids are 338 

completely removed from soil solution (Czaban et al., 2016b; Hill & Jones, 2019) and mineralized 339 

(Wilkinson et al., 2014; Hill & Jones, 2019) within minutes to hours. Clearly the >100 kDa organic 340 

N studied here was not respired at similar rates with the 14CO2 evolution expected to continue 341 

beyond the 48 hours chase period used here (Enggrob et al. subm.). The significant effect of pH 342 

level on accumulated 14CO2 in soil without maize disappeared in the presence of maize. This change 343 

in respiration pattern in especially the low pH soil must be related to higher microbial or plant 344 

respiration. Increased microbial respiration could be due to a more active microbial biomass in 345 

rhizosphere soil (Blagodatskaya et al., 2014), a greater overall turnover of the >100 kDa organic N 346 

added, or a shift from anabolism to catabolism in the microbial community (Liang et al., 2017). 347 

Plants may also have contributed to the 14CO2 respiration organic compounds from the >100 kDa 348 

12  

organic N were taken up and used as energy source in the maize roots (Näsholm et al., 2009; Hill et 349 

al., 2011b; Warren, 2012; Hildebrandt et al., 2015). 350 

351 

4.2. Added >100 kDa organic N remaining in soil with and without maize 352 

In the soils without maize, 6 to 50% of individual amino acids added with the >100 kDa organic N 353 

remained in the soil 48 hours after the addition. Assuming an equal degradation of proteins in the 354 

added >100 kDa organic N, we use the individual amino acids with lowest proportions remaining as 355 

an estimate of the proportion of proteins in the >100 kDa organic N remaining intact as added. 356 

Thus, the present results strongly support that depolymerization of dissolved protein-sized organic 357 

N occurs rapidly (Enggrob et al. subm.). The lowest proportions of individual amino acids 358 

remaining in the soil without maize were 13, 16 and 6% of added at the low, medium and high soil 359 

pH levels, respectively (Fig. 3). We interpret amino acids remaining at higher proportions than the 360 

lowest (i.e. above the dashed line in Fig. 3) as representing 13C-labeled amino acids incorporated in 361 

the microbial biomass (Enggrob et al. subm.). This based on the higher proportions of simpler 362 

amino acids (fewer biosynthetic steps) remaining than the more complex amino acids (from left to 363 

right on Fig. 3). The amino acids with fewer biosynthetic steps (alanine, glutamine/glutamate, 364 

asparagine/aspartate, glycine) are among those known to be part of bacterial cell walls (Simelyte et 365 

al., 2003; Vollmer et al., 2008; Schneewind & Missiakas, 2012). Thus, the majority of the labeled 366 

amino acids recovered from soil were most likely build into microbial tissue in the soil without 367 

maize. 368 

369 

The finding of more amino acids remaining in soil with than without maize (Fig. 3) is surprising, 370 

since turnover is generally considered to be greater in rhizosphere than bulk soil (Godlewski & 371 

Adamczyk, 2007; Blagodatskaya et al., 2014). The higher proportions of amino acids remaining in 372 

the soil with than without maize point to a lower overall degradation of the added >100 kDa organic 373 

N. Plant exudation of C-rich compounds usually makes the rhizosphere N limited (Kuzyakov, 2002; 374 

Jones et al., 2013), which ought to have increased the microbial need for N mining (Kuzyakov, 375 

2010). Soil in the micro-lysimeters was expected to be low in available N as no fertilizers were 376 

added and the roots occupied the whole soil. However, N mining would lead to greater loss of C 377 

from the added >100 kDa organic N, which was evident for the 14CO2 respiration, but not for the 378 

loss of 13C-labeled amino acids from the >100 kDa organic N added. Instead, other nutrients than N 379 

may have been limited (Dijkstra et al., 2013), and thus reduced the microbial turnover of added 380 

13  

>100 kDa organic N. The higher proportion of amino acids with fewer biosynthetic steps remaining 381 

also in the soil with maize again indicate microbial incorporation of amino acids derived from the 382 

added >100 kDa organic N. This can be explained if maize exudation of C-rich compounds reduces 383 

the microbial need for using amino acid C-skeletons for energy. In addition, unrecovered root hairs 384 

and root fragments may contain 13C-labeled amino acids, which could contribute to a greater 385 

proportion of amino acids in the soil. 386 

387 

4.3. Bulk uptake and presence of intact amino acids in maize 388 

The uptake of 15N in maize after 48 hours reached up to 12 % of the added with the >100 kDa 389 

organic N. This is in the same range of the 0-26% of added 15N in alanine and tri-alanine in grass 390 

after 2.5 hours (Wilkinson et al., 2015), in the lower range of the 13-28% of added 15N in an amino 391 

acid mixture recovered in grass after 48 hours (Sauheitl et al., 2009b), and somewhat lower than the 392 

30% of added 15N in alanine recovered in wheat after 24 hours (Hill & Jones, 2019). The lower 393 

level of 15N uptake in the present study compared to studies with similar chase periods is probably 394 

due the the expected delay in production in bio-available organic N after depolymerization. The 15N 395 

uptake was markedly higher than 13C uptake (Fig. 4) showing uptake of inorganic 15N from 396 

mineralization of the >100 kDa and/or post-uptake metabolism of organic N (Näsholm et al., 2009; 397 

Warren, 2012). The presence of 13C in plant tissue indicates organic N uptake since uptake of 13C 398 

via e.g. pyruvate and bicarbonate usually cannot explain all 13C uptake from rapidly cycling amino 399 

acids. Moreover, the present setup with a sealed root system the fixation of 13C via photosynthesis 400 

must have been minimal as also shown by the lower 13C presence in shoots than roots (Fig. 4). 401 

402 

We analyzed root material for 13C-labeled amino acids by compound specific isotope analysis to 403 

confirm the uptake of amino acids derived from the >100 kDa organic N. Across the soil pH 404 

gradient, 13C-labeled amino acids were present in maize roots, which support the uptake of organic 405 

N. Interestingly, the average presence of 13C-labeled amino acids in maize roots accounted for one 406 

third of the bulk 13C uptake, which is a higher proportion than previously reported (Sauheitl et al., 407 

2009a). The 13C-labeled amino acids varied in their presence in roots, which can arise from 408 

differences in the actual bio-availability in soil, the actual uptake, and the post-uptake fate in roots 409 

of the specific amino acids. We cannot with the present data determine the uptake rates of amino 410 

acids from the >100 kDa organic N, neither can we deduce whether the organic N was taken up as 411 

free amino acids or bound in peptides (Hill et al., 2011b). However, the post-uptake fate of 412 

14  

individual amino acids may be indicated by the specific enrichment of amino acids in the maize 413 

roots (Fig. S3), where a greater specific enrichment indicate a greater recycling of amino acid C-414 

skeletons. We speculate that the post-uptake fate of amino acids is a balance between abundance of 415 

the amino acid in the plant tissue and the energy gain when using the amino acid in catabolism 416 

(Hildebrandt et al., 2015). For example, comparing leucine and tyrosine both with high energy gain 417 

in catabolism (Hildebrandt et al., 2015), we found greater presence of leucine than tyrosine both as 418 

% of added (Fig. 5) and as specific enrichment (Fig. S3), which could be related to a greater 419 

abundance of leucine in maize root tissue (Fig. S2). The difference could also be related to a greater 420 

uptake of leucine than tyrosine though. 421 

422 

4.4 Estimation of N uptake in organic form 423 

We found that 20-30% of N uptake was in organic form across the soil pH gradient assuming that 424 

the 15N uptake represents the net-plant-available N. We advocate that the N uptake in organic form 425 

should be based on the ratio between the net-uptake of 13C and 15N. We acknowledge that the 13C-426 

uptake in the present study could be affected by C-tracer uptake bias, but as shown by the 427 

compound specific isotope analysis then at least one third of the C-tracer uptake could be accounted 428 

for by 13C in amino acids. Furthermore, the C-tracer uptake in the present study does not take 13C 429 

loss from post-uptake metabolism into account. The recent study by (Hill & Jones, 2019) reports 430 

both 13C and 15N in ‘% of added’ alanine to wheat and our calculation of their data show a 4% 13C-431 

to-15N uptake-ratio after 24 hours. The higher organic N uptake in the present study is most likely 432 

related to a slower release of bio-available organic N, than with the pulse of alanine added in the 433 

Hill and Jones study. A pulse of organic N may saturate the microbial biomass (Czaban et al., 434 

2016b) and allow plants a better chance for competition (Näsholm et al., 2009). Interestingly, the 435 

20-30% net-organic N-uptake estimate in the present study is in line with the 13C-to-15N ratio-based 436 

estimates reported for e.g. glycine uptake in four grassland species (Näsholm et al., 2000), wheat 437 

(Näsholm et al., 2001), and tomato (Ge et al., 2009), and uptake of amino acid mixtures in grass 438 

(Sauheitl et al., 2009b). Hence, the present study shows that when N is added to a plant-soil system 439 

in large molecular sizes then the potential for intact organic N uptake is at least at the same level or 440 

higher than when organic N is added as individual amino acids. 441 

442 

Conclusion 443 

15  

We conducted an experiment in soils with and without maize at three pH levels with addition of 444 

triple-labeled >100 kDa organic N. Maize was grown for three weeks in micro-lysimeters prior to 445 

addition of the >100 kDa organic N creating rhizosphere soil in the whole soil volume in treatments 446 

with maize. In soil without maize, mineralization differed significantly between pH levels, whereas 447 

there was no difference in 14CO2 among soil pH levels in presence of maize. The >100 kDa organic 448 

N was rapidly turned over in soil both with and without maize, but surprisingly more amino acids 449 

derived from >100 kDa organic N remained in soil with than without maize most likely in the 450 

microbial biomass. Maize grew better at a soil pHCaCl2 of 5.4 and 7.1, and the total 15N uptake from 451 

>100 kDa organic N increased with higher soil pH reaching 12 % at a pH level of 7.1 48 hours after 452 

addition of the >100 kDa organic N. Based on the 13C-to-15N uptake-ratio we estimated that 20-30% 453 

of the N uptake occurred in organic form across the three pH levels. The presence of 13C-labeled 454 

amino acids in maize roots confirmed the organic N uptake. 455 

456 

Acknowledgement 457 

The study was financially supported by The Independent Research Fund Denmark – Technology 458 

and Production (Project no. 1335-00760B). The authors wish to thank Limagrain A/S for delivering 459 

seeds for the experiment. 460 

461 

Author contributions 462 

All authors designed and executed the experiment and analysis, K.L.E. ran the statistical analysis, 463 

K.L.E. and J.R. drafted the manuscript, and all authors revised the manuscript and approved the 464 

final version. 465 

466 

References 467 

468 

Blagodatskaya E, Blagodatsky S, Anderson TH, Kuzyakov Y. 2014. Microbial growth and 469 

carbon use efficiency in the rhizosphere and root-free soil. Plos ONE 9, e93282. 470 

471 

Chapin FS, Moilanen L, Kielland K. 1993. Preferential use of organic nitrogen for growth by a 472 

nonmycorrhizal arctic sedge. Nature 361, 150-153. 473 

474 

16  

Clarholm M. 1985. Interactions of bacteria, protozoa and plants leading to mineralization of soil-475 

nitrogen. Soil Biology & Biochemistry 17, 181-187. 476 

477 

Colas D, Doumeng C, Pontalier PY, Rigal L. 2013. Green crop fractionation by twin-screw 478 

extrusion: Influence of the screw profile on alfalfa (Medicago sativa) dehydration and protein 479 

extraction. Chemical Engineering and Processing 72, 1-9. 480 

481 

Czaban W, Jämtgard S, Näsholm T, Rasmussen J, Nicolaisen M, Fomsgaard IS. 2016a. Direct 482 

acquisition of organic N by white clover even in the presence of inorganic N. Plant and Soil 407, 483 

91-107. 484 

485 

Czaban W, Rasmussen J, Laursen BB, Vidkjær NH, Sapkota R, Nicolaisen M, Fomsgaard IS. 486 

2018. Multiple effects of secondary metabolites on amino acid cycling in white clover rhizosphere. 487 

Soil Biology & Biochemistry 123, 54-63. 488 

489 

Czaban W, Rasmussen J, Nicolaisen M, Fomsgaard IS, 2016b. Dissipation kinetics of 490 

asparagine in soil measured by compound-specific analysis with metabolite tracking. Biology and 491 

Fertility of Soils 52, 911-916. 492 

493 

Dijkstra FA, Carrillo Y, Pendall E, Morgan JA, 2013. Rhizosphere priming: a nutrient 494 

perspective. Frontiers in Microbiology 4, Article no. 216. 495 

496 

Enggrob KL, Larsen T, Larsen M, Elsgaard L, Rasmussen J. 2019. The influence of hydrolysis 497 

and derivatization on the determination of amino acid content and isotopic ratios in dual-labeled 498 

(13C, 15N) white clover. Rapid Communications in Mass Spectrometry 33, 21-30. 499 

500 

Enggrob KL, Larsen T, Rasmussen J. Molecular size doesn’t matter for turning over large 501 

organic N in soil. Nature, submitted. 502 

503 

Forsum O, Svennerstam H, Ganeteg U, Näsholm T. 2008. Capacities and constraints of amino 504 

acid utilization in Arabidopsis. New Phytologist 179, 1058-1069. 505 

506 

17  

Ge TD, Song SW, Roberts P, Jones DL, Huang DF, Iwasaki K. 2009. Amino acids as a nitrogen 507 

source for tomato seedlings: The use of dual-labeled (13C, 15N) glycine to test for direct uptake by 508 

tomato seedlings. Environmental and Experimental Botany 66, 357-361. 509 

510 

Godlewski M, Adamczyk B. 2007. The ability of plants to secrete proteases by roots. Plant 511 

Physiology and Biochemistry 45, 657-664. 512 

513 

Hildebrandt TM, Nesi AN, Araujo WL, Braun HP. 2015. Amino acid catabolism in plants. 514 

Molecular Plant 8, 1563-1579. 515 

516 

Hill PW, Farrar J, Roberts P, Farrell M, Grant H, Newsham KK, Hopkins DW, Bardgett RD, 517 

Jones DL. 2011a. Vascular plant success in a warming Antarctic may be due to efficient nitrogen 518 

acquisition. Nature Climate Change 1, 50-53. 519 

520 

Hill PW, Jones DL. 2019. Plant–microbe competition: does injection of isotopes of C and N into 521 

the rhizosphere effectively characterise plant use of soil N? New Phytologist 221, 796-806. 522 

523 

Hill PW, Quilliam RS, Deluca TH, Farrar J, Farrell M, Roberts P, Newsham KK, Hopkins 524 

DW, Bardgett RD, Jones DL. 2011b. Acquisition and assimilation of nitrogen as peptide-bound 525 

and D-enantiomers of amino acids by wheat. Plos ONE 6, e19220. 526 

527 

Jämtgard S, Näsholm T, Huss-Danell K. 2008. Characteristics of amino acid uptake in barley. 528 

Plant and Soil 302, 221-231. 529 

530 

Jan MT, Roberts P, Tonheim SK, Jones DL. 2009. Protein breakdown represents a major 531 

bottleneck in nitrogen cycling in grassland soils. Soil Biology & Biochemistry 41, 2272-2282. 532 

533 

Jones DL, Clode PL, Kilburn MR, Stockdale EA, Murphy DV. 2013. Competition between 534 

plant and bacterial cells at the microscale regulates the dynamics of nitrogen acquisition in wheat 535 

(Triticum aestivum). New Phytologist 200, 796-807. 536 

537 

18  

Jones DL, Healey JR, Willett VB, Farrar JF, Hodge A. 2005a. Dissolved organic nitrogen 538 

uptake by plants - an important N uptake pathway? Soil Biology & Biochemistry 37, 413-423. 539 

540 

Jones DL, Shannon D, Junvee-Fortune T, Farrar JF. 2005b. Plant capture of free amino acids is 541 

maximized under high soil amino acid concentrations. Soil Biology & Biochemistry 37, 179-181. 542 

543 

Kielland K, McFarland J, Olson K. 2006. Amino acid uptake in deciduous and coniferous taiga 544 

ecosystems. Plant and Soil 288, 297-307. 545 

546 

Knicker H. 2011. Soil organic N - An under-rated player for C sequestration in soils? Soil Biology 547 

& Biochemistry 43, 1118-1129. 548 

549 

Kuzyakov Y. 2002. Review: Factors affecting rhizosphere priming effects. Journal of Plant 550 

Nutrition and Soil Science 165, 382-396. 551 

552 

Kuzyakov Y. 2010. Priming effects: Interactions between living and dead organic matter. Soil 553 

Biology & Biochemistry 42, 1363-1371. 554 

555 

Lassaletta L, Billen G, Grizzetti B, Anglade J, Garnier J. 2014. 50 year trends in nitrogen use 556 

efficiency of world cropping systems: the relationship between yield and nitrogen input to cropland. 557 

Environmental Research Letters 9, Article no. 105011. 558 

559 

Liang C, Schimel JP, Jastrow JD. 2017. The importance of anabolism in microbial control over 560 

soil carbon storage. Nature Microbiology 2, Article no. 17105. 561 

562 

Meier U. 2001. Growth stages of mono-and dicotyledonous plants. BBCH Monographs 2nd edition. 563 

Federal Biological Research Centre for Agriculture and Forestry, Berlin, Germany. 564 

565 

Näsholm T, Ekblad A, Nordin A, Giesler R, Högberg M, Högberg P. 1998. Boreal forest plants 566 

take up organic nitrogen. Nature 392, 914-916. 567 

568 

19  

Näsholm T, Huss-Danell K, Högberg P. 2000. Uptake of organic nitrogen in the field by four 569 

agriculturally important plant species. Ecology 81, 1155-1161. 570 

571 

Näsholm T, Huss-Danell K, Högberg P. 2001. Uptake of glycine by field grown wheat. New 572 

Phytologist 150, 59-63. 573 

574 

Näsholm T, Kielland K, Ganeteg U, 2009. Uptake of organic nitrogen by plants. New Phytologist 575 

182, 31-48. 576 

577 

Paungfoo-Lonhienne C, Schenk PM, Lonhienne TGA, Brackin R, Meier S, Rentsch D, 578 

Schmidt S. 2009. Nitrogen affects cluster root formation and expression of putative peptide 579 

transporters. Journal of Experimental Botany 60, 2665-2676. 580 

581 

Paungfoo-Lonhienne C, Visser J, Lonhienne TGA, Schmidt S. 2012. Past, present and future of 582 

organic nutrients. Plant and Soil 359, 1-18. 583 

584 

Rasmussen J, Gjettermann B, Eriksen J, Jensen ES, Høgh-Jensen H. 2008. Fate of 15N and 14C 585 

from labelled plant material: Recovery in perennial ryegrass-clover mixtures and in pore water of 586 

the sward. Soil Biology & Biochemistry 40, 3031-3039. 587 

588 

Rasmussen J, Kuzyakov Y. 2009. Carbon isotopes as proof for plant uptake of organic nitrogen: 589 

Relevance of inorganic carbon uptake. Soil Biology & Biochemistry 41, 1586-1587. 590 

591 

Rasmussen J, Sauheitl L, Eriksen J, Kuzyakov Y. 2010. Plant uptake of dual-labeled organic N 592 

biased by inorganic C uptake: Results of a triple labeling study. Soil Biology & Biochemistry 42, 593 

524-527. 594 

595 

RCoreTeam. 2018. R: A language and environment for statistical computing. R Foundation for 596 

Statistical Computing. 597 

598 

Rousk J, Baath E. 2011. Growth of saprotrophic fungi and bacteria in soil. FEMS Microbiology 599 

Ecology 78, 17-30. 600 

20  

601 

Royston JP. 1982. An extension of Sharpiro and Wilk-W test for normality of large samples. 602 

Journal of the Royal Statistical Society. Series C Applied Statistics 31, 115-124. 603 

604 

Rubaek GH. 2008. Effects of liming and P fertilisation. In: Long continued agricultural soil 605 

experiments: a Nordic platform – a catalogue. Editors Petersen J, Mattsson L, Reiley H, Sala T, 606 

Thorvaldsson G, Christensen BT. DJF Plant Science 16, 12-14, Aarhus University, Denmark. 607 

608 

Sauheitl L, Glaser B, Weigelt A. 2009a. Advantages of compound-specific stable isotope 609 

measurements over bulk measurements in studies on plant uptake of intact amino acids. Rapid 610 

Communications in Mass Spectrometry 23, 3333-3342. 611 

612 

Sauheitl L, Glaser B, Weigelt A. 2009b. Uptake of intact amino acids by plants depends on soil 613 

amino acid concentrations. Environmental and Experimental Botany 66, 145-152. 614 

615 

Schimel JP, Bennett J. 2004. Nitrogen mineralization: Challenges of a changing paradigm. 616 

Ecology 85, 591-602. 617 

618 

Schneewind O, Missiakas DM. 2012. Protein secretion and surface display in Gram-positive 619 

bacteria. Philosophical Transactions of the Royal Society B-Biological Sciences 367, 1123-1139. 620 

621 

Simelyte E, Rimpilainen M, Zhang X, Toivanen P. 2003. Role of peptidoglycan subtypes in the 622 

pathogenesis of bacterial cell wall arthritis. Annals of the Rheumatic Diseases 62, 976-982. 623 

624 

Sinsabaugh RL, Lauber CL, Weintraub MN, Ahmed B, Allison SD, Crenshaw C, Contosta 625 

AR, Cusack D, Frey S, Gallo ME, Gartner TB, Hobbie SE, Holland K, Keeler BL, Powers JS, 626 

Stursova M, Takacs-Vesbach C, Waldrop MP, Wallenstein MD, Zak DR, Zeglin LH. 2008. 627 

Stoichiometry of soil enzyme activity at global scale. Ecology Letters 11, 1252-1264. 628 

629 

Soil Survey Staff. 1999. Soil Taxonomy: A Basic System of Soil Classification for Making and 630 

Interpreting Soil Surveys. Natural Resources Conservation Service, United States Department of 631 

Agriculture, Washington D.C., USA. 632 

21  

633 

Soper FM, Paungfoo-Lonhienne C, Brackin R, Rentsch D, Schmidt S, Robinson N. 2011. 634 

Arabidopsis and Lobelia anceps access small peptides as a nitrogen source for growth. Functional 635 

Plant Biology 38, 788-796. 636 

637 

Vollmer W, Blanot D, de Pedro MA. 2008. Peptidoglycan structure and architecture. FEMS 638 

Microbiology Reviews 32, 149-167. 639 

640 

Vranova V, Rejsek K, Formanek P. 2013. Proteolytic activity in soil: A review. Applied Soil 641 

Ecology 70, 23-32. 642 

643 

Warren CR. 2012. Post-uptake metabolism affects quantification of amino acid uptake. New 644 

Phytologist 193, 522-531. 645 

646 

Wilkinson A, Hill PW, Farrar JF, Jones DL, Bardgett RD. 2014. Rapid microbial uptake and 647 

mineralization of amino acids and peptides along a grassland productivity gradient. Soil Biology & 648 

Biochemistry 72, 75-83. 649 

650 

Wilkinson A, Hill PW, Vaieretti MV, Farrar JF, Jones DL, Bardgett RD. 2015. Challenging 651 

the paradigm of nitrogen cycling: no evidence of in situ resource partitioning by coexisting plant 652 

species in grasslands of contrasting fertility. Ecology and Evolution 5, 275-287. 653 

Table 1. Basic properties of the three Jyndevad soils.

Name Liming pH1 C N [t ha-1] [g kg-1 soil]

Low pH 0 3.6 11.7 0.6 Medium pH 4 5.4 10.5 0.6 High pH 12 7.1 13.4 0.8

1 pH measured in 0.01 M CaCl2 in a 1:2.5 soil:solution ratio.  

 

Table 2. Composition of the >100 kDa organic N used in the experiment; 13C and 15N enrichment expressed as atom fraction (AF) of the isotope. Data is given as mean ± standard error (n = 4).

Fraction C quantity N quantity C/N ratio 14C-activity AF 13C AF 15N [mg ml-1] [mg ml-1] [Bq ml-1]

>100 kDa 0.84 ±0.01 0.16 ±0.001 5.0 8.11 ±0.14 0.083 ±7.6E-5 0.014 ±0.1E-5

Table 3. Dry matter yield of maize grown in Jyndevad soils at three pH levels. Mean ± standard error (n = 4). Letters show significant differences among soil pH levels within each column.

Shoot Root Total

[mg DM micro-lysimeter-1]

Low pH 87 ± 9 b 65 ± 4 b 153 ± 13 b

Medium pH 117 ± 6 ab 109 ± 14 a 226 ± 19 a

High pH 127 ± 7 a 90 ± 5 ab 216 ± 12 a

 

 

Figure 1.

Figure 1. Conceptual figure showing the routes of 15N and 13C entry into plant from protein-sized dissolved organic N (DON, >100 kDa in the present study). Initially, protein-sized organic N needs to be depolymerized to lower molecular weight (Mw) DON, which can either be directly assimilated by plants or be mineralized to inorganic N forms before plant uptake. Thus, plant 15N-enrichment is the result of either organic N or inorganic N uptake, and plant 13C-enrichment is the result of either organic C uptake or inorganic C assimilation via photosynthesis or dark fixation. The lower box indicates that at present we lack knowledge of the proportion of total N uptake occurring via organic N forms.

15N, 13C, 14CLabeled

DON>100 kDa

LowerMw DON

Plant 15N + 13C

Depolymerization

CO2 [g]HCO3

-

Inorganic NNH4

+, NO3-Mineralization

Bio-available N

Plant 15NPlant 13C

Organic N uptake

Inorganic N uptake

??

Figure 2.

Figure 2. Mineralization of >100 kDa organic N to 14CO2 in Jyndevad soils. (a) Temporal development of mineralization and (b) accumulated mineralization after 48 hours soil without and with maize. The three pH levels are low at pHCaCl2 3.6, medium at pHCaCl2 5.4, and high at pHCaCl2 7.1. Statistical differences among soil pH levels in accumulated 14CO2 after 48 hours are indicated by different letters above the bars (n = 4).

b

Without maize With maize

Min

era

liza

tio

n t

o C

O2 (

ac

cu

mu

late

d 1

4C

O2 o

f a

dd

ed

14C

, %)

0

2

4

6

8

10

12

a

Time of incubation (hours)0 10 20 30 40 50

0

2

4

6

8

10

12 Low pH (3.6)Medium pH (5.4)High pH (7.1)

aabb

- a -

Figure 3.

Figure 3. Bound amino acids from added >100 kDa organic N remaining in Jyndevad soils after 48 hours without and with maize in soil at (a) low pH, (b) medium pH, and (c) high pH. Significant differences between unplanted soil and soil with maize in 13C remaining for individual amino acids are marked by an asterisk above the bars (n=4). Amino acids are organized from left on right with increasing steps in their biosynthesis. The amino acids: asparagine and aspartate (Asx), glutamine and glutamate (Glx), and Proline and Threonine (Pro/Thr) elute together in the GC-C-IRMS analysis of acid hydrolyzed samples. The red dashed line indicate the lowest proportion of an individual amino acid remaining in soil without maize.

a: Low pH

0

20

40

60

80

100Without maizeWith maize

b: Medium pH

Bo

un

d a

min

o a

cid

rem

ain

ing

in s

oil

(%

of

13 C

ad

de

d i

n in

div

idu

al a

min

o a

cid

s)

0

20

40

60

80

100

c: High pH

Ala

As

x

Glx

Ser

Gly

Pro

/Th

r

Va

l

Ile

Leu Lys Tyr

Ph

e

0

20

40

60

80

100

*

*

*

* *

*

*

**

**

*

*

*

* *

*

**

*

*

*

*

*

*

*

**

*

*

Figure 4.

Figure 4. Bulk uptake of 13C and 15N from >100 kDa organic N 48 hours after addition in maize (a) shoots, (b) roots, and (c) the whole plant. Significant differences in uptake among soils with different pH are marked by different letters above the bars (n =4).

a: Shoot

Rec

ove

ry o

f tr

acer

in

pla

nt

tiss

ue

(% o

f ad

ded

wit

h >

100

kDa

org

anic

N)

0

2

4

6

8Low pHMedium pHHigh pH

b: Root

0

2

4

6

8

c: Whole plant

0

5

10

15

b a a

b a a

b a a

- a -

- a -

- a -

15N 13C

Figure 5.

Figure 5. Presence of 13C-labeled bound amino acids from added >100 kDa organic N in maize roots after 48 hours in Jyndevad soils in soil with (a) low pH, (b) medium pH, and (c) high pH. Significant differences between presence among individual amino acids within each soil pH level are marked by different letters above the bars (n=4). Amino acids are organized from left on right with increasing steps in their biosynthesis. The amino acids: asparagine and aspartate (Asx), glutamine and glutamate (Glx), and Proline and Threonine (Pro/Thr) elute together in the GC-C-IRMS analysis of the acid hydrolyzed samples.

a: Low pH

0.0

0.5

1.0

1.5

2.0

b: Medium pH

Pre

sen

ce o

f la

bel

ed b

ou

nd

am

ino

aci

d i

n m

aize

ro

ots

(%

of

13 C

ad

ded

in

ind

iviu

al a

min

o a

cid

s)

0.0

0.5

1.0

1.5

2.0

c: High pH

Ala

As

x

Glx

Ser

Gly

Pro

/Th

r

Va

l

Ile

Leu Lys Tyr

Ph

e

0.0

0.5

1.0

1.5

2.0

b bc

cd

aa

a

b

cd cd

bc

d

de de

ef

b

a

bc

cd

df ef

d

fgg

bcbc

ce

a a

b

cb

defeg

cd

fgg

Supporting Material for Enggrob et al.

 

Figure S1. Concentration, 13C and 15N atom fractions of individual amino acids bound in the >100 kDa organic N added to the soils. Mean ± standard error (n = 4).

Am

ino

aci

d c

on

cen

trat

ion

(mg

ml-

1)

0.00

0.05

0.10

0.15

0.20

Am

ino

aci

d 13

C a

tom

fra

cti

on

0.00

0.05

0.10

Ala

As

x

Glx

Ser

Gly

Pro

/Th

r

Val Ile

Leu Lys Tyr

Ph

e

Am

ino

aci

d 15

N a

tom

fra

ctio

n

0.00

0.01

0.02

 

 

Figure S2. Concentration of bound amino acids after 48 hours in soil with maize (a, c, e) and maize root tissue (b, d, f) from control added water and soil added protein-sized organic N (>100 kDa). Low pH (a, b), Medium pH (c, d), and High pH (e, f) in Jyndevad soils. Bars show mean ± standard error (n = 4). Amino acids are organized from left on right with increasing steps in their biosynthesis. The amino acids: asparagine and aspartate (Asx), glutamine and glutamate (Glx), and Proline and Threonine (Pro/Thr) elute together in the GC-C-IRMS analysis of the acid hydrolyzed samples.

a: Low pH

0.00

0.05

0.10

0.15

0.20

0.25

0.30

Control>100 kDa

c: Medium pH

Bo

un

d a

min

o a

cid

co

nce

ntr

atio

n (

mg

AA

g-1

)

0.00

0.05

0.10

0.15

0.20

0.25

0.30

e: High pH

Ala

Asx Glx

Se

r

Gly

Pro

/Th

r

Va

l

Ile

Leu Lys Tyr

Ph

e

0.00

0.05

0.10

0.15

0.20

0.25

0.30

b: Low pH

0

2

4

6

8

10

12

d: Medium pH

0

2

4

6

8

10

12

f: High pH

Ala

Asx Glx

Se

r

Gly

Pro

/Th

r

Va

l

Ile

Leu Lys Tyr

Ph

e

0

2

4

6

8

10

12

--- Soil --- --- Root ---

nn

nn

nn

nn

*

*

*

*

Figure S3. Specific enrichment of 13C-labeled bound amino acids from added protein-sized organic N (>100 kDa) in maize roots after 48 hours in Jyndevad soils at (a) low pH, (b) medium pH, and (c) high pH. Significant differences between specific enrichment among individual amino acids within each soil pH level are marked by different letters above the bars (n=4). Amino acids are organized from left on right with increasing steps in their biosynthesis. The amino acids: asparagine and aspartate (Asx), glutamine and glutamate (Glx), and Proline and Threonine (Pro/Thr) elute together in the GC-C-IRMS analysis of the acid hydrolyzed samples.

a: Low pH

0.0

0.2

0.4

0.6

0.8

1.0

b: Medium pH

Sp

ecif

ic e

nri

chm

ent

of

bo

un

d a

min

o a

cid

s i

n m

aiz

e r

oo

ts (

µg

13C

mg

-1 t

ota

l C

)

0.0

0.2

0.4

0.6

0.8

1.0

c: High pH

Ala

As

x

Glx

Ser

Gly

Pro

/Th

r

Va

l

Ile

Leu Lys Tyr

Ph

e

0.0

0.2

0.4

0.6

0.8

1.0

cc

d

aba

ab

a

d

d

c

fde

f

ab

cc

ba

d

g

ef

g

c

c

c

aa

aa

a

bc

dc

d