You are on page 1of 16

Organic Geochemistry 44 (2012) 21–36

Contents lists available at SciVerse ScienceDirect

Organic Geochemistry
journal homepage: www.elsevier.com/locate/orggeochem

Structural characterization of gilsonite bitumen by advanced nuclear magnetic


resonance spectroscopy and ultrahigh resolution mass spectrometry revealing
pyrrolic and aromatic rings substituted with aliphatic chains
John R. Helms a, Xueqian Kong b, Elodie Salmon a, Patrick G. Hatcher a, Klaus Schmidt-Rohr b,⇑,
Jingdong Mao a,⇑
a
Old Dominion University, Department of Chemistry and Biochemistry, 4541 Hampton Blvd., Norfolk, VA 23529, United States
b
Iowa State University, Department of Chemistry, Gilman Hall, Ames, IA 50011, United States

a r t i c l e i n f o a b s t r a c t

Article history: Gilsonite, a naturally occurring asphaltite bitumen, consists of a complex mixture of organic compounds.
Received 2 March 2011 In the present study, advanced one and two dimensional solid state and solution 1H, 13C and 15N nuclear
Received in revised form 29 November 2011 magnetic resonance (NMR) and electrospray ionization Fourier transform ion cyclotron resonance mass
Accepted 7 December 2011
spectrometry (ESI-FT-ICR-MS) were employed to investigate its composition and structure. 13C NMR
Available online 14 December 2011
yielded a carbon aromaticity of 27%. Aromatic moieties in gilsonite were primarily single rings or small
clusters of fused rings. Half of the aromatic carbons of gilsonite can be accounted for by pyrroles. 15N and
13
C cross polarization-magic angle spinning (CP-MAS) NMR showed that most nitrogen in gilsonite was
pyrrolic. The aromatic rings were heavily substituted with alkyl chains, as evidenced by 1HA13C correla-
tion spectra. Advanced solid state NMR spectral editing techniques clearly identified specific functional
groups such as CCH3, CCH2, and C@CH2 (exomethylene). 1HA13C wideline separation (WISE) NMR helped
identify mobile and non-protonated alkyl carbons. FT-ICR-MS indicated that 64% of calculated formulae
generated by ESI were aliphatic, while only about 0.8–2.5% of formulae contained possible aromatic rings.
All of the assigned formulae contained at least one heteroatom (N, O or S), indicating that ionization by
ESI was selective for the polar fraction of gilsonite and potentially less reflective of the overall chemical
character of gilsonite than NMR spectroscopy. By combining the information obtained from advanced
NMR and ultrahigh resolution MS we propose a structural model for gilsonite as a mixture of many pyr-
rolic and a few fused aromatic rings highly substituted with and connected by mobile aliphatic chains.
Ó 2011 Elsevier Ltd. All rights reserved.

1. Introduction

Bitumen refers to solid or liquid hydrocarbon deposits soluble


Abbreviations: AI, aromatic index; AImod, modified aromatic index; BC, black
carbon; CARS, condensed aromatic ring structures; CP, cross polarization; COSY, in organic solvents (Killops and Killops, 2005). Gilsonite is a natu-
correlation spectroscopy (2D-NMR); CSA, chemical shift anisotropy; DBE, double rally occurring bituminous asphalt and formally classified as an
bond equivalents; DEPT, distortionless enhancement by polarization transfer; DP, asphaltite bitumen (Abraham, 1945). It occurs primarily along
direct polarization; DPEP, deoxyphylloerythroetioporphyrin ; EA, elemental anal-
the Colorado–Utah border and was found in the early 1860s in
ysis; ESCA, electron spectroscopy for chemical applications (X-ray photoelectron
spectroscopy or XPS); ESI, electrospray ionization; FT, Fourier transform; HETCOR,
the Uinta Basin in northeastern Utah (Bell and Hunt, 1963). Gilson-
heteronuclear correlation (2D-NMR); HH-CP, Hartmann–Hahn cross polarization; ite has been widely used in mining and industry and sold all over
HMQC, heteronuclear multi-quantum correlation (2D-NMR); ICR, ion cyclotron the world. For example, it has been used as an additive in oil dril-
resonance; KMDA, Kendrick mass defect analysis; LG-CP, Lee-Goldburg cross ling fluids and for oil well cementing and has a long history of use
polarization; MAS, magic angle spinning; MS, mass spectrometry; MREV-8,
as both a pigment and binding agent in paints, enamels and inks
Mansfield, Rhim, Elleman and Vaughn’s 8 pulse line narrowing sequence; NMR,
nuclear magnetic resonance spectroscopy; NOM, natural organic matter; ppm, parts (Tripp and White, 2006). Natural bitumens are also under investi-
per million units of frequency variation from a standard; T1, spin–lattice relaxation gation as exploitable petroleum fuels (Meyer et al., 2007). Further-
time; TOCSY, total correlation spectroscopy (2D-NMR); TOSS, total sideband more, gilsonite and other natural bitumens are important
suppression; TPPM, two-pulse phase modulated decoupling; WISE NMR, wideline
intermediates and/or byproducts in the geological formation of
separation nuclear magnetic resonance.
⇑ Corresponding authors. Tel.: +1 515 294 6105; fax: +1 515 294 0105 (K. crude oils that can be analyzed to improve our understanding of
Schmidt-Rohr), tel.: +1 757 683 6874; fax: +1 757 683 4628 (J. Mao). petroleum geochemistry (Bell and Hunt, 1963; Killops and Killops,
E-mail addresses: srohr@iastate.edu (K. Schmidt-Rohr), jmao@odu.edu (J. Mao). 2005).

0146-6380/$ - see front matter Ó 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.orggeochem.2011.12.001
22 J.R. Helms et al. / Organic Geochemistry 44 (2012) 21–36

Despite the widespread industrial use and petrochemical po- has supported this. Solid state NMR can address this question by
tential of gilsonite, its molecular structure has not been well char- estimating the size of clusters of fused aromatic rings, by long
acterized due to its complexity and heterogeneity. Past studies range CAH dipolar dephasing and quantification of the ring edge
have primarily focused on the investigations of biomarkers such carbon fractions (Brewer et al., 2009; Mao et al., 2010b).
as petroporphyrins (Treibs, 1934, 1936; Sugihara and McGee, The structural characterization of gilsonite is critical for under-
1957; Corwin, 1959; Quirke and Maxwell, 1980; Quirke et al., standing its properties, its applications in industry, and its geolog-
1980a, 1980b; Hajibrahim et al., 1981; Gill et al., 1985), hopanes, ical significance. Solid state NMR has been widely applied in the
steranes and carotenoids (Ruble et al., 1994; Schoell et al., 1994). study of natural organic matter (NOM) such as coal, oil shale, hu-
Quirke et al. (1980b) found that the aetio-type porphyrins and mic materials and peats (Hatcher et al., 1981, 1983; Dennis et al.,
deoxyphylloerythroetioporphyrin (DPEP) were products of reduc- 1982; Miknis et al., 1982; Wilson, 1987; Solum et al., 1989; Ander-
tive degradation of chlorophylls. The porphyrins present in gilson- son et al., 1992; Maciel et al., 1993; Wilson et al., 1993; Preston,
ite were most likely formed from naturally occurring chlorophylls 1996; Nanny et al., 1997; Clifford et al., 1999; Hu et al., 2000; Conte
through this degradation (Quirke and Maxwell, 1980; Quirke et al., et al., 2004; Smernik et al., 2006). Advanced NMR techniques, espe-
1980b; Hajibrahim et al., 1981). Compound specific isotopic analy- cially spectral editing techniques, have increased the amount of
sis of carotenoids indicated a strong algal or cyanobacterial source information obtainable from NOM samples (Wu and Zilm, 1993;
for the diagenetic starting material for gilsonite (Ruble et al., 1994; Wu et al., 1994; Hu et al., 2000; Keeler and Maciel, 2000). We have
Schoell et al., 1994). Additionally, isotopic values for hopanes (pen- developed and applied many new, advanced solid state NMR tech-
tacyclic membrane lipids) and moretanes suggested the influence niques, especially spectral editing (Schmidt-Rohr and Mao, 2002;
of mid-water bacteria and potentially some methanotrophs (Scho- Mao and Schmidt-Rohr, 2003), for the investigations of NOM
ell et al., 1994). (Mao et al., 2007a, 2007c). While the broad and heavily overlapped
However, biomarkers such as the porphyrin compounds only ac- bands of complex NOM allow traditional 13C NMR to identify only
count for a minor fraction of gilsonite (Treibs, 1936). The bulk of the about 10 types of chemical groups, our spectral editing techniques,
gilsonite consists of aliphatic and aromatic hydrocarbons with few which selectively retain peaks of certain types of functional groups,
olefinic groups, as indicated by the iodine number (Baker et al., can identify more than two dozen different moieties (Mao et al.,
1967; Quirke and Maxwell, 1980). Its bulk chemical characteristics 2007a, 2007c). Moreover, two dimensional 1HA13C heteronuclear
have been investigated using electron spectroscopy for chemical correlation (HETCOR) NMR is used to detect proximities and con-
applications (ESCA; usually referred to as X-ray photoelectron spec- nectivities of different functional groups. 15N CP-MAS NMR is ap-
troscopy or XPS), infrared spectroscopy, X-ray diffraction spectros- plied to characterize forms of nitrogen in gilsonite and combined
copy, and 1H and 13C solution NMR analysis. ESCA provided with recoupled dipolar dephasing to determine which resonances
verification of elemental analysis results. ESCA also indicated that are from N bonded to hydrogen (Fang and Schmidt-Rohr, 2009).
carbon was mainly present as aliphatic hydrocarbons, that nitro- Gilsonite is soluble in a range of organic solvents such as chloro-
gen–carbon bonding environments such as aliphatic and aromatic form; therefore we have also studied it by solution NMR techniques.
amines as well as nitrogen heterocycles such as pyrrole were pres- Solution 13C NMR spectra appear well resolved; however, the signals
ent, that CAO bonding environments were present and that a very of many components are broad and almost invisible due to low
small amount of sulfur was present as organic sulfides or heterocy- mobility of some components and resulting short spin–spin relaxa-
cles such as thiophenes (Clark et al., 1983). Infrared spectroscopy tion times (T2). Thus, advanced solid state and solution NMR tech-
showed that CAH (88% of IR signal) and CAO (10% of IR signal) niques provide complementary structural information.
were the dominant vibrational modes present in gilsonite with Mass spectrometry is used to determine the molecular weight
acids and esters each representing <1% of the vibrational signal of ions. In organic chemistry, advanced solution NMR and mass
(Clark et al., 1983). Organic acids were present in gilsonite in trace spectrometry data have been routinely collected to elucidate the
amounts (Grantham and Douglas, 1977; Clark et al., 1983). structures of organic compounds. However, the approach has not
Yen et al. (1961) used X-ray diffraction to estimate the aroma- been frequently used to elucidate the structures of natural organic
ticity of gilsonite as well as several petroleum asphaltenes. They matter in organic geochemistry. Despite the availability of high
reported a much lower aromaticity for gilsonite (14%) than for resolution mass spectrometry for the analysis of fossil fuels for
petroleum asphaltenes (22–53%). However, Bell and Hunt (1963) over two decades (Comisarow and Marshall, 1974; Grigsby,
described gilsonite as ‘‘predominantly aromatic’’ based on liquid 1989; Hsu et al., 1994; Amster, 1996; Marshall et al., 1998), to date,
chromatography, elemental analysis and infrared measurements. there has been no detailed, comprehensive characterization of gil-
Wen et al. (1978) report the aromaticity of gilsonite asphaltene, sonite using Fourier transform ion cyclotron mass spectrometry
which is the pentane insoluble, benzene soluble fraction, as 38% (FT-ICR-MS) reported in the literature. The high mass accuracy of
based on 1H solution NMR. Yet the perception of gilsonite as a FT-ICR-MS makes it possible to differentiate ions in complex mix-
‘‘predominantly aromatic substance’’ (North, 1985) persisted until tures without employing a separation method prior to infusion
Clark et al. (1983) applied solution 1H and 13C NMR to gilsonite into the mass spectrometer (Klein et al., 2006).
samples and showed a variety of aliphatic carbon environments Mass separation is only carried out on charged particles (ions)
and a single broad aromatic peak. They reported an aromatic:ali- and therefore the quality of MS data strongly depends on the type
phatic carbon ratio of 1:4 for a gilsonite sample with limited atmo- of ionization used during sample introduction. Electrospray ioniza-
spheric exposure, which was more in line with the X-ray tion (ESI) coupled with FT-ICR-MS yields a detailed mass spectrum
diffraction results (Yen et al., 1961) and the absence of a ‘shake- of ionic organic molecules with highly accurate mass to charge (m/
up satellite’ peak, indicative of conjugated systems, in the ESCA z) determination. Positive ionization and negative ionization modes
spectrum (Clark et al., 1983). To resolve this controversy, reliable are selective in favor of chemical moieties capable of bearing posi-
quantification of aromaticity is required. Direct polarization 13C tive or negative formal charges, respectively. As anticipated, posi-
NMR, after a sufficiently long relaxation delay (Mao et al., 2000; tive ionization exhibits higher selectivity/sensitivity for nitrogen
Mao and Schmidt-Rohr, 2004a), provides accurate aromaticity val- containing molecules such as amines, amides and nitrogen hetero-
ues. Gilsonite has often been described as containing condensed cycles, while negative ionization shows higher selectivity/sensitiv-
polyaromatic ‘sheets’ or ‘asphaltene sheets’ (Yen et al., 1961; Bell ity to acidic moieties such as fatty acids and phenols.
and Hunt, 1963; Wen et al., 1978; Yen, 2000) based on its high car- Smith et al. (2008) used ESI-FT-ICR-MS to characterize heavy
bon content and its high refractive index, but little direct evidence vacuum gas oil distillation cuts from Athabasca bitumen and found
J.R. Helms et al. / Organic Geochemistry 44 (2012) 21–36 23

that the technique could rapidly and reproducibly analyze the po- bile aliphatic components, a 13C DP/MAS spectrum with a short re-
lar fraction of petroleum distillates and provide unique molecular cycle delay of 1.5 s was also recorded, with 2000 scans. Recoupled
formulae assigned to each peak. It has been shown that chemically dipolar dephasing for 13C DP-MAS NMR at a spinning speed of
isolated petroporphyrins can be analyzed by ESI-MS (Van Berkel 14 kHz was employed to obtain quantitative information on the
et al., 1993), however, it has not been determined whether they non-protonated carbons and carbons of mobile groups (Mao and
are effectively ionized by ESI in the more complex matrix of unal- Schmidt-Rohr, 2003). The dipolar dephasing time was 68 ls, the re-
tered bitumen. cycle delay was 150 s, and 512 scans were collected.
It is the objective of this study to characterize the structure of Qualitative composition information was obtained with good
gilsonite bitumen by advanced NMR and FT-ICR-MS techniques. sensitivity by 13C cross polarization-total sideband suppression
Based on the detailed data provided by the two techniques, we (CP-TOSS) NMR experiments at a spinning speed of 6.5 kHz and a
propose a structural model for gilsonite. We believe this is the cross polarization (CP) time of 1 ms, with a 1H 90° pulse length
most comprehensive chemical characterization to date of a fossil of 4 ls and a 2 s recycle delay. The short recycle delay for CP meth-
fuel sample using a combination of advanced spectroscopic tech- ods is predicated on the shorter 1H spin–lattice relaxation time
niques. The protocol used here can be applied to the study of other (0.5 s for gilsonite) compared to 13C and allows for many more
fossil fuels and related samples, including their extracts, coals, ker- scans to be completed within a given time frame. Four-pulse total
ogens, algaenans and other bitumens. Finally, the core structures suppression of sidebands (TOSS) (Dixon, 1982) was employed be-
identified in preserved geological samples such as gilsonite will fore detection and two pulse phase modulated (TPPM) decoupling
help gain insights into the structures of potentially refractory com- was applied for optimum resolution. One thousand scans were col-
ponents in nature. lected. The corresponding subspectrum with signals of non-pro-
tonated carbons and mobile groups such as rotating CH3 was
obtained by 13C CP-TOSS combined with 40 ls dipolar dephasing.
2. Materials and methods
The number of scans was 1024, with a 2 s recycle delay.
In order to separate the signals of anomeric carbons (OACAO)
2.1. Geological setting
from those of aromatic carbons, both of which resonate between
90 and 120 ppm, the aromatic carbon signals were selectively sup-
The sample of the bitumen, gilsonite, used in this study was col-
pressed by a five pulse 13C chemical shift anisotropy (CSA) filter
lected in the Uinta Basin near the town of Bonanza, Utah and was
with a CSA filter time of 47 ls (Mao and Schmidt-Rohr, 2004b).
provided by Dr. Gary Thompson at Rocky Mountain College, Mon-
The number of scans was 1024, with a 2 s recycle delay.
tana. The Uinta Basin covers an area of ca. 24,000 square kilome-
The combined spectrum of immobile CH2 and CH groups was
ters (Osmond, 1964). Over 2200 m of lacustrine sediments,
obtained with good sensitivity in a simple spectral editing experi-
primarily siliciclastic and carbonate, were deposited there between
ment. First, a 13C CP-TOSS spectrum was recorded using a short CP
the late Cretaceous and middle Eocene (Picard and High, 1968; Ru-
time of 50 ls. It showed predominantly protonated carbons in
ble et al., 1994). The source beds of gilsonite are the calcareous oil
immobile segments, but residual peaks of quaternary carbons re-
shales of the Parachute Creek Member (Mahogany Zone) of the
sulted from two-bond magnetization transfer. Second, a 13C CP-
Green River Formation, which are rich in organic matter (Bell
TOSS spectrum was recorded using a short CP of 50 ls and 40 ls
and Hunt, 1963).
dipolar dephasing. It contained only the residual signals of quater-
nary carbons or mobile segments (including CH3 groups with >50%
2.2. NMR spectroscopy efficiency). For both spectra, 2000 scans were collected using a re-
cycle delay of 1 s. The difference of the two spectra was the spec-
Most solid state NMR experiments, except several spectral edit- trum of immobile CH2 and CH carbons, with a small CH3
ing techniques and long range dipolar dephasing, were performed contribution (Mao et al., 2007a).
on finely ground, but otherwise unaltered gilsonite using a Bruker Spectral editing of CH2 signals was achieved by selection of the
DSX spectrometer at 100 MHz for 13C, with magic angle spinning three spin coherence of CH2 groups, using a 13C 90° pulse and 1H
(MAS) in a 7 mm double resonance probehead. The high speed 0°/180° pulses applied after the first quarter of one rotation period
quantitative direct polarization-magic angle spinning (DP-MAS) with MREV-8 decoupling (Mao and Schmidt-Rohr, 2005). A total of
experiments were run in a 4 mm double resonance probehead. So- 56,728 scans were collected with a 1 s recycle delay and the spin-
lid state 15N cross polarization-magic angle spinning (CP-MAS) and ning speed was 5.787 kHz.
CP-MAS with recoupled dipolar dephasing were performed at For CH (methine) selection, a robust method based on CAH
40 MHz for 15N in a 7 mm double resonance probehead. All the multiple quantum coherence (Schmidt-Rohr and Mao, 2002) was
solution NMR experiments were conducted by dissolving gilsonite used at 4 kHz MAS. CH group multiple quantum coherence was
in deuterated chloroform (CDCl3) in a 5 mm triple resonance pro- not dephased by the spin-pair CH dipolar coupling while CH2
behead using a Bruker DRX 400 spectrometer, with 1H and 13C coherence was dephased by dipolar coupling of the carbons to
experiments performed at 400 and 100 MHz, respectively. Chemi- the two protons. The first of a pair of recorded spectra contains sig-
cal shifts were calculated using ACD/lab v.12.0. nals of CH, as well as residual quaternary carbon and CH3 peaks
that were removed by taking the difference with a second spec-
2.2.1. Solid state NMR trum acquired with the same pulse sequence except for additional
High speed quantitative 13C direct polarization-magic angle 40 ls dipolar dephasing before detection. Each spectrum was
spinning (DP-MAS) NMR and high speed quantitative 13C DP-MAS based on 5888 scans with a 1 s recycle delay.
NMR with recoupled dipolar dephasing provide quantitative struc- Two dimensional (2D) 1HA13C heteronuclear correlation (HET-
tural information. These were run at a spinning speed of 14 kHz. COR) NMR experiments (Mao and Schmidt-Rohr, 2006) were per-
The 90° 13C pulse length was 4 ls. Recycle delays were tested by formed at a spinning speed of 6.5 kHz. The scale on which
1
the cross polarization-spin lattice relaxation time-total suppression HA13C proximities were probed was chosen by the cross polariza-
of sidebands (CP-T1-TOSS) technique to ensure that all carbon sites tion method and by 1H spin diffusion before cross polarization. Pri-
were fully relaxed (Mao et al., 2000). The recycle delay was 150 s marily one-bond 1HA13C connectivities were revealed by 0.1 ms of
and 512 scans were collected. This technique was fully described Lee-Goldburg cross polarization (LG-CP), which suppressed 1HA1H
elsewhere (Mao and Schmidt-Rohr, 2003). In order to highlight mo- spin diffusion during CP. A total of 512 scans were collected using a
24 J.R. Helms et al. / Organic Geochemistry 44 (2012) 21–36

1.4 s recycle delay. Prolonging LG-CP to 0.5 ms allowed for mostly Tesla Bruker Apex-Qe mass spectrometer. For positive ionization,
one-and two-bond 1HA13C connectivities. Scans (448) were col- the sample was dissolved in pyridine and diluted in 1:1 tetrahy-
lected with a 0.8 s recycle delay. A 40 ls dipolar dephasing delay drofuran (THF):methanol (MeOH) with 0.1% trifluoroacetic acid
was inserted to reveal proximities for non-protonated aromatic (TFA). For negative ionization the sample was dissolved in pyridine
carbons to alkyl substituents (Mao et al., 2007a). The number of and diluted in 1:1 THF:MeOH with 0.1% ammonium hydroxide
scans was 512, with a 1.4 s recycle delay. The correlations of the (NH4OH). TFA and NH4OH were added as pH adjusting buffers
carbons to protons within a 4 Å radius were shown with standard and were chosen because of their high volatility.
Hartmann–Hahn CP (HH-CP) of 0.5 ms, which allowed for some 1H Prior to both analyses the mass analyzer was calibrated using a
spin diffusion. For each t1 increment, 192 scans were collected with mixture of polyethylene glycols (PEG). Positive ionization mode
a recycle delay of 1.4 s. We used 96 t1 increments. mass spectra were calibrated internally using a homologous series
Fused aromatic rings were identified by the large number of of aliphatic amines, while the negative ionization mode was cali-
carbons located far from the protons on the periphery of the struc- brated internally using a homologous series of fatty acids (Sleigh-
ture. The signals of these carbons were selected by a long range ter et al., 2008). In both cases the internal calibrants were
recoupled CAH dipolar dephasing technique (Mao and Schmidt- naturally occurring components of the gilsonite.
Rohr, 2003). In short, two 1H 180° pulses per rotation period pre- Gilsonite is a complicated mixture of compounds and thus gen-
vent magic angle spinning from averaging out weak CH dipolar erates a very complicated spectrum. ‘Data mining’ techniques, as
couplings. After 0.9 ms of recoupled dipolar dephasing time, the outlined by several previous studies (Hsu et al., 1994; Hughey
signals of most individual aromatic rings are dephased while those et al., 2001; Kujawinski, 2002; Kim et al., 2003; Herniman et al.,
of fused rings remain at the 95% level. In order to detect non-pro- 2005; Hertkorn et al., 2006; Hockaday et al., 2006; Koch and Ditt-
tonated carbons with good relative efficiency, direct polarization- mar, 2006; Sleighter and Hatcher, 2007), were therefore used to
total sideband suppression (DP-TOSS) as described above was used summarize the mass spectral results. Kendrick mass defect analy-
at a spinning speed of 7 kHz. The ‘‘c-integral’’ was employed to sis (KMDA) (Kendrick, 1963; Grigsby, 1989) was used to identify
suppress sidebands up to the fourth order (DeAzevedo et al., prominent homologous series within each spectrum to internally
2000). The 13C 90° and 180° pulse lengths were 4 ls and 8.1 ls, calibrate the exact m/z assignments. Once the spectra were cali-
respectively. The details of this technique have been described brated to within 1.0 ppm error (1 ppm = 0.0001%), relatively
elsewhere (Mao and Schmidt-Rohr, 2003). For each gilsonite spec- straightforward computational tools were used to obtain exact ele-
trum, the number of scans and recycle delay were 1024 and 20 s, mental formula assignments (C, H, N, O, S) for the majority of peaks
respectively. in the spectrum (Kim et al., 2006). Elemental atom ratios (e.g. O/C,
Two-dimensional 1HA13C wide line separation (WISE) NMR H/C, N/C, etc.) were calculated for all of those peaks that were as-
measurements were performed with 13C decoupling during evolu- signed formulae. From these, we plotted each peak within van
tion (Schmidt-Rohr et al., 1992; Tekely et al., 1993; Mao and Krevelen space (Kim et al., 2003), providing graphical representa-
Schmidt-Rohr, 2006). Two different CP times, 0.1 ms or 1 ms, were tions that allow a fingerprint characterization of the distribution
used. A dipolar dephasing delay of 40 ls was employed to selec- of ions generated by ESI. Several chemical indices were calculated,
tively observe signals of non-protonated and mobile carbons. All which indicate the likely structural character of the molecules that
WISE experiments used 512 scans per t1 increment and 0.5 s re- can be inferred from their formulae. These indices include double
cycle delay, except the experiment with 0.1 ms CP and 40 ls dipo- bond equivalents (DBE) (Hockaday et al., 2006; Koch and Dittmar,
lar dephasing, which used 2000 scans. We used 96 t1 increments of 2006),
5 ls.
15 1
N CP-MAS was conducted with a contact time of 1 ms and DBE ¼ 1 þ C  O  S  H ð1Þ
2
0.5 s recycle delay (262,144 scans). In order to observe non-proton-
ated and mobile nitrogen functional groups, 15N CP-MAS with a condensed aromatic ring structures (CARS) (Hockaday et al., 2006),
recoupled dipolar dephasing time of 291 ls (Fang and Schmidt-
DBE
Rohr, 2009) was performed. A total of 53,760 scans were recorded CARS  > 0:7 ð2Þ
C
with a 0.5 s recycle delay. The spinning speed was 6.5 kHz.
aromatic index (AI) and modified AI (AImod) (Koch and Dittmar,
2.2.2. Solution NMR 2006),
1
H NMR of gilsonite with Hahn echo was conducted with a 2 s
1 þ C  O  S  0:5H
recycle delay. The interval s in the Hahn echo sequence (p/2–s– AI  P 0:67;
COSNP
p–s-detection) was varied for a series of experiments. As s in-
1 þ C  0:50O  S  0:5H
creased, the mobile components with long T2 relaxation time sur- AImod  P 0:5 ð3Þ
C  0:50O  S  N  P
vived, but immobile components with short T2 relaxation time
were dephased. Sixteen scans were recorded for each spectrum. aliphatics (Perdue, 1984),
A 13C solution NMR spectrum was collected consisting of 12,000
DBE
scans with a recycle delay of 1 s. Gradient enhanced heteronuclear aliphatics  < 0:3 ð4Þ
C
multi-quantum correlation (HMQC) NMR (Vuister et al., 1991) was
used to show direct 1HA13C J-couplings. Correlation spectroscopy and black carbon (BC) (Kim et al., 2004).
(COSY) was used to show 1HA1H interactions across three bonds
BC  H=C < 0:8 and 0:3 < O=C > 0:6 ð5Þ
(Aue et al., 1976). Total correlation spectroscopy (TOCSY) NMR
(Bax and Davis, 1985) was used to show correlation of proton spins Kendrick mass defect analysis (KMDA) was used to identify and
over a relatively long range (multiple bonds) within a given spin enumerate series of structural homologues and commonly ob-
system. served repeating units within structural compound classes. KMDA
is based on the fact that ions within a homologous series have the
2.3. Fourier transform ion cyclotron resonance mass spectrometry same difference between the nominal mass and Kendrick mass.
The Kendrick mass of the ion was calculated by multiplying the
Gilsonite was analyzed using electrospray ionization (ESI) in measured mass of the ion by the quotient of the IUPAC mass and
both positive ionization and negative ionization modes on a 12 nominal mass of the repeating unit (Kendrick, 1963; Grigsby,
J.R. Helms et al. / Organic Geochemistry 44 (2012) 21–36 25

1989; Hughey et al., 2001; Stenson et al., 2003). It is important to highlighted the spectral region characteristic of non-protonated
note that since tandem MS was not used, FT-ICR-MS did not distin- aromatic carbons two bonds from O or N.
guish between structural isomers (Stenson, 2008) and ESI shows The low shoulder from 145–165 ppm on the other side of the
considerable selectivity during ionization; thus the chemical indi- aromatic peak, highlighted in Fig. 1a, can be assigned to aromatic
ces and KMD series were used to simply constrain the contents of CAO and possibly pyridinic CAN sites. Given that two carbons
the sample and may have given quantitatively misleading values or would be bonded to each pyridinic N, the spectral area fraction be-
reflected isomeric overlap between compound classes. tween 150 and 165 ppm (2%) provides an upper limit of 2%/2 = 0.01
to the ratio of all C to pyridinic N; the likely value is smaller, due to
contributions from aromatic CAO in this spectral range. Given the
2.4. Elemental analysis
3.3:80 = 0.04 C:N ratio of our sample, pyridinic N accounts for less
than 1=4 of all nitrogen. The likely fraction of pyridinic N is smaller,
Triplicate solid gilsonite samples were analyzed for weight per-
between 1/5 and 1/10 of all N.
cent carbon and nitrogen using a Thermo-Finnegan Flash 1112 ser-
The assignments of signals and the corresponding percentages
ies elemental analyzer (EA). The EA was calibrated for both carbon
of different functional groups are listed in Table 1. Specific assign-
and nitrogen using nicotinamide.
ments are difficult because the bands were broad and overlapping;
however, the interpretation was assisted by spectral editing as de-
3. Results scribed below and in Table 1.

13
3.1. Quantitative C NMR spectra 3.2. NMR spectral editing

Fig. 1a shows the 13C DP-MAS spectrum obtained at a spinning Fig. 2 shows a series of 13C CP-MAS NMR spectra acquired with
speed of 14 kHz, which provided quantitative structural informa- suitably designed radio frequency pulse sequences to selectively
tion on the whole sample. Spin counting (Smernik and Oades, detect signals from specific types of chemical groups. Fig. 2a is
2000) yields a 13C observability of 110 ± 15% in DP-MAS NMR, the 13C CP-TOSS spectrum, which shows qualitative structural
which confirms that all gilsonite carbon is observed in this spec- information and was used as the reference for the selected sub-
trum. Similar to the NMR spectra of other post-diagenetic organic spectra (Fig. 2b–g). The corresponding 13C CP-TOSS spectrum after
matter samples such as kerogen (Werner-Zwanzinger et al., 2005; 40 ls of dipolar dephasing (Fig. 2b) exhibited only signals of non-
Smernik et al., 2006; Mao et al., 2010a), the DP-MAS NMR spec- protonated carbons and carbons of mobile groups, including rotat-
trum of gilsonite has mainly two broad bands: the signals of sp3- ing CH3 groups, which have a reduced CAH dipolar coupling due to
hybridized alkyl C from 5–60 ppm and the band of sp2-hybridized their fast motions. This spectrum provided similar structural infor-
carbons such as aromatics around 100–165 ppm. The O-alkyls mation as in Fig. 1b but with better sensitivity for some peaks. In
(50–100 ppm) were almost undetectable, as expected for heavily particular, comparison of Fig. 2a and b (as well as Fig. 2d) showed
degraded geological samples. Fig. 1b shows the spectrum of 13C that the possible pyridine CAN carbons resonating near 160 ppm
DP-MAS with recoupled dipolar dephasing of 68 ls, providing are mostly not protonated. As in the spectra of Fig. 1, the maxima
quantitative structural information on non-protonated carbons of the aromatic band in Fig. 2a and b did not align exactly. Two
and carbons of mobile groups such as CH3. Note that the highest bands from mobile CCH3 were observed around 17 and 22 ppm,
peak of the sp2 hybridized aromatic band around 100–150 ppm and one from mobile CCH2 groups around 31 ppm. Some non-pro-
of Fig. 1b did not match that of Fig. 1a: the former at 133 ppm tonated quaternary carbons resonating around 40 ppm were also
was primarily from non-protonated aromatic carbons and the lat- present.
ter at 128 ppm more attributed to protonated aromatic carbons The 13C CP-TOSS spectrum after a 13C CSA filter of 47 ls, which
(see further discussion below). In the alkyl C region, we observed exhibited only sp3-hybridized carbon signals, is displayed in
two bands from mobile CCH3 around 17 and 22 ppm, and one from Fig. 2c. This technique separates signals of anomerics (OACAO)
mobile CCH2C groups around 31 ppm. In Fig. 1b, we have from overlapping bands of aromatics (Mao et al., 2007a, 2007b,
2007c). No contribution from anomeric carbon was observed for
the gilsonite sample, indicating that sugar rings were insignificant.
Fig. 2c indicates that the 13C signals above 90 ppm all belonged to
sp2-hybridized carbons.
Using a short CP time of 50 ls, primarily protonated carbons
were selected; while mobile CH2 and CH3 resonances were signif-
icantly suppressed because of their low CP efficiencies (Fig. 2d). In
order to suppress residual signals of carbons two bonds from 1H, a
spectrum after dipolar dephasing of 40 ls was also acquired. The
short CP spectrum subtracted by the spectrum with a double filter
of short CP and dipolar dephasing showed primarily signals of
immobile protonated carbons. There was a small signal from pro-
tonated aromatics or C@CH2 around 123 ppm.
Fig. 2e shows the CH2 only spectrum of gilsonite obtained by
three spin coherence selection (Mao and Schmidt-Rohr, 2005).
CH2 contributed a small amount of signal in the sp2 hybridized re-
gion, observed around 123 ppm; this band was assigned to exo-
methylene –C@CH2. Exomethylene was previously identified in
natural resins (Anderson et al., 1992; Clifford et al., 1999). Signifi-
cant CCH2C (non-polar alkyl) signals were dominant in this selec-
Fig. 1. Quantitative spectra of (a) DP-MAS and (b) DP-MAS with 68 ls recoupled
dipolar dephasing showing non-protonated carbons and mobile segments such as
tive spectrum. Clearly, most of the carbons that resonated in the
CH3. Shoulders assigned to aromatic CAO and possible pyridinic CAN are alkyl region belonged to CCH2C groups. Fig. 2f shows the CH only
highlighted in (a), those of non-protonated aromatic C two bonds from O or N in (b). spectrum, acquired using dipolar distortionless enhancement by
26 J.R. Helms et al. / Organic Geochemistry 44 (2012) 21–36

Table 1
13 13 13
Quantitative integration results from C DP-MAS, C DP-MAS with recoupled dipolar dephasing, C CP-TOSS, and spectral editing.

Integration range (ppm) Predominant structural moieties DP-MAS (%) Refined using DP dipolar dephasing and spectral editing
5–27 CH3 + CH2 25
27–60 Alkyl CH2 and CH 45 38% CH2
6.2% CH (CH selection)
1.1% CQuaternary (dipolar dephasing)
60–100 Alkyl OC 0.3 Non-protonated C (dipolar dephasing)
100–165 Aromatic + alkene 29 19% Non-protonated aromatic CAC (dipolar dephasing)
5.7% Aromatic CH
2.3% Exomethylene @CH2
2.2% (150–165 ppm) Aromatic CAO and pyridine CAN
165–185 Amides, carboxylic acids, esters (N/OAC@O) 0.4 Non-protonated C (dipolar dephasing)
185–220 Ketones, quinones (C@O) 0.4 Non-protonated C (dipolar dephasing)

Fig. 2g shows a 13C DP-MAS spectrum with a short recycle delay,


which accentuated the signal from mobile alkyl moieties. The short
recycle delay selected signals with short 13C T1 (spin–lattice relax-
ation) times. Rigid segments have relatively long 13C T1 values
compared with mobile ones because mobility drives relaxation.
The spectrum of the mobile segments only exhibited signals of
sp3-hybridized carbons, indicating that many of the alkyls were
more mobile than the sp2-hybridized carbons. This conclusion
was consistent with the result of the dipolar dephased spectrum
(Fig. 1b); signals of highly mobile alkyls were also retained in this
spectrum.

3.3. Structural information from short range 1HA13C HETCOR NMR

Fig. 3 shows several 2D 1HA13C HETCOR spectra acquired under


various conditions. 1H slices at various 13C chemical shifts were ex-
tracted to observe the correlations more clearly. Fig. 3a is the 2D
HETCOR spectrum with LG-CP of 0.1 ms, in which primarily one-
bond connectivities were observed. In order to assign the 13C signal
between 51 and 61 ppm and at 30 ppm, we extracted 1H slices
from this 2D spectrum (Fig. 3b). We integrated over 13C chemical
shifts between 51 and 61 ppm, associated with the highest peak
in the CH-only spectrum, and took a slice at the 30 ppm resonance
position of CCH2C groups for reference. The 1H spectrum associated
with the 51–61 ppm 13C resonance showed various alkyl contribu-
tions, including a shoulder near 4 ppm that was attributed to NCH
groups (e.g. in peptides). Nevertheless, this component was not
dominant, which indicated that peptides were not the main con-
tributors to this carbon signal.
In order to observe correlations between aromatic C and alkyl H,
which are separated by at least two bonds, we recorded a 2D HET-
COR spectrum with a longer 0.5 ms LGCP contact time (Fig. 3c). It
mostly showed one-and two-bond 1HA13C connectivities, i.e. the
peaks belonged to 13C and 1H nuclei within 0.25 nm of one another.
1
HA13C 2D HETCOR successfully offered better separation between
two types of aromatic moieties, namely (i) non-protonated aro-
matics bonded to alkyls at 1H chemical shift of 0.8–1.2 ppm and
13
Fig. 2. Spectral editing for identification of functional groups in 13C NMR. (a) CP-
C chemical shift of 133 ppm, and (ii) primarily protonated aro-
TOSS spectrum for reference. (b) Dipolar-dephased CP-TOSS spectrum showing matics or aromatics closer to aromatic protons than alkyl protons
non-protonated carbons and mobile segments such as CH3. (c) Selection of sp3- with a 1H chemical shift of 7.3 ppm and a 13C chemical shift of
hybridized carbon signals by a chemical shift anisotropy filter. (d) Selection of 123 ppm. The assignment of the two types of aromatic signals
immobile CH and CH2 signals with residual CH3. (e) Signals of immobile CH2 groups
was also based on spectral editing results described above. 1H
selected based on three-spin coherence. (f) CH-only spectrum (by dipolar DEPT),
and (g) DP spectrum with a short, 1.5 s, recycle delay. Vertical dashed lines at 134, slices extracted at the alkyl 13C chemical shifts of 17, 31 and
127, 122, 31, 22, and 15 ppm mark important peaks or shoulders in the spectra. 40 ppm showed correlations primarily with their own alkyl pro-
tons; 1H slices extracted near the aromatic 13C chemical shifts of
127 and 135 ppm showed correlations with alkyl protons, indicat-
polarization transfer (DEPT) (Schmidt-Rohr and Mao, 2002); small ing that these were closely associated with alkyl side chains
CCH and aromatic CH signals were observed. Both bands were (Fig. 3d). Fig. 3e displays the 2D HETCOR spectrum with 0.5 ms
broad, indicating their wide range of chemical environments. We LGCP and 40 ls dipolar dephasing, which provided spectra selec-
have highlighted the ppm region where NCH could resonate tively of the non-protonated (or mobile carbons). The 1H spectrum
(Fig. 2f). associated with 13C chemical shifts of 125–145 ppm (Fig. 3f)
J.R. Helms et al. / Organic Geochemistry 44 (2012) 21–36 27

Fig. 4. (a) Series of DP-TOSS spectra of gilsonite after recoupled dipolar dephasing
of the indicated durations. (b) Corresponding long-range dipolar dephasing curves
for non-protonated aromatic carbons in gilsonite, integrated between 108 and
145 ppm. S/S0 refers to the aromatic signal from the dipolar dephased spectrum as a
fraction of the signal in the reference spectrum. Data for wood charcoal, lignin, and
type II kerogen (Mao and Schmidt-Rohr, 2003; Mao et al., 2010a) are shown for
reference. Filled diamonds: gilsonite; open hexagons: wood char; open triangles:
lignin and open squares: type II kerogen.

3.4. Information on fused aromatic rings based on 1HA13C recoupled


long range dipolar dephasing NMR

Fig. 4 shows the use of the 1HA13C recoupled long range dipolar
dephasing technique (Mao and Schmidt-Rohr, 2003) to investigate
the presence of fused ring carbons in gilsonite. Fig. 4a shows a
series of 13C direct polarization/total suppression of sidebands
(DP-TOSS) NMR spectra of gilsonite with increasing recoupled
Fig. 3. 1HA13C 2D HETCOR spectra of gilsonite (a) with LG-CP of 0.1 ms whose dephasing times. In order to estimate the aromatic cluster size,
extracted proton slices between 51–61 ppm and at 30 ppm are listed in (b); (c) with we compared the dephasing curve for the aromatic region (108–
LG-CP of 0.5 ms and its extracted proton slices at 17, 31, 40, 127, and 135 ppm in 145 ppm) of gilsonite with published data of lignin (Brewer et al.,
(d); (e) with LG-CP of 0.5 ms and dipolar dephasing of 40 ls and extracted proton 2009), an immature type II kerogen (Mao et al., 2010a) and char-
slices between 125–145 ppm in (f); and (g) with HH-CP of 0.5 ms and its slices
coal (Mao and Schmidt-Rohr, 2003) (Fig. 4b). The dephasing of gil-
between 115–142 ppm in (h).
sonite was faster than that of kerogen, which has been estimated to
contain an average of six rings per fused aromatic structure (Mao
showed a dominant proton contribution from alkyl H, rather than et al., 2010a) but slower than that of lignin, which contains mostly
from aromatic H, indicating that these aromatics were highly single benzene rings. This indicated small clusters of perhaps 4
substituted. This was further confirmed by a 2D HETCOR spectrum fused aromatic rings in gilsonite. Three-and four-ring structures
with 0.5 ms Hartmann–Hahn CP with an effective spin diffusion in petroleum deposits have been attributed to aromatization of dit-
time of 0.125 ms (Fig. 3g); the dominant proton contribution was erpenoid, triterpenoid, sterol, hopane and similar moieties (Killops
also from alkyl H. Fig. 3h shows the proton cross section between and Killops, 2005).
115–142 ppm. Thus, by using 1HA13C 2D HETCOR NMR, we have
shown that a significant fraction of non-protonated aromatic car- 3.5. 2D WISE NMR
bon signal is from substituted ring carbons connected to alkyls
1
(rather than from fused aromatic rings, the other possible source HA13C WISE NMR can identify mobile segments, in terms of
of non-protonated carbon). motional narrowing of 1H wideline spectra. In gilsonite, endgroups
28 J.R. Helms et al. / Organic Geochemistry 44 (2012) 21–36

of alkyl chains are mobile. WISE NMR helped distinguish mobile lar dephasing (thin line). The difference between them shows that
and non-protonated carbons, both of which had weak CAH dipolar the lineshape is a superposition of a motionally narrowed and a
couplings and contributed to 13C spectra after dipolar dephasing. more or less rigid component. For reference, the broad lineshape
Further, combined with 1H spin diffusion it provided information of the rigid CH and CH2 groups that resonated between 53 and
on the distance of mobile endgroups from various spectrally re- 44 ppm (with 0.1 ms CP) is also shown (dashed line). Fig. 5c shows
solved segments, in particular aromatic rings. Fig. 5 shows the 1H the corresponding set of 1H wide line spectra associated with 13C
wide-line patterns extracted from WISE spectra at various 13C peak signal at 39 ppm (thick line). Again, differences between the spec-
positions. Two different CP times, 0.1 and 1 ms, were used in order tra without (thick line) and with dipolar dephasing (thin line)
to vary the spin diffusion times and also detectability of protonated prove that there is a superposition of mobile and rigid components,
and non-protonated carbons. Dipolar dephasing of 40 ls was in- but with a much smaller contribution from the mobile segments.
serted to selectively observe the signals of mobile or non-proton- Finally, Fig. 5d displays 1H wide line spectra of carbons resonat-
ated carbons. ing around 140–123 ppm with 0.1 ms CP (thick line), 1 ms CP (thin
Fig. 5a shows the 1H wide-line spectra at 13C chemical shift of line), and 1 ms CP and 40 ls dipolar dephasing (dashed line). The
30 ppm (thick line) and between 44 and 53 ppm (thin line), both sharp peaks of mobile alkyl segments became visible, though with
with 0.1 ms CP and 40 ls dipolar dephasing. The signal at low intensity, for the non-protonated aromatic carbons at long
30 ppm is attributed to mobile CCH2C groups, which survived after cross polarization times of 1 ms. This shows that aromatic cores
dipolar dephasing; the high mobility reduced HAH dipolar cou- and mobile alkyl segments are within approximately 2 nm of one
plings and resulted in a narrower 1H spectrum. The broad 1H spec- another.
trum associated with the carbons resonating between 44 and
15
53 ppm indicates large proton–proton dipolar couplings. There- 3.6. N NMR, with recoupled dipolar dephasing
fore, the weak CAH dipolar coupling associated with these carbons
cannot be attributed to motional narrowing and must be explained Fig. 6a shows the 15N CP-MAS spectrum of gilsonite without
by assigning the 44–53 ppm signal after dipolar dephasing to non- dipolar dephasing, while Fig. 6b shows the 15N CP-MAS spectrum
protonated (quaternary) carbon. of gilsonite with 291 ls recoupled dipolar dephasing. The 15N CP-
WISE NMR can differentiate overlapping contributions from ri- MAS spectrum shows only one broad band ranging from 140–
gid and mobile segments. Fig. 5b shows 1H wide line spectra at 110 ppm (Fig. 6a). Protonated pyrrole N (three bonding partners)
30 ppm with 0.1 ms CP (thick line), and the same after 40 ls dipo- in five-membered rings resonates between 160 and 130 ppm in
15
N spectra, overlapping with amide N, e.g. in peptides, that reso-
nates between 140 and 110 ppm (Thorn et al., 1996). Resonances
of pyrrolic and amide N overlap heavily, but based on the very
low (<1%) amide signal in the solid state 13C NMR spectra, we
can attribute the majority of the signal observed here to pyrrolic
N. Significant 15N NMR signals of amines, which resonate at
<80 ppm, are not observed. No significant signals were detected
in the 15N CP-MAS spectrum after recoupled dipolar dephasing
(Fig. 6b), indicating that all of the detectable N forms are proton-
ated, consistent with pyrrolic NAH groups. Pyridinic N, which is
not protonated and has unfavorable chemical shift positions and
anisotropies, is not detected efficiently in these experiments. From
13
C NMR, we estimated its fraction to be between 10% and 25% of
all N (see above).

3.7. Solution NMR

Fig. 7 shows 1H NMR spectra of gilsonite after various T2 filter


times ranging from 0.2–60 ms. The interval t in the Hahn echo se-
quence (p/2–s–p–s-detection) was adjusted for each experiment.
As s increases, the signals of highly mobile components with long
T2 relaxation time will survive, but those of less mobile

Fig. 5. 1H wideline spectra extracted from 2D 1HA13C wideline separation (WISE)


NMR at 13C chemical shift of (a) 53–44 ppm (thin line) and 30 ppm (thick line), both
with 0.1 ms CP and 40 ls dipolar dephasing delay; (b) 30 ppm (thick line) and 44–
53 ppm (dashed line) with 0.1 ms CP, and 30 ppm with 0.1 ms CP and 40 ls dipolar
dephasing delay (thin line; same data as that in (a) but with less line broadening);
(c) 39 ppm (thick line) and 44–53 ppm (dashed line) with 0.1 ms CP, and 39 ppm
with 0.1 ms CP and 40 ls dipolar dephasing delay (thin line); and (d) 140– Fig. 6. 15N CP-MAS spectra (a) without dipolar dephasing and (b) with 291 ls
123 ppm, acquired under the conditions of 0.1 ms CP (thick line), 1 ms CP (thin recoupled 1HA15N dipolar dephasing, which suppresses the signals of protonated
line), and 1 ms CP combined with 40 ls dipolar dephasing delay (dashed line). nitrogen.
J.R. Helms et al. / Organic Geochemistry 44 (2012) 21–36 29

On the basis of these data and those of Table 1, the carbon fractions
listed in Table 3 were obtained.
Fig. 8 shows the alkyl portion of the 13C solution NMR spectrum
of gilsonite, which matched the spectrum published by Clark et al.
(1983). We have been able to assign the 10 highest peaks to three
specific structures, namely unbranched chain ends, chain segments
with a methyl branch and chain ends with two methyl groups (see
Fig. 8a and b). Comparison with the solid state 13C NMR spectrum
in Fig. 1 and the CH only spectrum in Fig. 2f indicates that a large
fraction of the alkyl signals were invisible in 13C solution NMR. For
instance, the solid state NMR spectra showed significant intensity
above 40 ppm, which was invisible in solution NMR, probably
due to large line widths that resulted from limited mobility.
We also recorded two dimensional COSY and TOCSY 1H NMR as
well as HMQC 1HA13C NMR spectra (not shown). The 1H spectra
confirmed that various alkyl species are chemically bonded. How-
ever, the broad spectral components that account for most of the
signal are not visible and alkyl-aromatic cross peaks were not seen
in these spectra because the scalar couplings between them are too
weak.

3.8. Elemental analysis

The gilsonite sample contained 80.0 ± 0.7% carbon, and


3.3 ± 0.1% nitrogen by weight. For comparison, average values from

Table 3
Functional group quantification from combined NMR data of Tables 1 and 2.

Functional group C percentage NMR


Alkyl CH3 13
Alkyl CH2 50
Fig. 7. 1H solution NMR spectra with various T2 filter times ranging from 0.2–60 ms. Alkyl CH 7
The sharp peak at 7.3 ppm was caused by proton exchange with the solvent, @CH2 2
deuterated chloroform. Aromatic CH 6
Aromatic CAC (non-protonated) 19
Aromatic CAO 2.2
NC@O 0.4
components with short T2 relaxation time will dephase. Clearly, the C@O 0.4
sharp peaks at 1.2 and 0.8 ppm belonged to highly mobile CH2
and CH3, respectively. The broad components at around 2–3 ppm
were assigned to CH2 or CH directly attached to aromatic rings,
which have reduced flexibility. The significant intensity of this
band (17% of all H) provides evidence of abundant alkyl-aromatic
connections. The resonances of NH and CH protons of the pyrrole
and other aromatic rings overlapped at around 7–8 ppm. The sharp
peak at 7.3 ppm was due to exchanged protons on the deuterated
chloroform solvent.
The relative 1H NMR signal areas and assignments are listed in
Table 2. We also calculated the corresponding carbon signal areas.
First, the signal intensity of a CHn group was divided by n, the num-
ber of protons per carbon in the group. The resulting numbers were
then normalized to a total of 100% and finally corrected for the
fraction fnp of non-protonated carbons, which are invisible in 1H
NMR but easily determined by 13C NMR, by dividing each entry
by (1 + fnp). The results are given in the last column of Table 2.

Table 2
Signal intensities in solution 1H NMR, and corresponding carbon fractions, given the
fraction of non-protonated carbons from 13C NMR.
1
ppm Functional group H NMR intensity (%) Corresponding
C fraction (%)
0–1 Alkyl CH3 26 13
1–2 Alkyl CH1.8 52 43
2–5 CH1.8 bonded to aryl 17 14 Fig. 8. (a) Alkyl region of the 13C solution NMR spectrum of gilsonite. Assignments
6–9 Aromatic CH + NH 4.5 7 of the 10 highest peaks to three structures, shown in (b), are indicated above the
– Non-protonated C – 23 spectrum. The chemical shifts for these structures were calculated using the ACD
Lab v.12.0 software.
30 J.R. Helms et al. / Organic Geochemistry 44 (2012) 21–36

Table 4 weight for the combined positive and negative mode spectra (with
Comparison of elemental analysis with average molecular formulas determined from redundant peaks removed) was 533 Da, while the peak intensity
NMR and electrospray Fourier transform ion cyclotron mass spectra (ESI-FT-ICR-MS).
Carbon and nitrogen were measured using a Thermo-Finnegan Flash 1112 series
weighted average molecular weight was 355 Da.
elemental analyzer. N/A: Not measured. The large numbers of peaks observed for gilsonite (Fig. 9) com-
plicated our interpretation of the mass spectral data. However, we
Elemental Solid state ESI-FT-ICR-MS ESI-FT-ICR-MS
analysisa NMR number intensity
have summarized our data and used some relatively simple visual-
weighted weighted ization techniques to characterize the sample. The van Krevelen
a
Carbon wt% 80 (84.0) 83% 79.61 80.10
plot in Fig. 10a was dominated by hydrocarbon peaks (low O/C)
a
Hydrogen wt% (10.67) 11% 10.57 11.15 and lipids (H/C = 1.7–2.25; O/C = 0.02–0.18), as well as a smaller
a
Nitrogen wt% 3.3 (2.40) 3% 3.11 3.25 number of peaks in the regions associated with peptides and pro-
a
Sulfur wt% (0.50) N/A 0.88 1.04 teins (H/C = 1.5–2.0; O/C = 0.2–0.5) (Kujawinski, 2002; Sleighter
a
Oxygen wt% (3.74) 3.2% 5.83 4.32
a
and Hatcher, 2007). Because all peaks in the spectra are singly
H/C atomic ratio (1.51) 1.49 1.59 1.67
charged and have low N/C ratios (<0.2) it is likely that the few
a
Values in parentheses are averages based on literature values (Bell and Hunt, peaks in the protein region are rather degraded proteins, degraded
1963; Clark et al., 1983; Jacob, 1989).
petroporphyrins or simply nitrogen heterocycles, amides, or
amines (Fig. 10b).
Both ESI modes were selective for substituted hydrocarbons. All
previous studies (Bell and Hunt, 1963; Clark et al., 1983; Jacob, of the assigned formulae included at least one heteroatom: N, O or
1989) are reported in Table 4. This table also compiles elemental S. The average (mode) formula included two heteroatoms (mean
compositions calculated from NMR and MS. The NMR elemental value = 3.3 heteroatoms, median value = 3, and no formulae con-
data were calculated from the fractions and CHNO compositions tained more than 16). The median formula (323 of the formulae
of the functional groups in Table 3 (Mao et al., 2000). accounting for 2.7% of the total peak intensity) contained two oxy-
gens and one nitrogen (mean = 1.17 N, 1.93 O, 0.145 S; maxi-
3.9. FT-ICR-MS mum = 8 N, 10 O, 3 S). There was a clear bias in favor of oxygen
containing moieties in the negative ionization mode (Fig. 10; Ta-
Negative ionization mode ESI produced a spectrum (Fig. 9a) ble 4). We found that the most populated group of peaks and the
consisting of 2576 resolved peaks (m/z: 203–730; average highest intensity peaks in negative ionization mode contained
MW = 458), 1914 of which were assigned unique molecular formu- two oxygen atoms and likely represented carboxylic acids, which
lae (C, H, N, O, S; error < 1 ppm). Positive ionization mode ESI gen- were readily ionized by ESI, but were quantitatively negligible in
erated a spectrum (Fig. 9b) of 3663 peaks (m/z: 210–908; average the sample according to NMR. Sorting the formulae (from both
Mw = 573), 2199 of which were assigned unique molecular formu-
lae (C, H, N, O, S; error < 1 ppm). Two hundred twenty-two of the
peaks with assigned formulae (5%) were present in both positive
and negative ionization spectra, indicating that the external and
internal calibrations for positive and negative mode were both suc-
cessful and mutually compatible. The number average molecular

Fig. 9. (a) Positive and (b) negative electrospray ionization FT-ICR mass spectra for Fig. 10. (a) Van Krevelen diagram and (b) 3-D van Krevelen diagram for formula
gilsonite. Inset spectra are shown with enhanced y-axis scales. assignments from positive (black) and negative (gray) electrospray ionization.
J.R. Helms et al. / Organic Geochemistry 44 (2012) 21–36 31

identified porphyrin molecules. However, 115 ions were observed


that had four or more nitrogen atoms, and several of these con-
tained at least 20 carbon atoms and zero, one or two metal ion(s).
These molecules likely represent partially degraded porphyrins
that have retained the tetrapyrrole metal binding core of the origi-
nal molecule. Pyrrolic compounds with alkyl or olefinic side chains
could account for most of the nitrogen containing molecules; how-
ever, they were not readily distinguishable from amines by mass
spectrometry alone and may represent structural isomers or mix-
tures of isomers.
Possible lipid biomarkers identified within ESI-FT-ICR-MS spec-
tra were suggestive of both algae and cyanobacteria. Polyunsatu-
rated fatty acids including chloroplast fatty acids were identified,
as well as some saturated and mono-unsaturated fatty acids typi-
cal of bacteria and cyanobacteria (Kenyon, 1972; Canuel et al.,
1995). Other potential biomarker ions found within the ICR-MS
spectra included phylloquinone (vitamin K; C31H46O2), brassicas-
terol (indicative of algae; C28H46O), and testosterone (indicates
bacteria or animals; C19H28O2).
Indices calculated based on the molecular formulae determined
for the MS peaks (Eqs. (1)–(5); Table 5) suggest that the predomi-
nant structural features of ionized molecules present in gilsonite
are aliphatic hydrocarbons. Nearly 65% of formulae, representing
more than 77% of the total peak intensity, satisfy the definition
of the aliphatic index. In fact, KMD analysis (Table 6) suggests that
Fig. 11. Incidence of heteroatoms expressed as (a) number of formulae containing a 9% of resolved peaks were part of a CH2 series and therefore con-
given number of heteroatoms (O, S, and N), and (b) percentage of total ESI-FT-ICR- tained some aliphatic character while slightly less than 0.1% were
MS peak intensity represented by formulae with a given number of heteroatoms. part of a CH series. The CH KMD series could have included tertiary
aliphatic carbons, as well as protonated aromatic carbons. The per-
centage of peaks identified as containing aromatic rings was be-
positive and negative modes) according to the number of each het-
tween 0.08% and 2.5%, depending on whether the AI > 0.67 or the
eroatom present (Fig. 11) shows that most of the ions contained
less conservative AImod > 0.5 criterion was used. All three formulae
four or fewer oxygen atoms, three or fewer nitrogen atoms and
identified as CARS (0.08%) also satisfy the criterion for AImod > 0.5
one or no sulfur atoms. While this suggests that intensity is not
and BC simultaneously. No formula indices were identified for
an accurate indicator of abundance within the sample, because it
32% of formulae. As previously mentioned, oxygen containing com-
is strongly impacted by variations in ionization efficiency, the bulk
pounds were heavily selected during negative mode ESI and there-
elemental composition of gilsonite based on the molecular formu-
fore COO, OCH2, H2O and O KMDA series make up a much larger
lae calculated for the ESI ionizable fraction of gilsonite accurately
fraction of the analytical window than they represented as a frac-
reflects the bulk elemental composition determined by combus-
tion of the gilsonite sample.
tion elemental analysis (Table 4).
No peaks were observed that were clearly identifiable as previ-
ously reported petroporphyrins (Quirke and Maxwell, 1980; Haji- 4. Discussion
brahim et al., 1981; Gill et al., 1985; Clezy et al., 1989; Qian
et al., 2008; McKenna et al., 2009). Porphyrins were probably not 4.1. Composition from NMR, FT-ICR-MS and elemental analysis
efficiently ionized by ESI within the complex matrix of gilsonite.
Likewise, expanding the molecular formula calculations to include Tables 1–3 indicate that solid state NMR provided non-
vanadium, cobalt, magnesium, and nickel yielded no previously destructive, comprehensive, quantitative structural information

Table 5
Summary of peaks conforming to 5 formula indices: condensed aromatic ring structures (CARS), aromatic index (AI), aliphatic, and black carbon.

CARS (Hockaday AI > 0.67 (Koch and Modified AI > 0.5 (Koch Aliphatic Black carbon
et al., 2006) Dittmar, 2006) and Dittmar, 2006) (Perdue, 1984) (Kim et al., 2004)
No. of peaks 3 3 3 2505 5
% of peaks 0.08 0.08 0.08 64 0.13
% Peak height 0.04 0.04 0.04 78 0.07
Number ave. MW 549 549 549 540 507
Height ave. MW 518 518 518 505 483

Table 6
Homologous series identified by Kendrick mass defect analysis (KMDA).

CH2 CH OCH2 COO H2 H2O O


# of series 343 3 956 633 591 1020 993
# of peaks in series 3649 6 2539 1356 3514 2624 2659
% of total peaks 8.82 0.08 24.57 16.27 15.19 26.21 25.52
# of series with >4 peaks 208 0 64 2 331 33 68
Average # of peaks per series 11 2 3 2 6 3 3
32 J.R. Helms et al. / Organic Geochemistry 44 (2012) 21–36

on gilsonite whereas 1H solution NMR had severe limitations. 13C on the C/N ratio and the aromatic carbon fraction determined by
solid state NMR showed that 70% of carbon nuclei in gilsonite are quantitative 13C NMR, we determined an upper limit to the abun-
in alkyl groups and 27% in aromatics. By contrast, 1H solution dance of pyrrole rings. There are 3.5 N per 100 C in gilsonite and for
NMR provided little useful aromaticity information, since most each N in a pyrrole ring, there are four aromatic carbons. Thus,
aromatic carbons in this sample are not protonated. The aromatics there are no more than 3.5  4 = 14 pyrrole carbons per 100 C
also did not show up well in solution 13C NMR spectra, due to their (14%). Given the 27% aromatic fraction, at least 27–14 = 13% of car-
low mobility and resulting large line widths. Standard 1H solution bon are in non-pyrrolic aromatics.
NMR could not determine the CH2 to CH ratio in gilsonite, which The preceding analysis gives only an upper limit (14%) for the
was reliably estimated from solid state NMR with spectral editing. pyrrolic carbon fraction. The fraction of carbons that are in fact
On the other hand, the 1H NMR spectrum gave the ratio of CH3 to pyrrolic is estimated based on the signal of non-protonated C at
CH2, which cannot be obtained from solid state NMR in gilsonite <133 ppm. Aromatic carbons resonate in this range only if they
since many CH2 groups have high mobilities similar to those of are at a two-bond distance from N or O (and not bonded to O them-
CH3 groups. The good agreement of elemental composition be- selves) (Bovey et al., 1988). Thus, the signal of non-protonated car-
tween elemental analysis and atom ratios calculated from NMR bons at <133 ppm (11%, shaded in Fig. 1b) is due to C substituted
data confirms the reliability of NMR (Table 4). phenols, pyridines and pyrroles. There are two such carbons per
Mass spectral data based on ESI for gilsonite were specific for the heteroatom and consequently the contributions from phenols to
polar (ionic) fraction of gilsonite, but in many respects confirmed the non-protonated carbon signal at <130 ppm must be twice that
key NMR results. The aliphatic chemical index accounted for 64% of the aromatic CAO and pyridinic CAN observed between 150 and
of assigned formulas or 78% of the total peak intensity (Table 5) 165 ppm (2.2%, shaded in Fig. 1a). Thus, the pyrrolic carbons two
and 9% of assigned formulae were part of a CH2 homologous series bonds from N accounts for approximately 11–4.4% = 7% of all C.
Table 6). Solid state NMR showed that aromatic CAO accounts for Therefore, the pyrrolic carbons accounted for 2  7% = 14% of all
most of the organically bound oxygen, which is only 2.2% of C (Ta- C, and other aromatics for 27–14 = 13%. Further, 14% C in pyrrole
ble 1). Both NMR and MS data showed that gilsonite is primarily ali- rings corresponds to 14/4 = 3.5 N per 100 C. This means that nearly
phatic. Although the MS data suggests organic acids are present, the all the N is in pyrrole rings.
NMR results show that carboxylic carbons are extremely minor
constituents of gilsonite (less than 0.5%; Table 1), highlighting the 4.3. Highly substituted aromatic rings: refractory organic matter
need for a more appropriate and representative ionization method
for quantitative characterization of this type of sample, but con- The aromatic rings in gilsonite are highly substituted by alkyls.
firming the conclusions of Klein et al. (2006) that indicated that Three independent lines of evidence supported this conclusion.
the polar fraction of petroleum samples can be successfully charac- First, quantitative DP-MAS and DP-MAS results (Fig. 1; Table 1)
terized using ESI-FT-ICR-MS with no pre-infusion fractionation. indicated that 77% aromatics were non-protonated and long range
Despite the presence of mass spectral peaks with H/C and O/C dipolar dephasing data (Fig. 4) clearly showed that the aromatics in
ratios typically associated with protein (Fig. 10a; Sleighter and gilsonite did not form large clusters of fused aromatic rings. Solu-
Hatcher, 2007), we found no quantitative NMR evidence for their tion and solid state NMR agree that aromatic H only accounts for a
being present in the sample. If protein were present in 13C NMR small fraction (5%) of total protons, see Tables 1–3. Specifically,
detectable quantities, it would have shown a peak near 172 ppm Fig. 1 showed only minor signals from aromatic CAH two bonds
due to NAC@O, which was not observed significantly. 15N and from N, which typically resonate at <115 ppm. This demonstrates
13
C NMR indicated that much of the organic nitrogen is present that the carbons two bonds from N in the pyrrole ring are mostly
as five-membered pyrrolic rings (Fig. 6; Thorn et al., 1996). This non-protonated and likely substituted by alkyl segments. Second,
1
is consistent with the molecular formula assignments for nitrogen HA13C 2D HETCOR with 40 ls dipolar dephasing (Fig. 3) showed
containing peaks. Chlorophylls and their anoxic degradation prod- that the protons near non-protonated aromatics are mostly in alkyl
ucts consist largely of macrocyclic moieties containing up to four groups. Third, the broad band of 1H solution NMR around 2–3 ppm
pyrrolic rings. The fact that nearly all of the FT-ICR-MS formulae could be attributed to CH2 or CH directly bonded to aromatic rings,
in the three-dimensional van Krevelen diagram (Fig. 10b) fell along which accounts for 1/5 of all alkyl groups, providing further evi-
rays which intersect at (N/C = 0, O/C = 0, H/C = 2), suggests dence of abundant alkyl-aromatic connections.
that gilsonite consists largely of a complex array of intermediates Our previous studies of type II kerogen (Mao et al., 2010a) and
of multiple diagenetic/catagenetic processes with chemically and coal (Mao et al., 2010b) indicated that most of the aromatics in
structurally varied starting material (biomolecules) and a far less these geologically preserved organic matter samples are non-pro-
diverse end point (primarily aliphatic hydrocarbons). tonated and in moderately sized clusters of fused rings. Coupled
If we consider that elemental analysis (Table 4) is suggestive of with the results in the present study, we hypothesize that non-pro-
the molecular formula of some arbitrarily ‘average’ gilsonite mole- tonated aromatics, either in the form of fused rings or highly
cule, we calculate a formula of C40H61N1O1 with an additional sul- substituted rings, are refractory components in nature. Further-
fur in 10% of molecules and an additional oxygen in 35% of more, our study of a fulvic acid from Antarctica showed that dia-
molecules. This gives a DBE of approximately 9, suggesting that genesis resulted in a significant fraction of non-protonated alkyl
no more than two aromatic ring structures are expected per mole- and aromatic carbons (Mao et al., 2007a), consistent with our
cule (benzene DBE = 4). This also yields a molecular weight of hypothesis here. Therefore, we propose that non-protonated car-
571.5 Da, which is consistent with the average MW determined bons can be used as a humification index.
by ESI-MS, but well below the molecular weight reported in Dickie
and Yen (1967) for gilsonite and other bitumens. Similarly, the 4.4. Structural model
median C, H, N, O and S values from the MS data yield a ‘typical’
formula of C34H54N1O2. Our results showed that gilsonite bitumen is composed of pyr-
role and other aromatic rings highly substituted with alkyl chains.
4.2. Pyrrole rings The 1H solution NMR indicated that approximately 14/70 = 20% of
the alkyl groups (1 in 5) are bonded to aromatic rings (Table 2). If
Aromatic carbons in gilsonite can be part of five-membered most alkyl chains link aromatic rings (Figs. 3 and 5), the linker
(mostly pyrrole) or ‘‘regular’’ six-membered aromatic rings. Based between the two rings consists of 2  5 = 10 alkyl carbons on
J.R. Helms et al. / Organic Geochemistry 44 (2012) 21–36 33

Fig. 12. Model of the main components of gilsonite structure. A typical ‘‘monomer unit’’ containing one aromatic ring and 21 carbons in total is outlined by the dashed ellipse.
Gilsonite exists as a highly diverse mixture of molecules of various size and structure. The model structure shown here represents the high molecular weight fraction, while
smaller fragments are also common.

average. Also, we concluded that 14/27 or about 50% of all aromatic 4.5. Comparison with literature
carbons are bonded to alkyl segments (Table 2). If we assume
about 6 aromatic carbons per isolated ring or cluster to be typical, ESI-FT-ICR-MS detected a vanishingly small number of ions sug-
50% or 3 carbons are bonded to segments of 5 alkyl carbons, giving gestive of condensed hydrocarbons unlike earlier studies that sug-
an average ‘‘monomer’’ of 15 alkyl and 6 aromatic carbons, with al- gested that gilsonite contains significant quantities of condensed
kyl CH3:alkyl CH2:alkyl CH:aromatic C proportions of 3:11:1.5:6 aromatics or so-called asphaltene sheets (Yen et al., 1961; Wen
(Table 3). The majority of organic oxygen in our gilsonite sample et al., 1978; Clark et al., 1983). Also, 13C DP-TOSS long range dipolar
was apparently present as aromatic CAO with only traces of car- dephasing curves (Fig. 4) indicated that gilsonite contains primar-
boxyl and carbonyl groups detected (Fig. 1; Tables 1–3). ily clusters of only a few fused aromatic rings, which shows that
Based on the information obtained in this study, we propose a large fused ring molecules such as asphaltene sheets are not pres-
structural model in which five-membered pyrrole rings and small ent at NMR detectable concentrations in gilsonite. This finding may
clusters of fused six-membered rings are connected to, and linked invalidate much of the structural information reported previously
by, alkyl chains of an average length of 10 carbons, see Fig. 12. (Yen et al., 1961) that assumed asphaltene sheets were a major
Whether the pyrrole rings are linked to the fused rings by alkyl structural component of the mixture and indicates the need for
chains was not clear from our data. Since pyrrole rings contain reinterpretation of the X-ray diffraction data (Yen et al., 1961). This
fewer carbons than fused six-membered aromatic rings, a larger finding also differentiates gilsonite, and possibly other similar na-
number of pyrrole rings was needed to match the number of car- tive bitumens, structurally from process asphalts that are byprod-
bons in the fused rings. The large fraction of CH2 and small amount ucts of oil refinement. Our results do not rule out the existence of
of CH groups showed that linear alkyl chains predominate. Methyl aromatic sheets in gilsonite, but suggest they are a small fraction of
end and side groups were detected by solution 13C and 1H NMR the aromatics or are largely composed of nitrogen-containing mac-
(Fig. 8), and the methyl branches could account for a large fraction rocycles rather than carbonaceous condensed aromatics.
of the observed alkyl CH groups. Earlier studies involving petroporphyrins in gilsonite suggested
Clearly, not every molecule of gilsonite resembles the structure an anoxic and therefore relatively rapid and comprehensive pres-
in Fig. 12, but molecules exhibiting similar structural features must ervation of algal material prior to burial and catagenesis as the ori-
be very common. We reiterate that gilsonite is a complicated mix- gin of gilsonite (Quirke et al., 1980b; Ruble et al., 1994). Our results
ture with a broad ranging continuum of fragments with varying are consistent with this view. The presence of aromatic molecules
molecular weights. The model structure is depicted as a polymeric in gilsonite suggests that either (i) a more comprehensive suite of
structure because a significant number of the pyrrole and/or aro- biomolecules survived diagenesis than a typical algaenan (e.g. Bot-
matic subunits appear to be connected in an amorphous polymer ryococcus braunii), (ii) the dominant species of algae present during
or network structure. Based on the molecular weight distribution sedimentation produced algaenan similar to Chlorella marina,
reported here (Tables 5 and 6) and elsewhere, the degree of inter- which were shown to yield aromatic hydrocarbons upon pyrolysis
connection and the length of the connecting aliphatic chains are (Derenne et al., 1996), or (iii) that the pyrrolic moieties within the
probably highly variable. Macrocycles such as petroporphyrins porphyrin core of algal pigments was efficiently preserved. It is not
and refractory biomolecules represent a small fraction of the gil- surprising that we observed six-membered aromatic rings since it
sonite as shown here by FT-ICR-MS and determined previously has been commonly reported that six-membered aromatic rings
(Quirke and Maxwell, 1980; Ruble et al., 1994; Schoell et al., are preserved and/or produced during diagenesis and catagenesis
1994). Oxygen is present mainly as aromatic CAO and some (North, 1985; Killops and Killops, 2005). Our results did not indi-
C@O, according to NMR. cate a significant contribution from lignin, preserved structural
34 J.R. Helms et al. / Organic Geochemistry 44 (2012) 21–36

features of which are observed in coal samples (Behar and Hatcher, Acknowledgments
1995). Earlier studies have also suggested that gilsonite and other
native bitumens experienced relatively low temperature catagene- We would like to thank the National Science Foundation (EAR-
sis based on the preservation of thermally labile biomarkers (Eglin- 0843996 and CBET-0853950) and the Donors of the Petroleum Re-
ton et al., 1979). Our finding of extremely low abundance or search Fund, administered by the American Chemical Society Grant
absence of condensed aromatics, which form during high temper- 46373-G2 for financial support. We also thank Dr. Gary G. Thomp-
ature catagenesis and charring, supports this as well. Exomethyl- son for technical support and Dr. Rachel Sleighter for assistance
ene, which was detected and identified by 13C NMR, is often with FT-ICR-MS. Dr. Hussain Abdulla, Dr. Sylvie Derenne, and
associated with cyclization and cross linking reactions that occur two anonymous reviewers provided helpful comments.
during diagenesis and catagenesis and remove olefinic character
from biogenic resins (Clifford et al., 1999).
Possible environmental sources of pyrrole rings include de-
Associate Editor—Sylvie Derenne
graded pigments (Quirke and Maxwell, 1980; Quirke et al.,
1980b; Hajibrahim et al., 1981; Leenheer, 2003, 2009) and Mail-
lard reaction products from protein and carbohydrate starting
References
materials (Hodge, 1953; Hayase et al., 1989; Fang and Schmidt-
Rohr, 2009). The bulk structure of gilsonite does not resemble Abraham, H., 1945. Asphalts and Allied Substances. Van Nostrand, New York.
that of modern chlorophylls or their degradation products in Amster, J.I., 1996. Fourier transform mass spectrometry. Journal of Mass
plants or algae (Brown et al., 1984; Takamiya et al., 2000). For Spectrometry 31, 1325–1337.
Anderson, K.B., Winans, R.E., Botto, R.E., 1992. The nature and fate of natural resins
chlorophylls, there are methyl or ethyl groups on the pyrrole in the geosphere-II. Identification, classification and nomenclature of resinites.
rings, the CH3 of which would have resonated at around 9 ppm Organic Geochemistry 18, 829–841.
for 13C (2 ppm for 1H) and 16 ppm for 13C (1.3 ppm for 1H), Aue, W.P., Bartholdi, E., Ernst, R.R., 1976. Two dimensional spectroscopy application
to nuclear magnetic resonance. Journal of Chemical Physics 64, 2229–2246.
respectively. These signals were not observed in gilsonite. While Baker, E.W., Yen, T.F., Dickie, J.P., Rhodes, R.E., Clark, L.F., 1967. Mass spectrometry of
some chlorophyll products such as petroporphyrins may have porphyrins. II. Characterization of petroporphyrins. Journal of the American
been preserved with the chlorin ring intact, our results suggest Chemical Society 89, 3631–3639.
Bax, A., Davis, D.G., 1985. MLEV-17-based two-dimensional homonuclear
that the majority of the macrocycles must have been opened magnetization transfer spectroscopy. Journal of Magnetic Resonance 65, 355–
(Leenheer, 2003) and most of the unsaturated bonds must have 360.
been hydrogenated. It might be suggested that Maillard or other Behar, F., Hatcher, P.G., 1995. Artificial coalification of a fossilwood from brown coal
by confined system pyrolysis. Energy and Fuels 9, 984–994.
similar biomolecular reaction products might have provided sig-
Bell, K.G., Hunt, J.M., 1963. Native bitumens associated with oil shales. In: Breger,
nificant sources of pyrrole rings in gilsonite; however, it was I.A. (Ed.), Organic Geochemistry. Elsevier, Amsterdam.
shown recently that the Maillard reaction produces mostly non- Bovey, F.A., Jelinski, L., Mirau, P.A., 1988. Nuclear Magnetic Resonance Spectroscopy.
protonated N (Fang and Schmidt-Rohr, 2009), while N in gilsonite Academic Press, San Diego, CA.
Brewer, C.E., Schmidt-Rohr, K., Satrio, J.A., Brown, R.C., 2009. Characterization of
is mostly protonated. biochar from fast pyrolysis and gasification systems. Environmental Progress
Electrospray FT-ICR-MS provided an extensive database of and Sustainable Energy 28, 386–396.
molecular formulae for polar constituents of gilsonite. ESI pro- Brown, C.E., Spenser, R.B., Burger, V.T., Katz, J.J., 1984. Cross-polarization/magic-
angle sample-spinning 13C NMR spectroscopic study of chlorophyll a in the
vided supporting evidence for the NMR detected presence of pri- solid state. Proceedings of the National Academy of Sciences of the United
marily pyrrolic nitrogen compounds in gilsonite and revealed States of America 81, 641–644.
several peaks that were likely structurally very similar to petro- Canuel, E.A., Cloern, J.E., Ringelberg, D.B., Guckert, J.B., Rau, G.H., 1995. Molecular
and isotopic tracers used to examine sources of organic matter and its
porphyrins. Petroporphyrins, hydrocarbons, and other non-polar incorporation into the food webs of San Francisco Bay. Limnology and
molecules may be more readily ionizable by employing liquid Oceanography 40, 67–81.
injection field desorption ionization (Smith et al., 2008), laser in- Clark, D., Wilson, R., Quirke, J.M.E., 1983. An evaluation of the potential of ESCA
(electron spectroscopy for chemical applications) (and other spectroscopic
duced acoustic desorption (Pinkston et al., 2009), or atmospheric techniques) in the surface and bulk characterisation of kerogens, brown coal
pressure photo-ionization (Purcell et al., 2007; Qian et al., 2008; and gilsonite. Chemical Geology 39, 215–239.
McKenna et al., 2009) in place of ESI, and these modes of ioniza- Clezy, P.S., Fookes, C.J.R., Prashar, J.K., 1989. The chemistry of pyrrolic compounds
LXI. Petroporphyrins from the Julia Creek Oil Shale: further evidence for the
tion should be used in addition to ESI in the characterization stud-
derivation of aetiotype petroporphyrins from chlorophyll. Australian Journal of
ies of mostly hydrophobic natural organic mixtures in the future. Chemistry 42, 775–786.
The lack of mass spectral differentiation between structural Clifford, D.J., Hatcher, P.G., Botto, R.E., Muntean, J.V., Anderson, K.B., 1999. The
isomers using FT-ICR-MS suggests that a physical or chemical sep- nature and fate of natural resins in the geosphere IX. Structure and maturation
similarities of soluble and insoluble polylabdanoids isolated from tertiary class I
aration method such as extraction, reversed phase liquid chroma- resinites. Organic Geochemistry 30, 635–650.
tography, or gel permeation chromatography would be useful Comisarow, M.B., Marshall, A.G., 1974. Frequency-sweep Fourier transform ion
prior to infusion into the mass spectrometer. The need for further cyclotron resonance spectroscopy. Chemical Physics Letters 26, 489–490.
Conte, P., Spaccini, R., Piccolo, A., 2004. State of the art of CPMAS 13C-NMR
development of tandem MS methods that are compatible with FT- spectroscopy applied to natural organic matter. Progress in Nuclear Magnetic
ICR-MS (Sleighter, 2009) is similarly indicated. A more representa- Resonance Spectroscopy 44, 214–223.
tive mass spectral database would provide further insights into Corwin, A.H., 1959. Petro porphyrins. In: Proceedings of the 5th World Petroleum
Congress, New York, 5(109).
the molecular formulae present in natural organic mixtures, DeAzevedo, E.R., Hu, W.-G., Bonagamba, T.J., Schmidt-Rohr, K., 2000. Principles of
would better complement the NMR techniques demonstrated in centerband-only detection of exchange in solid-state nuclear magnetic
this study, and would better facilitate the proposal of model resonance, and extension to four-time centerband-only detection of exchange.
Journal of Chemical Physics 112, 8988–9001.
molecular structures. Dennis, L.W., Maciel, G.E., Hatcher, P.G., Simoneit, B.R.T., 1982. 13C nuclear magnetic
The new structural information obtained using advanced NMR resonance studies of kerogen from Cretaceous black shales thermally altered by
techniques, especially the structural model proposed, represents basaltic intrusions and laboratory simulations. Geochimica et Cosmochimica
Acta 46, 901–907.
a significant step forward in the comprehensive chemical charac-
Derenne, S., Largeau, C., Berkaloff, C., 1996. First example of an algaenan yielding an
terization of gilsonite bitumen. While the use of ultrahigh resolu- aromatic-rich pyrolysate. Possible geochemical implications on marine kerogen
tion mass spectrometry is hampered somewhat by the selectivity formation. Organic Geochemistry 24, 617–627.
of ionization methods currently available, it is clear that the ability Dickie, J.P., Yen, T.F., 1967. Macrostructure of the asphaltic fractions by various
instrumental methods. Analytical Chemistry 39, 1847–1852.
to accurately calculate molecular formulae based on mass spectral Dixon, W.T., 1982. Spinning-sideband-free and spinning-sideband-only NMR
data has extraordinary potential as an analytical tool. spectra in spinning samples. Journal of Chemical Physics 77, 1800–1809.
J.R. Helms et al. / Organic Geochemistry 44 (2012) 21–36 35

Eglinton, G., Hajibrahim, S.K., Maxwell, J.R., Quirke, J.M.E., Shaw, G.T., Volkman, J.K., Leenheer, J.A., 2009. Systematic approaches to comprehensive analyses of natural
Wardroper, A.M.K., 1979. Lipids of aquatic sediments, recent and ancient. organic matter. Annals of Environmental Science 3, 1–130.
Transactions of the Royal Society 293, 69–91. Maciel, G.E., Bronniman, C.E., Ridenour, C.F., 1993. H-1-NMR spectroscopy –
Fang, X., Schmidt-Rohr, K., 2009. Fate of the amino acid in glucose-glycine approaches for carbonaceous solids. Advances in Chemistry Series 229, 27–44.
melanoidins investigated by solid-state nuclear magnetic resonance (NMR). Mao, J.-D., Schmidt-Rohr, K., 2003. Recoupled long-range C–H dipolar dephasing in
Journal of Agricultural and Food Chemistry 57, 10701–10711. solid-state NMR, and its use for selection of fused aromatic rings. Journal of
Gill, J.P., Evershed, R.P., Chicarelli, M.I., Wolff, G.A., Maxwell, J.R., Eglinton, G., 1985. Magnetic Resonance 162, 217–227.
Capillary gas chromatographic–mass spectrometric studies of the Mao, J.-D., Schmidt-Rohr, K., 2004a. Accurate quantification of aromaticity and
petroporphyrins of the gilsonite bitumen (Eocene USA). Journal of nonprotonated aromatic carbon fraction in natural organic matter by 13C solid-
Chromatography 350, 37–62. state nuclear magnetic resonance. Environmental Science and Technology 38,
Grantham, P.J., Douglas, A.G., 1977. Carboxylic acids in some Uinta Basin bitumens. 2680–2684.
In: Campos, R., Goñi, J. (Eds.), Advances in Organic Geochemistry. Enadimsa, Mao, J.-D., Schmidt-Rohr, K., 2004b. Separation of aromatic-carbon 13C NMR signals
Madrid, pp. 193–207. from di-oxygen alkyl bands by a chemical-shift-anisotropy filter. Solid State
Grigsby, R.D., 1989. Processing of high resolution mass spectral data by use of Nuclear Magnetic Resonance 26, 36–45.
Kendrick masses in a rectangular array. In: Ashe, T.R., Wood, K.V. (Eds.), Novel Mao, J.-D., Schmidt-Rohr, K., 2005. Methylene spectral editing in solid-state 13C
Techniques in Fossil Fuel Mass Spectrometry. American Society for Testing and NMR by three-spin coherence selection. Journal of Magnetic Resonance 176, 1–
Materials, Philadelphia, pp. 172–193. 6.
Hajibrahim, S.K., Quirke, J.M.E., Eglinton, G., 1981. Petroporphyrins V. Structurally- Mao, J.-D., Schmidt-Rohr, K., 2006. Absence of mobile carbohydrate domains in dry
related porphyrin series in bituments, shales and petroleums – evidence from humic substances proven by NMR, and implications for organic-contaminant
HPLC and mass spectrometry. Chemical Geology 32, 173–188. sorption models. Environmental Science and Technology 40, 1751–1756.
Hatcher, P.G., Schnitzer, M., Dennis, L.W., Maciel, G.E., 1981. Aromaticity of humic Mao, J.-D., Hu, W.-G., Schmidt-Rohr, K., Davies, G., Ghabbour, E.A., Xing, B., 2000.
substances in soils. Soil Society of America Journal 45, 1089–1094. Quantitative characterization of humic substances by solid-state carbon-13
Hatcher, P.G., Spiker, E.C., Szeverenyi, N.M., Maciel, G.E., 1983. Selective nuclear magnetic resonance. Soil Science Society of America Journal 64, 873–
preservation and origin of petroleum-forming aquatic kerogen. Nature 305, 884.
498–501. Mao, J.-D., Cory, R.M., McKnight, D.M., Schmidt-Rohr, K., 2007a. Characterization of
Hayase, F., Nagaraj, R.H., Miyata, S., Njoroge, F.G., Monnier, V.M., 1989. Aging of a nitrogen-rich fulvic acid and its precursor algae from solid state NMR. Organic
proteins: immunological detection of a glucose-derived pyrrole formed during Geochemistry 38, 1277–1292.
Maillard reaction in vivo. Journal of Biological Chemistry 263, 3758–3764. Mao, J.-D., Fang, X., Schmidt-Rohr, K., Carmo, A.M., Hundal, L.S., Thompson, M.L.,
Herniman, J.M., Langley, G.J., Bristow, T.W.T., O’Connor, G., 2005. The validation of 2007b. Molecular-scale heterogeneity of humic acid in particle-size fractions of
exact mass measurements for small molecules using FT-ICRMS for improved two Iowa soils. Geoderma 140, 17–29.
confidence in the selection of elemental formulas. Journal of the American Mao, J.-D., Tremblay, L., Gagne, J.-P., Kohl, S., Rice, J., Schmidt-Rohr, K., 2007c. Humic
Society of Mass Spectrometry 16, 1100–1108. acids from particulate organic matter in the Saguenay Fjord and the St.
Hertkorn, N., Benner, R., Frommberger, M., Schmitt-Kopplin, P., Witt, M., Kaiser, K., Lawrence Estuary investigated by advanced solid-state NMR. Geochimica et
Kettrup, A., Hedges, J.I., 2006. Characterization of a major refractory component Cosmochimica Acta 71, 5483–5499.
of marine dissolved organic matter. Geochimica et Cosmochimica Acta 70, Mao, J.-D., Fang, X., Lan, Y., Schimmelmann, A., Mastalerz, M., Xu, L., Schmidt-Rohr,
2990. K., 2010a. Chemical and nanometer-scale structure of kerogen and its change
Hockaday, W.C., Grannas, A.M., Kim, S., Hatcher, P.G., 2006. Direct molecular during thermal maturation investigated by advanced solid-state 13C NMR
evidence for the degradation and mobility of black carbon in soils from spectroscopy. Geochimica et Cosmochimica Acta 74, 2110–2127.
ultrahigh-resolution mass spectral analysis of dissolved organic matter from a Mao, J.-D., Schimmelmann, A., Mastalerz, M., Hatcher, P.G., Li, Y., 2010b. Structural
fire-impacted forest soil. Organic Geochemistry 37, 501–510. features of a bituminous coal and their changes during low-temperature
Hodge, J.E., 1953. Dehydrated foods: chemistry of browning reactions in model oxidation and loss of volatiles investigated by advanced solid-state NMR
systems. Journal of Agricultural and Food Chemistry 1, 928–943. spectroscopy. Energy and Fuels 24, 2536–2544.
Hsu, C.S., Llang, Z., Campanas, J.E., 1994. Hydrocarbon characterization by ultrahigh Marshall, A.G., Hendrickson, C.L., Jackson, G.S., 1998. Fourier transform ion
resolution Fourier transform ion cyclotron resonance mass spectrometry. cyclotron resonance mass spectrometry: a primer. Mass Spectrometry
Analytical Chemistry 66, 850–855. Reviews 17, 1–35.
Hu, J.Z., Harper, J.K., Taylor, C., Pugmire, R.J., Grant, D.M., 2000. Modified spectral McKenna, A.M., Purcell, J.M., Rodgers, R.P., Marshall, A.G., 2009. Identification of
editing methods for 13C CP/MAS experiments in solids. Journal of Magnetic vanadyl porphyrins in a heavy crude oil and raw asphaltene by atmospheric
Resonance 142, 326–330. pressure photoionization Fourier transform ion cyclotron (FT-ICR) mass
Hughey, C.A., Hendrickson, C.L., Rodgers, R.P., Marshall, A.G., 2001. Kendrick mass spectrometry. Energy and Fuels 23, 2122–2128.
defect spectrum: a compact visualanalysis for ultrahigh-resolution broadband Meyer, R.F., Attanasi, E.D., Freeman, P.A., 2007. Heavy Oil and Natural Bitumen
mass spectra. Analytical Chemistry 73, 4676–4681. Resources in Geological Basins of the World, Open Form Document No. 2007-
Jacob, H., 1989. Classification, structure, genesis and practical importance of natural 1084. US Geological Survey, Reston, VA, pp. 1–42.
solid oil bitument (‘‘migrabitumen’’). International Journal of Coal Geology 11, Miknis, F.P., Smith, J.W., Maughan, E.K., Maciel, G.E., 1982. Nuclear magnetic
65–79. resonance: a technique for direct evaluation of source-rock potential. American
Keeler, C., Maciel, G.E., 2000. 13C NMR spectral editing of humic materials. Journal of Association of Petroleum Geologists Bulletin 66, 1396–1401.
Molecular Structure 550–551, 297–305. Nanny, M.A., Minear, R.A., Leenheer, J.A., 1997. Nuclear magnetic resonance
Kendrick, E., 1963. A mass scale based on CH2 = 14.0000 for high resolution mass spectroscopy in environmental chemistry. In: Topics in Environmental
spectrometry of organic compounds. Analytical Chemistry 35, 2146–2154. Chemistry. Oxford University Press, New York.
Kenyon, C.N., 1972. Fatty acid composition of unicellular strains of blue-green algae. North, F.K., 1985. Petroleum Geology. Allen & Unwin, Boston.
Journal of Bacteriology 109, 827–834. Osmond, J.C., 1964. Guidebook to the geology and mineral resources of the uinta
Killops, S., Killops, V., 2005. Introduction to Organic Geochemistry. Blackwell basin. In: Sabatka, E.F. (Ed.), 13th Annual Field Conference of the Intermountain
Publishing, Boston, MA. Association of Petroleum Geologists. Intermountain Association of Petroleum
Kim, S., Kramer, R.W., Hatcher, P.G., 2003. Graphical method for analysis of Geologists, pp. 47–58.
ultrahigh-resolution broadband mass spectra of natural organic matter, the van Perdue, E.M., 1984. Analytical constraints on the structural features of humic
Krevelen diagram. Analytical Chemistry 75, 5336–5344. substances. Geochimica et Cosmochimica Acta 48, 1435–1442.
Kim, S., Kaplan, L.A., Benner, R., Hatcher, P.G., 2004. Hydrogen-deficient molecules Picard, M.D., High, L.R., 1968. Sedimentary cycles in the Green River Formation
in natural riverine water samples evidence for the existence of black carbon in (Eocene), Uinta Basin, Utah. Journal of Sedimentary Research 38, 378–383.
DOM. Marine Chemistry 92, 225–234. Pinkston, D.S., Duan, P., Gallardo, V.A., Habicht, S.C., Tan, X., Qian, K., Gray, M.,
Kim, S., Rodgers, R.P., Marshal, A.G., 2006. Truly ‘exact’ mass: elemental composition Mullen, K., Kenttamaa, H.I., 2009. Analysis of asphaltenes and asphaltene model
can be determined uniquely from molecular mass measurements at 0.1 mDa compounds by laser-induced acoustic desorption/Fourier transform ion
accuracy for molecules up to 500 Da. International Journal of Mass cyclotron resonance mass spectrometry. Energy and Fuels 23, 5564–5570.
Spectrometry 251, 260–265. Preston, C.M., 1996. Application of NMR to soil organic matter analysis: history and
Klein, G.C., Angstrom, A., Rodgers, R.P., Marshal, A.G., 2006. Use of saturates/ prospects. Soil Science 16, 144–166.
aromatics/resins/asphaltenes (SARA) fractionation to determine matrix effects Purcell, J.M., Hendrickson, C.L., Rodgers, R.P., Marshall, A.G., 2007. Atmospheric
in crude oil analysis by electrospray ionization Fourier transform ion cyclotron pressure photoionization proton transfer for complex organic mixtures
resonance mass spectrometry. Energy and Fuels 20, 668–672. investigated by Fourier transform ion cyclotron resonance mass spectrometry.
Koch, B.P., Dittmar, T., 2006. From mass to structure: an aromaticity index for high- Journal of the American Society of Mass Spectrometry 18, 1682–1689.
resolution mass data of natural organic matter. Rapid Communications in Mass Qian, K., Mennito, A.S., Edwards, K.E., Ferrughelli, D.T., 2008. Observation of vanadyl
Spectrometry 20, 926–932. porphyrins and sulfur-containing vanadyl porphyrins in a petroleum
Kujawinski, E.B., 2002. Electrospray ionization Fourier transform ion cyclotron asphaltene by atmospheric pressure photoionization Fourier transform ion
resonance mass spectrometry (ESI FT-ICR MS): characterization of complex cyclotron resonance mass spectrometry. Rapid Communications in Mass
environmental mixtures. Environmental Forensics 3, 207–216. Spectrometry 22, 2153–2160.
Leenheer, J.A., 2003. Comprehensive Characterization of Dissolved and Colloidal Quirke, J.M.E., Maxwell, J.R., 1980. Petrophyrins III. Characterisation of a C32
Organic Matter in Waters Associated with Groundwater Recharge at the Orange aetioporphyrin from gilsonite as the bis[porphyrinato-mercurcy(II)
County Water District, Final Report to the Orange County Water District, acetato]mercury(II) complex. Origin and significance. Tetrahedron 36, 3453–
Fountain Valley, California, US Geological Survey, Washington, DC. 3456.
36 J.R. Helms et al. / Organic Geochemistry 44 (2012) 21–36

Quirke, J.M.E., Maxwell, J.R., Eglinton, G., 1980a. Petrophyrins IV. Nuclear electrospray ionization fourier transform ion cyclotron resonance mass spectra.
overhauser enhancement 1H NMR studies of deoxophylloerythroetio Analytical Chemistry 75, 1275–1284.
porphyrins from gilsonite. Tetrohedron Letters 21, 2987–2990. Sugihara, J.M., McGee, L.R., 1957. Porphyrins in gilsonite. Journal of Organic
Quirke, J.M.E., Shaw, G.J., Soper, P.D., Maxwell, J.R., 1980b. Petroporphyrins II. The Chemistry 22, 795.
presence of porphyrins with extended alkyl substituents. Tetrahedron 36, Takamiya, K.-I., Tsuchiya, T., Ohta, H., 2000. Degradation pathway(s) of chlorophyll:
3261–3267. what has gene cloning revealed? Trends in Plant Science 5, 426–431.
Ruble, T.E., Bakel, A.J., Philip, R.P., 1994. Compound specific isotopic variability in Tekely, P., Palmas, P., Mutzenhardt, P., 1993. Elimination of hetero-nuclear dipolar
Uinta Basin native bitumens: paleoenviromental implications. Organic interactions from carbon-13-detected proton spectra in wide-line separation
Geochemistry 21, 661–671. nuclear magnetic resonance spectroscopy. Macromolecules 26, 7363–7365.
Schmidt-Rohr, K., Mao, J.-D., 2002. Efficient CH-group selection and identification in Thorn, K.A., Goldenburg, W.S., Younger, S.J., Weber, E.J., 1996. Covalent binding of
13
C solid-state NMR by dipolar DEPT and 1H chemical-shift filtering. Journal of aniline to humic substances: comparison of nucleophilic addition, enzyme-, and
the American Chemical Society 124, 13938–13948. metal-catalyzed reactions by 15N NMR. In: Gaffney, J.S., Marley, N.A., Clark, S.B.
Schmidt-Rohr, K., Clauss, J., Spiess, H.W., 1992. Correlation of structure, mobility, (Eds.), Humic and Fulvic Acids: Isolation, Structure, and Environmental Role.
and morphological information in heterogeneous polymer materials by two- American Chemical Society, Washington, DC, pp. 299–326.
dimensional wideline-separation NMR spectroscopy. Macromolecules 25, Treibs, A., 1934. Organic minerals III. Chlorophyll and hemin derivatives in
3273–3277. bituminous rocks, petroleums, mineral waxes and asphalts. Origin of
Schoell, M., Hwang, R.J., Carlson, R.M.K., Welton, J.E., 1994. Carbon isotopic petroleum. Justus Liebigs Annalen der Chemie 510, 42–62.
composition of individual biomarkers in gilsonites. Organic Geochemistry 21, Treibs, A., 1936. Chlorophyll and hemin derivatives in organic mineral substances.
673–683. Angewandte Chemie 49, 682–686.
Sleighter, R.L., 2009. Export of terrestrial dissolved organic matter along a river to Tripp, B.T., White, E.R., 2006. Gilsonite. In: Kogel, J.E., Trivedi, N.C., Barker, J.M.,
ocean transect of the lower Chesapeake Bay investigated by advanced analytical Krukowski, S.T. (Eds.), Industrial Minerals and Rocks: Commodities, Markets,
techniques. In: Chemistry and Biochemistry PhD, Old Dominion University, and Uses. Society for Mining, Metallurgy, and Exploration, Littleton, CO, pp.
Norfolk, VA, p. 241. 481–493.
Sleighter, R.L., Hatcher, P.G., 2007. The application of electrospray ionization Van Berkel, G.J., Quinones, M.A., Quirke, J.M.E., 1993. Geoporphyrin analysis using
coupled to ultrahigh resolution mass spectrometry for the molecular electrospray ionization-mass spectrometry. Energy and Fuels 7, 411–419.
characterization of natural organic matter. Journal of Mass Spectrometry 42, Vuister, G.W., Boelens, R., Kaptein, R., 1991. Gradient-enhanced HMQC and HSQC
559–574. spectroscopy – applications to N-15-labeled MNT repressor. Journal of the
Sleighter, R.L., McKee, G.A., Liu, Z., Hatcher, P.G., 2008. Naturally present fatty acids American Chemical Society 113, 9688–9690.
as internal calibrants for Fourier transform mass spectra of dissolved organic Wen, C.S., Chilingarian, G.V., Yen, T.F., 1978. Properties and structure of bitumens.
matter. Limnology and Oceanography Methods 6, 246–253. In: Chilingarian G.V., Yen, T.F. (Eds.), Developments in Petroleum Science:
Smernik, R.J., Oades, J.M., 2000. The use of spin counting for determining Bitumens, Asphalts, and Tar Sands, vol. 7. Elsevier, Amsterdam.
quantitation in solid state 13C NMR spectra of natural organic matter: 2. HF- Werner-Zwanzinger, U., Lis, G., Mastalerz, M., Schimmelmann, A., 2005. Thermal
treated soil fractions. Geoderma 96, 159–171. maturity of type II kerogen from the New Albany Shale assessed by 13C CP/MAS
Smernik, R.J., Schwark, L., Schmidt, M.W.I., 2006. Assessing the quantitative NMR. Solid State Nuclear Magnetic Resonance 27, 140–148.
reliability of solid-state C-13 NMR spectra of kerogens across a Wilson, M.A., 1987. NMR Techniques and Applications in Geochemistry and Soil
gradient of thermal maturity. Solid State Nuclear Magnetic Resonance 29, Chemistry. Pergamon Press, Oxford, UK.
312–321. Wilson, M.A., Hanna, J.V., Anderson, K.B., Botto, R.E., 1993. 1H CRAMPS NMR derived
Smith, D.F., Rahimi, P., Teclemariam, A., Rodgers, R.P., Marshall, A.G., 2008. hydrogen distributions in various coal macerals. Organic Geochemistry 20,
Characterization of athabasca bitumen heavy vacuum gas oil distillation cuts 985–999.
by negative/positive electrospray ionization and automated liquid injection Wu, X., Zilm, K.W., 1993. Complete spectral editing in CP/MAS NMR. Journal of
field desorption ionization Fourier transform ion cyclotron resonance mass Magnetic Resonance A 102.
spectrometry. Energy and Fuels 22, 3118–3125. Wu, X., Burns, S.T., Zilm, K.W., 1994. Spectral editing in CP/MAS NMR. Generating
Solum, M.S., Pugmire, R.J., Grant, D.M., 1989. C-13 solid-state NMR of Argonne subspectra based on proton multiplicities. Journal of Magnetic Resonance 111,
premium coals. Energy and Fuels 3, 187–193. 29–36.
Stenson, A.C., 2008. Reversed-phase chromatography fractionation tailored to mass Yen, T.F., 2000. The realms and definitions of asphaltenes. In: Yen, T.F., Chilingarian,
spectral characterization of humic substances. Environmental Science and G.V. (Eds.), Developments in Petroleum Science 40B: Asphaltenes and Asphalts,
Technology 42, 2060–2065. vol. 2. Elsevier, Amsterdam, pp. 7–27.
Stenson, A.C., Marshall, A.G., Cooper, W.T., 2003. Exact masses and chemical Yen, T.F., Erdman, J.G., Pollack, S.S., 1961. Investigation of the structure of petroleum
formulas of individual suwannee river fulvic acids from ultrahigh resolution asphaltenes by X-ray diffraction. Analytical Chemistry 33, 1587–1594.

You might also like