SlideShare a Scribd company logo
 
	
  
	
  
	
   	
  
	
  
Department of Chemistry
Thesis submitted for the degree of MSc in
Analytical Science: Methods and Instrumental
Techniques
August 2014
Ultrahigh resolving power tandem mass spectrometry of
petroleum components
By: Simona Gherghel
Supervisor: Dr Mark P. Barrow
	
  
	
  
	
  
	
  
	
  
	
  
	
  
	
  
  	
   	
   	
  
	
  
	
  
Table	
  of	
  Contents	
  
1	
   Introduction	
  ................................................................................................................	
  1	
  
1.1	
   Crude	
  oil	
  ..................................................................................................................................................................	
  1	
  
1.2	
   Naphthenic	
  acid	
  chemistry	
  ..............................................................................................................................	
  2	
  
1.3	
   Formation	
  of	
  naphthenic	
  acid	
  dimers	
  ........................................................................................................	
  5	
  
1.4	
   Analytical	
  techniques	
  for	
  looking	
  at	
  crude	
  oils	
  ......................................................................................	
  6	
  
1.5	
   Fourier	
  transform	
  ion	
  cyclotron	
  resonance	
  mass	
  spectrometry	
  ....................................................	
  7	
  
1.6	
   High	
  mass	
  accuracy	
  .........................................................................................................................................	
  10	
  
1.7	
   Kendrick	
  mass	
  defects:	
  the	
  key	
  to	
  unlocking	
  chemical	
  formulas	
  ................................................	
  10	
  
1.8	
   Double-­‐bond	
  equivalent	
  versus	
  carbon	
  number	
  plot	
  .......................................................................	
  12	
  
1.9	
   Main	
  aims	
  .............................................................................................................................................................	
  13	
  
2	
   Experimental	
  .............................................................................................................	
  14	
  
2.1	
   Sample	
  preparation	
  .........................................................................................................................................	
  14	
  
2.2	
   Fourier	
  Transform-­‐Infrared	
  Spectroscopy	
  (FT-­‐IR)	
  ...........................................................................	
  15	
  
2.3	
   Mass	
  Spectrometry	
  Analysis	
  ........................................................................................................................	
  15	
  
2.4	
   Calibration	
  ...........................................................................................................................................................	
  17	
  
2.5	
   Data	
  analysis	
  .......................................................................................................................................................	
  17	
  
3	
   Results	
  and	
  discussion	
  ...............................................................................................	
  19	
  
3.1	
   Proof	
  of	
  naphthenic	
  acid	
  dimers	
  existence	
  in	
  solution-­‐phase	
  ......................................................	
  19	
  
3.2	
   2	
  year	
  old	
  Kodak	
  naphthenic	
  acid	
  sample	
  .............................................................................................	
  20	
  
3.2.1	
   ISD	
  experiment	
  0	
  V	
  ...........................................................................................................................................	
  20	
  
3.2.2	
   ISD	
  experiment	
  (60	
  V)	
  ....................................................................................................................................	
  25	
  
3.3	
   Fresh	
  sample	
  ......................................................................................................................................................	
  29	
  
3.3.1	
   ISD	
  (0	
  V)	
  of	
  serial	
  dilution	
  ............................................................................................................................	
  29	
  
3.3.2	
   Importance	
  of	
  ion	
  accumulation	
  time	
  (IAT)	
  .........................................................................................	
  35	
  
3.4	
   CID	
  on	
  the	
  2	
  year	
  old	
  Kodak	
  NA	
  sample	
  .................................................................................................	
  37	
  
3.5	
   IRMPD	
  experiment	
  of	
  the	
  fresh	
  Kodak	
  NA	
  sample	
  ............................................................................	
  40	
  
3.6	
   Fresh	
  Kodak	
  NA	
  sample	
  doped	
  with	
  salts	
  .............................................................................................	
  44	
  
3.7	
   EID	
  experiment	
  on	
  a	
  NIST	
  crude	
  oil	
  sample	
  .........................................................................................	
  47	
  
4	
   Conclusion	
  ................................................................................................................	
  52	
  
5	
   References	
  ................................................................................................................	
  54	
  
6	
   Acknowledgments	
  ....................................................................................................	
  58	
  
7	
   Appendix	
  ..................................................................................................................	
  59	
  
7.1	
   Known	
  ions	
  mass	
  list	
  ......................................................................................................................................	
  59	
  
7.2	
   Mass	
  error	
  distribution	
  histograms	
  .........................................................................................................	
  65	
  
7.2.1	
   2	
  year	
  old	
  doped	
  Kodak	
  NA	
  sample-­‐ISD	
  0	
  V	
  .........................................................................................	
  65	
  
7.2.2	
   2	
  year	
  old	
  doped	
  Kodak	
  NA	
  sample-­‐ISD	
  60	
  V	
  .......................................................................................	
  66	
  
7.2.3	
   Fresh	
  sample:	
  serial	
  dilution	
  (ISD	
  0	
  V	
  .....................................................................................................	
  67	
  
7.2.4	
   CID	
  on	
  2	
  year	
  old	
  Kodak	
  NA	
  .........................................................................................................................	
  71	
  
7.2.5	
   IRMPD	
  on	
  fresh	
  Kodak	
  NA	
  sample	
  ............................................................................................................	
  73	
  
	
  
	
  
	
  
  	
   	
   	
  
Abbreviations	
  
	
  
APCI Atmospheric Pressure Chemical Ionization
CID Collision Induced Dissociation
DBE Double Bond Equivalent
EI Electron Ionization
EID Electron Induced Dissociation
ESI- Electrospray ionization
FAB Fast Atom Bombardment
FT-ICR MS Fourier Transform Ion Cyclotron Resonance Mass Spectrometry
FT-IR Fourier Transform- Infrared Spectroscopy
GC-MS Gas Chromatography Mass Spectrometry
IAT Ion Accumulation Time
ICR Ion Cyclotron Resonance
IRMPD Infrared Multiphoton Dissociation
ISD In Source Dissociation
KMD Kendrick Mass Defect
MS/MS Tandem mass spectrometry
NA Naphthenic Acid
SARA Saturate, aromatic, resin and asphaltene
SORI Sustained Off-Resonance Irradiation
	
  
	
  
  	
   	
   	
  
Summary	
  
	
  
	
  
The selective ionization of acidic compounds from a commercial naphthenic acid mixture and
of nitrogen-containing compounds from a crude oil sample was carried out using Fourier
transform ion cyclotron resonance mass spectrometry (FT-ICR MS) in negative-ion and
positive-ion electrospray ionization (ESI), respectively. Naphthenic acids are of great concern
to the petroleum industry as they are responsible for various unwanted phenomena such as
pipeline corrosion, formation of sodium naphthenate deposits and they also are toxic to
aquatic organisms. Naphthenic acids have been previously demonstrated to form O4 class
naphthenic acid dimers. During this research, the formation of sodium bound naphthenic acids
(NaO4 class) has been observed. The ultrahigh resolving power and mass accuracy of FT-ICR
MS offers accurate mass assignments and information on compositional differences of
naphthenic acid species. Several tandem MS methods such as collision-induced dissociation
(CID) and infrared multiphoton dissociation (IRMPD) were further employed for structural
characterization of the naphthenic acid dimers. CID and IRMPD of the same precursor ions
produced very similar fragmentation patterns it has been shown that dimers reflect the most
abundant monomer species present. Electron-induced dissociation (EID) was used for
structural characterization of crude oil and it provided valuable fragmentation information
showing successive losses of alkyl chains leading back to the stable cores made of fused
rings.
	
  
Introduction	
  
1	
  
	
  
1 Introduction	
  
1.1 Crude	
  oil	
  	
  
Crude oil is a fundamental energy source that has been playing a key role in the progress of
modern human society. Even under very optimistic estimates about the development of
alternative energy sources, it will remain one of the most heavily used energy sources for
many decades to come.1
Global dependence on crude oil originates from its high versatility; it
can be used as fuel for vehicles and airplanes, as a home heating source and almost all
chemical products, such as plastics, pesticides, detergents, dyes and even medicines begin
with oil-derived feedstock.2
Nevertheless crude oil is a limited resource and while its production is decreasing, its
consumption only keeps increasing. Therefore, lately there has been an increased focus on the
use of heavier, lower quality sources.3
Heavy oils are one of the most complex mixtures in
nature and they can contain thousands to millions of various components.4
Typically crude
oils have a high proportion of light hydrocarbons (CcHh) and other minor ingredients. The
hydrocarbons can be divided into four types: paraffin based (15 to 60%), naphthenes based
(30 to 60%), aromatic based (3 to 30%) and asphaltic based (remainder). Carbon accounts for
about 85% (83-87%) of the crude oil mass, hydrogen for about 12% (10-14%) and the H/C
ratio is about 1.8.5
The minor ingredients of crude oil include nitrogen (0.1-2%), oxygen (0.05-1.5%), sulfur
(0.005-6%) and traces of metal (0.1%) such as are sodium (Na), calcium (Ca), magnesium
(Mg), aluminium (Al), iron (Fe), vanadium (V), and nickel (Ni). Even though petroleum
contains more than 90% hydrocarbons, it is the heteroatom-containing compounds (NnOoSs)
the ones that cause most problems regarding pollution, poisoning of catalysts, corrosion and
formation of emulsions.6
Low quality crude oils contain higher amounts of sulfur, acidic
components and other hydrocarbons. Sulfur-containing compounds naturally occurring in
crude oils act as poisons to the catalysts used in the conversion of feedstock to useable
intermediate and end products.7
Sweet and sour are terms used to refer to a crude oil’s
approximate sulfur content. As a rule, when the sulfur content is in excess of 0.5% the crude
oil is considered sour, and when it is 0.5% or less it is considered sweet.8
Nitrogen-containing
Introduction	
  
2	
  
	
  
compounds also represent a concern as they too can play a role in the catalyst deactivation.9
Oxygen-containing components can include phenols, ketones and carboxylic acids. The
presence of acidic substances in crude oil was first observed in a Romanian oil in 1874 by
Hell and Medinger10
. Later on, in 1890 Aschan11
named them naphthenic acids. Nowadays, it
is well known that naphthenic acids cause corrosion in pipes and equipment used in the
processing plants and that they are toxic to aquatic environments.12
1.2 Naphthenic	
  acid	
  chemistry	
  
Naphthenic acids (NAs) comprise of a large group of saturated monocyclic, acyclic and
aromatic carboxylic acids. They are naturally occurring compounds in most petroleum
sources, including crude oil and bitumen from the oil sands.13
The classical formula for NAs
is CnH2n+ZO2, where n indicates the carbon number and Z is referred to as the “hydrogen
deficiency” and is zero or a negative even integer number. More negative Z values represent
an increase in the number of hydrogen atoms lost as the structure gets more compact and as
rings are formed14
. The Z value is equal to 0 for saturated linear hydrocarbon chains and it
becomes -2 for monocyclic NAs, -4 for bicyclic, -6 for tricyclic and so forth.15
Sample
structures of naphthenic acids are shown in Figure 1.
	
  
Figure 1. Representative structures of naphthenic acids where Z is the hydrogen deficiency, R
represents an alkyl chain and n indicates the number of CH2 groups
Introduction	
  
3	
  
	
  
The chemical, physical and toxicological characteristics of different NA mixtures depend on
their differences in molecular structure, composition, volatility and polarity. NAs dissolve in
organic solvents and their pH influence their water solubility. Generally, their pKa values are
between 5 and 6.15
Because NAs are water soluble to some degree, their release to
wastewaters must be monitored.16
Naphthenic acid make up to 4 wt% of the crude oil composition17
and their presence in crude
oils is of great concern to the petroleum industry. NAs are responsible for corrosion in
refineries, pollution in refinery wastewaters and in oil sands extraction waters, formation of
calcium and sodium naphthenate deposits during production and processing, and formation of
emulsions.18
Naphthenate deposition can obstruct pipelines (Figure 2), causing production
irregularities and in some cases production shutdowns that can be very expensive for the oil
companies. Formation of emulsions and naphthenate deposits results in millions of dollars
expenses for the petroleum industry.
Figure 2. Naphthenate deposit causing obstruction of a hydrocyclone19
Introduction	
  
4	
  
	
  
Naphthenic acids are also responsible for the corrosion of pipeline carbon steel alloys that are
otherwise resistant to corrosion from compounds containing sulfides.20
An example of
pipeline affected by NA corrosion is shown in Figure 3. Corrosion by naphthenic acid is not
fully understood but it is known that it involves the chelation of the metal ion by the
carboxylate group with the release of hydrogen gas.21
Qu and his colleagues showed that
corrosivity of NAs depends greatly on the molecular mass and structure; NAs with higher
averaged molecular weight and higher averaged number of ring structures were less
corrosive.22
Temperatures between 220 and 400° C facilitate corrosion whereas at
temperatures above 400 °C the NAs	
  decompose forming a film that protects the alloy.23
Figure 3. Naphthenic acid corrosion in crude oil unit piping 24
With the increase for oil mining, transportation and application, more and more soils are
susceptible to crude oil contamination. At the same time, more soils contaminated with crude
oil enter the aquatic environment through surface runoff.25
Naphthenic acids are
environmentally significant because they are known to be toxic to a variety of aquatic
organisms, algae and mammals by acting as endocrine disruptors.26
The extraction processes
of bitumen, an extra heavy oil subgroup, from oil sands produce tailing waters that contain
naphthenic acids. As a consequence, oil sands companies need to monitor and report the
concentration of NAs in the waters on and near their leases.27
Introduction	
  
5	
  
	
  
Nonetheless, NAs also represent a resource. They and their metal carboxylates are used as
dryers in dyes, as rubber plasticizer and as fungicides for wood preservatives.28
By separating
naphthenic acids and other acidic species present in crude oils, the quality of the crude oil will
improve, and on the other hand the NAs can be used as feedstock.
1.3 Formation	
  of	
  naphthenic	
  acid	
  dimers	
  
	
  
Naphthenic acids are both hydrogen-bond acceptors because of the –C=O carbonyl group and
hydrogen bond donors because of the –OH hydroxyl group; therefore they can participate in
dimer formation by hydrogen bonding. Smith and his colleagues29
used Fourier	
  transform	
  
ion	
  cyclotron	
  resonance	
  mass	
  spectrometry	
  (FT-­‐ICR	
   MS) to show that naphthenic acids
self-associate in the gas-phase at high enough concentration (1 to 10 mg/mL), forming
naphthenic acids dimers (general formula CnH2n+zO4). Moreover, they were able to
demonstrate that the higher the concentration of naphthenic acids, the higher their tendency to
form aggregates.
Although dimer formation is an accepted phenomenon in the world of petroleomics, there
have not been many studies focused on this topic. Da Campo et al.30
have taken a further look
at CnH2n+zO4 aggregates. Fourier transform infrared spectroscopy of these dimers revealed
that they are also present in the solution-phase. Using an FT-ICR MS instrument, they
demonstrated that these compounds are bound by noncovalent bonds and they reported a
possible general structure as the one shown in Figure 4, where a proton is shared between two
oxygen atoms. The formation of these species is dependent on the accumulation time of the
ions in the hexapole of the ion source. It was noticed that by increasing the accumulation
time, and thus increasing the number of collisions taking place, there was a decrease in the
signal intensity of these species, which was related to the weak noncovalent bonds that hold
together the dimers.
Introduction	
  
6	
  
	
  
Figure 4. Possible structure for the singly charged noncovalent naphthenic acids dimers30
Mapolelo et al31
studied calcium naphthenate deposits and sodium naphthenate emulsion
sample using FT-ICR MS, only after preparation in the laboratory to first remove the metals
from the naphthenic acids and thus creating free acids.
Naphthenic acids dimers will affect the behaviour of the naphthenic acids in the solution-
phase and so the problems associated with them. In this paper, a better understanding of
naphthenic acid dimers was pursued through the use of ultrahigh resolving power mass
spectrometry.
1.4 Analytical	
  techniques	
  for	
  looking	
  at	
  crude	
  oils	
  
	
  
Petroleomics is referred to as the detailed study for elucidating the chemical composition of
the components naturally occurring in petroleum and crude oil samples. High resolving power
mass spectrometry, especially FT-ICR MS is nowadays the central petroleomics-grade
technique.
How effective the oil recovery methods are depends on the composition of the oil of interest.
Analysis of the composition of crude oil can be highly complex and separation of the major
crude oil components can facilitate their characterization. Saturate, aromatic, resin and
asphaltene (SARA) analysis is a separation technique that has been widely used in
petroleomics. It involves the separation of crude oil components based on their solubility and
polarity into four main classes: saturate, aromatic, resin and asphaltene fractions. The first
step is the precipitation of asphaltenes in n-alkane solvents, followed by liquid
Introduction	
  
7	
  
	
  
chromatography using silica or alumina columns to separate the three remaining fractions by
their polarity.32
Elucidating the chemical structure of heavy components in crude oil can allow a better
understanding and prediction of the petroleum behaviour during processing. Because the
organic composition of heavy crude oil is so complex, its characterization has been in the past
limited to bulk properties (such as viscosity, density, electric conductivity, light scattering,
UV-visible and infrared spectroscopy, 13
C nuclear magnetic resonance NMR, X-ray
diffraction)33
and various chemical separation methods based on solubility (like SARA
analysis), boiling point, gas and liquid chromatography.34
NMR and X-ray diffraction have
been employed in determining the molecular structure of crude oil compounds, but both
techniques are limited in the sense that they can only provide an average molecular structure
of components. 35
At present, Fourier transform ion cyclotron resonance mass spectrometry is an indispensable
method of choice for the analysis of complex samples such as petroleum. FT-ICR MS
dominates the petroleomics field because of its high mass accuracy and ultrahigh resolving
power, making it possible to calculate the elemental composition of the compounds with great
accuracy.
1.5 Fourier	
  transform	
  ion	
  cyclotron	
  resonance	
  mass	
  spectrometry	
  
All types of mass spectrometer allow compounds to be identified by the production of gas-
phase ions from a neutral sample and their subsequent sorting and detection based on their
mass-to-charge ratio (m/z). The sample can be solid, liquid or gas, depending upon the type of
ion source used. The ions are transferred into the mass analyser (ICR cell) and the differences
in masses of the fragments enable the mass analyser to separate the ions according to their
mass to charge (m/z) ratio.36
FT-ICR MS instruments determine the m/z of an ion by measuring the cyclotron frequency of
the ion trapped in a fixed magnetic field. The magnetic field is usually generated using a
superconducting magnet. The equation that relates frequency to the m/z of the ions is the
following:
Introduction	
  
8	
  
	
  
𝑓 =
!"
!!"
(eq. 1),
where f is the cyclotron frequency (Hz), m is the mass of the ion (kg), q is the charge on the
charged particle (coulombs, C) and B is the magnetic field (tesla, T).37
Due to the inherently ultrahigh resolving power (for example 400 000 (full width at half
maximum) at m/z 400) and accurate mass measurement, Fourier transform ion cyclotron
resonance mass spectrometry has enabled analysts to study heavy crude oil at a molecular
level.38
FT-ICR MS has been playing a key role for petroleomics ever since its first
application to petroleum distillates39
as it has allowed a better understanding of the structure
of the components that make up this highly complex mixture and thus an insight into their
behaviour.
Most of the characterization of naphthenic acids has been performed using different mass
spectrometric (MS) techniques, including gas chromatography mass spectrometry (GC-MS)40
,
two-dimensional GC-MS41
and liquid chromatography (LC) MS.27
Also, different ionization
approaches such as electron ionization (EI)42
, fast atom bombardment (FAB)43
, atmospheric
pressure chemical ionization (APCI)44
, and lately electrospray ionization (ESI)17
have been
successfully used for the study of naphthenic acids.
ESI is a soft ionization technique that has gained great popularity in the past few years and it
has been shown that is it able to selectively ionize acidic and basic components in petroleum
samples.45
The analyte dissolved in a suited solvent is sprayed from a small capillary tube into
a strong electric field in the presence of a drying gas (typically nitrogen) and thus, generating
a fine “mist” of highly charged droplets as illustrated in Figure 5.46
Under negative-ion ESI
acids are deprotonated generating negative ions and under positive mode ESI (basic) neutral
are protonated forming positive ions.34
Although crude oil contains less than 10% heteroatom-
containing compounds, they represent the main concern for the petroleum industry. Because
of their high polarity, ESI is specific and especially efficient in generating their gas-phase
ions. Moreover, negative-ion ESI is a “soft” ionization techniques that generates [M − H]-
without extensive ion fragmentation or matrix interference.
Introduction	
  
9	
  
	
  
Figure 5. Schematic representation of ion formation by electrospray ionization (ESI)
Tandem mass spectrometry (MS/MS) techniques allow insight on structure of naphthenic
acids by offering further information about specific ions. The ions of interest from the ion
source are selected and then subjected to fragmentation through different dissociation
techniques such as collision induced dissociation, infrared multiphoton dissociation and
electron induced dissociation.
Collision-induced dissociation (CID), also known as collisionally activated dissociation
(CAD), is by far the most common MS/MS technique and it involves the acceleration of the
ions by electrical potential to high kinetic energy followed by the collision of the ions with
neutral gas molecules, resulting in further dissociation of the ions.
Central to infrared multiphoton dissociation (IRMPD), is the absorption of multiple infrared
photons, leading to a continuous build-up of internal energy by the ions. When the internal
energy overcomes the barrier against dissociation, fragmentation of the ions occurs.47
The
irradiation of ions is typically carried out using a pulsed CO2 laser.
Electron induced dissociation (EID) involves the fragmentation of singly-charged ions
succeeding their interaction with high-energy electrons (>10 eV).48
EID has been used in the
past for the analysis of peptide ions, metabolites and sodium clusters cations49
but as to date,
there are no publications testing the viability of EID for petroleum samples. EID would allow
Introduction	
  
10	
  
	
  
going beyond the analysis of petroleum using elemental composition alone by obtaining
greater structural insight.
1.6 High	
  mass	
  accuracy	
  
	
  
In mass spectrometry, mass accuracy is defined as the ratio of the m/z measurement error to
the true m/z, generally expressed in parts per million (ppm) and mass resolving power
represents the ability to distinguish between two peaks varying slightly in m/z.50
Each isotope of each chemical element has a different mass defect and they are known to a
great accuracy. By using the power of the inherent high mass accuracy and ultrahigh
resolving power of Fourier transform mass spectrometry, it is possible to compare petroleum
samples on a molecular level with high accuracy. The importance of ultrahigh resolving
power can be illustrated in the case of CH4 and O as they both have a nominal mass of 16, but
their exact mass is 16.0313 Da and 15.9949 Da respectively, resulting in a mass difference of
36.4 mDa. Similarly SH4 and C3 have a nominal mass of 36, but their exact masses (36.0034
Da and 36.0000 Da, respectively) differ by 3.4m Da.4
But even with a high mass accuracy and knowledge of the ion’s charge state, the chemical
complexity of naphthenic acids increases greatly with molecular weight and the assignment of
the elemental composition can be unambiguously done only to about 500 Da.51
Nevertheless,
by using Kendrick mass defect plot the mass range can be extended up to three times.52
	
  
1.7 Kendrick	
  mass	
  defects:	
  the	
  key	
  to	
  unlocking	
  chemical	
  formulas	
  
Kendrick mass defect (KMD) plots have been successfully applied for the study of petroleum
samples using ultrahigh resolving power mass spectrometry as they have allowed data
interpretation and visualization of complex data sets. KMD plots are particularly useful for
indicating the characteristics and behaviour of the species of interest.53
Kendrick mass defect calculation transforms the IUPAC mass scale (12
C=12.000 00 Da) to
the Kendrick mass scale (CH2=14.00000 Da instead of 14.01565 Da) using the following
equation:
Introduction	
  
11	
  
	
  
Kendrick Mass=IUPAC Mass×
!"
!".!"#$#
(eq. 2).
Then the Kendrick mass defect is defined by:
Kendrick Mass Defect (KMD) = Nominal Kendrick Mass-Kendrick Mass (eq. 3),
where Kendrick mass is calculated as shown above and nominal Kendrick Mass is the
Kendrick mass rounded to the nearest integer.
For samples containing a high number of hydrocarbons, such as petroleum, KMD plots
simplify the display by showing peak patterns lined up in the horizontal direction. The
effectiveness of KMD plots relies on the fact that members of a homologous series from
crude oils, i.e. compounds that have the same constitution of heteroatoms and number of rings
plus double bonds but different alkyl chain length, will have an identical Kendrick mass
defect. In a KMD plot, a homologous series will appear on a horizontal row, making it easy to
distinguish from species of other class and type. Another great advantage is that “outlier” data
can be immediately determined if they fall outside the main patterns54
.
Furthermore, unambiguous assignment of a series of related compounds at low m/z serves to
determine all other higher mass members of that species. As a result, KMD scaling extends
elemental composition assignment to masses up to three times higher than would be possible
using mass-measurement accuracy only.
For illustrative purposes, a KMD plot of naphthenic acids is shown in Figure 6. The primary
trend horizontally is a constant difference in Kendrick nominal mass of 14.01565 (indicating
an increase with a CH2 group) and vertically there is a constant Kendrick mass defect
difference of 0.013455
(indicating a difference of two hydrogen atoms and thus an increase in
the degree of unsaturation).
Introduction	
  
12	
  
	
  
Figure 6. Kendrick Mass Defect plot of monomeric naphthenic acids
1.8 Double-­‐bond	
  equivalent	
  versus	
  carbon	
  number	
  plot	
  
	
  
The first step to evaluate the structure of a compound is to determine its elemental
composition. The exact mass measured with MS can be used to calculate the elemental
formulae (CcHhNnOoSs). Once the elemental composition is known, double-bond equivalent
(DBE), also known as degree of unsaturation, plays a key role in predicting the chemical
structure of the compounds present in crude oil samples. DBE represents the number of
double bonds and rings in a given molecule. DBE and Z values are different ways of thinking
about the same thing, i.e. the hydrogen deficiency (rings and double bonds and degree of
unsaturation).
The DBE values calculated from the elemental composition can be plotted against carbon
number, resulting in a “DBE plot”. Species of the same DBE, i.e. from the same homologous
series, will be plotted on a horizontal line. The DBE plots may be rendered more informative
Introduction	
  
13	
  
	
  
by colour-coding each m/z ion depending on its measured relative abundance. Herein, the
gradient goes from red for the least abundant ions to yellow for the most abundant ions.
Additionally, the size of each plot dot can differ with the abundance of that ion. Such a plot of
naphthenic acids is shown in Figure 7.
Figure 7. Example of DBE plot for O2 class species, where the most abundant species have a
DBE value of 3.5	
  
1.9 Main	
  aims	
  
The aim of the following study is to use ultrahigh resolving power FT-ICR MS and tandem
MS techniques for characterization of petroleum components.
The initial focus is on aggregation of naphthenic acids in the solution phase: understanding
their composition and whether these aggregates form selectively or depending on the
available monomers. Then electron-induced ionization will be applied for the first time to
petroleum in order to obtain preliminary data for proof-of-concept of a new approach for
structural insights into petroleum. The interaction of singly-charged ions with high-energy
electrons (>10 eV) is expected to result in mass spectra with more extended ion fragmentation
than slow-heating dissociation techniques such as CID or IRMPD.48
Experimental	
  
14	
  
	
  
2 Experimental	
  
2.1 Sample	
  preparation	
  
	
  
A commercial Kodak naphthenic acid mixture (The Eastman Kodak Company, Rochester,
NY) was used for all the analysis carried out for the characterization of naphthenic acids
using negative-ion ESI.
The water used in the experiments was purified using a Direct-Q 3 Ultrapure Water System
from Millipore (Billerica, MA, USA). An already prepared two years old sample containing 4
mg of Kodak NA mixture dissolved in 5 mL of acetonitrile (WVR, Leuven, Belgium) and 5
mL of Milli-Q water was now doped with 50µL Agilent #1969 ESI-L Low Concentration HP
calibrating mix (Agilent, Palo Alto, CA). The concentration of the sample is 0.4 mg/mL and
is expressed in mass/volume instead of molarity as naphthenic acid is a mixture. This sample
was then subjected to in-source dissociation (ISD) at 0 and 60 V and to CID fragmentation at
2, 6, 8 and 12 V. ISD is a variation of CID, where an increase in the difference in potential
between the two ion funnels gives ions higher kinetic energy at an early stage of the
instrument. ISD is a non-selective technique useful for dissociation of noncovalent species.
To compare the results obtained from the 2 year old Kodak NA sample, a fresh solution of
Kodak NA sample was prepared. This fresh solution containing 19.9 mg of Kodak NA
mixture was dissolved in 5 mL of acetonitrile and 5 mL of Milli-Q water (~2 mg/mL). Using
this stock solution a serial dilution with the following concentrations: 0.5, 0.1, 0.05, 0.001,
0.005 and 0.001 mg/mL was created and subjected to ISD (0 V) experiments. In the previous
CID experiments, the quadrupole was used to isolate the m/z 503.4 ion (with a 1 Da wide
isolation window). The ions in the isolation window pass though the collision cell, where CID
experiments were performed and then the ions were passed to the ICR cell. To improve the
quality of the isolation and the resolving power of the resulting spectra, the quadrupole was
used to isolate the species of interest then a correlated sweep (in the cell) plus correlated shots
(1% clean up shots power, pulse 0.15846 sec) in the cell was used. The fragmentation was
carried out using IRMPD rather than SORI-CID as the later will raise the pressure in the cell
and decrease performance.	
  
Experimental	
  
15	
  
	
  
Doping of a fresh Kodak NA sample with sodium chloride and analysing the effects salt has
on NAs was the next step. A fresh 0.05 mg/mL stock solution Kodak NA mixture was
prepared in 10 mL of acetonitrile and 10 mL of Milli-Q water. A 600µM stock solution of
sodium chloride was prepared in 20 mL of Milli-Q water (0.7 mg NaCl). Three samples
containing the same naphthenic acid mass concentration (0.05 mg/mL) and varying salt
concentration values were prepared. The concentration of salt in the three samples was as
follows: 0.6 µM, 54.55 µM and 300 µM, respectively. The blank stock solution of Kodak NA
sample, as well as the three samples of Kodak NA doped with salt was subjected to MS
analysis. A 0.05 mg/mL solution of Kodak NA was also doped with significant higher
amounts of NaCl to mirror the seawater salinity, i.e. approximately 600 mM, but it was not
run thought the FT-ICR MS.
A light-sour crude oil sample, which is a standard reference material from the National
Institute of Standards and Technology (NIST, Gaithersburg, Maryland, U.S.A.) was used to
obtain a positive-ion ESI spectrum. Then isolation was performed using a window with the
width of 60 Da so that the isolation window spans ions of two higher/lower carbon numbers
each side of the central ion. EID experiments were carried out and a single EID spectrum was
obtained. Because of the difficulty faced by tuning with low signal, the spectrum collected
was the average of 500 scans and it took about an hour to acquire.
2.2 Fourier	
  Transform-­‐Infrared	
  Spectroscopy	
  (FT-­‐IR)	
  
	
  
FT-IR analysis was carried out on a JASCO FT/IR-470 plus. The Kodak NA mixture was
analysed directly, with no prior dilution in solvents. The baseline correction was produced in
a nitrogen flow atmosphere, followed by the acquisition of the sample FT-IR spectrum as an
average of 200 scans.
2.3 Mass	
  Spectrometry	
  Analysis	
  
All mass spectrometry experiments were carried out on a Bruker solariX Fourier transform
ion cyclotron resonance mass spectrometer fitted with an Apollo II ion source and a 12 tesla
superconducting, refrigerated and ultrashield magnet (Bruker Daltonik GmbH, Bremen). An
instrument schematic of the instrument used is shown Figure 8.
Experimental	
  
16	
  
	
  
Figure 8. Schematic of 12 T Bruker Solarix FT-ICR MS (courtesy of Bruker Daltonik,
GmbH, Bremen)
For the Kodak NA samples, the instrument was run in negative-ion conditions as negative
ESI-MS favours the ionization and detection of acidic compounds such as naphthenic acids.
These acidic sites give up a proton and thus creating negatively charged molecules. For the
crude oil sample, the instrument was run in positive-ion conditions due to the high abundance
of basic nitrogen species, which can be readily protonated and thus, providing a strong signal
for MS/MS experiments.
The conditions of ESI were the following: syringe flow of 120-300µL/hour, drying gas
(nitrogen) temperature of 220° C and a flow rate of 4L/min. The nebulizer gas (also nitrogen)
was kept at a pressure of 1.2 bar. The ion accumulation time (IAT) was tuned while
monitoring the transient lifetime. For all the CID experiments the IAT was set to 0.5 sec, for
IRMPD experiments it was 2 sec and for the ISD experiments of the old and fresh samples it
was 0.001s. A set of ISD experiments where the IAT varied between 0.1 and 0.4 was carried
out in order to obtain an understanding how the spectrum is affected by accumulating ions for
longer period of times.
The mass range was set between m/z 147.41 and 3000. Mass spectra consisted of 4M Word
data points the signal-to-noise ratio was enhanced by summing 100 time domain transients.
Transient length was approximately 1.67 s with a resolving power (m/Δm50%) of about 400
Experimental	
  
17	
  
	
  
000 at m/z 400 (Δm50% is the mass spectral full peak width at half height). After acquisition,
data were zero-filled56
once and sine-bell apodization57
was applied.
2.4 Calibration	
  
	
  
Mass lists were obtained using the processing software Data Analysis 4.2 (Bruker Daltonik
GmbH, Bremen) where the S/N threshold was set to be higher than 4. Calibration was carried
out externally for the old Kodak NA sample doped with Agilent #G1969 Low Concentration
HP calibrating mix. In this way, accurate m/z values were obtained over a wide mass range
that bracketed the monomer and dimer region of interest and thus improving the mass
accuracy and confidence in assignments.
For subsequent spectra, the calibrating material was not used in order to eliminate possible
peak contributions. The calibration was carried out internally using the O2, O4 and NaO4
homologous peaks (the list of known mass ion used for calibration is provided in the
Appendix).
2.5 Data	
  analysis	
  
	
  
Spectral interpretation was carried out using the Data Analysis 4.2 software. For the
calculation of the elemental formulas, the usual CcHhNnOoSs formula for petroleum was
considered, where c and h are unlimited, n and o are between 0 and 5 and s is between 0 and
4. Up to two 13
C atoms were also included in the calculation because of the high content of
carbon, along with 1 atom of Na. Other constraints included even electron species, maximum
H/C ratio of 3 and a minimum of -0.5 and a maximum of 40 rings plus double bonds.
Several approaches for data analysis and visualization of the large amount of data sets
acquired were used in this paper, including Kendrick mass defect plots, double bond
equivalent plots, bar charts of contributions from different type of species and distribution of
the intensity of various Z series versus carbon number. Data were processed with Excel 14
(Microsoft Corporation, Redmond, Washington, U.S.A.), Origin 9.1 (OriginLab,
Northampton, MA, USA) and GraphPad Prism 6 (GraphPad Software, LA Jolla California,
USA).
Experimental	
  
18	
  
	
  
The mass spectra gathered varied from 200 to 300 entries in the peak list for the CID and
IRMPD experiments and between 1000-3000 data points for the ISD experiments. The m/z,
intensity and resolving power values were the parameters imported from Data Analysis
software into Microsoft Excel.
Using eq. 2 and eq. 3, a method was created for the calculation of the Kendrick mass defects.
An example of the calculation is shown next.
For an ion with the composition C15H25O2
-
(the negatively singly charged ion of the C15H26O2
naphthenic acid with Z=-4) the IUPAC mass is 237.186004, but using the Kendrick mass
scale, the ion has a Kendrick mass of 237.186004*14/14.01565=236.921160. The Nominal
Kendrick mass is 237 and the Kendrick mass defect is 237-236.92116=0.07884.
Each data entry was manually assigned on the basis of Kendrick mass defects and of m/z
values within an error range typically of 1 ppm or less. Then each formula was used to
calculate the Z and DBE values, followed by various visualization methods. For all mass
spectra, mass error distribution histograms of the main class species present were created and
they can be found in the Appendix. The low mass errors obtained using FT-ICR MS allow
more confident assignments than by using other type of mass analyser.
Double bond equivalent was calculated using the following equation58
:
DBE= 𝑐 −
!
!
+
!
!
+ 1, for the elemental formulae CcHhNnOoSsNa1.
To ensure persistence in data analysis, all DBE values were calculated for the molecular ions.
So, for example, for the molecular ion C15H29O2
-
, DBE=15-29/2+1=1.5 (corresponding to a
saturated naphthenic acid compound, Z=0). Equivalently, all deprotonated ions of monomeric
naphthenic acids from the ESI experiments will have an odd DBE value.
For convenience purposes, assignments were categorized in terms of their carbon number,
number of double bonds and rings, and their constituent heteroatoms (O, N, S, Na). The
heteroatom(s) determines the compound class; for example, O or O1 denotes a composition of
CcHhO1 and O2 denotes a composition of CcHhNnO2.
Results	
  and	
  discussion	
  
19	
  
	
  
3 Results	
  and	
  discussion	
  
3.1 Proof	
  of	
  naphthenic	
  acid	
  dimers	
  existence	
  in	
  solution-­‐phase	
  
	
  
Carboxylic acids are well known to have a characteristic twin-peak infrared absorption band,
with a small peak between 1740 and 1750 cm-1
(monomer form) and a strong peak between
1700 and 1715 cm-1
(dimer form). The pure, undiluted in organic solvents, Kodak naphthenic
acid mixture was analysed using a Fourier Transform Infrared Spectroscopy. Its FT-IR
spectrum is shown in Figure 9 and it displays this pattern with an apex at 1703, associated
with the presence of NA dimers and a smaller apex at 1742 cm-1
, associated with the
monomeric NA structures. This readily proves that the components from the Kodak mixture
associate to form dimers in the solution-phase at a sufficiently high concentration.
Figure 9. Expanded region of the FT-IR spectrum of the Kodak NA mixture showing the
absorption band characteristic to the C=O stretching in the carbonyl group. The maxima at
1703 cm-1
corresponds to dimeric species and the maxima at 1742 to the monomeric species.
Figure based on data provided from previous study20
using the same Kodak NA mixture as in
this work
160016501700175018001850
0.0
0.1
0.2
0.3
0.4
0.5
Absorption(a.u.)
Wavenumber (cm-1
)
1742
(monomer)
1703
(dimer)
C=O stretch
(COOH)
Results	
  and	
  discussion	
  
20	
  
	
  
3.2 2	
  year	
  old	
  Kodak	
  naphthenic	
  acid	
  sample	
  
3.2.1 ISD	
  experiment	
  0	
  V	
  	
  
	
  
During an ISD experiment (0 V), using the 2 year old Kodak naphthenic acid sample doped
with Agilent #G1969 ESI-L Low Concentration HP calibrating mix, it was observed that
besides the monomer distribution in the m/z 170-390 region (attributed to the naphthenic acid
species, i.e. O2 class species), there are three other distinct distributions present in the m/z
365-675, 670-920 and 950-1200 m/z region along with clusters in the m/z 1080-1230, 1370-
1520 and 1700-1850 regions (Figure 10).
Figure 10. Mass spectrum of a 2 year old Kodak NA sample showing several distributions
(negative-ion ESI, ISD 0 V). The peaks marked with a star originate from the Agilent #G1969
ESI-L Low Concentration HP calibrating mix
For convenience purposes, the four distributions present were named using roman numerals
as shown in Table 1.
Results	
  and	
  discussion	
  
21	
  
	
  
Table 1. Assigned names for the four distributions in the mass spectrum of a 2 year old Kodak
NA sample (negative-ion ESI, ISD 0 V)
m/z	
  region	
   Distribution	
  name	
  
170-­‐390	
   I	
  
365-­‐675	
   II	
  
670-­‐920	
   III	
  
950-­‐1200	
   IV	
  
A very effective visual method to illustrate features and behaviour of the species of interest is
the use of Kendrick plots. The m/z for all the peaks in the spectrum were rescaled to the
Kendrick Mass scale and the nominal Kendrick Mass was plotted against KMD to reveal the
four distributions present (Figure 11). Even though two-dimensional KMD plots do not
typically show the relative intensity of the compounds, they are very useful in illustrating the
degree of unsaturation. It can be clearly noticed that distribution II has the highest Kendrick
mass defect variation and thus its components extend over a wider range of degrees of
unsaturation when compared to the other distribution present such as distribution I, i.e. the O2
class species. Also an overlap of species can be observed in distribution II.
Figure 11. Kendrick Mass plot of the four distributions present in the 2 year old Kodak NA
sample (negative-ion ESI, ISD 0V)
Results	
  and	
  discussion	
  
22	
  
	
  
Carrying out data analysis, the majority of the peaks in distribution II were assigned to O4 and
NaO4 class species. Distribution III and IV have tentative assignments of NaOS and NaO3,
respectively. Within acceptable mass errors, the compositions were not naphthenic acid
multimers, and so further work would be necessary to comprehensively characterize these
distributions.
The 2 year old Kodak NA sample exhibits mainly oxygen-containing species such as: O2,
O3S, O4, NaO4, O4S and O5S. The OxS species are believed to be contaminants from sulfur-
containing compounds, such as sodium dodecyl sulfate, a common surfactant in many
cleaning and hygiene products with the general formula CH3(CH2)11OSO3Na.
The inherent ultrahigh resolving power and mass accuracy of the FT-ICR MS used in this
study for determining the elemental composition of naphthenic acids is readily evident, even
when operating in broadband (m/z 147.41 and 3000) as shown in Figure 12, which represents
a 20 Da and a 16 mDa mass scale-expended segment. An increase in the mass spectra
resolving power can be achieved by lengthening the transient acquisition (tacq). Longer tacq are
obtained by increasing the number of data points collected and/or by increasing the lowest
m/z value.
The m/z 468-501 zoom in spectrum, reveals single charged molecules, separated by
approximately 14 Da for members of the same homologous series (m/z=473.36, m/z=487.38
and m/z=501.4) and by approximately 2 Da for compounds that have a DBE difference of 1
(m/z=473.36 and m/z=475.38). The 16 mDa zoom-in region (m/z	
   469.324-­‐469.340	
  
window) reveals a resolved doublet of molecules: a NaO4 class species on the left hand side
and an O4 class species on the right side. The	
  two	
  peaks	
  are	
  being	
  separated	
  by	
  about	
  2.4	
  
mDa,	
  i.e.	
  1	
  mDa	
  less	
  than	
  between	
  SH4 and C3. This pair of peaks repeats throughout the
dimer distribution, pointing out the importance of ultrahigh resolving power and mass
accuracy.
Results	
  and	
  discussion	
  
23	
  
	
  
Figure 12. Mass scale-expanded segments (20 Da and 16 mDa wide) of negative ESI FT-ICR
mass spectra of the 2 year old Kodak NA mixture. The bottom spectrum is a zoom in on a
NaO4/O4 pair.
Results	
  and	
  discussion	
  
24	
  
	
  
The abundance values of the species present in distribution I and II have been calculated as
the sum of the intensities of all species from the same class (Table 2). O4 class dominates
distribution II, fitting with previous work reported by Da Campo30
. At a smaller contribution,
it is shown for the first time the formation of sodium bound naphthenic acid dimers.
Table 2. Summed absolute intensity and normalized relative intensity of the species present in
the monomer and dimer distributions for the old 0.4 mg/mL Kodak NA sample (ISD 0 V)
Species Intensity Relative intensity (%)
Monomer O2 1701126842 23.60
O3S 7415251 0.10
Dimer O4 4783158046 66.34
NaO4 612119306 8.49
O4S 11957646 0.17
O5S 93864521 1.30
The relative intensity of the compounds present in the monomer and dimer distributions are
illustrated as graph in Figure 13. O4 class species represent approximately 66% of the species
present in the dimer distribution and NaO4 represent approximately 8%, respectively.
Figure 13. Relative intensity of the species observed in the monomer and dimer distribution in
the 2 year old Kodak naphthenic acid mixture (negative-ion ESI, ISD 0 V)
0	
  
10	
  
20	
  
30	
  
40	
  
50	
  
60	
  
70	
  
O2	
   O3S	
   O4	
   NaO4	
   O4S	
   O5S	
  
Relative	
  Intensity	
  (%)	
  
Species	
  present	
  
Old	
  Kodak	
  naphthenic	
  acid	
  sample	
  composition	
  (ISD	
  0	
  V)	
  	
  
	
  	
  	
  	
  	
  	
  	
  	
  	
  	
  	
  	
  	
  	
  	
  
	
  	
  	
  	
  	
  	
  	
  	
  	
  	
  	
  	
  monomer	
  
	
  
	
  dimer	
  
Results	
  and	
  discussion	
  
25	
  
	
  
3.2.2 ISD	
  experiment	
  (60	
  V)	
  
By increasing the in source fragmentation energy to 60 V, most of the dimeric species
dissociate back to monomeric species, and thus a decrease in intensity of the dimer
distribution was noticed while the intensity of the monomer increases as shown in Figure 14.
Figure 14. Mass spectrum of 2 year old Kodak naphthenic acid sample showing several
distributions (negative-ion ESI, ISD 60 V)
The main event noticed when increasing the fragmentation power to 60 V was that no O4, O4S
and O5S class species were present anymore and the dimer distribution consisted of only
NaO4 class species. The summed absolute and relative intensities of the species present in the
old Kodak NA sample are gathered in Table 3.
Results	
  and	
  discussion	
  
26	
  
	
  
Table 3. Summed absolute intensity and the normalized relative intensity of the species
present in the monomer and dimer distributions for the old Kodak NA sample (ISD 60 V)
Species Intensity Relative intensity (%)
Monomer O2 4786823693 80.36
O2S 5123729 0.09
O3S 24629987 0.41
Dimer NaO4 1139835910 19.14
The relative intensity of the compounds present in the monomer and dimer distributions are
illustrated as bar chart graph in Figure 15.
Figure 15. Relative intensity of the species observed in the monomer and dimer distribution in
the 2 year old Kodak naphthenic acid mixture (negative-ion ESI, ISD 60 V)
For a visual comparison, the content of O2, O4 and NaO4 class species in the Kodak NA
sample at the two different ISD experiments (0 and 60 V) is illustrated as stacked bar chart in
Figure 16. If at 0 V, O4 naphthenic acid dimers predominate the spectrum, at 60 V there are no
O4 species present, while the monomeric naphthenic acids dominate the spectrum. The
explanation behind it is that at higher voltages, O4 dimers dissociate back to the monomeric
form of NAs. The NaO4 dissociate less easily than the O4 dimers as they are tightly bound due
to the sodium atom. This result, determining the role of sodium, will later lead on to be
looked at using a fresh Kodak NA sample.
0	
  
10	
  
20	
  
30	
  
40	
  
50	
  
60	
  
70	
  
80	
  
90	
  
O2	
   O2S	
   O3S	
   NaO4	
  
Relative	
  Intensity	
  (%)	
  
Species	
  present	
  
Old	
  Kodak	
  NA	
  sample	
  composition	
  (ISD	
  
60	
  V)	
  
	
  	
  	
  	
  	
  	
  	
  	
  	
  dimer
	
  
	
  	
  	
  	
  	
  	
  	
  	
  	
  	
  	
  	
  	
  	
  	
  
	
  	
  	
  	
  	
  	
  	
  	
  	
  	
  	
  	
   	
  monomer	
  
Results	
  and	
  discussion	
  
27	
  
	
  
Figure 16. The intensity of main class species present in the monomer and dimer region for
the 2 year old Kodak NA mixture during ISD experiments at 0 and 60 V. At 60 V ISD, there
are no O4 dimers present
The DBE plots of the three main components from the ISD 0 V and ISD 60 V spectra of the 2
year old Kodak NA sample are shown in Figure 17. The DBE plots at ISD 0 V and at ISD 60
V are very similar for O2 and NaO4 class species, respectively. The carbon number varies
between 11 and 26 for O2 class compounds, between 22 and 42 for O4 class compounds and
between 24 and 41 for NaO4 class compounds. The DBE values observed in negative-ion ESI
are half integers due to the change in electron availability attributed to loss of a proton. The
saturate species (Z=0) will have a DBE of 1.5 due to the carboxylic acid group. All higher
DBE values are associated with rings (or double bonds). The DBE ranges from 1.5 to 8.5 for
O2 species, from 2.5 to 12.5 for O4 species and from 4 to 12 for NaO4 class species. The most
abundant species have a DBE value of 3.5 for O2 class, of 5.5 for O4 class and of 6 for NaO4
class.
0.E+00	
  
1.E+09	
  
2.E+09	
  
3.E+09	
  
4.E+09	
  
5.E+09	
  
6.E+09	
  
7.E+09	
  
8.E+09	
  
0V	
   60V	
  
Contributions	
  of	
  the	
  O2,O4	
  and	
  NaO4	
  class	
  
species	
  (ISD	
  experiments)	
  
NaO4	
  
O4	
  
O2	
  
Results	
  and	
  discussion	
  
28	
  
	
  
	
  
	
  
	
  
	
  
Figure	
  17.	
  DBE	
  plots	
  of	
  O2,	
  O4	
  and	
  NaO4	
  species	
  (from	
  top	
  to	
  bottom)	
  at	
  0	
  V	
  (right	
  side)	
  
and	
  60	
  V	
  (left	
  side)	
  
	
  
	
  
	
  
	
  
	
  
10 12 14 16 18 20 22 24 26 28 30 32 34 36 38 40 42
0
1
2
3
4
5
6
7
8
9
10
11
12
13
14 Old Kodak NA sample (ISD 0 V)
O2 class DBE plotDBE
Carbon number
10 12 14 16 18 20 22 24 26 28 30 32 34 36 38 40 42
0
1
2
3
4
5
6
7
8
9
10
11
12
13
14 Old Kodak NA sample (ISD 60 V)
O2 class DBE plot
DBE
Carbon number
10 12 14 16 18 20 22 24 26 28 30 32 34 36 38 40 42
0
2
4
6
8
10
12
14
Old Kodak NA sample (ISD 0 V)
O4 class DBE plot
DBE
Carbon number
10 12 14 16 18 20 22 24 26 28 30 32 34 36 38 40 42
0
2
4
6
8
10
12
14
Old Kodak NA sample (ISD 0 V)
NaO4 class DBE plot
DBE
Carbon number
10 12 14 16 18 20 22 24 26 28 30 32 34 36 38 40 42
0
2
4
6
8
10
12
14
Old Kodak NA sample (ISD 60 V)
NaO4 class DBE plot
DBE
Carbon number
Results	
  and	
  discussion	
  
29	
  
	
  
3.3 Fresh	
  sample	
  
The existence of sodium naphthenates in the two year old Kodak NA sample was believed to
be caused by the presence of sodium from glassware. In order to investigate this hypothesis a
serial dilution of fresh Kodak NA was prepared and analysed using ISD at 0 V to mirror
previous experimental conditions.
3.3.1 ISD	
  (0	
  V)	
  of	
  serial	
  dilution	
  
	
  
The data analysis revealed that the main compounds in the dimer region of the fresh Kodak
NA sample are mainly O4 class species with a very small contribution from NaO4 class
species. Because the interaction timescale is minutes rather than years, there is less time to
interact with sodium from glassware and so much less of the NaO4 class species are observed
when compared with the 2 year old Kodak NA sample.
The work carried by Smith et al29
using negative ESI FT-ICR MS on a 375-400° C distillation
cut from Athabasca bitumen showed that high sample concentration must be used in order to
observe multimers aggregation. At 1 mg/mL, multimers formation was evident and as
concentration increased to 5 mg/mL and 10 mg/mL the relative abundance of multimers was
higher than that of monomers. In this paper, the dimer formation is shown to occur at lower
concentration of Kodak NA sample (0.001 to 0.5 mg/mL).
It was noticed that by increasing the concentration of the naphthenic acid sample, there was a
large increase in the relative abundance of the O4 class species, whereas the O2 monomers
decrease significantly (highlighted data in Table 4). Increasing the concentration from 0.001
mg/mL to 0.5 mg/mL, the abundance of monomeric naphthenic acids drops from ~82% to
~4.5%, while the abundance of O4 dimers increases from ~10% to ~82%.
Results	
  and	
  discussion	
  
30	
  
	
  
Table 4. Relative abundance of the species present in the fresh Kodak NA sample as a factor
of concentration (ISD 0 V).
Comparing the abundance of NaO4 class species from the old fresh Kodak NA sample with
that from the fresh sample, a decrease in the latter was noticed. Where the relative abundance
of NaO4 class species in the 2 year old 0.4 mg/mL Kodak NA sample was ~8.5% (Table 2), in
the fresh 0.5 mg/mL Kodak NA sample it is ~2.8%.
The relative abundance of naphthenic acid monomers and dimers as a factor of concentration
(on a logarithmical scale) is shown in Figure 18. The crossover between the O2 and O4 class
species occurs between 0.005 mg/mL to 0.01 mg/mL.
Figure 18. The relative abundance of O2, O4 and NaO4 class species present in the fresh
Kodak NA sample depending on the sample concentration (ISD, 0 V)
Species/Concentration
(mg/mL)
0.001 0.005 0.01 0.05 0.1 0.5
Monomer O2 81.84734 48.54421 41.65119 17.33382 9.2603 4.42647
O3 0 0.06375 0.10108 0.04236 0 0.14678
O4 1.61367 1.2567 1.17202 0.34764 0.03462 0.02243
O2S 1.55263 1.70097 1.46321 0.84632 0.24235 0.40252
O3S 4.86567 0.1608 0.1229 0.01219 0.1459 0
Dimer NaO4 0 0.44595 0.68072 0.95765 1.90841 2.8396
O4 9.74035 45.27648 52.36459 74.5034 84.42081 82.30006
O5 0 0 0.1012 0.4483 0.16802 0.76221
O4S 0.30642 1.99654 2.34308 5.46864 3.81959 8.82549
O5S 0.07392 0 0 0.03969 0 0.27444
Results	
  and	
  discussion	
  
31	
  
	
  
Plotting out the data from Table 4 into a bar chart (Figure 19) can provide useful overview on
how concentration affects the overall distribution trend in the naphthenic acid sample. By
increasing concentration, naphthenic acid monomers decrease in abundance, whereas
noncovalent O4 dimers and sodium bound naphthenic acid dimers increase in abundance.
Figure 19. The relative abundance of the species present in the fresh Kodak NA sample at
various concentrations (ISD, 0 V)
Further information can be provided by using DBE plot of different compound classes at
various concentrations. DBE plots of the O2 and O4 species at 0.001mg/mL, 0.01mg/mL and
0.1 mg/mL dilution samples are illustrated in Figure 20. Although DBE plots are usually
created for a single class, for this set of data it was decided to include the O2 and O4 classes in
the same DBE plot. The resulting DBE plots help visualise how an increase in concentration
by an order of magnitude changes a shift from high content of monomeric species to high
content of dimeric aggregates. In the 0.001mg/mL sample, the naphthenic acids with a DBE
number of 3.5 are predominant, while in the 0.1mg/mL sample, the naphthenic acid dimers
with a DBE of 5.5 are most abundant.
Results	
  and	
  discussion	
  
32	
  
	
  
Figure 20. DBE plots of O2 and O4 species from spectra of a fresh Kodak NA sample at
various concentrations
Results	
  and	
  discussion	
  
33	
  
	
  
The effect concentration has on the ratio of NaO4/O4 species from the fresh Kodak NA sample
was explored next (Table 5). With increasing concentration the NaO4/O4 ratio increases.
Table 5. The absolute abundance of NaO4 and O4 class species and their ratio in the fresh
Kodak NA sample at various concentrations (ISD, 0 V)
Class/	
  Concentration	
  	
  
(mg/mL)	
  
0.001	
   0.005	
   0.01	
   0.05	
   0.1	
   0.5	
  
NaO4	
   0	
   1.77E+07	
   3.31E+07	
   9.71E+07	
   6.58E+08	
   8.07E+08	
  
O4	
   1.79E+08	
   1.79E+09	
   2.55E+09	
   7.56E+09	
   2.91E+10	
   2.34E+10	
  
NaO4/O4	
   0	
   0.010	
   0.013	
   0.013	
   0.023	
   0.035	
  
By plotting the absolute abundance of the NaO4 and O4 class species from Table 5 into a bar
chart (Figure 21), the change in overall intensity is displayed.
Figure 21. The absolute abundance of the O4 and NaO4 compounds in a fresh Kodak NA
sample at different concentrations. The lowest concentration analysed, i.e. 0.001mg/mL
contains no NaO4 class species
0.0E+00	
  
5.0E+09	
  
1.0E+10	
  
1.5E+10	
  
2.0E+10	
  
2.5E+10	
  
3.0E+10	
  
3.5E+10	
  
0.001	
   0.005	
   0.01	
   0.05	
   0.1	
   0.5	
  
Absolute	
  intensity	
  
Concentration	
  (mg/mL)	
  
O4	
  and	
  NaO4	
  absolute	
  intensity	
  as	
  a	
  factor	
  
of	
  concentration	
  
NaO4	
  
O4	
  
Results	
  and	
  discussion	
  
34	
  
	
  
For a better visualization of the NaO4/O4 ratio at different Kodak NA concentrations, a
stacked bar chart normalized to 100% was created (Figure 22). It can be observed that an
increase in concentration leads to the presence of more NaO4 class species relative to the
abundance of O4 species. Where there was no NaO4 class species at 0.001 mg/ml, at 0.5
mg/mL, they make up to ~3.3% of the summed relative abundance of NaO4 and O4 class
species. A possible explanation is that O4 species dissociate more readily, while NaO4 species
are more tightly bound.
Figure 22. The summed relative intensity of the O4 and NaO4 compounds in a fresh Kodak
NA sample at different concentrations. The y scale shows the 0-10% segment out of 0-100%.
0%	
  
1%	
  
2%	
  
3%	
  
4%	
  
5%	
  
6%	
  
7%	
  
8%	
  
9%	
  
10%	
  
0.001	
   0.005	
   0.01	
   0.05	
   0.1	
   0.5	
  
Relative	
  intensity	
  (%)	
  
Concentration	
  (mg/mL)	
  
O4	
  and	
  NaO4	
  normalized	
  to	
  100%	
  
contribution	
  as	
  a	
  factor	
  of	
  concentration	
  
O4	
  
NaO4	
  
Results	
  and	
  discussion	
  
35	
  
	
  
3.3.2 Importance	
  of	
  ion	
  accumulation	
  time	
  (IAT)	
  
	
  
Accumulating ions in an rf-only multipole for longer period of times before the mass analysis,
also referred to as multipole storage assisted dissociation (MSAD), is a well-known
phenomenon to affect the degree of fragmentation.59
The space-charge limit is defined as the largest amount of charge that can be stored within a
trap.60
	
   It is hypothesised that once the ion density exceeds the space-charge limit, the
coulombic-repulsive forces between the charges will push the ion to larger radii. As a
consequence of the expansion, the ions are subject to higher amplitude oscillations and
therefore they reach higher kinetic energy. Collisions with the background gas molecules in
the hexapole at these higher energies results in fragmentation.61
As illustrated in Figure 23, increasing ion accumulation time from 0.1 to 0.4 s facilitates the
number of undergoing collisions, and thus causing a decrease in the relative intensity of the
naphthenic acid dimers, as well as a slight shift towards higher m/z.
Therefore, ion accumulation time is an important parameter that needs to be carefully
controlled throughout the same set of experiments, otherwise there is a significant change in
monomer/dimer distribution.
Results	
  and	
  discussion	
  
36	
  
	
  
	
  
Figure	
  23.FT-­‐ICR	
  mass	
  spectra	
  (negative-­‐ion	
  ESI,	
  ISD	
  0	
  V)	
  of	
  a	
  1	
  mg/mL	
  NA	
  Kodak	
  mix	
  in	
  
1:1	
  acetonitrile	
  and	
  water	
  at	
  different	
  ion	
  accumulation	
  times	
  	
  
	
  
	
  
	
  
	
  
	
  
	
  
	
  
	
  
	
  
	
  
	
  
	
  
	
  
Results	
  and	
  discussion	
  
37	
  
	
  
3.4 CID	
  on	
  the	
  2	
  year	
  old	
  Kodak	
  NA	
  sample	
  
To obtain an initial insight into the naphthenic acid dimers and how they dissociate back to
naphthenic acids, a collision induced dissociation experiment was carried out. The molecular
ion at m/z=503.410574, corresponding to C32H55O4
-
(Z=-8, DBE=5.5), was chosen as the
precursor ion because of its high intensity. The isolation window was set to be 1 m/z wide.
Within the 1 m/z window, besides the O4 class compound, there was also present a NaO4 class
compound (C31H44NaO4
-
, z=-16, DBE=10), as well as a small number of other very low
intensity ions.
The intensity of the two species during the 2, 6, 8 and 12 V CID experiments are shown in
Table 6. At 2 V, the intensity of the NaO4 compound was significantly less than the intensity
of the O4 compound. It was noticed that by increasing the collision cell energy, there was a
decrease in the intensity of the O4 species, while the intensity of the NaO4 species was slightly
increasing. The significant decrease in intensity of the O4 species occurs from going from 2 V
to 6 V collision cell energy. By 12 V, C31H44NaO4
-
is more intense than C32H55O4
-
.
Table 6. The intensity of the precursor ions during the 2, 6, 8 and 12 V CID experiments
Precursor ion / Collision
cell energy (V)
2 6 8 12
C32H55O4
-
2.05E+09 4.01E+08 7.63E+07 2.09E+06
C31H44NaO4
-
2.72E+07 3.53E+07 3.51E+07 4.00E+07
The absolute intensity of the precursor ions were normalized to relative intensity and then
plotted against the collision cell energy for a better visualization (Figure 24). Higher collision
cell voltages, i.e. higher kinetic energy collisions, result in an increased chance for the weakly
bound O4 class species to undergo fragmentation. The crossover between the two species
happens between 8 V and 12 V.
Results	
  and	
  discussion	
  
38	
  
	
  
Figure 24. The relative intensity of the O4 and NaO4 precursor ion as a function of the
collision cell voltage. By increasing the collision cell voltage, the intensity of the O4 class
species decreases
An example of a CID spectrum is illustrated in figure 25, where the parent ion (m/z=503.4)
and the fragmented species can be observed.
Figure 25. FT-ICR MS spectrum (CID 8 V)
Results	
  and	
  discussion	
  
39	
  
	
  
The resulting spectra for the four different CID experiments, gave rise to monomeric
naphthenic acid species with very similar characteristics in regards to the number of carbon
atoms, the degree on unsaturation and the relative intensity distribution of the NAs with
different z series (Figure 26).
Figure 26. Distribution of the NA homologous series fragments produced from the
dissociation of C32H55O4
-
(Z=-8) during different CID experiments (2, 6, 8 and 12 V)
	
  
The main difference between the four resulting spectra was an increase in signal intensity of
the NA monomers with the increase in collision cell energy. Increasing collision energy
increases abundance of fragments, but does not change fragments produced. The most intense
NA homologous series produced during CID experiments has a Z number of -4, as a result of
the dissociation in half of the O4 dimer precursor ion with a Z number of -8. NA monomers
with Z=-6 and Z=-2 are also very intense. The low intensity monomeric NAs with a Z value
between -10 and -18 are believed to originate from the NaO4 dimer precursor ion.
0.0E+00	
  
5.0E+07	
  
1.0E+08	
  
1.5E+08	
  
2.0E+08	
  
2.5E+08	
  
3.0E+08	
  
12	
   14	
   16	
   18	
   20	
   22	
  
Intensity	
  
Carbon	
  number	
  
Distribution	
  O2-­‐6	
  V	
  (CID)	
  	
  
0.E+00	
  
1.E+07	
  
2.E+07	
  
3.E+07	
  
4.E+07	
  
5.E+07	
  
6.E+07	
  
12	
   14	
   16	
   18	
   20	
   22	
  
Intensity	
  
Carbon	
  number	
  
Distribution	
  O2-­‐	
  2	
  V	
  (CID)	
  
0	
  
-­‐2	
  
-­‐4	
  
-­‐6	
  
-­‐8	
  
-­‐10	
  
-­‐12	
  
-­‐16	
  
-­‐18	
  
0.0E+00	
  
5.0E+07	
  
1.0E+08	
  
1.5E+08	
  
2.0E+08	
  
2.5E+08	
  
3.0E+08	
  
3.5E+08	
  
4.0E+08	
  
12	
   14	
   16	
   18	
   20	
   22	
  
Intensity	
  
Carbon	
  number	
  
Distribution	
  O2-­‐	
  12	
  V	
  (CID)	
  
0.0E+00	
  
5.0E+07	
  
1.0E+08	
  
1.5E+08	
  
2.0E+08	
  
2.5E+08	
  
3.0E+08	
  
12	
   14	
   16	
   18	
   20	
   22	
  
Intensity	
  
Carbon	
  number	
  
Distribution	
  O2-­‐	
  8	
  V	
  (CID)	
  
Results	
  and	
  discussion	
  
40	
  
	
  
3.5 IRMPD	
  experiment	
  of	
  the	
  fresh	
  Kodak	
  NA	
  sample	
  
	
  
During CID experiments, the precursor ions are selected using the quadrupole and then the
ions are passed to collision cell for fragmentation with argon gas (pressure in collision cell is
~7×10-6
mbar). As seen before leakage can happen through the isolation window.
To improve the quality of the isolation, besides isolation in the quadrupole, a correlated sweep
and correlated shots in the ICR cell were used.62
Zoom-in insets on the precursor regions are
shown in Figure 27, where it can be easily observed that carrying out just the isolation in the
quadrupole allows leakage of other ions into the isolation window, and when the correlated
sweep and correlated shots are added the isolation window is cleaner.
Figure 27. Enlarged mass spectrum of the precursor region for the quadrupole isolation (upper
spectrum) and for the quadrupole isolation plus correlated sweep plus correlated shots
(bottom spectrum)
After performing higher resolving power isolation, the options for dissociation in the ICR cell
are sustained off-resonance irradiation - collision induced dissociation (SORI-CID) or
infrared multiphoton dissociation (IRMPD). IRMPD has a great advantage over SORI-CID
because it does not require the use of collision gas. In this way, the pressure in the analyser
Results	
  and	
  discussion	
  
41	
  
	
  
cell is maintained at an optimal level while the pumpdown time is eliminated and thus there is
no diminish in the performance of the instrument (peak shape, resolving power).
IRMPD on the fresh solution was carried out on a series of ions of same carbon number and
class, but differing DBE. This allows a study of fragmentation patterns as a function of DBE.
A difference in Z number by 2 can mean a double bond but also the formation of a new ring.
Firstly the dissociation of the m/z 503.4 precursor ion (C32H55O4
-
, Z=-8) was carried out to
compare with the CID data of the old solution. The resulted spectrum is shown in Figure 28.
Figure 28. FT-ICR MS spectrum (IRMPD) showing the precursor ion at m/z 503.4 and the
fragmented ions
From Figure 29, it can be observed that the monomer distribution obtained from the
dissociation of the m/z 503.4 precursor using IRMPD is very similar with the dissociation of
the same ion using CID (Figure 26). From the fragmentation of the molecular ion with Z=-8,
once again Z=-4 is the most intense fragment series, followed by Z=-2 and Z=-6. While for
the CID data, the apex for Z=-6 (at carbon number 17) is more intense than the apex for Z=-2
(at C16), for the IRMPD data, the apex for Z=-6 (at C17) is less intense than the apex for Z=-
2 (at C16). Also, to be noticed the apex for the series with z=-4 has a maxima at carbon
number 16, i.e. half of the number of carbons the precursor ion has. The fragments obtained
Results	
  and	
  discussion	
  
42	
  
	
  
from both CID and IRMPD data show that the dimers reflect the most abundant monomer
species present and there is no preferential selectivity during dimer formation.
Figure 29. Distribution of the naphthenic acids homologous series fragments produced from
the dissociation of the molecular ion C32H55O4
-
(z=-8) using IRMPD
Next on, IRMPD was carried out on two less saturated precursor ions: C32H53O4
-
(m/z=501.4,
Z=-10) and C32H51O4
-
(m/z=499.4, Z=-12) (Figure 30). By increasing the Z number (index of
hydrogen deficiency), there are more possible combinations for the fragments to form. It
appears to be a trend with respect to the relative intensities of the different homologous series
as a function of the Z value of the selected dimer. A more negative Z value leads to Z series
of monomers being more evenly spread, rather than one or two Z series predominating.	
  
0.E+00	
  
1.E+08	
  
2.E+08	
  
3.E+08	
  
4.E+08	
  
5.E+08	
  
6.E+08	
  
7.E+08	
  
8.E+08	
  
10	
   12	
   14	
   16	
   18	
   20	
   22	
   24	
  
Intensity	
  
Carbon	
  number	
  
O2	
  distribution-­‐framentation	
  of	
  m/z	
  503.4,	
  
Z=-­‐8	
  	
  (IRMPD)	
  
0	
  
-­‐2	
  
-­‐4	
  
-­‐6	
  
-­‐8	
  
Results	
  and	
  discussion	
  
43	
  
	
  
Figure 30. Distribution of the naphthenic acids homologous series fragments produced from
the dissociation of the molecular ion C32H53O4
-
(z=-10) on the left graph and of the molecular
ion C32H51O4
-
(z=-12) on the right graph (IRMPD)
The most negative Z value for the monomer region does not exceed the Z value of the dimer
which proves that there is no additional fragmentation is occurring.
0.0E+00	
  
2.0E+07	
  
4.0E+07	
  
6.0E+07	
  
8.0E+07	
  
1.0E+08	
  
1.2E+08	
  
1.4E+08	
  
1.6E+08	
  
1.8E+08	
  
2.0E+08	
  
10	
   12	
   14	
   16	
   18	
   20	
   22	
  
Intensity	
  
Carbon	
  number	
  
O2	
  distribution	
  -­‐framentation	
  of	
  
m/z	
  501.4,	
  Z=-­‐10	
  (	
  IRMPD)	
  
0	
  
-­‐2	
  
-­‐4	
  
-­‐6	
  
-­‐8	
  
-­‐10	
  
0.0E+00	
  
2.0E+07	
  
4.0E+07	
  
6.0E+07	
  
8.0E+07	
  
1.0E+08	
  
1.2E+08	
  
10	
   12	
   14	
   16	
   18	
   20	
   22	
  
Intensity	
  
Carbon	
  number	
  
O2	
  distribution	
  -­‐framentation	
  
of	
  m/z	
  499.4,	
  Z=-­‐12	
  (	
  IRMPD)	
  
0	
  
-­‐2	
  
-­‐4	
  
-­‐6	
  
-­‐8	
  
-­‐10	
  
-­‐12	
  
Results	
  and	
  discussion	
  
44	
  
	
  
3.6 Fresh	
  Kodak	
  NA	
  sample	
  doped	
  with	
  salts	
  
	
  
It was noticed in previous experiments that salt affects the dimer formation. For further
investigating of this matter, a fresh Kodak NA mixture was doped with sodium chloride. The
oil industry has long used a technique of pumping seawater into an oil reservoir to increase
the pressure and push oil toward producing wells. Understanding the ability of naphthenic
acid to form sodium bond dimers after coming in contact with seawater will help understand
deposition formation in industry.
The average concentration of salt in seawater is about 35 g/L or 599mM, but because such
high salt concentration would overwhelm the ESI process, the concentration of salt used in
this study for doping the naphthenic acid samples was much lower, in the µM range.
Nevertheless, for illustrative reasons, a 0.05 mg/mL Kodak NA sample was doped with
enough sodium chloride to imitate the actual seawater linearity (left hand side vial in Figure
31). At seawater salt concentration level, there is a dramatic change in the appearance of
naphthenic acid the sample when compared to a regular Kodak NA sample that has not been
doped with salt (right hand side vial).
Figure 31. Photography of a Kodak NA mixture (0.05 mg/mL) at seawater concentration of
salt (600 mM) (left hand side) and of a Kodak NA mixture (0.05 mg/mL) with no salt added
(right hand side)
Results	
  and	
  discussion	
  
45	
  
	
  
A fresh 0.05 mg/mL Kodak NA sample that contained no added salts was used as a blank
sample. Mass spectra zoom-ins on the NaO4/O4 pair of peaks from the blank Kodak NA
sample and the three Kodak NA samples doped with sodium chloride are shown in Figure 32.
Figure 32. Expanded mass range of the m/z 431.314 C25H44NaO4
-
and of the m/z 431.317
C27H43O4
-
. Increase in the concentration of salt, increases the relative intensity of NaO4 class
The data analysis carried showed that NaO4 class species make up to 7.23% of the dimeric
distribution in the NA blank sample (Table 7). As expected, by increasing the amount of salt
added to the NA sample, there is an increase in the NaO4 contribution at the expense of
dissociation of the weakly bound O4 species.
Table 7. The relative abundance of the species present in the dimer distribution at none or
different concentration of salt
Species/Concentration salt (µM) 0 0.6 54.55 300
O4 90.63 87.98 86.89 82.73
NaO4 7.23 11.11 11.94 16.20
O4S 2.14 0.92 1.17 1.07
Results	
  and	
  discussion	
  
46	
  
	
  
The relative abundance of the O4, NaO4 and O4S classes in the blank and in the salt doped
Kodak NA samples was plotted as line graph to show the general trend (Figure 33). By
increasing the salt concentration, there is a decrease in the contribution of the O4 class and an
increase in the contribution of the NaO4 class. Adding enough NaCl so the salt concentration
of the Kodak NA sample is 300 µM will increase the NaO4 class contribution to the dimer
from ~7% (the natural state, when no salt is added) to about 16%. A 300 µg/mL salt
concentration is 2000 times lower than the 600 mM seawater salinity.
Figure 33. Relative abundance of O4, NaO4 and O4S species in the dimer distribution of
Kodak NA mixture containing none or various concentration of salt
	
  
	
  
	
  
	
  
	
  
	
  
Results	
  and	
  discussion	
  
47	
  
	
  
3.7 EID	
  experiment	
  on	
  a	
  NIST	
  crude	
  oil	
  sample	
  
	
  
As an additional line of work, further development of methods for understanding structures of
petroleum molecules was pursued. Determining the elemental composition for petroleum
samples has improved greatly the field of petroleomics, but understanding the structures of
the tens of thousands of components present represents the next challenge. In this work, EID
has been applied for the first time to crude oil samples.
Firstly, a positive-ion mode ESI spectrum of a NIST crude oil sample was obtained (Figure
34). The multitude of peaks in the mass spectrum proves the complexity of petroleum
samples.
Figure 34. Mass spectrum of the NIST crude oil (positive mode ESI)
The main contributor to the NIST crude oil mass spectrum is the N1 class (~66%), followed
by OS class (~26%) (Table 8). Other species such as NO, NO2, O4S and OS2 are present as
minor components.
Results	
  and	
  discussion	
  
48	
  
	
  
Table 8. The relative abundance of the species present in the positive-mode ESI spectrum of
NIST crude oil
Species	
   Relative	
  abundance	
  (%)	
  
N	
   66.35	
  
NO	
   3.11	
  
NO2	
   0.35	
  
O4S	
   0.34	
  
OS	
   25.54	
  
OS2	
   1.18	
  
For a better visualization, the values from Table 8 are plotted as a bar chart in Figure 35.
Figure 35. The abundance of the species present in NIST crude oil (positive-ion conditions
ESI)
Because of the high content of N present in the petroleum sample and the known problems
associated with the N-containing compounds in crude oil, such as pollution and catalyst
poisoning, data analysis was focused on N class species. The DBE values of the components
of N1 class were plotted against their carbon number as shown in the DBE plot from Figure
36. DBE values extend from 3.5 to 22.5 with, DBE between 7.5 and 9.5 being most
abundant.
0	
  
10	
  
20	
  
30	
  
40	
  
50	
  
60	
  
70	
  
N	
   NO	
   NO2	
   O4S	
   OS	
   OS2	
  
Relative	
  abundance	
  (%)	
  
Species	
  present	
  
NIST	
  crude	
  oil	
  composition	
  (positive	
  mode	
  
ESI)	
  
Results	
  and	
  discussion	
  
49	
  
	
  
Figure 36. DBE plot of the N1 class present in a NIST crude oil (positive-ion ESI)
The isolation window was chosen to be 60 Da wide in order to accommodate five data points
along homologues series. A DBE plot of the N1 class species from the isolation window is
shown in Figure 37. The DBE values extend from 3.5 to 19.5 while the most abundant species
have DBE values of 6.5 to 9.5. The carbon number ranged between 27 and 32.
Figure 37 DBE plot of the N1 class for the isolation window spectrum
A single EID spectrum (Figure 38) was obtained after one hour experimental time in order to
increase the signal to noise ratio. It was determined that a 30 V cathode bias produces the
0 10 20 30 40 50 60 70 80
0
5
10
15
20
25
NIST crude oil
N1 class DBE plot
DBE
Carbon number
0 5 10 15 20 25 30 35
0
2
4
6
8
10
12
14
16
18
20
22
Isolation
N1 Class DBE plot
DBE
Results	
  and	
  discussion	
  
50	
  
	
  
most extensive fragmentation for the NIST crude oil sample. The cathode bias controls the
energy of the electrons used for irradiation. Lower cathode biases were also tested, but it was
noticed that at values higher than 20 V, the degree of fragmentation improved.
Figure 38. EID spectrum of a NIST crude oil sample. The isolation window extends from m/z
375 to m/z 435
The DBE plot of the N1 class species from the EID spectrum (Figure 39) reveals that DBE
values extend from 3.5 to 20.5, showing a similar distribution as the N1 class species from the
isolation spectrum. When comparing the carbon number of N1 class species between the two
spectra, it is clear that additional fragmentation occurs during EID experiments. Where the
lowest carbon number during isolation was 27, the minimum carbon number after EID was
11. So, fragments smaller than the smallest intact components are seen.
Results	
  and	
  discussion	
  
51	
  
	
  
	
  
Figure	
  39. DBE plot of the N1 class for the EID spectrum	
  
Fragmentation causes successive loss of alkyl chains from the core (fused rings) rather than
alkyl chains plus aromatic carbon, as proved by the horizontal lines to the left of the isolated
peaks in the DBE plot. However, fragmentation does not cause a reduction in aromaticity.
In this research, it was demonstrated that EID is a viable tandem mass spectrometry method
for petroleum analysis. Through EID significant fragmentation was produced with no increase
of pressure in the ICR cell. EID dissociates alkyl chain and thus it provides information about
the core of molecules. It could potentially be used to distinguish archipelago structures
(condensed aromatic cores connected by aliphatic chains) from island structures (one
aromatic core with aliphatic chains). For archipelago structures, DBE plot will show stable
structures favoured along the homologous series, while for island structures it will show loss
of alkyl chains evenly as seen with the NIST crude oil sample.
Therefore, it is possible to use EID on petroleum samples to remove alkyl chains and identify
the minimum number of carbons needed for just the stable core for a particular class species.
	
  
	
  
0 5 10 15 20 25 30 35
2
4
6
8
10
12
14
16
18
20
22
DBE
Carbon number
EID-30 V cathode bias
N1 class DBE plot
Conclusion	
  
52	
  
	
  
4 Conclusion	
  
In this work, negative-ion ESI FT-ICR mass spectrometry reveals two forms of naphthenic
acid dimers to be present in a commercial naphthenic acid sample, instead of one as
previously believed: the O4 class species that have been formerly described by other groups
and the NaO4 class species. The power of the inherent ultrahigh resolving power and mass
accuracy associated with FTICR-MS made it possible to resolve NaO4/O4 pair of peaks,
separated by only 2.4 mDa.
	
  
Formation	
  of	
  naphthenic	
  acid	
  dimers	
  may	
  simply	
  be	
  the	
  route	
  that	
  naturally	
  occurs,	
  but	
  
understanding	
  any	
  preferential	
  dimer	
  aggregation	
  would	
  help	
  understand	
  if	
  there	
  are	
  
any	
  particular	
  culprits	
  or	
  whether	
  all	
  species	
  present	
  are	
  equally	
  guilty	
  for	
  corrosion	
  in	
  
refineries	
   and	
   toxicity	
   to	
   aquatic	
   environment.	
   How easily naphthenic acids can form
dimers will impact the behaviour of these NAs and the availability of those protons to have
consequences for corrosion.
Comparing a 2 year old Kodak naphthenic acid solution with a freshly prepared one, it was
noticed that the dimers present depend on the solution age, and if a sample is preserved in
glassware for long enough time, the sodium bound naphthenic acid will be more abundant
(~8.5% contribution compared to ~2.8% contribution for the fresh sample).
The O4 class compounds have been demonstrated to dissociate readily as they are bound by a
weak noncovalent bond that breaks when increasing the collision power to higher voltages
such as 60 V for ISD experiments and 12 V for CID experiments. On the other hand, the
abundance of NaO4 class compounds remained relatively stable and it is believed that they are
bound by much stronger bonds.
A proposed structure is shown in Figure 40, where the sodium atom is shared between the
four oxygen atoms.
Conclusion	
  
53	
  
	
  
Figure 40. Possible molecular structure of the singly charged NaO4 class species present in
naphthenic acid sample, with delocalized electron density between the carbon atom and the
two oxygen atoms
A fresh Kodak naphthenic acid solution was doped with NaCl in order to monitor the effect of
sodium abundance upon the contribution of the NaO4 class species and to compare with the
aged naphthenic acid sample. This is highly relevant to the extraction of oil from reservoirs
process, where seawater is pumped down to help extract oil. Sodium bound dimers can form
precipitates when seawater comes in contact with naphthenic acids, leading to naphthenates
deposition inside pipelines. Using Kodak NA samples doped with sodium chloride, it was
shown that the more sodium is available, the higher the contribution of the NaO4 class is.
Doping the naphthenic acid sample with a salt concentration 2000 times lower than the actual
seawater salt concentration, increased the NaO4 contribution to the dimer from ~7% to ~16%.
Electron induced ionization was applied successfully for the first time to petroleum providing
proof of concept data. In petroleomics, fs. The EID experiment carried on a crude oil sample
demonstrated the great potential of EID MS/MS in structure characterization.
Further work can include possible tandem mass spectrometry methods for fragmentation of
species with tentative assignments such as the ones in distribution III and IV from the fresh
Kodak NA sample. Additional experiments can be carried out on doping naphthenic acid
samples with higher amounts of salt to mirror the actual salinity level of seawater and on
study of real world sodium naphthenate deposits. Further EID experiments can be carried out
using different cathode potentials on various class compounds from different petroleum
samples in order to obtain greater insight into molecular structures, including information
about favoured structures (archipelago versus island structures). EID MS/MS via FT-ICR MS
could be coupled with chromatography for structure information of isomeric compounds from
petroleum samples.
References	
  
54	
  
	
  
5 References	
  
	
  
1.	
   Cho,	
  Y.;	
  Ahmed,	
  A.;	
  Kim,	
  S.,	
  Application	
  of	
  Atmospheric	
  Pressure	
  Photo	
  Ionization	
  
Hydrogen/Deuterium	
  Exchange	
  High-­‐Resolution	
  Mass	
  Spectrometry	
  for	
  the	
  Molecular	
  
Level	
  Speciation	
  of	
  Nitrogen	
  Compounds	
  in	
  Heavy	
  Crude	
  Oils.	
  Anal.	
  Chem.	
  2013,	
  85	
  (20),	
  
9758-­‐9763.	
  
2.	
   Griffiths,	
  M.	
  T.;	
  Da	
  Campo,	
  R.;	
  O'Connor,	
  P.	
  B.;	
  Barrow,	
  M.	
  P.,	
  Throwing	
  light	
  on	
  
petroleum:	
   simulated	
   exposure	
   of	
   crude	
   oil	
   to	
   sunlight	
   and	
   characterization	
   using	
  
atmospheric	
  pressure	
  photoionization	
  fourier	
  transform	
  ion	
  cyclotron	
  resonance	
  mass	
  
spectrometry.	
  Anal.	
  Chem.	
  2014,	
  86	
  (1),	
  527-­‐34.	
  
3.	
   Colati,	
   K.	
   A.	
   P.;	
   Dalmaschio,	
   G.	
   P.;	
   De	
   Castro,	
   E.	
   V.	
   R.;	
   Gomes,	
   A.	
   O.;	
   Vaz,	
   B.	
   G.;	
  
Romão,	
   W.,	
   Monitoring	
   the	
   liquid/liquid	
   extraction	
   of	
   naphthenic	
   acids	
   in	
   brazilian	
  
crude	
  oil	
  using	
  electrospray	
  ionization	
  FT-­‐ICR	
  mass	
  spectrometry	
  (ESI	
  FT-­‐ICR	
  MS).	
  Fuel	
  
2013,	
  108,	
  647-­‐655.	
  
4.	
   Marshall,	
   A.	
   G.;	
   Rodgers,	
   R.	
   P.,	
   Petroleomics:	
   The	
   Next	
   Grand	
   Challenge	
   for	
  
Chemical	
  Analysis.	
  Acc.	
  Chem.	
  Res.	
  2004,	
  37	
  (1),	
  53-­‐59.	
  
5.	
   Boduszynski,	
   M.	
   M.,	
   Composition	
   of	
   heavy	
   petroleums.	
   1.	
   Molecular	
   weight,	
  
hydrogen	
   deficiency,	
   and	
   heteroatom	
   concentration	
   as	
   a	
   function	
   of	
   atmospheric	
  
equivalent	
  boiling	
  point	
  up	
  to	
  1400.degree.F	
  (760.degree.C).	
  Energy	
  Fuels	
  1987,	
  1	
  (1),	
  2-­‐
11.	
  
6.	
   Mapolelo,	
  M.	
  M.	
  Extraction,	
  Chromatographic	
  Separation	
  and	
  Characterization	
  of	
  
Polar	
   Acidic	
   Species	
   in	
   Crude	
   Oils	
   and	
   Naphthenate	
   Deposits	
   by	
   Ultrahigh	
   Resolution	
  
Fourier	
   Transform	
   Ion	
   Cyclotron	
   Resonance	
   Mass	
   Spectrometry.	
   Florida	
   State	
  
University,	
  2010.	
  
7.	
   Liu,	
   P.;	
   Shi,	
   Q.;	
   Chung,	
   K.	
   H.;	
   Zhang,	
   Y.;	
   Pan,	
   N.;	
   Zhao,	
   S.;	
   Xu,	
   C.,	
   Molecular	
  
Characterization	
  of	
  Sulfur	
  Compounds	
  in	
  Venezuela	
  Crude	
  Oil	
  and	
  Its	
  SARA	
  Fractions	
  by	
  
Electrospray	
  Ionization	
  Fourier	
  Transform	
  Ion	
  Cyclotron	
  Resonance	
  Mass	
  Spectrometry.	
  
Energy	
  Fuels	
  2010,	
  24	
  (9),	
  5089-­‐5096.	
  
8.	
   McKetta,	
   J.	
   J.,	
   Encyclopedia	
   of	
   Chemical	
   Processing	
   and	
   Design.	
   Marcel	
   Dekker:	
  
United	
  States	
  of	
  America,	
  1990;	
  Vol.	
  35.	
  
9.	
   Rodgers,	
   R.	
   P.,	
   Hughey,	
   C.	
   A.;	
   Hendrickson,	
   C.	
   L.;	
   	
   Marshall,	
   A.	
   G.,	
   Advanced	
  
Characterization	
  of	
  Petroleum	
  Crude	
  and	
  Products	
  by	
  High	
  Field	
  Fourier	
  Transform	
  Ion	
  
Cyclotron	
  Resonance	
  Mass	
  Spectrometry.	
  Fuel	
  Chemistry	
  Division	
  Reprints	
  2002,	
  47	
  (2).	
  
10.	
   Hell,	
   C.;	
   Medinger,	
   E.,	
   Ueber	
   das	
   Vorkommen	
   und	
   die	
   Zusammensetzung	
   von	
  
Säuren	
  im	
  Rohpetroleum.	
  Ber.	
  Dtsch.	
  Chem.	
  Ges.	
  1874,	
  7	
  (2),	
  1216-­‐1223.	
  
11.	
   Aschan,	
   O.,	
   Ueber	
   die	
   in	
   dem	
   Erdöl	
   aus	
   Baku	
   vorkommenden	
   Säuren	
   mit	
  
niedrigerem	
  Kohlenstoffgehalt.	
  Ber.	
  Dtsch.	
  Chem.	
  Ges.	
  1891,	
  24	
  (2),	
  2710-­‐2724.	
  
12.	
   Kannel,	
  P.	
  R.;	
  Gan,	
  T.	
  Y.,	
  Naphthenic	
  acids	
  degradation	
  and	
  toxicity	
  mitigation	
  in	
  
tailings	
  wastewater	
  systems	
  and	
  aquatic	
  environments:	
  a	
  review.	
  J	
  Environ	
  Sci	
  Health	
  A	
  
Tox	
  Hazard	
  Subst	
  Environ	
  Eng	
  2012,	
  47	
  (1),	
  1-­‐21.	
  
13.	
   Scott,	
  A.	
  C.;	
  Mackinnon,	
  M.	
  D.;	
  Fedorak,	
  P.	
  M.,	
  Naphthenic	
  Acids	
  in	
  Athabasca	
  Oil	
  
Sands	
   Tailings	
   Waters	
   Are	
   Less	
   Biodegradable	
   than	
   Commercial	
   Naphthenic	
   Acids.	
  
Environ.	
  Sci.	
  Technol.	
  2005,	
  39	
  (21),	
  8388-­‐8394.	
  
14.	
   Dzidic,	
   I.;	
   Somerville,	
   A.	
   C.;	
   Raia,	
   J.	
   C.;	
   Hart,	
   H.	
   V.,	
   Determination	
   of	
   naphthenic	
  
acids	
  in	
  California	
  crudes	
  and	
  refinery	
  wastewaters	
  by	
  fluoride	
  ion	
  chemical	
  ionization	
  
mass	
  spectrometry.	
  Anal.	
  Chem.	
  1988,	
  60	
  (13),	
  1318-­‐1323.	
  
References	
  
55	
  
	
  
15.	
   Headley,	
   J.	
   V.;	
   Peru,	
   K.	
   M.;	
   McMartin,	
   D.	
   W.;	
   Winkler,	
   M.,	
   Determination	
   of	
  
dissolved	
  naphthenic	
  acids	
  in	
  natural	
  waters	
  by	
  using	
  negative-­‐ion	
  electrospray	
  mass	
  
spectrometry.	
  J	
  AOAC	
  Int	
  2002,	
  85	
  (1),	
  182-­‐7.	
  
16.	
   Qian,	
  K.;	
  Robbins,	
  W.	
  K.;	
  Hughey,	
  C.	
  A.;	
  Cooper,	
  H.	
  J.;	
  Rodgers,	
  R.	
  P.;	
  Marshall,	
  A.	
  G.,	
  
Resolution	
   and	
   Identification	
   of	
   Elemental	
   Compositions	
   for	
   More	
   than	
   3000	
   Crude	
  
Acids	
   in	
   Heavy	
   Petroleum	
   by	
   Negative-­‐Ion	
   Microelectrospray	
   High-­‐Field	
   Fourier	
  
Transform	
   Ion	
   Cyclotron	
   Resonance	
   Mass	
   Spectrometry.	
   Energy	
   Fuels	
   2001,	
   15	
   (6),	
  
1505-­‐1511.	
  
17.	
   Barrow,	
  M.	
  P.;	
  McDonnell,	
  L.	
  A.;	
  Feng,	
  X.;	
  Walker,	
  J.;	
  Derrick,	
  P.	
  J.,	
  Determination	
  of	
  
the	
  nature	
  of	
  naphthenic	
  acids	
  present	
  in	
  crude	
  oils	
  using	
  nanospray	
  Fourier	
  transform	
  
ion	
   cyclotron	
   resonance	
   mass	
   spectrometry:	
   the	
   continued	
   battle	
   against	
   corrosion.	
  
Anal.	
  Chem.	
  2003,	
  75	
  (4),	
  860-­‐6.	
  
18.	
   Simon,	
   S.;	
   Reisen,	
   C.;	
   Bersås,	
   A.;	
   Sjöblom,	
   J.,	
   Reaction	
   between	
   tetrameric	
   acids	
  
and	
  Ca	
  2+	
  in	
  oil/water	
  system.	
  Ind.	
  Eng.	
  Chem.	
  Res.	
  2012,	
  51	
  (16),	
  5669-­‐5676.	
  
19.	
   Scaled	
   Solutions	
   LTD,	
   Organic	
   Deposits-­‐	
   Napthenates	
  
http://www.scaledsolutions.com/Performance%20Tests/Organic%20Deposits	
  
(accessed	
  18	
  Aug).	
  
20.	
   El	
   Mahdy,	
   G.	
   A.;	
   Atta,	
   A.	
   M.;	
   Dyab,	
   A.	
   K.	
   F.;	
   Al-­‐Lohedan,	
   H.	
   A.,	
   Protection	
   of	
  
Petroleum	
   Pipeline	
   Carbon	
   Steel	
   Alloys	
   with	
   New	
   Modified	
   Core-­‐Shell	
   Magnetite	
  
Nanogel	
  against	
  Corrosion	
  in	
  Acidic	
  Medium.	
  J.	
  Chem.	
  2013,	
  2013,	
  9.	
  
21.	
   Slavcheva,	
  E.;	
  Shone,	
  B.;	
  Turnbull,	
  A.,	
  Review	
  of	
  naphthenic	
  acid	
  corrosion	
  in	
  oil	
  
refining.	
  Br.	
  Corrros.	
  J	
  1999,	
  34	
  (2),	
  125-­‐131.	
  
22.	
   Qu,	
   D.	
   R.;	
   Zheng,	
   Y.	
   G.;	
   Jiang,	
   X.;	
   Ke,	
   W.,	
   Correlation	
   between	
   the	
   corrosivity	
   of	
  
naphthenic	
  acids	
  and	
  their	
  chemical	
  structures.	
  Anti-­‐Corros	
  Method	
  M	
  2007,	
  54	
  (4),	
  211-­‐
218.	
  
23.	
   Clemente,	
  J.	
  S.;	
  Fedorak,	
  P.	
  M.,	
  A	
  review	
  of	
  the	
  occurrence,	
  analyses,	
  toxicity,	
  and	
  
biodegradation	
  of	
  naphthenic	
  acids.	
  Chemosphere	
  2005,	
  60	
  (5),	
  585-­‐600.	
  
24.	
   GE-­‐Measurement	
   and	
   control,	
   Naphthenic	
   Acid	
   Corrosion.	
   http://www.ge-­‐
mcs.com/en/ndt-­‐corrosion-­‐inspection/naphthenic-­‐acid-­‐corrosion.html	
   (accessed	
   17	
  
Aug).	
  
25.	
   Wang,	
   Y.;	
   Zhou,	
   Q.;	
   Peng,	
   S.;	
   Ma	
   Lena,	
   Q.;	
   Niu,	
   X.,	
   Toxic	
   effects	
   of	
   crude-­‐oil-­‐
contaminated	
   soil	
   in	
   aquatic	
   environment	
   on	
   Carassius	
   auratus	
   and	
   their	
   hepatic	
  
antioxidant	
  defense	
  system.	
  J	
  Environ	
  Sci	
  2009,	
  21	
  (5),	
  612-­‐617.	
  
26.	
   Mohamed,	
  M.	
  H.;	
  Wilson,	
  L.	
  D.;	
  Peru,	
  K.	
  M.;	
  Headley,	
  J.	
  V.,	
  Colloidal	
  properties	
  of	
  
single	
   component	
   naphthenic	
   acids	
   and	
   complex	
   naphthenic	
   acid	
   mixtures.	
   J.	
   Colloid	
  
Interface	
  Sci.	
  2013,	
  395	
  (1),	
  104-­‐110.	
  
27.	
   Yen,	
  T.	
  W.;	
  Marsh,	
  W.	
  P.;	
  MacKinnon,	
  M.	
  D.;	
  Fedorak,	
  P.	
  M.,	
  Measuring	
  naphthenic	
  
acids	
   concentrations	
   in	
   aqueous	
   environmental	
   samples	
   by	
   liquid	
   chromatography.	
  
Journal	
  Chromatogr.	
  A	
  2004,	
  1033	
  (1),	
  83-­‐90.	
  
28.	
   Hemmingsen,	
  P.	
  V.;	
  Kim,	
  S.;	
  Pettersen,	
  H.	
  E.;	
  Rodgers,	
  R.	
  P.;	
  Sjöblom,	
  J.;	
  Marshall,	
  A.	
  
G.,	
   Structural	
   characterization	
   and	
   interfacial	
   behavior	
   of	
   acidic	
   compounds	
   extracted	
  
from	
  a	
  North	
  Sea	
  oil.	
  Energy	
  Fuels	
  2006,	
  20	
  (5),	
  1980-­‐1987.	
  
29.	
   Smith,	
  D.	
  F.;	
  Schaub,	
  T.	
  M.;	
  Rahimi,	
  P.;	
  Teclemariam,	
  A.;	
  Rodgers,	
  R.	
  P.;	
  Marshall,	
  A.	
  
G.,	
  Self-­‐association	
  of	
  organic	
  acids	
  in	
  petroleum	
  and	
  Canadian	
  bitumen	
  characterized	
  by	
  
low-­‐	
  and	
  high-­‐resolution	
  mass	
  spectrometry.	
  Energy	
  Fuels	
  2007,	
  21	
  (3),	
  1309-­‐1316.	
  
30.	
   Da	
   Campo,	
   R.;	
   Barrow,	
   M.	
   P.;	
   Shepherd,	
   A.	
   G.;	
   Salisbury,	
   M.;	
   Derrick,	
   P.	
   J.,	
  
Characterization	
   of	
   Naphthenic	
   Acid	
   Singly	
   Charged	
   Noncovalent	
   Dimers	
   and	
   Their	
  
Dependence	
   on	
   the	
   Accumulation	
   Time	
   within	
   a	
   Hexapole	
   in	
   Fourier	
   Transform	
   Ion	
  
Cyclotron	
  Resonance	
  Mass	
  Spectrometry.	
  Energy	
  Fuels	
  2009,	
  23	
  (11),	
  5544-­‐5549.	
  
Simona Gherghel-MSc Thesis
Simona Gherghel-MSc Thesis
Simona Gherghel-MSc Thesis
Simona Gherghel-MSc Thesis
Simona Gherghel-MSc Thesis
Simona Gherghel-MSc Thesis
Simona Gherghel-MSc Thesis
Simona Gherghel-MSc Thesis
Simona Gherghel-MSc Thesis
Simona Gherghel-MSc Thesis
Simona Gherghel-MSc Thesis
Simona Gherghel-MSc Thesis
Simona Gherghel-MSc Thesis
Simona Gherghel-MSc Thesis
Simona Gherghel-MSc Thesis
Simona Gherghel-MSc Thesis
Simona Gherghel-MSc Thesis
Simona Gherghel-MSc Thesis

More Related Content

Viewers also liked

Bhutanese culture influence learning skill
Bhutanese culture influence learning skillBhutanese culture influence learning skill
Bhutanese culture influence learning skillSonam Dendup
 
雙橡園 銷售報告書 4.01-4.26
雙橡園 銷售報告書 4.01-4.26雙橡園 銷售報告書 4.01-4.26
雙橡園 銷售報告書 4.01-4.26Jeff Ke
 
Herramientas colaborativas
Herramientas colaborativasHerramientas colaborativas
Herramientas colaborativasdannasaavedra12
 

Viewers also liked (6)

arvind laxmichandra
arvind laxmichandraarvind laxmichandra
arvind laxmichandra
 
Brown shantel 3.3
Brown shantel 3.3Brown shantel 3.3
Brown shantel 3.3
 
EVON ABDELBASET Cv
EVON ABDELBASET CvEVON ABDELBASET Cv
EVON ABDELBASET Cv
 
Bhutanese culture influence learning skill
Bhutanese culture influence learning skillBhutanese culture influence learning skill
Bhutanese culture influence learning skill
 
雙橡園 銷售報告書 4.01-4.26
雙橡園 銷售報告書 4.01-4.26雙橡園 銷售報告書 4.01-4.26
雙橡園 銷售報告書 4.01-4.26
 
Herramientas colaborativas
Herramientas colaborativasHerramientas colaborativas
Herramientas colaborativas
 

Similar to Simona Gherghel-MSc Thesis

Matthew Koen - MSc Thesis - 2014 - Hydrocracking of Long Chain n-Paraffins un...
Matthew Koen - MSc Thesis - 2014 - Hydrocracking of Long Chain n-Paraffins un...Matthew Koen - MSc Thesis - 2014 - Hydrocracking of Long Chain n-Paraffins un...
Matthew Koen - MSc Thesis - 2014 - Hydrocracking of Long Chain n-Paraffins un...Matthew Koen
 
Application of-chitosan-a-natural-aminopolysaccharide-for-dye-removal-from-aq...
Application of-chitosan-a-natural-aminopolysaccharide-for-dye-removal-from-aq...Application of-chitosan-a-natural-aminopolysaccharide-for-dye-removal-from-aq...
Application of-chitosan-a-natural-aminopolysaccharide-for-dye-removal-from-aq...Linh Ve
 
Graduation Thesis TUe- Michael Beljaars
Graduation Thesis TUe-  Michael BeljaarsGraduation Thesis TUe-  Michael Beljaars
Graduation Thesis TUe- Michael BeljaarsMichael Beljaars
 
2013 abstract book ASNFC-Shanghai
2013 abstract book ASNFC-Shanghai2013 abstract book ASNFC-Shanghai
2013 abstract book ASNFC-ShanghaiKonstantin German
 
MasterThesis-YimingHu
MasterThesis-YimingHuMasterThesis-YimingHu
MasterThesis-YimingHuYiming Hu
 
M1 - Photoconductive Emitters
M1 - Photoconductive EmittersM1 - Photoconductive Emitters
M1 - Photoconductive EmittersThanh-Quy Nguyen
 
2012thesisBennoMeier
2012thesisBennoMeier2012thesisBennoMeier
2012thesisBennoMeierBenno Meier
 
Master thesis - Nanologica AB
Master thesis - Nanologica ABMaster thesis - Nanologica AB
Master thesis - Nanologica ABMartin A. Olsson
 
Droplet Behaviour and Thermal Separation in Ranque-Hilsch Vortex Tubes
Droplet Behaviour and Thermal Separation in Ranque-Hilsch Vortex TubesDroplet Behaviour and Thermal Separation in Ranque-Hilsch Vortex Tubes
Droplet Behaviour and Thermal Separation in Ranque-Hilsch Vortex TubesRaoul Liew
 
1997 epa health_effects_of_mercury_v5___great_resource
1997 epa health_effects_of_mercury_v5___great_resource1997 epa health_effects_of_mercury_v5___great_resource
1997 epa health_effects_of_mercury_v5___great_resourceDo No Harm, llc
 
ICSA17 Imunologia - Manual eletroforese
ICSA17 Imunologia - Manual eletroforeseICSA17 Imunologia - Manual eletroforese
ICSA17 Imunologia - Manual eletroforeseRicardo Portela
 
Dlugogorski and Balucan_2014_Dehydroxylation of serpentine minerals_Implicati...
Dlugogorski and Balucan_2014_Dehydroxylation of serpentine minerals_Implicati...Dlugogorski and Balucan_2014_Dehydroxylation of serpentine minerals_Implicati...
Dlugogorski and Balucan_2014_Dehydroxylation of serpentine minerals_Implicati...Reydick D Balucan
 
AAS-GTA-reading-material.pdf
AAS-GTA-reading-material.pdfAAS-GTA-reading-material.pdf
AAS-GTA-reading-material.pdfDrkalaivani2
 
AAS-GTA-reading-material.pdf material reading
AAS-GTA-reading-material.pdf material readingAAS-GTA-reading-material.pdf material reading
AAS-GTA-reading-material.pdf material readingAbeerRagab2
 
Fulltext01
Fulltext01Fulltext01
Fulltext01Ngoc Ngk
 

Similar to Simona Gherghel-MSc Thesis (20)

Matthew Koen - MSc Thesis - 2014 - Hydrocracking of Long Chain n-Paraffins un...
Matthew Koen - MSc Thesis - 2014 - Hydrocracking of Long Chain n-Paraffins un...Matthew Koen - MSc Thesis - 2014 - Hydrocracking of Long Chain n-Paraffins un...
Matthew Koen - MSc Thesis - 2014 - Hydrocracking of Long Chain n-Paraffins un...
 
Application of-chitosan-a-natural-aminopolysaccharide-for-dye-removal-from-aq...
Application of-chitosan-a-natural-aminopolysaccharide-for-dye-removal-from-aq...Application of-chitosan-a-natural-aminopolysaccharide-for-dye-removal-from-aq...
Application of-chitosan-a-natural-aminopolysaccharide-for-dye-removal-from-aq...
 
Graduation Thesis TUe- Michael Beljaars
Graduation Thesis TUe-  Michael BeljaarsGraduation Thesis TUe-  Michael Beljaars
Graduation Thesis TUe- Michael Beljaars
 
2013 abstract book ASNFC-Shanghai
2013 abstract book ASNFC-Shanghai2013 abstract book ASNFC-Shanghai
2013 abstract book ASNFC-Shanghai
 
fjgervai2004
fjgervai2004fjgervai2004
fjgervai2004
 
MasterThesis-YimingHu
MasterThesis-YimingHuMasterThesis-YimingHu
MasterThesis-YimingHu
 
M1 - Photoconductive Emitters
M1 - Photoconductive EmittersM1 - Photoconductive Emitters
M1 - Photoconductive Emitters
 
2012thesisBennoMeier
2012thesisBennoMeier2012thesisBennoMeier
2012thesisBennoMeier
 
HASMasterThesis
HASMasterThesisHASMasterThesis
HASMasterThesis
 
Master thesis - Nanologica AB
Master thesis - Nanologica ABMaster thesis - Nanologica AB
Master thesis - Nanologica AB
 
Droplet Behaviour and Thermal Separation in Ranque-Hilsch Vortex Tubes
Droplet Behaviour and Thermal Separation in Ranque-Hilsch Vortex TubesDroplet Behaviour and Thermal Separation in Ranque-Hilsch Vortex Tubes
Droplet Behaviour and Thermal Separation in Ranque-Hilsch Vortex Tubes
 
1997 epa health_effects_of_mercury_v5___great_resource
1997 epa health_effects_of_mercury_v5___great_resource1997 epa health_effects_of_mercury_v5___great_resource
1997 epa health_effects_of_mercury_v5___great_resource
 
Dissertation
Dissertation Dissertation
Dissertation
 
ICSA17 Imunologia - Manual eletroforese
ICSA17 Imunologia - Manual eletroforeseICSA17 Imunologia - Manual eletroforese
ICSA17 Imunologia - Manual eletroforese
 
iosdft
iosdftiosdft
iosdft
 
Dlugogorski and Balucan_2014_Dehydroxylation of serpentine minerals_Implicati...
Dlugogorski and Balucan_2014_Dehydroxylation of serpentine minerals_Implicati...Dlugogorski and Balucan_2014_Dehydroxylation of serpentine minerals_Implicati...
Dlugogorski and Balucan_2014_Dehydroxylation of serpentine minerals_Implicati...
 
AAS-GTA-reading-material.pdf
AAS-GTA-reading-material.pdfAAS-GTA-reading-material.pdf
AAS-GTA-reading-material.pdf
 
AAS-GTA-reading-material.pdf material reading
AAS-GTA-reading-material.pdf material readingAAS-GTA-reading-material.pdf material reading
AAS-GTA-reading-material.pdf material reading
 
Fulltext01
Fulltext01Fulltext01
Fulltext01
 
Ref#11.pdf
Ref#11.pdfRef#11.pdf
Ref#11.pdf
 

Simona Gherghel-MSc Thesis

  • 1.             Department of Chemistry Thesis submitted for the degree of MSc in Analytical Science: Methods and Instrumental Techniques August 2014 Ultrahigh resolving power tandem mass spectrometry of petroleum components By: Simona Gherghel Supervisor: Dr Mark P. Barrow                
  • 2.             Table  of  Contents   1   Introduction  ................................................................................................................  1   1.1   Crude  oil  ..................................................................................................................................................................  1   1.2   Naphthenic  acid  chemistry  ..............................................................................................................................  2   1.3   Formation  of  naphthenic  acid  dimers  ........................................................................................................  5   1.4   Analytical  techniques  for  looking  at  crude  oils  ......................................................................................  6   1.5   Fourier  transform  ion  cyclotron  resonance  mass  spectrometry  ....................................................  7   1.6   High  mass  accuracy  .........................................................................................................................................  10   1.7   Kendrick  mass  defects:  the  key  to  unlocking  chemical  formulas  ................................................  10   1.8   Double-­‐bond  equivalent  versus  carbon  number  plot  .......................................................................  12   1.9   Main  aims  .............................................................................................................................................................  13   2   Experimental  .............................................................................................................  14   2.1   Sample  preparation  .........................................................................................................................................  14   2.2   Fourier  Transform-­‐Infrared  Spectroscopy  (FT-­‐IR)  ...........................................................................  15   2.3   Mass  Spectrometry  Analysis  ........................................................................................................................  15   2.4   Calibration  ...........................................................................................................................................................  17   2.5   Data  analysis  .......................................................................................................................................................  17   3   Results  and  discussion  ...............................................................................................  19   3.1   Proof  of  naphthenic  acid  dimers  existence  in  solution-­‐phase  ......................................................  19   3.2   2  year  old  Kodak  naphthenic  acid  sample  .............................................................................................  20   3.2.1   ISD  experiment  0  V  ...........................................................................................................................................  20   3.2.2   ISD  experiment  (60  V)  ....................................................................................................................................  25   3.3   Fresh  sample  ......................................................................................................................................................  29   3.3.1   ISD  (0  V)  of  serial  dilution  ............................................................................................................................  29   3.3.2   Importance  of  ion  accumulation  time  (IAT)  .........................................................................................  35   3.4   CID  on  the  2  year  old  Kodak  NA  sample  .................................................................................................  37   3.5   IRMPD  experiment  of  the  fresh  Kodak  NA  sample  ............................................................................  40   3.6   Fresh  Kodak  NA  sample  doped  with  salts  .............................................................................................  44   3.7   EID  experiment  on  a  NIST  crude  oil  sample  .........................................................................................  47   4   Conclusion  ................................................................................................................  52   5   References  ................................................................................................................  54   6   Acknowledgments  ....................................................................................................  58   7   Appendix  ..................................................................................................................  59   7.1   Known  ions  mass  list  ......................................................................................................................................  59   7.2   Mass  error  distribution  histograms  .........................................................................................................  65   7.2.1   2  year  old  doped  Kodak  NA  sample-­‐ISD  0  V  .........................................................................................  65   7.2.2   2  year  old  doped  Kodak  NA  sample-­‐ISD  60  V  .......................................................................................  66   7.2.3   Fresh  sample:  serial  dilution  (ISD  0  V  .....................................................................................................  67   7.2.4   CID  on  2  year  old  Kodak  NA  .........................................................................................................................  71   7.2.5   IRMPD  on  fresh  Kodak  NA  sample  ............................................................................................................  73        
  • 3.         Abbreviations     APCI Atmospheric Pressure Chemical Ionization CID Collision Induced Dissociation DBE Double Bond Equivalent EI Electron Ionization EID Electron Induced Dissociation ESI- Electrospray ionization FAB Fast Atom Bombardment FT-ICR MS Fourier Transform Ion Cyclotron Resonance Mass Spectrometry FT-IR Fourier Transform- Infrared Spectroscopy GC-MS Gas Chromatography Mass Spectrometry IAT Ion Accumulation Time ICR Ion Cyclotron Resonance IRMPD Infrared Multiphoton Dissociation ISD In Source Dissociation KMD Kendrick Mass Defect MS/MS Tandem mass spectrometry NA Naphthenic Acid SARA Saturate, aromatic, resin and asphaltene SORI Sustained Off-Resonance Irradiation    
  • 4.         Summary       The selective ionization of acidic compounds from a commercial naphthenic acid mixture and of nitrogen-containing compounds from a crude oil sample was carried out using Fourier transform ion cyclotron resonance mass spectrometry (FT-ICR MS) in negative-ion and positive-ion electrospray ionization (ESI), respectively. Naphthenic acids are of great concern to the petroleum industry as they are responsible for various unwanted phenomena such as pipeline corrosion, formation of sodium naphthenate deposits and they also are toxic to aquatic organisms. Naphthenic acids have been previously demonstrated to form O4 class naphthenic acid dimers. During this research, the formation of sodium bound naphthenic acids (NaO4 class) has been observed. The ultrahigh resolving power and mass accuracy of FT-ICR MS offers accurate mass assignments and information on compositional differences of naphthenic acid species. Several tandem MS methods such as collision-induced dissociation (CID) and infrared multiphoton dissociation (IRMPD) were further employed for structural characterization of the naphthenic acid dimers. CID and IRMPD of the same precursor ions produced very similar fragmentation patterns it has been shown that dimers reflect the most abundant monomer species present. Electron-induced dissociation (EID) was used for structural characterization of crude oil and it provided valuable fragmentation information showing successive losses of alkyl chains leading back to the stable cores made of fused rings.  
  • 5. Introduction   1     1 Introduction   1.1 Crude  oil     Crude oil is a fundamental energy source that has been playing a key role in the progress of modern human society. Even under very optimistic estimates about the development of alternative energy sources, it will remain one of the most heavily used energy sources for many decades to come.1 Global dependence on crude oil originates from its high versatility; it can be used as fuel for vehicles and airplanes, as a home heating source and almost all chemical products, such as plastics, pesticides, detergents, dyes and even medicines begin with oil-derived feedstock.2 Nevertheless crude oil is a limited resource and while its production is decreasing, its consumption only keeps increasing. Therefore, lately there has been an increased focus on the use of heavier, lower quality sources.3 Heavy oils are one of the most complex mixtures in nature and they can contain thousands to millions of various components.4 Typically crude oils have a high proportion of light hydrocarbons (CcHh) and other minor ingredients. The hydrocarbons can be divided into four types: paraffin based (15 to 60%), naphthenes based (30 to 60%), aromatic based (3 to 30%) and asphaltic based (remainder). Carbon accounts for about 85% (83-87%) of the crude oil mass, hydrogen for about 12% (10-14%) and the H/C ratio is about 1.8.5 The minor ingredients of crude oil include nitrogen (0.1-2%), oxygen (0.05-1.5%), sulfur (0.005-6%) and traces of metal (0.1%) such as are sodium (Na), calcium (Ca), magnesium (Mg), aluminium (Al), iron (Fe), vanadium (V), and nickel (Ni). Even though petroleum contains more than 90% hydrocarbons, it is the heteroatom-containing compounds (NnOoSs) the ones that cause most problems regarding pollution, poisoning of catalysts, corrosion and formation of emulsions.6 Low quality crude oils contain higher amounts of sulfur, acidic components and other hydrocarbons. Sulfur-containing compounds naturally occurring in crude oils act as poisons to the catalysts used in the conversion of feedstock to useable intermediate and end products.7 Sweet and sour are terms used to refer to a crude oil’s approximate sulfur content. As a rule, when the sulfur content is in excess of 0.5% the crude oil is considered sour, and when it is 0.5% or less it is considered sweet.8 Nitrogen-containing
  • 6. Introduction   2     compounds also represent a concern as they too can play a role in the catalyst deactivation.9 Oxygen-containing components can include phenols, ketones and carboxylic acids. The presence of acidic substances in crude oil was first observed in a Romanian oil in 1874 by Hell and Medinger10 . Later on, in 1890 Aschan11 named them naphthenic acids. Nowadays, it is well known that naphthenic acids cause corrosion in pipes and equipment used in the processing plants and that they are toxic to aquatic environments.12 1.2 Naphthenic  acid  chemistry   Naphthenic acids (NAs) comprise of a large group of saturated monocyclic, acyclic and aromatic carboxylic acids. They are naturally occurring compounds in most petroleum sources, including crude oil and bitumen from the oil sands.13 The classical formula for NAs is CnH2n+ZO2, where n indicates the carbon number and Z is referred to as the “hydrogen deficiency” and is zero or a negative even integer number. More negative Z values represent an increase in the number of hydrogen atoms lost as the structure gets more compact and as rings are formed14 . The Z value is equal to 0 for saturated linear hydrocarbon chains and it becomes -2 for monocyclic NAs, -4 for bicyclic, -6 for tricyclic and so forth.15 Sample structures of naphthenic acids are shown in Figure 1.   Figure 1. Representative structures of naphthenic acids where Z is the hydrogen deficiency, R represents an alkyl chain and n indicates the number of CH2 groups
  • 7. Introduction   3     The chemical, physical and toxicological characteristics of different NA mixtures depend on their differences in molecular structure, composition, volatility and polarity. NAs dissolve in organic solvents and their pH influence their water solubility. Generally, their pKa values are between 5 and 6.15 Because NAs are water soluble to some degree, their release to wastewaters must be monitored.16 Naphthenic acid make up to 4 wt% of the crude oil composition17 and their presence in crude oils is of great concern to the petroleum industry. NAs are responsible for corrosion in refineries, pollution in refinery wastewaters and in oil sands extraction waters, formation of calcium and sodium naphthenate deposits during production and processing, and formation of emulsions.18 Naphthenate deposition can obstruct pipelines (Figure 2), causing production irregularities and in some cases production shutdowns that can be very expensive for the oil companies. Formation of emulsions and naphthenate deposits results in millions of dollars expenses for the petroleum industry. Figure 2. Naphthenate deposit causing obstruction of a hydrocyclone19
  • 8. Introduction   4     Naphthenic acids are also responsible for the corrosion of pipeline carbon steel alloys that are otherwise resistant to corrosion from compounds containing sulfides.20 An example of pipeline affected by NA corrosion is shown in Figure 3. Corrosion by naphthenic acid is not fully understood but it is known that it involves the chelation of the metal ion by the carboxylate group with the release of hydrogen gas.21 Qu and his colleagues showed that corrosivity of NAs depends greatly on the molecular mass and structure; NAs with higher averaged molecular weight and higher averaged number of ring structures were less corrosive.22 Temperatures between 220 and 400° C facilitate corrosion whereas at temperatures above 400 °C the NAs  decompose forming a film that protects the alloy.23 Figure 3. Naphthenic acid corrosion in crude oil unit piping 24 With the increase for oil mining, transportation and application, more and more soils are susceptible to crude oil contamination. At the same time, more soils contaminated with crude oil enter the aquatic environment through surface runoff.25 Naphthenic acids are environmentally significant because they are known to be toxic to a variety of aquatic organisms, algae and mammals by acting as endocrine disruptors.26 The extraction processes of bitumen, an extra heavy oil subgroup, from oil sands produce tailing waters that contain naphthenic acids. As a consequence, oil sands companies need to monitor and report the concentration of NAs in the waters on and near their leases.27
  • 9. Introduction   5     Nonetheless, NAs also represent a resource. They and their metal carboxylates are used as dryers in dyes, as rubber plasticizer and as fungicides for wood preservatives.28 By separating naphthenic acids and other acidic species present in crude oils, the quality of the crude oil will improve, and on the other hand the NAs can be used as feedstock. 1.3 Formation  of  naphthenic  acid  dimers     Naphthenic acids are both hydrogen-bond acceptors because of the –C=O carbonyl group and hydrogen bond donors because of the –OH hydroxyl group; therefore they can participate in dimer formation by hydrogen bonding. Smith and his colleagues29 used Fourier  transform   ion  cyclotron  resonance  mass  spectrometry  (FT-­‐ICR   MS) to show that naphthenic acids self-associate in the gas-phase at high enough concentration (1 to 10 mg/mL), forming naphthenic acids dimers (general formula CnH2n+zO4). Moreover, they were able to demonstrate that the higher the concentration of naphthenic acids, the higher their tendency to form aggregates. Although dimer formation is an accepted phenomenon in the world of petroleomics, there have not been many studies focused on this topic. Da Campo et al.30 have taken a further look at CnH2n+zO4 aggregates. Fourier transform infrared spectroscopy of these dimers revealed that they are also present in the solution-phase. Using an FT-ICR MS instrument, they demonstrated that these compounds are bound by noncovalent bonds and they reported a possible general structure as the one shown in Figure 4, where a proton is shared between two oxygen atoms. The formation of these species is dependent on the accumulation time of the ions in the hexapole of the ion source. It was noticed that by increasing the accumulation time, and thus increasing the number of collisions taking place, there was a decrease in the signal intensity of these species, which was related to the weak noncovalent bonds that hold together the dimers.
  • 10. Introduction   6     Figure 4. Possible structure for the singly charged noncovalent naphthenic acids dimers30 Mapolelo et al31 studied calcium naphthenate deposits and sodium naphthenate emulsion sample using FT-ICR MS, only after preparation in the laboratory to first remove the metals from the naphthenic acids and thus creating free acids. Naphthenic acids dimers will affect the behaviour of the naphthenic acids in the solution- phase and so the problems associated with them. In this paper, a better understanding of naphthenic acid dimers was pursued through the use of ultrahigh resolving power mass spectrometry. 1.4 Analytical  techniques  for  looking  at  crude  oils     Petroleomics is referred to as the detailed study for elucidating the chemical composition of the components naturally occurring in petroleum and crude oil samples. High resolving power mass spectrometry, especially FT-ICR MS is nowadays the central petroleomics-grade technique. How effective the oil recovery methods are depends on the composition of the oil of interest. Analysis of the composition of crude oil can be highly complex and separation of the major crude oil components can facilitate their characterization. Saturate, aromatic, resin and asphaltene (SARA) analysis is a separation technique that has been widely used in petroleomics. It involves the separation of crude oil components based on their solubility and polarity into four main classes: saturate, aromatic, resin and asphaltene fractions. The first step is the precipitation of asphaltenes in n-alkane solvents, followed by liquid
  • 11. Introduction   7     chromatography using silica or alumina columns to separate the three remaining fractions by their polarity.32 Elucidating the chemical structure of heavy components in crude oil can allow a better understanding and prediction of the petroleum behaviour during processing. Because the organic composition of heavy crude oil is so complex, its characterization has been in the past limited to bulk properties (such as viscosity, density, electric conductivity, light scattering, UV-visible and infrared spectroscopy, 13 C nuclear magnetic resonance NMR, X-ray diffraction)33 and various chemical separation methods based on solubility (like SARA analysis), boiling point, gas and liquid chromatography.34 NMR and X-ray diffraction have been employed in determining the molecular structure of crude oil compounds, but both techniques are limited in the sense that they can only provide an average molecular structure of components. 35 At present, Fourier transform ion cyclotron resonance mass spectrometry is an indispensable method of choice for the analysis of complex samples such as petroleum. FT-ICR MS dominates the petroleomics field because of its high mass accuracy and ultrahigh resolving power, making it possible to calculate the elemental composition of the compounds with great accuracy. 1.5 Fourier  transform  ion  cyclotron  resonance  mass  spectrometry   All types of mass spectrometer allow compounds to be identified by the production of gas- phase ions from a neutral sample and their subsequent sorting and detection based on their mass-to-charge ratio (m/z). The sample can be solid, liquid or gas, depending upon the type of ion source used. The ions are transferred into the mass analyser (ICR cell) and the differences in masses of the fragments enable the mass analyser to separate the ions according to their mass to charge (m/z) ratio.36 FT-ICR MS instruments determine the m/z of an ion by measuring the cyclotron frequency of the ion trapped in a fixed magnetic field. The magnetic field is usually generated using a superconducting magnet. The equation that relates frequency to the m/z of the ions is the following:
  • 12. Introduction   8     𝑓 = !" !!" (eq. 1), where f is the cyclotron frequency (Hz), m is the mass of the ion (kg), q is the charge on the charged particle (coulombs, C) and B is the magnetic field (tesla, T).37 Due to the inherently ultrahigh resolving power (for example 400 000 (full width at half maximum) at m/z 400) and accurate mass measurement, Fourier transform ion cyclotron resonance mass spectrometry has enabled analysts to study heavy crude oil at a molecular level.38 FT-ICR MS has been playing a key role for petroleomics ever since its first application to petroleum distillates39 as it has allowed a better understanding of the structure of the components that make up this highly complex mixture and thus an insight into their behaviour. Most of the characterization of naphthenic acids has been performed using different mass spectrometric (MS) techniques, including gas chromatography mass spectrometry (GC-MS)40 , two-dimensional GC-MS41 and liquid chromatography (LC) MS.27 Also, different ionization approaches such as electron ionization (EI)42 , fast atom bombardment (FAB)43 , atmospheric pressure chemical ionization (APCI)44 , and lately electrospray ionization (ESI)17 have been successfully used for the study of naphthenic acids. ESI is a soft ionization technique that has gained great popularity in the past few years and it has been shown that is it able to selectively ionize acidic and basic components in petroleum samples.45 The analyte dissolved in a suited solvent is sprayed from a small capillary tube into a strong electric field in the presence of a drying gas (typically nitrogen) and thus, generating a fine “mist” of highly charged droplets as illustrated in Figure 5.46 Under negative-ion ESI acids are deprotonated generating negative ions and under positive mode ESI (basic) neutral are protonated forming positive ions.34 Although crude oil contains less than 10% heteroatom- containing compounds, they represent the main concern for the petroleum industry. Because of their high polarity, ESI is specific and especially efficient in generating their gas-phase ions. Moreover, negative-ion ESI is a “soft” ionization techniques that generates [M − H]- without extensive ion fragmentation or matrix interference.
  • 13. Introduction   9     Figure 5. Schematic representation of ion formation by electrospray ionization (ESI) Tandem mass spectrometry (MS/MS) techniques allow insight on structure of naphthenic acids by offering further information about specific ions. The ions of interest from the ion source are selected and then subjected to fragmentation through different dissociation techniques such as collision induced dissociation, infrared multiphoton dissociation and electron induced dissociation. Collision-induced dissociation (CID), also known as collisionally activated dissociation (CAD), is by far the most common MS/MS technique and it involves the acceleration of the ions by electrical potential to high kinetic energy followed by the collision of the ions with neutral gas molecules, resulting in further dissociation of the ions. Central to infrared multiphoton dissociation (IRMPD), is the absorption of multiple infrared photons, leading to a continuous build-up of internal energy by the ions. When the internal energy overcomes the barrier against dissociation, fragmentation of the ions occurs.47 The irradiation of ions is typically carried out using a pulsed CO2 laser. Electron induced dissociation (EID) involves the fragmentation of singly-charged ions succeeding their interaction with high-energy electrons (>10 eV).48 EID has been used in the past for the analysis of peptide ions, metabolites and sodium clusters cations49 but as to date, there are no publications testing the viability of EID for petroleum samples. EID would allow
  • 14. Introduction   10     going beyond the analysis of petroleum using elemental composition alone by obtaining greater structural insight. 1.6 High  mass  accuracy     In mass spectrometry, mass accuracy is defined as the ratio of the m/z measurement error to the true m/z, generally expressed in parts per million (ppm) and mass resolving power represents the ability to distinguish between two peaks varying slightly in m/z.50 Each isotope of each chemical element has a different mass defect and they are known to a great accuracy. By using the power of the inherent high mass accuracy and ultrahigh resolving power of Fourier transform mass spectrometry, it is possible to compare petroleum samples on a molecular level with high accuracy. The importance of ultrahigh resolving power can be illustrated in the case of CH4 and O as they both have a nominal mass of 16, but their exact mass is 16.0313 Da and 15.9949 Da respectively, resulting in a mass difference of 36.4 mDa. Similarly SH4 and C3 have a nominal mass of 36, but their exact masses (36.0034 Da and 36.0000 Da, respectively) differ by 3.4m Da.4 But even with a high mass accuracy and knowledge of the ion’s charge state, the chemical complexity of naphthenic acids increases greatly with molecular weight and the assignment of the elemental composition can be unambiguously done only to about 500 Da.51 Nevertheless, by using Kendrick mass defect plot the mass range can be extended up to three times.52   1.7 Kendrick  mass  defects:  the  key  to  unlocking  chemical  formulas   Kendrick mass defect (KMD) plots have been successfully applied for the study of petroleum samples using ultrahigh resolving power mass spectrometry as they have allowed data interpretation and visualization of complex data sets. KMD plots are particularly useful for indicating the characteristics and behaviour of the species of interest.53 Kendrick mass defect calculation transforms the IUPAC mass scale (12 C=12.000 00 Da) to the Kendrick mass scale (CH2=14.00000 Da instead of 14.01565 Da) using the following equation:
  • 15. Introduction   11     Kendrick Mass=IUPAC Mass× !" !".!"#$# (eq. 2). Then the Kendrick mass defect is defined by: Kendrick Mass Defect (KMD) = Nominal Kendrick Mass-Kendrick Mass (eq. 3), where Kendrick mass is calculated as shown above and nominal Kendrick Mass is the Kendrick mass rounded to the nearest integer. For samples containing a high number of hydrocarbons, such as petroleum, KMD plots simplify the display by showing peak patterns lined up in the horizontal direction. The effectiveness of KMD plots relies on the fact that members of a homologous series from crude oils, i.e. compounds that have the same constitution of heteroatoms and number of rings plus double bonds but different alkyl chain length, will have an identical Kendrick mass defect. In a KMD plot, a homologous series will appear on a horizontal row, making it easy to distinguish from species of other class and type. Another great advantage is that “outlier” data can be immediately determined if they fall outside the main patterns54 . Furthermore, unambiguous assignment of a series of related compounds at low m/z serves to determine all other higher mass members of that species. As a result, KMD scaling extends elemental composition assignment to masses up to three times higher than would be possible using mass-measurement accuracy only. For illustrative purposes, a KMD plot of naphthenic acids is shown in Figure 6. The primary trend horizontally is a constant difference in Kendrick nominal mass of 14.01565 (indicating an increase with a CH2 group) and vertically there is a constant Kendrick mass defect difference of 0.013455 (indicating a difference of two hydrogen atoms and thus an increase in the degree of unsaturation).
  • 16. Introduction   12     Figure 6. Kendrick Mass Defect plot of monomeric naphthenic acids 1.8 Double-­‐bond  equivalent  versus  carbon  number  plot     The first step to evaluate the structure of a compound is to determine its elemental composition. The exact mass measured with MS can be used to calculate the elemental formulae (CcHhNnOoSs). Once the elemental composition is known, double-bond equivalent (DBE), also known as degree of unsaturation, plays a key role in predicting the chemical structure of the compounds present in crude oil samples. DBE represents the number of double bonds and rings in a given molecule. DBE and Z values are different ways of thinking about the same thing, i.e. the hydrogen deficiency (rings and double bonds and degree of unsaturation). The DBE values calculated from the elemental composition can be plotted against carbon number, resulting in a “DBE plot”. Species of the same DBE, i.e. from the same homologous series, will be plotted on a horizontal line. The DBE plots may be rendered more informative
  • 17. Introduction   13     by colour-coding each m/z ion depending on its measured relative abundance. Herein, the gradient goes from red for the least abundant ions to yellow for the most abundant ions. Additionally, the size of each plot dot can differ with the abundance of that ion. Such a plot of naphthenic acids is shown in Figure 7. Figure 7. Example of DBE plot for O2 class species, where the most abundant species have a DBE value of 3.5   1.9 Main  aims   The aim of the following study is to use ultrahigh resolving power FT-ICR MS and tandem MS techniques for characterization of petroleum components. The initial focus is on aggregation of naphthenic acids in the solution phase: understanding their composition and whether these aggregates form selectively or depending on the available monomers. Then electron-induced ionization will be applied for the first time to petroleum in order to obtain preliminary data for proof-of-concept of a new approach for structural insights into petroleum. The interaction of singly-charged ions with high-energy electrons (>10 eV) is expected to result in mass spectra with more extended ion fragmentation than slow-heating dissociation techniques such as CID or IRMPD.48
  • 18. Experimental   14     2 Experimental   2.1 Sample  preparation     A commercial Kodak naphthenic acid mixture (The Eastman Kodak Company, Rochester, NY) was used for all the analysis carried out for the characterization of naphthenic acids using negative-ion ESI. The water used in the experiments was purified using a Direct-Q 3 Ultrapure Water System from Millipore (Billerica, MA, USA). An already prepared two years old sample containing 4 mg of Kodak NA mixture dissolved in 5 mL of acetonitrile (WVR, Leuven, Belgium) and 5 mL of Milli-Q water was now doped with 50µL Agilent #1969 ESI-L Low Concentration HP calibrating mix (Agilent, Palo Alto, CA). The concentration of the sample is 0.4 mg/mL and is expressed in mass/volume instead of molarity as naphthenic acid is a mixture. This sample was then subjected to in-source dissociation (ISD) at 0 and 60 V and to CID fragmentation at 2, 6, 8 and 12 V. ISD is a variation of CID, where an increase in the difference in potential between the two ion funnels gives ions higher kinetic energy at an early stage of the instrument. ISD is a non-selective technique useful for dissociation of noncovalent species. To compare the results obtained from the 2 year old Kodak NA sample, a fresh solution of Kodak NA sample was prepared. This fresh solution containing 19.9 mg of Kodak NA mixture was dissolved in 5 mL of acetonitrile and 5 mL of Milli-Q water (~2 mg/mL). Using this stock solution a serial dilution with the following concentrations: 0.5, 0.1, 0.05, 0.001, 0.005 and 0.001 mg/mL was created and subjected to ISD (0 V) experiments. In the previous CID experiments, the quadrupole was used to isolate the m/z 503.4 ion (with a 1 Da wide isolation window). The ions in the isolation window pass though the collision cell, where CID experiments were performed and then the ions were passed to the ICR cell. To improve the quality of the isolation and the resolving power of the resulting spectra, the quadrupole was used to isolate the species of interest then a correlated sweep (in the cell) plus correlated shots (1% clean up shots power, pulse 0.15846 sec) in the cell was used. The fragmentation was carried out using IRMPD rather than SORI-CID as the later will raise the pressure in the cell and decrease performance.  
  • 19. Experimental   15     Doping of a fresh Kodak NA sample with sodium chloride and analysing the effects salt has on NAs was the next step. A fresh 0.05 mg/mL stock solution Kodak NA mixture was prepared in 10 mL of acetonitrile and 10 mL of Milli-Q water. A 600µM stock solution of sodium chloride was prepared in 20 mL of Milli-Q water (0.7 mg NaCl). Three samples containing the same naphthenic acid mass concentration (0.05 mg/mL) and varying salt concentration values were prepared. The concentration of salt in the three samples was as follows: 0.6 µM, 54.55 µM and 300 µM, respectively. The blank stock solution of Kodak NA sample, as well as the three samples of Kodak NA doped with salt was subjected to MS analysis. A 0.05 mg/mL solution of Kodak NA was also doped with significant higher amounts of NaCl to mirror the seawater salinity, i.e. approximately 600 mM, but it was not run thought the FT-ICR MS. A light-sour crude oil sample, which is a standard reference material from the National Institute of Standards and Technology (NIST, Gaithersburg, Maryland, U.S.A.) was used to obtain a positive-ion ESI spectrum. Then isolation was performed using a window with the width of 60 Da so that the isolation window spans ions of two higher/lower carbon numbers each side of the central ion. EID experiments were carried out and a single EID spectrum was obtained. Because of the difficulty faced by tuning with low signal, the spectrum collected was the average of 500 scans and it took about an hour to acquire. 2.2 Fourier  Transform-­‐Infrared  Spectroscopy  (FT-­‐IR)     FT-IR analysis was carried out on a JASCO FT/IR-470 plus. The Kodak NA mixture was analysed directly, with no prior dilution in solvents. The baseline correction was produced in a nitrogen flow atmosphere, followed by the acquisition of the sample FT-IR spectrum as an average of 200 scans. 2.3 Mass  Spectrometry  Analysis   All mass spectrometry experiments were carried out on a Bruker solariX Fourier transform ion cyclotron resonance mass spectrometer fitted with an Apollo II ion source and a 12 tesla superconducting, refrigerated and ultrashield magnet (Bruker Daltonik GmbH, Bremen). An instrument schematic of the instrument used is shown Figure 8.
  • 20. Experimental   16     Figure 8. Schematic of 12 T Bruker Solarix FT-ICR MS (courtesy of Bruker Daltonik, GmbH, Bremen) For the Kodak NA samples, the instrument was run in negative-ion conditions as negative ESI-MS favours the ionization and detection of acidic compounds such as naphthenic acids. These acidic sites give up a proton and thus creating negatively charged molecules. For the crude oil sample, the instrument was run in positive-ion conditions due to the high abundance of basic nitrogen species, which can be readily protonated and thus, providing a strong signal for MS/MS experiments. The conditions of ESI were the following: syringe flow of 120-300µL/hour, drying gas (nitrogen) temperature of 220° C and a flow rate of 4L/min. The nebulizer gas (also nitrogen) was kept at a pressure of 1.2 bar. The ion accumulation time (IAT) was tuned while monitoring the transient lifetime. For all the CID experiments the IAT was set to 0.5 sec, for IRMPD experiments it was 2 sec and for the ISD experiments of the old and fresh samples it was 0.001s. A set of ISD experiments where the IAT varied between 0.1 and 0.4 was carried out in order to obtain an understanding how the spectrum is affected by accumulating ions for longer period of times. The mass range was set between m/z 147.41 and 3000. Mass spectra consisted of 4M Word data points the signal-to-noise ratio was enhanced by summing 100 time domain transients. Transient length was approximately 1.67 s with a resolving power (m/Δm50%) of about 400
  • 21. Experimental   17     000 at m/z 400 (Δm50% is the mass spectral full peak width at half height). After acquisition, data were zero-filled56 once and sine-bell apodization57 was applied. 2.4 Calibration     Mass lists were obtained using the processing software Data Analysis 4.2 (Bruker Daltonik GmbH, Bremen) where the S/N threshold was set to be higher than 4. Calibration was carried out externally for the old Kodak NA sample doped with Agilent #G1969 Low Concentration HP calibrating mix. In this way, accurate m/z values were obtained over a wide mass range that bracketed the monomer and dimer region of interest and thus improving the mass accuracy and confidence in assignments. For subsequent spectra, the calibrating material was not used in order to eliminate possible peak contributions. The calibration was carried out internally using the O2, O4 and NaO4 homologous peaks (the list of known mass ion used for calibration is provided in the Appendix). 2.5 Data  analysis     Spectral interpretation was carried out using the Data Analysis 4.2 software. For the calculation of the elemental formulas, the usual CcHhNnOoSs formula for petroleum was considered, where c and h are unlimited, n and o are between 0 and 5 and s is between 0 and 4. Up to two 13 C atoms were also included in the calculation because of the high content of carbon, along with 1 atom of Na. Other constraints included even electron species, maximum H/C ratio of 3 and a minimum of -0.5 and a maximum of 40 rings plus double bonds. Several approaches for data analysis and visualization of the large amount of data sets acquired were used in this paper, including Kendrick mass defect plots, double bond equivalent plots, bar charts of contributions from different type of species and distribution of the intensity of various Z series versus carbon number. Data were processed with Excel 14 (Microsoft Corporation, Redmond, Washington, U.S.A.), Origin 9.1 (OriginLab, Northampton, MA, USA) and GraphPad Prism 6 (GraphPad Software, LA Jolla California, USA).
  • 22. Experimental   18     The mass spectra gathered varied from 200 to 300 entries in the peak list for the CID and IRMPD experiments and between 1000-3000 data points for the ISD experiments. The m/z, intensity and resolving power values were the parameters imported from Data Analysis software into Microsoft Excel. Using eq. 2 and eq. 3, a method was created for the calculation of the Kendrick mass defects. An example of the calculation is shown next. For an ion with the composition C15H25O2 - (the negatively singly charged ion of the C15H26O2 naphthenic acid with Z=-4) the IUPAC mass is 237.186004, but using the Kendrick mass scale, the ion has a Kendrick mass of 237.186004*14/14.01565=236.921160. The Nominal Kendrick mass is 237 and the Kendrick mass defect is 237-236.92116=0.07884. Each data entry was manually assigned on the basis of Kendrick mass defects and of m/z values within an error range typically of 1 ppm or less. Then each formula was used to calculate the Z and DBE values, followed by various visualization methods. For all mass spectra, mass error distribution histograms of the main class species present were created and they can be found in the Appendix. The low mass errors obtained using FT-ICR MS allow more confident assignments than by using other type of mass analyser. Double bond equivalent was calculated using the following equation58 : DBE= 𝑐 − ! ! + ! ! + 1, for the elemental formulae CcHhNnOoSsNa1. To ensure persistence in data analysis, all DBE values were calculated for the molecular ions. So, for example, for the molecular ion C15H29O2 - , DBE=15-29/2+1=1.5 (corresponding to a saturated naphthenic acid compound, Z=0). Equivalently, all deprotonated ions of monomeric naphthenic acids from the ESI experiments will have an odd DBE value. For convenience purposes, assignments were categorized in terms of their carbon number, number of double bonds and rings, and their constituent heteroatoms (O, N, S, Na). The heteroatom(s) determines the compound class; for example, O or O1 denotes a composition of CcHhO1 and O2 denotes a composition of CcHhNnO2.
  • 23. Results  and  discussion   19     3 Results  and  discussion   3.1 Proof  of  naphthenic  acid  dimers  existence  in  solution-­‐phase     Carboxylic acids are well known to have a characteristic twin-peak infrared absorption band, with a small peak between 1740 and 1750 cm-1 (monomer form) and a strong peak between 1700 and 1715 cm-1 (dimer form). The pure, undiluted in organic solvents, Kodak naphthenic acid mixture was analysed using a Fourier Transform Infrared Spectroscopy. Its FT-IR spectrum is shown in Figure 9 and it displays this pattern with an apex at 1703, associated with the presence of NA dimers and a smaller apex at 1742 cm-1 , associated with the monomeric NA structures. This readily proves that the components from the Kodak mixture associate to form dimers in the solution-phase at a sufficiently high concentration. Figure 9. Expanded region of the FT-IR spectrum of the Kodak NA mixture showing the absorption band characteristic to the C=O stretching in the carbonyl group. The maxima at 1703 cm-1 corresponds to dimeric species and the maxima at 1742 to the monomeric species. Figure based on data provided from previous study20 using the same Kodak NA mixture as in this work 160016501700175018001850 0.0 0.1 0.2 0.3 0.4 0.5 Absorption(a.u.) Wavenumber (cm-1 ) 1742 (monomer) 1703 (dimer) C=O stretch (COOH)
  • 24. Results  and  discussion   20     3.2 2  year  old  Kodak  naphthenic  acid  sample   3.2.1 ISD  experiment  0  V       During an ISD experiment (0 V), using the 2 year old Kodak naphthenic acid sample doped with Agilent #G1969 ESI-L Low Concentration HP calibrating mix, it was observed that besides the monomer distribution in the m/z 170-390 region (attributed to the naphthenic acid species, i.e. O2 class species), there are three other distinct distributions present in the m/z 365-675, 670-920 and 950-1200 m/z region along with clusters in the m/z 1080-1230, 1370- 1520 and 1700-1850 regions (Figure 10). Figure 10. Mass spectrum of a 2 year old Kodak NA sample showing several distributions (negative-ion ESI, ISD 0 V). The peaks marked with a star originate from the Agilent #G1969 ESI-L Low Concentration HP calibrating mix For convenience purposes, the four distributions present were named using roman numerals as shown in Table 1.
  • 25. Results  and  discussion   21     Table 1. Assigned names for the four distributions in the mass spectrum of a 2 year old Kodak NA sample (negative-ion ESI, ISD 0 V) m/z  region   Distribution  name   170-­‐390   I   365-­‐675   II   670-­‐920   III   950-­‐1200   IV   A very effective visual method to illustrate features and behaviour of the species of interest is the use of Kendrick plots. The m/z for all the peaks in the spectrum were rescaled to the Kendrick Mass scale and the nominal Kendrick Mass was plotted against KMD to reveal the four distributions present (Figure 11). Even though two-dimensional KMD plots do not typically show the relative intensity of the compounds, they are very useful in illustrating the degree of unsaturation. It can be clearly noticed that distribution II has the highest Kendrick mass defect variation and thus its components extend over a wider range of degrees of unsaturation when compared to the other distribution present such as distribution I, i.e. the O2 class species. Also an overlap of species can be observed in distribution II. Figure 11. Kendrick Mass plot of the four distributions present in the 2 year old Kodak NA sample (negative-ion ESI, ISD 0V)
  • 26. Results  and  discussion   22     Carrying out data analysis, the majority of the peaks in distribution II were assigned to O4 and NaO4 class species. Distribution III and IV have tentative assignments of NaOS and NaO3, respectively. Within acceptable mass errors, the compositions were not naphthenic acid multimers, and so further work would be necessary to comprehensively characterize these distributions. The 2 year old Kodak NA sample exhibits mainly oxygen-containing species such as: O2, O3S, O4, NaO4, O4S and O5S. The OxS species are believed to be contaminants from sulfur- containing compounds, such as sodium dodecyl sulfate, a common surfactant in many cleaning and hygiene products with the general formula CH3(CH2)11OSO3Na. The inherent ultrahigh resolving power and mass accuracy of the FT-ICR MS used in this study for determining the elemental composition of naphthenic acids is readily evident, even when operating in broadband (m/z 147.41 and 3000) as shown in Figure 12, which represents a 20 Da and a 16 mDa mass scale-expended segment. An increase in the mass spectra resolving power can be achieved by lengthening the transient acquisition (tacq). Longer tacq are obtained by increasing the number of data points collected and/or by increasing the lowest m/z value. The m/z 468-501 zoom in spectrum, reveals single charged molecules, separated by approximately 14 Da for members of the same homologous series (m/z=473.36, m/z=487.38 and m/z=501.4) and by approximately 2 Da for compounds that have a DBE difference of 1 (m/z=473.36 and m/z=475.38). The 16 mDa zoom-in region (m/z   469.324-­‐469.340   window) reveals a resolved doublet of molecules: a NaO4 class species on the left hand side and an O4 class species on the right side. The  two  peaks  are  being  separated  by  about  2.4   mDa,  i.e.  1  mDa  less  than  between  SH4 and C3. This pair of peaks repeats throughout the dimer distribution, pointing out the importance of ultrahigh resolving power and mass accuracy.
  • 27. Results  and  discussion   23     Figure 12. Mass scale-expanded segments (20 Da and 16 mDa wide) of negative ESI FT-ICR mass spectra of the 2 year old Kodak NA mixture. The bottom spectrum is a zoom in on a NaO4/O4 pair.
  • 28. Results  and  discussion   24     The abundance values of the species present in distribution I and II have been calculated as the sum of the intensities of all species from the same class (Table 2). O4 class dominates distribution II, fitting with previous work reported by Da Campo30 . At a smaller contribution, it is shown for the first time the formation of sodium bound naphthenic acid dimers. Table 2. Summed absolute intensity and normalized relative intensity of the species present in the monomer and dimer distributions for the old 0.4 mg/mL Kodak NA sample (ISD 0 V) Species Intensity Relative intensity (%) Monomer O2 1701126842 23.60 O3S 7415251 0.10 Dimer O4 4783158046 66.34 NaO4 612119306 8.49 O4S 11957646 0.17 O5S 93864521 1.30 The relative intensity of the compounds present in the monomer and dimer distributions are illustrated as graph in Figure 13. O4 class species represent approximately 66% of the species present in the dimer distribution and NaO4 represent approximately 8%, respectively. Figure 13. Relative intensity of the species observed in the monomer and dimer distribution in the 2 year old Kodak naphthenic acid mixture (negative-ion ESI, ISD 0 V) 0   10   20   30   40   50   60   70   O2   O3S   O4   NaO4   O4S   O5S   Relative  Intensity  (%)   Species  present   Old  Kodak  naphthenic  acid  sample  composition  (ISD  0  V)                                                          monomer      dimer  
  • 29. Results  and  discussion   25     3.2.2 ISD  experiment  (60  V)   By increasing the in source fragmentation energy to 60 V, most of the dimeric species dissociate back to monomeric species, and thus a decrease in intensity of the dimer distribution was noticed while the intensity of the monomer increases as shown in Figure 14. Figure 14. Mass spectrum of 2 year old Kodak naphthenic acid sample showing several distributions (negative-ion ESI, ISD 60 V) The main event noticed when increasing the fragmentation power to 60 V was that no O4, O4S and O5S class species were present anymore and the dimer distribution consisted of only NaO4 class species. The summed absolute and relative intensities of the species present in the old Kodak NA sample are gathered in Table 3.
  • 30. Results  and  discussion   26     Table 3. Summed absolute intensity and the normalized relative intensity of the species present in the monomer and dimer distributions for the old Kodak NA sample (ISD 60 V) Species Intensity Relative intensity (%) Monomer O2 4786823693 80.36 O2S 5123729 0.09 O3S 24629987 0.41 Dimer NaO4 1139835910 19.14 The relative intensity of the compounds present in the monomer and dimer distributions are illustrated as bar chart graph in Figure 15. Figure 15. Relative intensity of the species observed in the monomer and dimer distribution in the 2 year old Kodak naphthenic acid mixture (negative-ion ESI, ISD 60 V) For a visual comparison, the content of O2, O4 and NaO4 class species in the Kodak NA sample at the two different ISD experiments (0 and 60 V) is illustrated as stacked bar chart in Figure 16. If at 0 V, O4 naphthenic acid dimers predominate the spectrum, at 60 V there are no O4 species present, while the monomeric naphthenic acids dominate the spectrum. The explanation behind it is that at higher voltages, O4 dimers dissociate back to the monomeric form of NAs. The NaO4 dissociate less easily than the O4 dimers as they are tightly bound due to the sodium atom. This result, determining the role of sodium, will later lead on to be looked at using a fresh Kodak NA sample. 0   10   20   30   40   50   60   70   80   90   O2   O2S   O3S   NaO4   Relative  Intensity  (%)   Species  present   Old  Kodak  NA  sample  composition  (ISD   60  V)                    dimer                                                          monomer  
  • 31. Results  and  discussion   27     Figure 16. The intensity of main class species present in the monomer and dimer region for the 2 year old Kodak NA mixture during ISD experiments at 0 and 60 V. At 60 V ISD, there are no O4 dimers present The DBE plots of the three main components from the ISD 0 V and ISD 60 V spectra of the 2 year old Kodak NA sample are shown in Figure 17. The DBE plots at ISD 0 V and at ISD 60 V are very similar for O2 and NaO4 class species, respectively. The carbon number varies between 11 and 26 for O2 class compounds, between 22 and 42 for O4 class compounds and between 24 and 41 for NaO4 class compounds. The DBE values observed in negative-ion ESI are half integers due to the change in electron availability attributed to loss of a proton. The saturate species (Z=0) will have a DBE of 1.5 due to the carboxylic acid group. All higher DBE values are associated with rings (or double bonds). The DBE ranges from 1.5 to 8.5 for O2 species, from 2.5 to 12.5 for O4 species and from 4 to 12 for NaO4 class species. The most abundant species have a DBE value of 3.5 for O2 class, of 5.5 for O4 class and of 6 for NaO4 class. 0.E+00   1.E+09   2.E+09   3.E+09   4.E+09   5.E+09   6.E+09   7.E+09   8.E+09   0V   60V   Contributions  of  the  O2,O4  and  NaO4  class   species  (ISD  experiments)   NaO4   O4   O2  
  • 32. Results  and  discussion   28             Figure  17.  DBE  plots  of  O2,  O4  and  NaO4  species  (from  top  to  bottom)  at  0  V  (right  side)   and  60  V  (left  side)             10 12 14 16 18 20 22 24 26 28 30 32 34 36 38 40 42 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 Old Kodak NA sample (ISD 0 V) O2 class DBE plotDBE Carbon number 10 12 14 16 18 20 22 24 26 28 30 32 34 36 38 40 42 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 Old Kodak NA sample (ISD 60 V) O2 class DBE plot DBE Carbon number 10 12 14 16 18 20 22 24 26 28 30 32 34 36 38 40 42 0 2 4 6 8 10 12 14 Old Kodak NA sample (ISD 0 V) O4 class DBE plot DBE Carbon number 10 12 14 16 18 20 22 24 26 28 30 32 34 36 38 40 42 0 2 4 6 8 10 12 14 Old Kodak NA sample (ISD 0 V) NaO4 class DBE plot DBE Carbon number 10 12 14 16 18 20 22 24 26 28 30 32 34 36 38 40 42 0 2 4 6 8 10 12 14 Old Kodak NA sample (ISD 60 V) NaO4 class DBE plot DBE Carbon number
  • 33. Results  and  discussion   29     3.3 Fresh  sample   The existence of sodium naphthenates in the two year old Kodak NA sample was believed to be caused by the presence of sodium from glassware. In order to investigate this hypothesis a serial dilution of fresh Kodak NA was prepared and analysed using ISD at 0 V to mirror previous experimental conditions. 3.3.1 ISD  (0  V)  of  serial  dilution     The data analysis revealed that the main compounds in the dimer region of the fresh Kodak NA sample are mainly O4 class species with a very small contribution from NaO4 class species. Because the interaction timescale is minutes rather than years, there is less time to interact with sodium from glassware and so much less of the NaO4 class species are observed when compared with the 2 year old Kodak NA sample. The work carried by Smith et al29 using negative ESI FT-ICR MS on a 375-400° C distillation cut from Athabasca bitumen showed that high sample concentration must be used in order to observe multimers aggregation. At 1 mg/mL, multimers formation was evident and as concentration increased to 5 mg/mL and 10 mg/mL the relative abundance of multimers was higher than that of monomers. In this paper, the dimer formation is shown to occur at lower concentration of Kodak NA sample (0.001 to 0.5 mg/mL). It was noticed that by increasing the concentration of the naphthenic acid sample, there was a large increase in the relative abundance of the O4 class species, whereas the O2 monomers decrease significantly (highlighted data in Table 4). Increasing the concentration from 0.001 mg/mL to 0.5 mg/mL, the abundance of monomeric naphthenic acids drops from ~82% to ~4.5%, while the abundance of O4 dimers increases from ~10% to ~82%.
  • 34. Results  and  discussion   30     Table 4. Relative abundance of the species present in the fresh Kodak NA sample as a factor of concentration (ISD 0 V). Comparing the abundance of NaO4 class species from the old fresh Kodak NA sample with that from the fresh sample, a decrease in the latter was noticed. Where the relative abundance of NaO4 class species in the 2 year old 0.4 mg/mL Kodak NA sample was ~8.5% (Table 2), in the fresh 0.5 mg/mL Kodak NA sample it is ~2.8%. The relative abundance of naphthenic acid monomers and dimers as a factor of concentration (on a logarithmical scale) is shown in Figure 18. The crossover between the O2 and O4 class species occurs between 0.005 mg/mL to 0.01 mg/mL. Figure 18. The relative abundance of O2, O4 and NaO4 class species present in the fresh Kodak NA sample depending on the sample concentration (ISD, 0 V) Species/Concentration (mg/mL) 0.001 0.005 0.01 0.05 0.1 0.5 Monomer O2 81.84734 48.54421 41.65119 17.33382 9.2603 4.42647 O3 0 0.06375 0.10108 0.04236 0 0.14678 O4 1.61367 1.2567 1.17202 0.34764 0.03462 0.02243 O2S 1.55263 1.70097 1.46321 0.84632 0.24235 0.40252 O3S 4.86567 0.1608 0.1229 0.01219 0.1459 0 Dimer NaO4 0 0.44595 0.68072 0.95765 1.90841 2.8396 O4 9.74035 45.27648 52.36459 74.5034 84.42081 82.30006 O5 0 0 0.1012 0.4483 0.16802 0.76221 O4S 0.30642 1.99654 2.34308 5.46864 3.81959 8.82549 O5S 0.07392 0 0 0.03969 0 0.27444
  • 35. Results  and  discussion   31     Plotting out the data from Table 4 into a bar chart (Figure 19) can provide useful overview on how concentration affects the overall distribution trend in the naphthenic acid sample. By increasing concentration, naphthenic acid monomers decrease in abundance, whereas noncovalent O4 dimers and sodium bound naphthenic acid dimers increase in abundance. Figure 19. The relative abundance of the species present in the fresh Kodak NA sample at various concentrations (ISD, 0 V) Further information can be provided by using DBE plot of different compound classes at various concentrations. DBE plots of the O2 and O4 species at 0.001mg/mL, 0.01mg/mL and 0.1 mg/mL dilution samples are illustrated in Figure 20. Although DBE plots are usually created for a single class, for this set of data it was decided to include the O2 and O4 classes in the same DBE plot. The resulting DBE plots help visualise how an increase in concentration by an order of magnitude changes a shift from high content of monomeric species to high content of dimeric aggregates. In the 0.001mg/mL sample, the naphthenic acids with a DBE number of 3.5 are predominant, while in the 0.1mg/mL sample, the naphthenic acid dimers with a DBE of 5.5 are most abundant.
  • 36. Results  and  discussion   32     Figure 20. DBE plots of O2 and O4 species from spectra of a fresh Kodak NA sample at various concentrations
  • 37. Results  and  discussion   33     The effect concentration has on the ratio of NaO4/O4 species from the fresh Kodak NA sample was explored next (Table 5). With increasing concentration the NaO4/O4 ratio increases. Table 5. The absolute abundance of NaO4 and O4 class species and their ratio in the fresh Kodak NA sample at various concentrations (ISD, 0 V) Class/  Concentration     (mg/mL)   0.001   0.005   0.01   0.05   0.1   0.5   NaO4   0   1.77E+07   3.31E+07   9.71E+07   6.58E+08   8.07E+08   O4   1.79E+08   1.79E+09   2.55E+09   7.56E+09   2.91E+10   2.34E+10   NaO4/O4   0   0.010   0.013   0.013   0.023   0.035   By plotting the absolute abundance of the NaO4 and O4 class species from Table 5 into a bar chart (Figure 21), the change in overall intensity is displayed. Figure 21. The absolute abundance of the O4 and NaO4 compounds in a fresh Kodak NA sample at different concentrations. The lowest concentration analysed, i.e. 0.001mg/mL contains no NaO4 class species 0.0E+00   5.0E+09   1.0E+10   1.5E+10   2.0E+10   2.5E+10   3.0E+10   3.5E+10   0.001   0.005   0.01   0.05   0.1   0.5   Absolute  intensity   Concentration  (mg/mL)   O4  and  NaO4  absolute  intensity  as  a  factor   of  concentration   NaO4   O4  
  • 38. Results  and  discussion   34     For a better visualization of the NaO4/O4 ratio at different Kodak NA concentrations, a stacked bar chart normalized to 100% was created (Figure 22). It can be observed that an increase in concentration leads to the presence of more NaO4 class species relative to the abundance of O4 species. Where there was no NaO4 class species at 0.001 mg/ml, at 0.5 mg/mL, they make up to ~3.3% of the summed relative abundance of NaO4 and O4 class species. A possible explanation is that O4 species dissociate more readily, while NaO4 species are more tightly bound. Figure 22. The summed relative intensity of the O4 and NaO4 compounds in a fresh Kodak NA sample at different concentrations. The y scale shows the 0-10% segment out of 0-100%. 0%   1%   2%   3%   4%   5%   6%   7%   8%   9%   10%   0.001   0.005   0.01   0.05   0.1   0.5   Relative  intensity  (%)   Concentration  (mg/mL)   O4  and  NaO4  normalized  to  100%   contribution  as  a  factor  of  concentration   O4   NaO4  
  • 39. Results  and  discussion   35     3.3.2 Importance  of  ion  accumulation  time  (IAT)     Accumulating ions in an rf-only multipole for longer period of times before the mass analysis, also referred to as multipole storage assisted dissociation (MSAD), is a well-known phenomenon to affect the degree of fragmentation.59 The space-charge limit is defined as the largest amount of charge that can be stored within a trap.60   It is hypothesised that once the ion density exceeds the space-charge limit, the coulombic-repulsive forces between the charges will push the ion to larger radii. As a consequence of the expansion, the ions are subject to higher amplitude oscillations and therefore they reach higher kinetic energy. Collisions with the background gas molecules in the hexapole at these higher energies results in fragmentation.61 As illustrated in Figure 23, increasing ion accumulation time from 0.1 to 0.4 s facilitates the number of undergoing collisions, and thus causing a decrease in the relative intensity of the naphthenic acid dimers, as well as a slight shift towards higher m/z. Therefore, ion accumulation time is an important parameter that needs to be carefully controlled throughout the same set of experiments, otherwise there is a significant change in monomer/dimer distribution.
  • 40. Results  and  discussion   36       Figure  23.FT-­‐ICR  mass  spectra  (negative-­‐ion  ESI,  ISD  0  V)  of  a  1  mg/mL  NA  Kodak  mix  in   1:1  acetonitrile  and  water  at  different  ion  accumulation  times                              
  • 41. Results  and  discussion   37     3.4 CID  on  the  2  year  old  Kodak  NA  sample   To obtain an initial insight into the naphthenic acid dimers and how they dissociate back to naphthenic acids, a collision induced dissociation experiment was carried out. The molecular ion at m/z=503.410574, corresponding to C32H55O4 - (Z=-8, DBE=5.5), was chosen as the precursor ion because of its high intensity. The isolation window was set to be 1 m/z wide. Within the 1 m/z window, besides the O4 class compound, there was also present a NaO4 class compound (C31H44NaO4 - , z=-16, DBE=10), as well as a small number of other very low intensity ions. The intensity of the two species during the 2, 6, 8 and 12 V CID experiments are shown in Table 6. At 2 V, the intensity of the NaO4 compound was significantly less than the intensity of the O4 compound. It was noticed that by increasing the collision cell energy, there was a decrease in the intensity of the O4 species, while the intensity of the NaO4 species was slightly increasing. The significant decrease in intensity of the O4 species occurs from going from 2 V to 6 V collision cell energy. By 12 V, C31H44NaO4 - is more intense than C32H55O4 - . Table 6. The intensity of the precursor ions during the 2, 6, 8 and 12 V CID experiments Precursor ion / Collision cell energy (V) 2 6 8 12 C32H55O4 - 2.05E+09 4.01E+08 7.63E+07 2.09E+06 C31H44NaO4 - 2.72E+07 3.53E+07 3.51E+07 4.00E+07 The absolute intensity of the precursor ions were normalized to relative intensity and then plotted against the collision cell energy for a better visualization (Figure 24). Higher collision cell voltages, i.e. higher kinetic energy collisions, result in an increased chance for the weakly bound O4 class species to undergo fragmentation. The crossover between the two species happens between 8 V and 12 V.
  • 42. Results  and  discussion   38     Figure 24. The relative intensity of the O4 and NaO4 precursor ion as a function of the collision cell voltage. By increasing the collision cell voltage, the intensity of the O4 class species decreases An example of a CID spectrum is illustrated in figure 25, where the parent ion (m/z=503.4) and the fragmented species can be observed. Figure 25. FT-ICR MS spectrum (CID 8 V)
  • 43. Results  and  discussion   39     The resulting spectra for the four different CID experiments, gave rise to monomeric naphthenic acid species with very similar characteristics in regards to the number of carbon atoms, the degree on unsaturation and the relative intensity distribution of the NAs with different z series (Figure 26). Figure 26. Distribution of the NA homologous series fragments produced from the dissociation of C32H55O4 - (Z=-8) during different CID experiments (2, 6, 8 and 12 V)   The main difference between the four resulting spectra was an increase in signal intensity of the NA monomers with the increase in collision cell energy. Increasing collision energy increases abundance of fragments, but does not change fragments produced. The most intense NA homologous series produced during CID experiments has a Z number of -4, as a result of the dissociation in half of the O4 dimer precursor ion with a Z number of -8. NA monomers with Z=-6 and Z=-2 are also very intense. The low intensity monomeric NAs with a Z value between -10 and -18 are believed to originate from the NaO4 dimer precursor ion. 0.0E+00   5.0E+07   1.0E+08   1.5E+08   2.0E+08   2.5E+08   3.0E+08   12   14   16   18   20   22   Intensity   Carbon  number   Distribution  O2-­‐6  V  (CID)     0.E+00   1.E+07   2.E+07   3.E+07   4.E+07   5.E+07   6.E+07   12   14   16   18   20   22   Intensity   Carbon  number   Distribution  O2-­‐  2  V  (CID)   0   -­‐2   -­‐4   -­‐6   -­‐8   -­‐10   -­‐12   -­‐16   -­‐18   0.0E+00   5.0E+07   1.0E+08   1.5E+08   2.0E+08   2.5E+08   3.0E+08   3.5E+08   4.0E+08   12   14   16   18   20   22   Intensity   Carbon  number   Distribution  O2-­‐  12  V  (CID)   0.0E+00   5.0E+07   1.0E+08   1.5E+08   2.0E+08   2.5E+08   3.0E+08   12   14   16   18   20   22   Intensity   Carbon  number   Distribution  O2-­‐  8  V  (CID)  
  • 44. Results  and  discussion   40     3.5 IRMPD  experiment  of  the  fresh  Kodak  NA  sample     During CID experiments, the precursor ions are selected using the quadrupole and then the ions are passed to collision cell for fragmentation with argon gas (pressure in collision cell is ~7×10-6 mbar). As seen before leakage can happen through the isolation window. To improve the quality of the isolation, besides isolation in the quadrupole, a correlated sweep and correlated shots in the ICR cell were used.62 Zoom-in insets on the precursor regions are shown in Figure 27, where it can be easily observed that carrying out just the isolation in the quadrupole allows leakage of other ions into the isolation window, and when the correlated sweep and correlated shots are added the isolation window is cleaner. Figure 27. Enlarged mass spectrum of the precursor region for the quadrupole isolation (upper spectrum) and for the quadrupole isolation plus correlated sweep plus correlated shots (bottom spectrum) After performing higher resolving power isolation, the options for dissociation in the ICR cell are sustained off-resonance irradiation - collision induced dissociation (SORI-CID) or infrared multiphoton dissociation (IRMPD). IRMPD has a great advantage over SORI-CID because it does not require the use of collision gas. In this way, the pressure in the analyser
  • 45. Results  and  discussion   41     cell is maintained at an optimal level while the pumpdown time is eliminated and thus there is no diminish in the performance of the instrument (peak shape, resolving power). IRMPD on the fresh solution was carried out on a series of ions of same carbon number and class, but differing DBE. This allows a study of fragmentation patterns as a function of DBE. A difference in Z number by 2 can mean a double bond but also the formation of a new ring. Firstly the dissociation of the m/z 503.4 precursor ion (C32H55O4 - , Z=-8) was carried out to compare with the CID data of the old solution. The resulted spectrum is shown in Figure 28. Figure 28. FT-ICR MS spectrum (IRMPD) showing the precursor ion at m/z 503.4 and the fragmented ions From Figure 29, it can be observed that the monomer distribution obtained from the dissociation of the m/z 503.4 precursor using IRMPD is very similar with the dissociation of the same ion using CID (Figure 26). From the fragmentation of the molecular ion with Z=-8, once again Z=-4 is the most intense fragment series, followed by Z=-2 and Z=-6. While for the CID data, the apex for Z=-6 (at carbon number 17) is more intense than the apex for Z=-2 (at C16), for the IRMPD data, the apex for Z=-6 (at C17) is less intense than the apex for Z=- 2 (at C16). Also, to be noticed the apex for the series with z=-4 has a maxima at carbon number 16, i.e. half of the number of carbons the precursor ion has. The fragments obtained
  • 46. Results  and  discussion   42     from both CID and IRMPD data show that the dimers reflect the most abundant monomer species present and there is no preferential selectivity during dimer formation. Figure 29. Distribution of the naphthenic acids homologous series fragments produced from the dissociation of the molecular ion C32H55O4 - (z=-8) using IRMPD Next on, IRMPD was carried out on two less saturated precursor ions: C32H53O4 - (m/z=501.4, Z=-10) and C32H51O4 - (m/z=499.4, Z=-12) (Figure 30). By increasing the Z number (index of hydrogen deficiency), there are more possible combinations for the fragments to form. It appears to be a trend with respect to the relative intensities of the different homologous series as a function of the Z value of the selected dimer. A more negative Z value leads to Z series of monomers being more evenly spread, rather than one or two Z series predominating.   0.E+00   1.E+08   2.E+08   3.E+08   4.E+08   5.E+08   6.E+08   7.E+08   8.E+08   10   12   14   16   18   20   22   24   Intensity   Carbon  number   O2  distribution-­‐framentation  of  m/z  503.4,   Z=-­‐8    (IRMPD)   0   -­‐2   -­‐4   -­‐6   -­‐8  
  • 47. Results  and  discussion   43     Figure 30. Distribution of the naphthenic acids homologous series fragments produced from the dissociation of the molecular ion C32H53O4 - (z=-10) on the left graph and of the molecular ion C32H51O4 - (z=-12) on the right graph (IRMPD) The most negative Z value for the monomer region does not exceed the Z value of the dimer which proves that there is no additional fragmentation is occurring. 0.0E+00   2.0E+07   4.0E+07   6.0E+07   8.0E+07   1.0E+08   1.2E+08   1.4E+08   1.6E+08   1.8E+08   2.0E+08   10   12   14   16   18   20   22   Intensity   Carbon  number   O2  distribution  -­‐framentation  of   m/z  501.4,  Z=-­‐10  (  IRMPD)   0   -­‐2   -­‐4   -­‐6   -­‐8   -­‐10   0.0E+00   2.0E+07   4.0E+07   6.0E+07   8.0E+07   1.0E+08   1.2E+08   10   12   14   16   18   20   22   Intensity   Carbon  number   O2  distribution  -­‐framentation   of  m/z  499.4,  Z=-­‐12  (  IRMPD)   0   -­‐2   -­‐4   -­‐6   -­‐8   -­‐10   -­‐12  
  • 48. Results  and  discussion   44     3.6 Fresh  Kodak  NA  sample  doped  with  salts     It was noticed in previous experiments that salt affects the dimer formation. For further investigating of this matter, a fresh Kodak NA mixture was doped with sodium chloride. The oil industry has long used a technique of pumping seawater into an oil reservoir to increase the pressure and push oil toward producing wells. Understanding the ability of naphthenic acid to form sodium bond dimers after coming in contact with seawater will help understand deposition formation in industry. The average concentration of salt in seawater is about 35 g/L or 599mM, but because such high salt concentration would overwhelm the ESI process, the concentration of salt used in this study for doping the naphthenic acid samples was much lower, in the µM range. Nevertheless, for illustrative reasons, a 0.05 mg/mL Kodak NA sample was doped with enough sodium chloride to imitate the actual seawater linearity (left hand side vial in Figure 31). At seawater salt concentration level, there is a dramatic change in the appearance of naphthenic acid the sample when compared to a regular Kodak NA sample that has not been doped with salt (right hand side vial). Figure 31. Photography of a Kodak NA mixture (0.05 mg/mL) at seawater concentration of salt (600 mM) (left hand side) and of a Kodak NA mixture (0.05 mg/mL) with no salt added (right hand side)
  • 49. Results  and  discussion   45     A fresh 0.05 mg/mL Kodak NA sample that contained no added salts was used as a blank sample. Mass spectra zoom-ins on the NaO4/O4 pair of peaks from the blank Kodak NA sample and the three Kodak NA samples doped with sodium chloride are shown in Figure 32. Figure 32. Expanded mass range of the m/z 431.314 C25H44NaO4 - and of the m/z 431.317 C27H43O4 - . Increase in the concentration of salt, increases the relative intensity of NaO4 class The data analysis carried showed that NaO4 class species make up to 7.23% of the dimeric distribution in the NA blank sample (Table 7). As expected, by increasing the amount of salt added to the NA sample, there is an increase in the NaO4 contribution at the expense of dissociation of the weakly bound O4 species. Table 7. The relative abundance of the species present in the dimer distribution at none or different concentration of salt Species/Concentration salt (µM) 0 0.6 54.55 300 O4 90.63 87.98 86.89 82.73 NaO4 7.23 11.11 11.94 16.20 O4S 2.14 0.92 1.17 1.07
  • 50. Results  and  discussion   46     The relative abundance of the O4, NaO4 and O4S classes in the blank and in the salt doped Kodak NA samples was plotted as line graph to show the general trend (Figure 33). By increasing the salt concentration, there is a decrease in the contribution of the O4 class and an increase in the contribution of the NaO4 class. Adding enough NaCl so the salt concentration of the Kodak NA sample is 300 µM will increase the NaO4 class contribution to the dimer from ~7% (the natural state, when no salt is added) to about 16%. A 300 µg/mL salt concentration is 2000 times lower than the 600 mM seawater salinity. Figure 33. Relative abundance of O4, NaO4 and O4S species in the dimer distribution of Kodak NA mixture containing none or various concentration of salt            
  • 51. Results  and  discussion   47     3.7 EID  experiment  on  a  NIST  crude  oil  sample     As an additional line of work, further development of methods for understanding structures of petroleum molecules was pursued. Determining the elemental composition for petroleum samples has improved greatly the field of petroleomics, but understanding the structures of the tens of thousands of components present represents the next challenge. In this work, EID has been applied for the first time to crude oil samples. Firstly, a positive-ion mode ESI spectrum of a NIST crude oil sample was obtained (Figure 34). The multitude of peaks in the mass spectrum proves the complexity of petroleum samples. Figure 34. Mass spectrum of the NIST crude oil (positive mode ESI) The main contributor to the NIST crude oil mass spectrum is the N1 class (~66%), followed by OS class (~26%) (Table 8). Other species such as NO, NO2, O4S and OS2 are present as minor components.
  • 52. Results  and  discussion   48     Table 8. The relative abundance of the species present in the positive-mode ESI spectrum of NIST crude oil Species   Relative  abundance  (%)   N   66.35   NO   3.11   NO2   0.35   O4S   0.34   OS   25.54   OS2   1.18   For a better visualization, the values from Table 8 are plotted as a bar chart in Figure 35. Figure 35. The abundance of the species present in NIST crude oil (positive-ion conditions ESI) Because of the high content of N present in the petroleum sample and the known problems associated with the N-containing compounds in crude oil, such as pollution and catalyst poisoning, data analysis was focused on N class species. The DBE values of the components of N1 class were plotted against their carbon number as shown in the DBE plot from Figure 36. DBE values extend from 3.5 to 22.5 with, DBE between 7.5 and 9.5 being most abundant. 0   10   20   30   40   50   60   70   N   NO   NO2   O4S   OS   OS2   Relative  abundance  (%)   Species  present   NIST  crude  oil  composition  (positive  mode   ESI)  
  • 53. Results  and  discussion   49     Figure 36. DBE plot of the N1 class present in a NIST crude oil (positive-ion ESI) The isolation window was chosen to be 60 Da wide in order to accommodate five data points along homologues series. A DBE plot of the N1 class species from the isolation window is shown in Figure 37. The DBE values extend from 3.5 to 19.5 while the most abundant species have DBE values of 6.5 to 9.5. The carbon number ranged between 27 and 32. Figure 37 DBE plot of the N1 class for the isolation window spectrum A single EID spectrum (Figure 38) was obtained after one hour experimental time in order to increase the signal to noise ratio. It was determined that a 30 V cathode bias produces the 0 10 20 30 40 50 60 70 80 0 5 10 15 20 25 NIST crude oil N1 class DBE plot DBE Carbon number 0 5 10 15 20 25 30 35 0 2 4 6 8 10 12 14 16 18 20 22 Isolation N1 Class DBE plot DBE
  • 54. Results  and  discussion   50     most extensive fragmentation for the NIST crude oil sample. The cathode bias controls the energy of the electrons used for irradiation. Lower cathode biases were also tested, but it was noticed that at values higher than 20 V, the degree of fragmentation improved. Figure 38. EID spectrum of a NIST crude oil sample. The isolation window extends from m/z 375 to m/z 435 The DBE plot of the N1 class species from the EID spectrum (Figure 39) reveals that DBE values extend from 3.5 to 20.5, showing a similar distribution as the N1 class species from the isolation spectrum. When comparing the carbon number of N1 class species between the two spectra, it is clear that additional fragmentation occurs during EID experiments. Where the lowest carbon number during isolation was 27, the minimum carbon number after EID was 11. So, fragments smaller than the smallest intact components are seen.
  • 55. Results  and  discussion   51       Figure  39. DBE plot of the N1 class for the EID spectrum   Fragmentation causes successive loss of alkyl chains from the core (fused rings) rather than alkyl chains plus aromatic carbon, as proved by the horizontal lines to the left of the isolated peaks in the DBE plot. However, fragmentation does not cause a reduction in aromaticity. In this research, it was demonstrated that EID is a viable tandem mass spectrometry method for petroleum analysis. Through EID significant fragmentation was produced with no increase of pressure in the ICR cell. EID dissociates alkyl chain and thus it provides information about the core of molecules. It could potentially be used to distinguish archipelago structures (condensed aromatic cores connected by aliphatic chains) from island structures (one aromatic core with aliphatic chains). For archipelago structures, DBE plot will show stable structures favoured along the homologous series, while for island structures it will show loss of alkyl chains evenly as seen with the NIST crude oil sample. Therefore, it is possible to use EID on petroleum samples to remove alkyl chains and identify the minimum number of carbons needed for just the stable core for a particular class species.     0 5 10 15 20 25 30 35 2 4 6 8 10 12 14 16 18 20 22 DBE Carbon number EID-30 V cathode bias N1 class DBE plot
  • 56. Conclusion   52     4 Conclusion   In this work, negative-ion ESI FT-ICR mass spectrometry reveals two forms of naphthenic acid dimers to be present in a commercial naphthenic acid sample, instead of one as previously believed: the O4 class species that have been formerly described by other groups and the NaO4 class species. The power of the inherent ultrahigh resolving power and mass accuracy associated with FTICR-MS made it possible to resolve NaO4/O4 pair of peaks, separated by only 2.4 mDa.   Formation  of  naphthenic  acid  dimers  may  simply  be  the  route  that  naturally  occurs,  but   understanding  any  preferential  dimer  aggregation  would  help  understand  if  there  are   any  particular  culprits  or  whether  all  species  present  are  equally  guilty  for  corrosion  in   refineries   and   toxicity   to   aquatic   environment.   How easily naphthenic acids can form dimers will impact the behaviour of these NAs and the availability of those protons to have consequences for corrosion. Comparing a 2 year old Kodak naphthenic acid solution with a freshly prepared one, it was noticed that the dimers present depend on the solution age, and if a sample is preserved in glassware for long enough time, the sodium bound naphthenic acid will be more abundant (~8.5% contribution compared to ~2.8% contribution for the fresh sample). The O4 class compounds have been demonstrated to dissociate readily as they are bound by a weak noncovalent bond that breaks when increasing the collision power to higher voltages such as 60 V for ISD experiments and 12 V for CID experiments. On the other hand, the abundance of NaO4 class compounds remained relatively stable and it is believed that they are bound by much stronger bonds. A proposed structure is shown in Figure 40, where the sodium atom is shared between the four oxygen atoms.
  • 57. Conclusion   53     Figure 40. Possible molecular structure of the singly charged NaO4 class species present in naphthenic acid sample, with delocalized electron density between the carbon atom and the two oxygen atoms A fresh Kodak naphthenic acid solution was doped with NaCl in order to monitor the effect of sodium abundance upon the contribution of the NaO4 class species and to compare with the aged naphthenic acid sample. This is highly relevant to the extraction of oil from reservoirs process, where seawater is pumped down to help extract oil. Sodium bound dimers can form precipitates when seawater comes in contact with naphthenic acids, leading to naphthenates deposition inside pipelines. Using Kodak NA samples doped with sodium chloride, it was shown that the more sodium is available, the higher the contribution of the NaO4 class is. Doping the naphthenic acid sample with a salt concentration 2000 times lower than the actual seawater salt concentration, increased the NaO4 contribution to the dimer from ~7% to ~16%. Electron induced ionization was applied successfully for the first time to petroleum providing proof of concept data. In petroleomics, fs. The EID experiment carried on a crude oil sample demonstrated the great potential of EID MS/MS in structure characterization. Further work can include possible tandem mass spectrometry methods for fragmentation of species with tentative assignments such as the ones in distribution III and IV from the fresh Kodak NA sample. Additional experiments can be carried out on doping naphthenic acid samples with higher amounts of salt to mirror the actual salinity level of seawater and on study of real world sodium naphthenate deposits. Further EID experiments can be carried out using different cathode potentials on various class compounds from different petroleum samples in order to obtain greater insight into molecular structures, including information about favoured structures (archipelago versus island structures). EID MS/MS via FT-ICR MS could be coupled with chromatography for structure information of isomeric compounds from petroleum samples.
  • 58. References   54     5 References     1.   Cho,  Y.;  Ahmed,  A.;  Kim,  S.,  Application  of  Atmospheric  Pressure  Photo  Ionization   Hydrogen/Deuterium  Exchange  High-­‐Resolution  Mass  Spectrometry  for  the  Molecular   Level  Speciation  of  Nitrogen  Compounds  in  Heavy  Crude  Oils.  Anal.  Chem.  2013,  85  (20),   9758-­‐9763.   2.   Griffiths,  M.  T.;  Da  Campo,  R.;  O'Connor,  P.  B.;  Barrow,  M.  P.,  Throwing  light  on   petroleum:   simulated   exposure   of   crude   oil   to   sunlight   and   characterization   using   atmospheric  pressure  photoionization  fourier  transform  ion  cyclotron  resonance  mass   spectrometry.  Anal.  Chem.  2014,  86  (1),  527-­‐34.   3.   Colati,   K.   A.   P.;   Dalmaschio,   G.   P.;   De   Castro,   E.   V.   R.;   Gomes,   A.   O.;   Vaz,   B.   G.;   Romão,   W.,   Monitoring   the   liquid/liquid   extraction   of   naphthenic   acids   in   brazilian   crude  oil  using  electrospray  ionization  FT-­‐ICR  mass  spectrometry  (ESI  FT-­‐ICR  MS).  Fuel   2013,  108,  647-­‐655.   4.   Marshall,   A.   G.;   Rodgers,   R.   P.,   Petroleomics:   The   Next   Grand   Challenge   for   Chemical  Analysis.  Acc.  Chem.  Res.  2004,  37  (1),  53-­‐59.   5.   Boduszynski,   M.   M.,   Composition   of   heavy   petroleums.   1.   Molecular   weight,   hydrogen   deficiency,   and   heteroatom   concentration   as   a   function   of   atmospheric   equivalent  boiling  point  up  to  1400.degree.F  (760.degree.C).  Energy  Fuels  1987,  1  (1),  2-­‐ 11.   6.   Mapolelo,  M.  M.  Extraction,  Chromatographic  Separation  and  Characterization  of   Polar   Acidic   Species   in   Crude   Oils   and   Naphthenate   Deposits   by   Ultrahigh   Resolution   Fourier   Transform   Ion   Cyclotron   Resonance   Mass   Spectrometry.   Florida   State   University,  2010.   7.   Liu,   P.;   Shi,   Q.;   Chung,   K.   H.;   Zhang,   Y.;   Pan,   N.;   Zhao,   S.;   Xu,   C.,   Molecular   Characterization  of  Sulfur  Compounds  in  Venezuela  Crude  Oil  and  Its  SARA  Fractions  by   Electrospray  Ionization  Fourier  Transform  Ion  Cyclotron  Resonance  Mass  Spectrometry.   Energy  Fuels  2010,  24  (9),  5089-­‐5096.   8.   McKetta,   J.   J.,   Encyclopedia   of   Chemical   Processing   and   Design.   Marcel   Dekker:   United  States  of  America,  1990;  Vol.  35.   9.   Rodgers,   R.   P.,   Hughey,   C.   A.;   Hendrickson,   C.   L.;     Marshall,   A.   G.,   Advanced   Characterization  of  Petroleum  Crude  and  Products  by  High  Field  Fourier  Transform  Ion   Cyclotron  Resonance  Mass  Spectrometry.  Fuel  Chemistry  Division  Reprints  2002,  47  (2).   10.   Hell,   C.;   Medinger,   E.,   Ueber   das   Vorkommen   und   die   Zusammensetzung   von   Säuren  im  Rohpetroleum.  Ber.  Dtsch.  Chem.  Ges.  1874,  7  (2),  1216-­‐1223.   11.   Aschan,   O.,   Ueber   die   in   dem   Erdöl   aus   Baku   vorkommenden   Säuren   mit   niedrigerem  Kohlenstoffgehalt.  Ber.  Dtsch.  Chem.  Ges.  1891,  24  (2),  2710-­‐2724.   12.   Kannel,  P.  R.;  Gan,  T.  Y.,  Naphthenic  acids  degradation  and  toxicity  mitigation  in   tailings  wastewater  systems  and  aquatic  environments:  a  review.  J  Environ  Sci  Health  A   Tox  Hazard  Subst  Environ  Eng  2012,  47  (1),  1-­‐21.   13.   Scott,  A.  C.;  Mackinnon,  M.  D.;  Fedorak,  P.  M.,  Naphthenic  Acids  in  Athabasca  Oil   Sands   Tailings   Waters   Are   Less   Biodegradable   than   Commercial   Naphthenic   Acids.   Environ.  Sci.  Technol.  2005,  39  (21),  8388-­‐8394.   14.   Dzidic,   I.;   Somerville,   A.   C.;   Raia,   J.   C.;   Hart,   H.   V.,   Determination   of   naphthenic   acids  in  California  crudes  and  refinery  wastewaters  by  fluoride  ion  chemical  ionization   mass  spectrometry.  Anal.  Chem.  1988,  60  (13),  1318-­‐1323.  
  • 59. References   55     15.   Headley,   J.   V.;   Peru,   K.   M.;   McMartin,   D.   W.;   Winkler,   M.,   Determination   of   dissolved  naphthenic  acids  in  natural  waters  by  using  negative-­‐ion  electrospray  mass   spectrometry.  J  AOAC  Int  2002,  85  (1),  182-­‐7.   16.   Qian,  K.;  Robbins,  W.  K.;  Hughey,  C.  A.;  Cooper,  H.  J.;  Rodgers,  R.  P.;  Marshall,  A.  G.,   Resolution   and   Identification   of   Elemental   Compositions   for   More   than   3000   Crude   Acids   in   Heavy   Petroleum   by   Negative-­‐Ion   Microelectrospray   High-­‐Field   Fourier   Transform   Ion   Cyclotron   Resonance   Mass   Spectrometry.   Energy   Fuels   2001,   15   (6),   1505-­‐1511.   17.   Barrow,  M.  P.;  McDonnell,  L.  A.;  Feng,  X.;  Walker,  J.;  Derrick,  P.  J.,  Determination  of   the  nature  of  naphthenic  acids  present  in  crude  oils  using  nanospray  Fourier  transform   ion   cyclotron   resonance   mass   spectrometry:   the   continued   battle   against   corrosion.   Anal.  Chem.  2003,  75  (4),  860-­‐6.   18.   Simon,   S.;   Reisen,   C.;   Bersås,   A.;   Sjöblom,   J.,   Reaction   between   tetrameric   acids   and  Ca  2+  in  oil/water  system.  Ind.  Eng.  Chem.  Res.  2012,  51  (16),  5669-­‐5676.   19.   Scaled   Solutions   LTD,   Organic   Deposits-­‐   Napthenates   http://www.scaledsolutions.com/Performance%20Tests/Organic%20Deposits   (accessed  18  Aug).   20.   El   Mahdy,   G.   A.;   Atta,   A.   M.;   Dyab,   A.   K.   F.;   Al-­‐Lohedan,   H.   A.,   Protection   of   Petroleum   Pipeline   Carbon   Steel   Alloys   with   New   Modified   Core-­‐Shell   Magnetite   Nanogel  against  Corrosion  in  Acidic  Medium.  J.  Chem.  2013,  2013,  9.   21.   Slavcheva,  E.;  Shone,  B.;  Turnbull,  A.,  Review  of  naphthenic  acid  corrosion  in  oil   refining.  Br.  Corrros.  J  1999,  34  (2),  125-­‐131.   22.   Qu,   D.   R.;   Zheng,   Y.   G.;   Jiang,   X.;   Ke,   W.,   Correlation   between   the   corrosivity   of   naphthenic  acids  and  their  chemical  structures.  Anti-­‐Corros  Method  M  2007,  54  (4),  211-­‐ 218.   23.   Clemente,  J.  S.;  Fedorak,  P.  M.,  A  review  of  the  occurrence,  analyses,  toxicity,  and   biodegradation  of  naphthenic  acids.  Chemosphere  2005,  60  (5),  585-­‐600.   24.   GE-­‐Measurement   and   control,   Naphthenic   Acid   Corrosion.   http://www.ge-­‐ mcs.com/en/ndt-­‐corrosion-­‐inspection/naphthenic-­‐acid-­‐corrosion.html   (accessed   17   Aug).   25.   Wang,   Y.;   Zhou,   Q.;   Peng,   S.;   Ma   Lena,   Q.;   Niu,   X.,   Toxic   effects   of   crude-­‐oil-­‐ contaminated   soil   in   aquatic   environment   on   Carassius   auratus   and   their   hepatic   antioxidant  defense  system.  J  Environ  Sci  2009,  21  (5),  612-­‐617.   26.   Mohamed,  M.  H.;  Wilson,  L.  D.;  Peru,  K.  M.;  Headley,  J.  V.,  Colloidal  properties  of   single   component   naphthenic   acids   and   complex   naphthenic   acid   mixtures.   J.   Colloid   Interface  Sci.  2013,  395  (1),  104-­‐110.   27.   Yen,  T.  W.;  Marsh,  W.  P.;  MacKinnon,  M.  D.;  Fedorak,  P.  M.,  Measuring  naphthenic   acids   concentrations   in   aqueous   environmental   samples   by   liquid   chromatography.   Journal  Chromatogr.  A  2004,  1033  (1),  83-­‐90.   28.   Hemmingsen,  P.  V.;  Kim,  S.;  Pettersen,  H.  E.;  Rodgers,  R.  P.;  Sjöblom,  J.;  Marshall,  A.   G.,   Structural   characterization   and   interfacial   behavior   of   acidic   compounds   extracted   from  a  North  Sea  oil.  Energy  Fuels  2006,  20  (5),  1980-­‐1987.   29.   Smith,  D.  F.;  Schaub,  T.  M.;  Rahimi,  P.;  Teclemariam,  A.;  Rodgers,  R.  P.;  Marshall,  A.   G.,  Self-­‐association  of  organic  acids  in  petroleum  and  Canadian  bitumen  characterized  by   low-­‐  and  high-­‐resolution  mass  spectrometry.  Energy  Fuels  2007,  21  (3),  1309-­‐1316.   30.   Da   Campo,   R.;   Barrow,   M.   P.;   Shepherd,   A.   G.;   Salisbury,   M.;   Derrick,   P.   J.,   Characterization   of   Naphthenic   Acid   Singly   Charged   Noncovalent   Dimers   and   Their   Dependence   on   the   Accumulation   Time   within   a   Hexapole   in   Fourier   Transform   Ion   Cyclotron  Resonance  Mass  Spectrometry.  Energy  Fuels  2009,  23  (11),  5544-­‐5549.