Tải bản đầy đủ (.pdf) (11 trang)

Precursor ion approach for simultaneous determination of nonethoxylated and ethoxylated alkylsulfate surfactants

Bạn đang xem bản rút gọn của tài liệu. Xem và tải ngay bản đầy đủ của tài liệu tại đây (2.39 MB, 11 trang )

Journal of Chromatography A 1653 (2021) 462421

Contents lists available at ScienceDirect

Journal of Chromatography A
journal homepage: www.elsevier.com/locate/chroma

Precursor ion approach for simultaneous determination of
nonethoxylated and ethoxylated alkylsulfate surfactants
Katarzyna Pawlak a, Kamil Wojciechowski a,b,∗
a
b

Faculty of Chemistry, Warsaw University of Technology, Noakowskiego 3, Warsaw 00-664, Poland
SaponLabs Ltd, Noakowskiego 3, Warsaw 00-664, Poland

a r t i c l e

i n f o

Article history:
Received 29 March 2021
Revised 16 July 2021
Accepted 16 July 2021
Available online 22 July 2021

a b s t r a c t
We present a new liquid chromatography–tandem mass spectrometry (LC-MS/MS) method for simultaneous determination of sodium lauryl sulfate and sodium laureth sulfate homologues in the range of alkyl
chain length C12 –C16 with 0–5 ethoxy groups. The method is based on scanning the precursor ions fragmenting to m/z 80 and 97 (Precursor Ion Scanning mode), which makes it specific for species with easily
cleavable sulfate groups. By monitoring fragmentation of thus discovered quasi-molecular ions we were
able to unequivocally identify all sulfate species present in complex mixtures of alkyl and alkyl-ether


sulfates with molecular weight ranging from 200 to 600 m/z. Because of the intrinsic sulfate-sensitivity,
the presented method can be also applied to non-sodium salts of alkyl- and alkyl-ether sulfates (e.g.
ammonium, mono- or triethanolamine, etc.), which are often used by cosmetic manufacturers to justify
the misleading SLS- and SLES-free claims (where SLS and SLES refer to sodium lauryl sulfate and sodium
laureth sulfate, respectively). The use of reversed phase liquid chromatography (RPLC) column with C4
instead of C18 shortened significantly the overall analysis time and allowed us to use a semiquantitative
method (based on single standard for Quantitative Analysis of Multi-component System, QAMS) to determine several SLS and SLES homologues in one run with the limit of quantification (LOQ) = 0.4 μg/mL and
of detection (LOD) in the range 0.12–0.97 μg/mL. The method was successfully applied to 17 commercially
available cosmetic/household products allowing verification of their manufacturers’ declarations.
© 2021 The Author(s). Published by Elsevier B.V.
This is an open access article under the CC BY license ( />
1. Introduction
Anionic surfactants are major constituents of most detergents
and cosmetic cleaning/washing products but are also ubiquitous in
many other formulations, where wetting, dispersing, emulsifying
or foaming activities are required [1]. Most of the currently available shampoos, shower gels, liquid soaps and dishwashing liquids
are based on alkyl and alkyl-ether sulfates produced by sulfonation of alkyl alcohols and ethoxylated alkyl alcohols. In industrial
practice pure alcohols are very rarely used, typically their mixtures
are employed instead (e.g. a mixture obtained from hydrolyzed and
hydrogenated coconut and palm oils, as in the case of the most
popular alkyl sulfate – sodium lauryl sulfate (SLS)). Its ethoxylated
analogue (sodium laureth sulfate, SLES) is produced analogously,
with an additional intermediate alcohol ethoxylation step [2]. The


Corresponding author at: Faculty of Chemistry, Warsaw University of Technology, Noakowskiego 3, Warsaw 00-664, Poland.
E-mail
addresses:

(K.

Pawlak),
(K. Wojciechowski).

“lauryl” name and its derivative “laureth” were introduced to highlight the ill-defined structures of the resulting mixtures, in contrast
to single species, e.g. sodium dodecylsulfate (SDS). The ethoxylation step introduces another heterogeneity of chemical structures
of SLES (number of the ethoxy units), in addition to the variable
length of the alkyl chain present in SLS. Consequently, the commercially available alkyl and alkyl-ether sulfates show wide distribution of chemical structures and surfactant properties [3,4]. Despite their great efficacity in lowering surface tension and sustaining foams, SLS and SLES may pose some environmental hazards
caused by their limited biodegradability and persistent foam formation [5]. Their high detergent power may also lead to excessive lipid and protein removal when used on a daily basis [6–8].
This prompts some consumers to avoid products with sodium lauryl and laureth sulfates and look for their “milder” alternatives. In
response, numerous cosmetic producers replace the sodium salts
with those of ammonium, lithium, ethanolamine, etc., or replace
SLS and SLES with their homologous mixtures (e.g. “sodium coco
sulfate”) to misguide the consumer with a different INCI (International Nomenclature of Cosmetic Ingredients) name [9]. Given

/>0021-9673/© 2021 The Author(s). Published by Elsevier B.V. This is an open access article under the CC BY license ( />

K. Pawlak and K. Wojciechowski

Journal of Chromatography A 1653 (2021) 462421

the variety of homologous alkyl and alkyl-ether sulfate surfactants, there is an urgent need for a robust and universal analytical
method capable of differentiation between not only the alkyl chain
length but also the extent of ethoxylation of SLES.
One of the major obstacles in quantitative analysis of complex
mixtures of natural or synthetic compounds is the lack of analytical standards. In contrast to many well-defined compounds with
distinct structures, no reference materials are available for most of
the complex molecules (like proteins, lipids or alkaloids) or mixtures of homologues (such as alkyl and alkyl-ether sulfates in the
present case). This problem can be circumvented using computational methods based on Quantitative Structure and Ionization Intensity Relationship (QSIIR). The method was successfully applied
to predict relative levels of 29 organic acids in complex matrices,
registered by product ion scanning. It offered accurate results (in

the range of 80–120%) for 16 organic acids for which the absolute
concentrations were quantified and used as reference. For the remaining organic acids, such accuracy could not be achieved due to
lack of standards or too low concentration of the acid in the samples [10]. Such approach enables development of methods relying
on a single standard for Quantitative es of Multi-component System (QAMS) [11,12].
Currently, most analytical methods for determination of alkyl
and alkyl-ether sulfate are based on spectrophotometric, electrochemical and chromatographic techniques [13–15]. The official EU
method for determination of anionic surfactants in detergents described in Regulation (EC) No. 648/2004 of the European Parliament and of the Council and in ISO 7875-1(1996) standard is based
on formation of blue-colored salts of anionic surfactants with
the methylene blue dye, which are determined spectrophotometrically after extraction to chloroform (Methylene Blue Active Substance test, MBAS). Although the method is continuously improved
[16,17], the ion-pair formation reaction does not provide sufficient
selectivity to enable distinction not only between the nature of the
anionic group (phosphate, carboxylic, sulfate or sulfonate) but also
the alkyl chain structure.
Another group of non-chain-length-selective methods is based
on formation of ionic associates between anionic surfactants and
cationic species using potentiometric sensors [18] or suppression
of ionic conductivity using ion chromatography [19]. Using a similar approach, Levine et al. were capable of determining concentration of ammonium lauryl sulfate, sodium laureth sulfate and
sodium alkyl (C10 SO4 − – C16 SO4 − ) ether sulfates present in commercially available detergents (dynamic linear ranges: 1.0–500,
2.5–550 and 3.0–630 mg/L, respectively) [20]. After chromatographic separation, a mixture of four linear alkylbenzene sulfonates
(C10 SO4 − - C13 SO4 − ) was detected by UV spectroscopy [21] and a
mixture of anionic and nonionic surfactants (including sodium lauryl sulfate and α -olefin sulfonate) - using an evaporative light scattering detector (ELSD) [22]. Nevertheless, selectivity is often when
such detectors are employed [21,23,24].
Regardless of the detection method, anionic surfactants can be
separated using ion exchange chromatography [19]. However, most
commonly the separation is achieved thanks to differences in the
affinity of aliphatic chains to a hydrophobic stationary phase. For
this purpose, deprotonated anionic surfactants (pH > 7) can be
separated as neutral ion pairs formed with quaternary ammonium cations (ion pair liquid chromatography, IPLC [25]) or in a
non-dissociated form (pH < 7), using reversed-phase liquid chromatography, RPLC. The former method is, however, not suitable
for mass spectrometry because of significant signal suppression

[26]. The advantage of lowering pH of the mobile phase is a reduction of hydrophilic interactions interfering with the chromatographic process. On the other hand, reduced pH may lower ionization efficiency in mass spectrometry [26]. Matthus and colleagues
[27] separated the alkylbenzene sulfonate (LAS), alcohol ethoxy-

lates (AE), and alcohol ethoxylated sulfates (AES) by reversedphase chromatography using a C18 column and ammonium acetate
to stabilize pH of the mobile phase. The partially separated compounds were detected using a fluorescence detection (FLD) and
a thermospray mass spectrometer working in the scanning mode
(m/z range 20 0–80 0). This provided more detailed structural information at the expense of sensitivity. The potential for quantitative
analy was exemplified by determination of the total concentration
of active substances in sewage in relation to the AES and LAS commercial mixtures. Dufour et al. employed ultra-high-performance
liquid chromatography with high-resolution mass spectrometry
(UPLC-HR-MS) to separate four homologues of alkylbenzene sulfonate (4-dodecylbenzenesulfonic acid, DBSA). The homologues differed in their alkyl chain lengths (decylbenzenesulfonate, undecylbenzenesulfonate, dodecylbenzenesulfonate, tridecylbenzenesulfonate) and could be separated in the RPLC mode using C4, C18
and C30 columns [28]. Levine et al. employed RPLC coupled with
electrospray ionization quadrupole ion trap mass spectrometry
(ESI-Q-IT-MS) to simultaneously determine three common surfactants: an amphoteric cocoamphoacetate, a nonionic alcohol ethoxylate and an anionic SLES (dynamic linear range 1.5–40 mg/L for total amount of SLES normalized against commercial mixture) [29].
Four alkyl sulfate and 2 ethoxymers of alkyl-ether sulfate homologues were determined in the SPE-preconcentrated wastewater
samples using a liquid chromatography–tandem mass spectrometry (LC–MS/MS) with electrospray ionization (ESI) in negative ion
mode. Based on the commercial mixture producer’s declarations,
the limits of quantification (LOQ) in the range 0.3–0.4 μg/L and
0.5–1.5 μg/L were reported for the SLS and SLES homologues, respectively [30]. Another interesting approach to separate surfactants present in commercial mixtures of SLS and SLES employed
the ion mobility mass spectrometry [31]. The ionized compounds
were distinguished based on their drift time in an electric field,
which depended on their molecular weight. The authors obtained
six peaks for the SLS mixture and twelve for the SLES one, although their identity was not determined due to lack of standards.
Nevertheless, the developed method could be applied to determine
the total SLS and SLES content (with respect to commercially available SLS and SLES mixtures) adsorbed by different dish surfaces.
In this contribution we further extend the analytical capabilities of LC-MS/MS technique by employing for the first time
an MS/MS scanning of the precursor ions to follow the sulfatebearing species. We propose a novel method for determination
of individual alkyl and alkyl-ether sulfates in mixtures of cosmetic/household ingredients and products using a single and easily
available standard (SDS) for quantitative analy of multi-component
system (QAMS). We show how the experimentally observed dependency of signal intensity on the retention, number of fragmentation ions obtained in a collision cell and recovery from a stationary

phase can be accounted for in QAMS methods to obtain good accuracy. SDS could be employed as a universal and sole standard, as
it was present in every sample. The new method allowed us to detect 5 non-ethoxylated and 20 ethoxylated sulfates in commercial
SLES products. To show the unprecedented application potential of
the new method for real-life samples analy we verified the manufacturers’ declarations about the presence/absence of SLS and SLES
ingredients in 17 cosmetic/household commercial products.
2. Experimental
Separation of the alkyl and alkyl ether sulfates was carried out
using a 1220 Infinity II LC Systems (Agilent Technologies, USA),
whereas their identification and quantitation was achieved with a
6460 Triple Quad tandem mass spectrometric detector with a Jet
Stream ion source (Agilent Technologies, USA). Analytes were separated by an Aeris Widepore C4 column (2.1 × 150 mm, 3.6 μm,
2


K. Pawlak and K. Wojciechowski

Journal of Chromatography A 1653 (2021) 462421

Fig. 1. Product ion mass spectra registered for dodecylsulfate anion (DS− , C12 SO4 − , m/z 265) and a selected anion present in SLES with their fragmentation ions.

˚ Phenomenex). The chromatographic separation employed a
300 A,
gradient elution from (40% acetonitrile: 60% water) to (95% acetonitrile: 5% water), both containing 0.15% (v/v) formic acid, over
8 min at a flow rate of 0.2 mL/min. Both the column and mobile
phase were thermostated at 45 °C.
The mass spectrometer was operated in negative ion mode established by ionization voltage of 30 0 0 V and nozzle voltage 0 V.
Heated (300 °C) nebulization gas flow of 8 L/min applied at 30 psi
was selected as the most appropriate to enhance the ionization of
low molecular weight compounds. The Precursor Ion (PrecI) scanning was applied for the discovery of SLS and SLES homologues. A
tandem mass spectrometer operating in the PrecI mode finds the

parent ions (the first quadrupole, Q1, operates in the m/z 20 0–80 0
scanning mode) for the m/z values of characteristic fragment ions
(the second quadrupole, Q3, operates in the selected ion monitoring mode). The negative fragmentation ions SO3 − at m/z 80 and
HSO4 − at m/z 97 were selected as characteristic of the sulfatebearing compounds (Fig. 1, Table SM1). Structurally similar sulfonate ions would produce the m/z 81 signals of HSO3 − ions in
addition to m/z 80. The m/z values of the parent ions were selectively discovered by the mass spectrometer. The analyses were
controlled and processed by a MassHunter Workstation software
(Agilent Technologies, USA). The employed chromatographic and
MS conditions are collected in Table SM1.
The commercial shampoos, hair conditioner and liquid soap
were purchased in a local cosmetic store in Warsaw (Poland). The
reference shampoo without SLS and SLES was provided by Saponlabs Ltd. (Poland). Three SLES and four SLS-type mixtures with the
specified amount of active substances (27–70% w/w) and ethoxylation degree were obtained from an industrial supplier. All SLES
were declared as sodium salts of ethoxylated sulfates of predominantly C12 -C14 alcohols with different ethoxylation degree and
average molecular weights (Mav ∼
= 340 g/mol, ethoxylation degree 1–2.5: SLES-340; Mav ∼
= 384 g/mol, ethoxylation degree 1–
2.5: SLES-384; Mav ∼
= 432 g/mol, ethoxylation degree > 2.5: SLES432). The following SLS-type products were used: sodium lauryl sulfate (SLS), ammonium lauryl sulfate (ALS), triethanolamine
lauryl sulfate (TEALS), monoethanolamine lauryl sulfate (MEALS).
Sodium dodecylsulfate, SDS (puriss ACS reagent, ≥ 99.0%) was pur-

chased from Sigma Aldrich (Poland). Acetonitrile (LC/MS purity)
from POCH (Gliwice, Poland); formic acid (LC/MS purity), from
Fisher Scientific (Fair Lawn, NJ, USA). Demineralized water from a
Milli-Q system Model Millipore Elix 3 (Molsheim, France) was employed.
Standard solution of SDS (1.0 mg/mL) was prepared in Milli Qwater. The surfactant SLES mixtures (0.2–0.5 mg/mL) and commercial cosmetic formulations (0.7–1.4 mg of cosmetic/mL) were prepared by dissolving them in Milli Q-water. For semi-quantitative
analyzes of alkyl and alkyl-ethoxy sulfates, to assure that the concentration of all sulfates is within the linear response range of calibration curve, all samples were diluted at two or more levels. All
solutions were filtered through 0.45 μm syringe filters prior RPLC
analy.
3. Results and discussion

3.1. Optimization of the detection conditions
In order to simultaneously quantify all alkyl and alkyl-ether sulfate homologues, the separation method should allow for selection of all species yielding hydrogen sulfate ion (HSO4 − , m/z = 97)
and/or radical sulfate anion (•SO3 − , m/z = 80) upon fragmentation. These species, resulting from dissociation of the C-O-S bond
in the sulfate group, can be conveniently selected by scanning ions
with m/z being reduced to 80 and 97 in the Precursor Ion Scanning
(PrecI) Mode. Although the latter offers additional intrinsic specificity in MS analy, until now the PrecI mode has been employed
almost solely in lipidomic and proteomics analyses [32] and its
potential in alkyl sulfate analy has been largely unexploited. Therefore, in the first part of the study, the detection conditions for electrospray tandem mass spectrometer (ESI-MS/MS) were optimized
using sodium dodecylsulfate (SDS) and a cosmetic/household ingredient Sodium Laureth Sulfate with declared average molecular
weight of 384 g/mol and average ethoxylation degree 1–2.5 (SLES384). To this aim, an SDS solution of 1 μg/ml was introduced
(5 μL) into the mobile phase stream (acetonitrile:water, 50:50
(v/v)) at a flow rate of 0.2 ml/min (FIA-ESI-MS). Using SDS, the
ionization voltage, gas flow rate and temperature, as well as ion
3


K. Pawlak and K. Wojciechowski

Journal of Chromatography A 1653 (2021) 462421

transmission voltage (fragmentor voltage) providing the highest
signals observed at m/z 265 (corresponding to dodecyl sulfate anion, CH3 -(CH2 )11 -SO4 − , (“C12 SO4 − ” or “DS−“ ), were optimized (see
Table SM1 in Supplementary Materials). As alkyl and alkyl-ether
sulfates tend to form stable anions also in acidic solutions due to
high dissociation constant of sulfate group (pKa = ~2.4) the mobile phase was acidified with formic acid (0.15% (v/v)) to sensitivity by elimination of adducts formation and to reduce the noise of
the mass spectrum [26]. Analogous analy employing the optimized
conditions was performed for SLES-384, where the highest signals
were observed at m/z 265, 309, 381 and 441. They corresponded
to the C12 anion (CH3 -(CH2 )11 -SO4 − ) previously found in SDS)
and to different laureth sulfate anions predominant in the mixture: CH3 -(CH2 )11 -(OCH2 )1 -SO4 − , CH3 -(CH2 )13 -(OCH2 )2 -SO4 − , CH3 (CH2 )11 -(OCH2 )4 -SO4 − , respectively (C12 EO1 SO4 − , C14 EO2 SO4 − and

C12 EO4 SO4 − , respectively). The selected dodecyl and laureth sulfate
ions were subjected to fragmentation (Product Ion Scanning) using
the 10, 20, 30, 40 eV collision energy. All resulting spectra featured
the m/z 97 signal, corresponding to HSO4 − anion, confirming that
all selected ions indeed contained the sulfate group. In some cases,
additional signal at m/z 80, corresponding to a radical anion SO3 − ,
was also observed. Other signals, if present, corresponded to the
breakdown of the C-O or C-C bonds within the ethoxylated part of
SLES molecules (Fig. 1). The highest signals were obtained for the
collision energy of 20 and 40 eV, and these conditions were employed in subsequent optimization of separation conditions.

temperature of the mobile phase and column, sample volume and
sample dilution (Fig. SM2).
In order to validate selectivity of the MS detection in PrecI
mode under optimized conditions, two other SLES mixtures (SLES340 with declared average molecular weight of 340 g/mol and average ethoxylation degree 1–2.5, and SLES-432 with declared average molecular weight of 432 g/mol and average ethoxylation
degree > 2.5) were analyzed using the method developed with
SLES-384. The chemical identity of the anions discovered by PrecI
was assigned using the Product Ion Scanning mode. Fragmentation
ions m/z 80 and 97, confirmed the presence of the sulfate group
in all cases. The difference between theoretical and experimentally established monoisotopic mass - M parameter (defined as
M =|Mtheoretical -Mexperimental |/Mtheoretical ·106 ) varied between 71
and 639 ppm (Table SM2), which is typical for low-molecularweight-compounds analyzed with quadrupole analyzers, due to the
intrinsic low resolution of the latter [36].
All tested SLES mixtures show similar chromatograms with the
proportion of more lipophilic derivatives increasing in order: SLES340 < SLES-384 < SLES-432. A more detailed analy showed that
retention time with the number of both alkyl and alkoxyl groups
(Fig. SM3), which was used to additionally confirm identity of the
homologues. This was especially useful when the compound signal was too small to obtain a rich Product Ion spectrum. All tested
SLES samples were abundant in ethoxylated species spanning the
alkyl chain lengths from 12 to 16 carbon atoms, and the number of

ethoxy groups from 1 to 5 (Fig. 2, Table SM2). Surprisingly, however, all samples featured also signals that could be assigned to
non-ethoxylated sulfates (C12 SO4 − - C16 SO4 − ), typical for SLS-type
products.

3.2. Optimization of the separation conditions by RPLC
Having selected the optimum conditions for detecting the
species releasing SO3 − and HSO4 − (m/z 80 and 97) in PrecI mode,
we optimized the chromatographic method, starting with a selection of mobile and stationary phases. Formic acid allowed separation of SLS and SLES homologues with a selectivity comparable to that achieved in the presence of trifluoroacetic acid (TFA)
in the mobile phase and better than that for ammonium acetate
[23,24]. The presence of formic acid in the mobile phase helps to
reduce the hydrophilic and enhance the hydrophobic interactions
of separated compounds with the alkylated silica stationary phase.
It should be noted that formic acid, in contrast to TFA, is not an
ion-pairing agent, hence its presence does not deteriorate the sensitivity of MS-based method. The ability of the mobile phase (0.15%
aqueous solution of formic acid with a linearly increasing amount
of acetonitrile from 0 to 98% in 20 min) to recover SDS and the
SLES-384 components from the surface of a hydrophobic stationary
phase was tested using two Phenomenex 150 × 2,1 mm columns.
One was loaded with fully porous silica particles modified with
C18 aliphatic chains (Luna) and the other - with core-shell silica
particles of wide pores modified with C4 aliphatic chains (Aeris).
Chromatographic performance during gradient elution was better
for the C4-bed core-shell column which produced significantly better shape of the peaks (intensity 3 to 10 times higher and lower
width, see Fig. SM1). The shape of the peak can be by: (1) too
strong interactions of the compound with the stationary phase not
balanced by the composition of the mobile phase solution, (2) the
rate of movement of the compounds in the column depending on
the temperature, (3) particle or pore size in the stationary phase
[33–35]. Considering that the column size, particle size and gradient elution method were the same in both cases, the observed reduction of the peak width must be related to the size of the particles’ pores. Due to better chromatographic efficiency (peaks’ shape)
without selectivity in comparison to the fully porous C18 column,

the core-shell C4 column was chosen for all subsequent experiments. Then, optimization of the chromatographic separation process was carried out by examining the gradient elution program,

3.3. LC-MS calibration using SDS
In order to allow for a semi-quantitative analy of the alkyl and
alkyl-ether sulfates in cosmetic/household ingredients and products, the LC-MS system was calibrated using SDS as a standard
(note that in contrast to SLES, SDS is a single molecule with known
and defined structure). As shown in Fig. 3 (inset), the calibration
curve for the m/z 265 signal in PrecI mode is linear up to at
least 10 μg/mL of SDS, confirming that the proposed method can
be used for quantitative determination of individual alkyl sulfates.
Nevertheless, an analogous quantitative analy of alkyl-ether sulfates (SLES-type) is more demanding for two reasons: (1) pure reference substances are not easily available, (2) the homologues may
differ not only in the alkyl chain length (Cn ) but also in ethoxylation degree (EOm ). Especially the latter fact may complicate quantitative analy since the alkyl and ethoxy chains contribute differently
to partitioning and fragmentation behavior of SLES molecules (see
Fig. SM3).
As the concentrations of individual components of the analyzed
SLES mixtures were not known, calibration was performed by a
series of dilutions to obtain concentrations ranging from 0.1 to
10 μg of raw material/SLES mixture in 1 ml of water. SLES-432
was used for this purpose, as it contains the highest number of
different SLS- and SLES-type molecules (Fig. 2). The diluted solutions were analyzed using the protocol described above (PrecI
mode). For each compound detected in the mixture, a good correlation was achieved between the peak area and the amount of
SLES-432 in solution (Fig. 3, Table SM3). For comparison, also the
curve for SDS is included in the same graph (m/z 265). The slope
of the resulting curves is proportional to the retention time for all
compounds detected in SLES-432, which is clearly a consequence
of different contributions of the -CH2 - and -OC2 H5 - groups to the
solubility and consequently – to the partition coefficient of different SLS and SLES homologues. The observed reduction of the sen4


K. Pawlak and K. Wojciechowski


Journal of Chromatography A 1653 (2021) 462421

Fig. 2. Extracted ion chromatograms (EIC) of SLES-384 (A) and SLES-432 (B) obtained for the precursor ions (indicated in A) discovered for the fragmentation ion m/z 97
(PrecI: ∗ ∗ →97 (40 eV)), see Table SM1 for experimental details). The colors visible in the on-line version correspond to the calibration curves in Fig. 3.

sitivity for the components eluted after longer time may be caused
by several reasons:





ture was selected for its highest number of detected SLS and SLES
homologues) were acquired by serial dilutions. The dependence of
signal height on concentration was then established using a linear
curve y = ax + b, where y is the peak area and x is the concentration of the sample (ingredient, cosmetic product, etc.) in the analyzed solution (expressed in μg/mL, see Fig. SM4). The sensitivity
coefficient, fi, for any given alkyl or alkyl-ether sulfate ion (i) in the
sample is calculated by normalizing the slope (a) for the i-th ion
to that for DS− (Eq. (1)):

reduced chromatographic recovery of analytes [37]
lower ion transport efficiency from the ionization to the vacuum chamber for higher molecular weight ions [38]
suppression of ion signal intensity by strong ion paring agents
and lower stability of ions in ESI chamber or collision cell for
highly ethoxylated homologues [26].

Without the use of analytical standards, the observed dependence of sensitivity on retention time excludes a direct quantitative determination of multiple components. To circumvent this
problem, we propose a simple correction scheme using the calibration data for SDS. As shown in Fig. 2, the latter is present in
all tested SLES mixtures, and its amount can be reliably quantified using the calibration curve from Fig. 3. Note that in a more

general case, for samples devoid of SDS, it can always be added
to the sample in known amount as an internal standard. In the
proposed scheme, each anion signal is normalized to that of DS−
(dodecylsulfate, C12 SO4 − ). First, a calibration curve for DS− anion
was obtained using solutions with the SDS standard. Next, changes
of signals for each ion detected in the mixture SLES-482 (this mix-

fi =

ai, sample
aDS, sample

(1)

The dependence of fi on retention time is shown in Fig. 4 and the
numerical values are collected in Table SM4. The retention time
significantly with the length of the alkyl chain for both SLS and
SLES. The sensitivity of the method is primarily dependent on the
number of -CH2 - groups in the SLS homologues and ethoxy groups
present in the SLES homologues. It has already been reported that
the length of alkyl chain influences the chromatographic recovery
[37] as well as the efficiency of ionization process and ion transmission from the electrospray to vacuum chamber [38]. However,
for SLES homologues, the decrease in sensitivity is more notable.
5


K. Pawlak and K. Wojciechowski

Journal of Chromatography A 1653 (2021) 462421


Fig. 3. Dilution correlation for SLES-432 obtained by LC-MS/MS method using Precursor Ion Scanning mode. Colors of points and trend lines in an electronic version
correspond to the peaks in Fig. 2. The inset shows a calibration graph for DS− obtained for SDS.

Fig. 4. Dependence of sensitivity coefficient on retention time depending on the number of -CH2 - (n) and ethoxy groups (m) in different SLS (Cn ) and SLES (Cn EOm SO4 − )
derivatives.

6


K. Pawlak and K. Wojciechowski

Journal of Chromatography A 1653 (2021) 462421

Table 1
Comparison of the declared and semi-quantitatively (LC-MS/MS) determined amount of active ingredients and the average molecular weight for 7 cosmetic/household products ingredients.
Product name

SLES-340
SLES-384
SLES-432
ALS
SLS
MEALS
TEALS

Sum of active ingredients

Average molecular weight and ethoxylation degree (in brackets)

Declared by manufacturer (%)


Determined by LC-MS/MS (%)

69.1
70.0
69.8
27.6
29.0
27.3
38.8

60.1
67.9
68.4
29.1
31.1
21.1
29.5

±
±
±
±
±
±
±

9.0
10.2
13.7

5.8
6.2
5.2
7.1

This is most likely related to the number of fragmentation products
for SLES homologues containing higher number of ethoxy groups,
which lowers the impact of the m/z 97 ion (Fig. 1). The change of
sensitivity with the number of ethoxy groups is linear for the same
length of alkyl chain and consequently a linear regression could be
applied. The latter provides a convenient statistical description of
the agreement between the fit and the experimental data, allowing
calculations of standard deviation of relative concentrations of SLS
and SLES homologues. The concentration of each surfactant ion (i)
can be then estimated using the sensitivity coefficient (Eq. (2)):

Ci =

( peak area )i − bDS, SDS
aDS, SDS · fi

·d

SDb,DS,SDS
· fi
aDS,SDS

Determined by LC-MS/MS

340 (1–2.5)

384 (1–2.5)
432 (> 2.5)
294
296
334
415

344
387
400
289
296
333
425

±
±
±
±
±
±
±

29 (1.1 ± 0.2)
35 (2.1 ± 0.2)
52 (2.5 ± 0.3)
14
11
58
83


ALS, TEALS and MEALS) are devoid of any ethoxylated impurities
(Fig. 5B) as expected, the SLES-type ingredients contain significant
amounts of SLS-type derivatives (Fig. 5A), as already noted during
the qualitative analy (see Fig. 2). Their presence can be probably
explained by incomplete ethoxylation of the raw materials used for
their production. When such a mixture is subjected to sulfonation
reaction, the corresponding mixture of alkyl and alkyl-ether sulfates is obtained.
The concentration of each alkyl sulfate and alkyl ethoxy sulfate anion was calculated using the semi-quantitative method described in the preceding section and their sum in each of the
products was compared with the active component content declared by the manufacturer (Table 1). The agreement between
the declared and determined sums is satisfactory, especially taking into account the semi-quantitative nature of our method. Only
for TEALS and MEALS the difference is more significant, probably
due to the lower dissociation degree of triethanolamine and monoethanolamine salts of the alkyl sulfates which may affect both
the chromatographic separation and subsequent MS detection. The
average ethoxylation degree (see Supporting Materials) of the three
SLES mixtures with increasing their average molecular weight and
agrees with the manufacturer’s declaration. The average molecular
weight of the alkyl and alkyl ether sulfates was also established
on the basis of their composition (Table 1). Also, in this case the
agreement with the manufacturer declarations is satisfactory, further validating our semi-quantitative method.
The present method has been developed and validated for detection of sulfate-based surfactants but can be extended to other
ionic surfactants. It offers good sensitivity and allows the determination of SLS and SLES homologues using a cheap and widely
available standard substance – sodium dodecyl sulfate (SDS). The
employed PrecI mode allows even for detection of (ethoxylated)
alkyl sulfates not yet described in the literature. In addition, the
method does not require any a priori knowledge of m/z values specific for each parent and fragmentation ion, which is an essential
requirement of the MRM method. Although the latter offers the
highest sensitivity [39], the PrecI mode still offers good sensitivity
and facilitates interpretation of the data as compared to the mass
spectra and chromatograms obtained in MS scanning mode.


(2)

where: aDS, SDS and bDS, SDS are the calibration curve parameters for
DS− ion in a standard SDS solution (external standard), and d is
the sample dilution.
The detection limits can be calculated using the sensitivity coefficients, fi , and calibration parameters for DS− (Eq. (3))

LODi = 3.3 ·

Declared by manufacturer

(3)

where SDb,DS, SDS is the standard deviation of calibration curve parameter b for DS− ion in a standard SDS solution.
Additionally, standard deviation (SD) for the peak areas in blank
samples (the shampoos devoid of SLS and SLES) was established
(SDSh , n = 5) and it was found significantly lower than SDb . The
LOD was in the range 0.12–0.97 mg/L, which is satisfactory for
determination of SLS and SLES homologues in cosmetic products.
Moreover, the sensitivity of the method could be further improved
by switching to a Multiple Reaction Monitoring (MRM) mode and
monitoring the specified pairs of parent/fragmentation ions.
3.4. Validation of the semi-quantitative method for alkyl and alkyl
ether sulfates determination
The method described above allows to correct for the experimentally observed dependence of sensitivity on retention time in
mixtures of practically unlimited number of SLS and SLES homologues and enables their semi-quantitative determination. In the
absence of reliable analytical standards for any other than SDS homologue of SLS and SLES, the analytical validity of the method was
critically assessed by comparing the sum of all semi-quantitatively
determined components with the total content of active substance

(manufacturer declaration) in seven cosmetic/household ingredient
products. Three SLES mixtures (SLES-340, SLES-384, SLES-432) and
four SLS-type derivatives: sodium lauryl sulfate (SLS), ammonium
lauryl sulfate (ALS), triethanolamine lauryl sulfate (TEALS) and monoethanolamine lauryl sulfate (MEALS) were employed for this
purpose. Figs. 5 and SM5 collect the extracted ion chromatograms
(EIC) showing the signal intensity from the non-ethoxylated alkyl
sulfate anions (SLS-type, Fig. 5A) and alkyl ethoxy sulfate anions
(SLES-type, Fig. 5B) species (all selected in PrecI mode). The first
striking observation is that while the SLS-type ingredients (SLS,

3.5. Alkyl (SLS-type) and alkyl ether sulfate (SLES-type) anions
determination in cosmetic/household products
Having established and validated the semi-quantitative method
for determination of SLS and SLES homologues, we analyzed 17
cosmetic/household products and compared their content of alkyl
and alkyl ethoxy sulfates with the declarations provided by manufacturers in the INCI (International Nomenclature of Cosmetic Ingredients) lists. The products were chosen in a way to represent
formulations with declared presence and absence of SLS- and SLEStype derivatives.
7


K. Pawlak and K. Wojciechowski

Journal of Chromatography A 1653 (2021) 462421

Fig. 5. Extracted multi-ion chromatograms (EIC) of: A (left panel) - alkyl (non-ethoxylated, SLS-type) and B (right panel) - alkyl ethoxy (ethoxylated, SLES-type) sulfates
detected in cosmetic/household ingredients using Precursor ion mode (PrecI: ∗ ∗ →97 (40 eV)).

would not be able to produce both m/z 80 and 97 signals. Therefore, they could not interfere with determination of the sulfatebased surfactants using our method based on PrecI mode. It should
be noted that the fragmentation ion observed at m/z 97 can also
be obtained for H2 PO4 − anion produced during fragmentation of

phospholipids. Nevertheless, such signals could be excluded on the
basis of different retention times due to longer aliphatic chains in

According to declarations, some of the selected products contain other anionic surfactants similar to SLS and SLES: sulfonate,
sulfosuccinate, sulfoacetate, taurate or isethionate, where the sulfur
atom in the headgroup is bound directly to C-atom (see Table 2).
Because of much higher energy required to dissociate an S-C bond,
under the presently employed fragmentation condition and due to
lack of the forth oxygen present in sulfate group, these moieties
8


K. Pawlak and K. Wojciechowski

Journal of Chromatography A 1653 (2021) 462421

Table 2
SLS and SLES homologues content determined semi-quantitatively in cosmetic and household products compared with the producer’s declarations in the INCI lists
(names in italics indicate alkyl (SLS-type) and alkyl-ether (SLES-type) sulfates).

No

Product description

1
2

Moisturizing shampoo for dry hair
Nutrifying shampoo for dry hair


3
4

Smoothing cleansing conditioner
Moisturizing shampoo for dry hair

5

Shampoo for dry and damaged hair

6
7

Shampoo for dyed hair
Moisturizing shampoo for normal and
dry hair
Strengthening shampoo for greasy and
falling out hair
Repair-shampoo
Shampoo without SLES
Anti-dandruff shampoo

8
9
10
11

12

Hypoallergenic shampoo for fair, dyed

and bleached hair

13
14
15
16

Hair-repairing shampoo
Liquid soap
Anti-dandruff shampoo
Shampoo for delicate and damaged
hair
Bath and shower gel

17

Sulfur-containing surfactants declared in
INCI list
Sodium Coco-Sulfate
Sodium Laureth Sulfate, Disodium Laureth
Sulfosuccinate, Sodium Lauryl Sulfoacetate
None
Sodium Methyl Cocoyl Taurate, Sodium
C14-16 Olefin Sulfonate
Sodium Lauryl Sulfate, Sodium Laureth
Sulfate, Sodium Xylenesulfonate
None
Sodium Coco-Sulfate

Alkyl sulfates [%]


Alkyl ethoxy
sulfates [%]

Compliance with
INCI (SLS/SLES)

14.69 ± 1.31
11.59 ± 0.53

< LOD
11.25 ± 0.50

+/+
-/+

< LOD
1.51 ± 0.79

< LOD
< LOD

+/+
-/+

28.75 ± 1.39

4.54 ± 0.91

+/+


23.16 ± 1.08
22.88 ± 1.03

19.18 ± 2.41
< LOD

-/+/+

12.14 ± 0.52

0.62 ± 0.14

+/+

30.51 ± 1.22
< LOD
9.22 ± 1.04

< LOD
< LOD
1.77 ± 0.42

+/+
+/+
+/+

Ammonium Lauryl Sulfate, Sodium Laureth
Sulfate
Sodium Coco-Sulfate

None
Sodium Laureth Sulfate, Sodium Lauryl
Sulfate, Sodium Xylenesulfonate,
TEA-Dodecylbenzenesulfonate
Sodium Laureth Sulfate, PEG-2
Dimeadowfoamamido-ethylmonium
methosulfate
Sodium Coco-Sulfate
Sodium Laureth Sulfate
Sodium Cocoyl Isethionate
Disodium Laureth Sulfosuccinate

5.10 ± 1.06

4.31 ± 0.49

-/+

42.10 ± 1.87
1.48 ± 0.43
< LOD
< LOD

< LOD
0.83 ± 0.19
< LOD
< LOD

+/+
-/+

+/+
+/+

Sodium Laureth Sulfate

5.42 ± 0.32

5.97 ± 0.62

-/+

phospholipids. An additional verification could be done by monitoring the PO3 − signal (m/z 79) [40,41].
The tested cosmetic/household products were diluted with water and to ensure that the resulting concentration lies within the
linear range, each sample was additionally diluted twice more. The
total SLS and SLES homologues concentrations determined semiquantitatively are collected in Table 2 and the respective chromatograms are shown in Fig. SM6. The majority of tested products indeed conforms to declarations and only in one case both
the alkyl and alkyl ethoxy sulfate anions were detected in significant amounts in a shampoo declared as devoid of any sulfurbased surfactants (shampoo 6). Two other products declared as
free from SLS- or SLES-type ingredients (a hair conditioner and a
shampoo prepared in our lab with no SLS or SLES added) did conform to their INCI declarations. It is worth stressing that one of
the products (shampoo 12) according to the producer’s declaration
contained PEG-2 dimeadowfoamamidoethylmonium methosulfate
(Meadowquat), where a sulfate group is present in the counterion
(methosulfate). The latter would fragment producing the m/z 80
and m/z 97 ions. However, our method is insensitive to false positive results of this type, since scanning for the PrecI-selected ions
is performed only for ions with m/z above 200. Thus, the amount
of alkyl sulfates determined using our method is not biased by the
presence of Meadowquat, and the product most likely indeed contains non-ethoxylated SLS-type surfactants not declared in the INCI
list.
In some cases (e.g. shampoo no. 4) the low amount of undeclared alkyl sulfates could be detected, suggesting their unintentional use – for example as an impurity present in other ingredients. Given the abundance of alkyl sulfates in SLES-type cosmetic/household ingredients depicted in Fig. 5, the undeclared
quantities of SLS homologues could have been even unintentionally introduced into final formulations (see e.g. shampoos no. 2,
12, 14, 17).


4. Conclusions
In this contribution we have developed a new MS/MS method
for selective detection of alkyl and alkyl ether sulfates based on
Precursor Ion Scanning (PrecI) Mode following their separation on
a C4 reversed-phase LC column. For this purpose, only ions releasing the SO3 − and HSO4 − moieties (m/z 80 and 97, respectively)
are selected and further monitored, providing selectivity to the organic sulfate species characteristic to homologues of sodium lauryl sulfate (SLS) and sodium laureth sulfate (SLES) with m/z in the
range 20 0–60 0. Other sulfur-bearing species commonly found in
SLS- and SLES-free products (sulfonate, sulfosuccinate, sulfoacetate,
taurate or isethionate), where the sulfur atom in the headgroup
is bound directly to C-atom, are excluded based on the absence
of HSO4 − signal (m/z 97). Further we have developed a quantitative method for determination of dodecylsulfate ions and a semiquantitative method for determination of any ethoxylated or nonethoxylated alkyl sulfates. The latter method is based on series
of dilutions providing the sensitivity coefficient (fi ) for each signal, which enables subsequent estimation of concentration of the
given species in the sample (all in one run using single standard
– easily accessible sodium dodecylsulfate (SDS)). The proposed LCMS/MS method allows simultaneous determination of SLS and SLES
homologues, confirmation of their identity, determination of average molecular weight of surfactants and degree of ethoxylation, as
well as discovery of new sulfate compounds. Conventionally, these
tasks require the use of three different methods, which can now be
replaced by a single one developed within this work. The overall
accuracy of the method strongly depends on the quality of correlation between fi and retention time for each group of homologues
established under the same detection conditions. For this reason,
we recommend to calibrate the method using mixtures with as
many SLS and SLES homologues as possible. Extension of the presented method to the environmental or food analy is also possible,
although because of higher required sensitivity, the Precursor Ion
9


K. Pawlak and K. Wojciechowski

Journal of Chromatography A 1653 (2021) 462421


Scanning (PrecI) should preferentially be replaced by the Multiple
Reaction Monitoring (MRM) mode.
Using the newly developed method we have assayed four SLStype and three SLES-type cosmetic/household ingredients for the
presence of Cn EOm SO4 − species in the range of n (alkyl chain
length) 12–16 and m (number of ethoxy groups) 0–5. While the
SLS-type ingredients are indeed devoid of any ethoxylated species,
the opposite is not true for the SLES-type ingredients which
are contaminated with non-ethoxylated derivatives. Finally, the
method was applied to 17 commercial cosmetic/household products to verify their consistency with the manufacturers’ declarations in terms of presence of alkyl (SLS-type) and alkyl ether (SLEStype) sulfates. While the SLES-like surfactants content was usually
consistent with the declarations, several formulations contained
undeclared SLS-like ingredients, most likely originating from impurities in SLES-type ingredients.

[10] L. Wu, Y. Wu, H. Shen, P. Gong, L. Cao, G. Wang, H. Hao, Quantitative structureion intensity relationship strategy to the prediction of absolute levels without
authentic standards, Anal. Chim. Acta 794 (2013) 67–75, doi:10.1016/j.aca.2013.
07.034.
[11] A. Stavrianidi, E. Stekolshchikova, A. Porotova, I. Rodin, O. Shpigun, Combination of HPLC–MS and QAMS as a new analytical approach for determination
of saponins in ginseng containing products, J. Pharm. Biomed. Anal. 132 (2017)
87–92, doi:10.1016/j.jpba.2016.09.041.
[12] S.P. Li, C.F. Qiao, Y.W. Chen, J. Zhao, X.M. Cui, Q.W. Zhang, X.M. Liu, D.J. Hu,
A novel strategy with standardized reference extract qualification and single
compound quantitative evaluation for quality control of panax notoginseng
used as a functional food, J. Chromatogr. A 1313 (2013) 302–307, doi:10.1016/j.
chroma.2013.07.025.
[13] E. Olkowska, Z. Polkowska, J. Namies´ nik, Analytical procedures for the determination of surfactants in environmental samples, Talanta 88 (2012) 1–13,
doi:10.1016/j.talanta.2011.10.034.
[14] E. Olkowska, Z. Polkowska, J. Namies´ nik, Analytics of surfactants in the environment: problems and challenges, Chem. Rev. 111 (2011) 5667–5700, doi:10.
1021/cr100107g.
[15] Y.R. Bazel, I.P. Antal, V.M. Lavra, Z.A. Kormosh, Methods for the determination of anionic surfactants, J. Anal. Chem. 69 (2014) 211–236, doi:10.1134/
S1061934814010043.

[16] B. Wyrwas, A. Zgoła-Grzes´ kowiak, Continuous flow methylene blue active substances method for the determination of anionic surfactants in river water
and biodegradation test samples, J. Surfactants Deterg. 17 (1) (2014) 191–198,
doi:10.1007/s11743- 013- 1469- x.
[17] K. Sini, M. Idouhar, A.C. Ahmia, A. Ferradj, A. Tazerouti, Spectrophotometric determination of anionic surfactants: optimization by response surface methodology and application to algiers bay wastewater, Environ. Monit. Assess. 189
(12) (2017) 1–12, doi:10.1007/s10661- 017- 6359- 7.
[18] N.M. Makarova, E.G. Kulapina, New potentiometric sensors based on ionic associates of sodium dodecylsulfate and cationic complexes of copper(II) with
some organic reagents, Electroanalysis 27 (3) (2015) 621–628, doi:10.1002/
elan.201400491.
[19] M.C. Bruzzoniti, R.M. De Carlo, C. Sarzanini, Determination of sulfonic acids
and alkylsulfates by ion chromatography in water, Talanta 75 (3) (2008) 734–
739, doi:10.1016/j.talanta.2007.12.026.
[20] L.H. Levine, J.E. Judkins, J.L. Garland, Determination of anionic surfactants during wastewater recycling process by ion pair chromatography with suppressed
conductivity detection, J. Chromatogr. A 874 (2) (20 0 0) 207–215, doi:10.1016/
S0 021-9673(0 0)0 0155-2.
[21] S. Wangkarn, P. Soisungnoen, M. Rayanakorn, K. Grudpan, Determination of
linear alkylbenzene sulfonates in water samples by liquid chromatography-UV
detection and confirmation by liquid chromatography-mass spectrometry, Talanta 67 (4) (2005) 686–695, doi:10.1016/j.talanta.2005.03.011.
[22] H.S. Park, C.K. Rhee, Simultaneous determination of nonionic and anionic
industrial surfactants by liquid chromatography combined with evaporative
light-scattering detection, J. Chromatogr. A 1046 (1–2) (2004) 289–291, doi:10.
1016/j.chroma.2004.06.061.
[23] S.H. Im, Y.H. Jeong, J.J. Ryoo, Simultaneous analysis of anionic, amphoteric,
nonionic and cationic surfactant mixtures in shampoo and hair conditioner by
RP-HPLC/ELSD and LC/MS, Anal. Chim. Acta 619 (1) (2008) 129–136, doi:10.
1016/j.aca.2008.03.058.
[24] S.H. Im, J.J. Ryoo, Characterization of sodium laureth sulfate by reversed-phase
liquid chromatography with evaporative light scattering detection and 1 H nuclear magnetic resonance spectroscopy, J. Chromatogr. A 1216 (12) (2009)
2339–2344, doi:10.1016/j.chroma.2009.01.005.
[25] M.Y. Ye, R.G. Walkup, K.D. Hill, Determination of surfactant sodium lauryl ether
sulfate by ion pairing chromatography with suppressed conductivity detection,

J. Liq. Chromatogr. 17 (19) (1994) 4087–4097, doi:10.1080/10826079408013602.
[26] M. Holcˇ apek, K. Volná, P. Jandera, L. Koláˇrová, K. Lemr, M. Exner, A. Církva,
Effects of ion-pairing reagents on the electrospray signal suppression of
sulphonated dyes and intermediates, J. Mass Spectrom. 39 (1) (2004) 43–50,
doi:10.1002/jms.551.
[27] E. Matthus, M.S. Holt, A. Kiewiet, G.B.J. Rijs, Environmental monitoring for
linear alkylbenzene sulfonate, alcohol ethoxylate, alcohol ethoxy sulfate, alcohol sulfate, and soap, Environ. Toxicol. Chem. 18 (11) (1999) 2634–2644,
doi:10.1002/etc.5620181133.
[28] A. Dufour, D. Thiébaut, L. Ligiero, M. Loriau, J. Vial, Chromatographic behavior
and characterization of polydisperse surfactants using ultra-high-performance
liquid chromatography hyphenated to high-resolution mass spectrometry, J.
Chromatogr. A 1614 (2020) 460731–460742, doi:10.1016/j.chroma.2019.460731.
[29] L.H. Levine, J.L. Garland, J.V. Johnson, Simultaneous quantification of polydispersed anionic, amphoteric and nonionic surfactants in simulated wastewater samples using C18 high-performance liquid chromatography-quadrupole
ion-trap mass spectrometry, J. Chromatogr. A 1062 (2) (2005) 217–225, doi:10.
1016/j.chroma.2004.11.038.
[30] C. Fernández-Ramos, O. Ballesteros, A. Zafra-Gómez, R. Blanc, A. Navalón,
J.L. Vílchez, Determination of alcohol sulfates and alcohol ethoxysulfates in
wastewater samples by liquid chromatography tandem mass spectrometry, Microchem. J. 106 (2013) 180–185, doi:10.1016/j.microc.2012.06.007.
[31] Y. Valadbeigi, M. Tabrizchi, Application of Ion mobility spectrometry in study
of surfactants adsorbed on different dish surfaces, Int. J. Ion Mobil. Spectrom.
17 (1) (2014) 35–41, doi:10.1007/s12127- 013- 0142- 4.
[32] I.H.K. Dias, R. Ferreira, F. Gruber, R. Vitorino, A. Rivas-Urbina, J.L. SanchezQuesada, J. Vieira Silva, M. Fardilha, V. De Freitas, A. Reis, Sulfate-based lipids:

Declaration of Competing Interest
The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.
CRediT authorship contribution statement
Katarzyna Pawlak: Conceptualization, Methodology, Validation,
Data curation, Visualization. Kamil Wojciechowski: Conceptualization, Writing – review & editing.
Acknowledgment

This work was financially supported by the Warsaw University
of Technology, Poland. Ms Aleksandra Chybicka is acknowledged
for technical assistance.
Supplementary materials
Supplementary material associated with this article can be
found, in the online version, at doi:10.1016/j.chroma.2021.462421.
References
[1] M.J. Rosen, J.T. Kunjappu, Surfactants and Interfacial Phenomena, 4th ed., John
Wiley and Sons, 2012, doi:10.1002/9781118228920.
[2] D. Myers, Surfactant Science and Technology, 3rd ed., John Wiley and Sons,
2005, doi:10.1002/047174607X.
[3] J. Chen, X. Hu, Y. Fang, G. Jin, Y. Xia, What dominates the interfacial properties
of extended surfactants: amphipathicity or surfactant shape? J. Colloid Interface Sci. 547 (2019) 190–198, doi:10.1016/j.jcis.2019.04.002.
[4] X. Liu, Y. Zhao, Q. Li, T. Jiao, J. Niu, Adsorption behavior of fatty alcohol ether
sulfonate at different interfaces, J. Surfactants Deterg. 20 (2) (2017) 401–409,
doi:10.1007/s11743- 016- 1918- 4.
[5] S.A. Mousavi, F. Khodadoost, Effects of detergents on natural ecosystems and
wastewater treatment processes: a review, Environ. Sci. Pollut. Res. 26 (2019)
26439–26448, doi:10.1007/s11356- 019- 05802- x.
[6] J. Bandier, B.C. Carlsen, M.A. Rasmussen, L.J. Petersen, J.D. Johansen, Skin reaction and regeneration after single sodium lauryl sulfate exposure stratified
by filaggrin genotype and atopic dermatitis phenotype, Br. J. Dermatol. 172 (6)
(2015) 1519–1529, doi:10.1111/bjd.13651.
´
[7] Jurek I., Góral I., Mierzynska
Z., Moniuszko-Szajwaj B., Wojciechowski K., Effect
of synthetic surfactants and soapwort (Saponaria Officinalis L.) extract on skinmimetic model lipid monolayers, Biochim. Biophys. Acta Biomembr. 1861 (3)
(2019) 556–564, doi: 10.1016/j.bbamem.2018.12.005.
[8] T. Bujak, M. Zagórska-Dziok, Z. Nizioł-Łukaszewska, Complexes of ectoine with
the anionic surfactants as active ingredients of cleansing cosmetics with reduced irritating potential, Molecules 25 (6) (2020) 1433–1446, doi:10.3390/
molecules25061433.

[9] V.C. Robinson, W.F. Bergfeld, D.V. Belsito, R.A. Hill, C.D. Klaassen, J.G. Marks,
R.C. Shank, T.J. Slaga, P.W. Snyder, F.A. Andersen, Final report of the amended
safety assessment of sodium laureth sulfate and related salts of sulfated
ethoxylated alcohols, Int. J. Toxicol. 29 (4) (2010) 151S–161S, doi:10.1177/
1091581810373151.
10


K. Pawlak and K. Wojciechowski

[33]

[34]
[35]

[36]
[37]

Journal of Chromatography A 1653 (2021) 462421

analysis of healthy human fluids and cell extracts, Chem. Phys. Lipids 221
(2019) 53–64, doi:10.1016/j.chemphyslip.2019.03.009.
R. Hayes, A. Ahmed, T. Edge, H. Zhang, Core-shell particles: preparation, fundamentals and applications in high performance liquid chromatography, J. Chromatogr. A 1357 (2014) 36–52, doi:10.1016/j.chroma.2014.05.010.
F. Gritti, G. Guiochon, Mass transfer kinetics, band broadening and column efficiency, J. Chromatogr. A 1221 (2012) 2–40, doi:10.1016/j.chroma.2011.04.058.
S. Fekete, E. Oláh, J. Fekete, Fast liquid chromatography: the domination of
core-shell and very fine particles, J. Chromatogr. A 1228 (2012) 57–71, doi:10.
1016/j.chroma.2011.09.050.
M. Smoluch, G. Grasso, P. Suder, J. Silberring, Mass Spectrometry: An Applied
Approach, 2nd ed., Wiley, 2019, doi:10.1002/9781119377368.
A. Beyaz, W. Fan, P.W. Carr, A.P. Schellinger, Instrument parameters controlling

retention precision in gradient elution reversed-phase liquid chromatography,
J. Chromatogr. A 1371 (2014) 90–105, doi:10.1016/j.chroma.2014.09.085.

[38] C. Patriarca, J.A. Hawkes, High molecular weight spectral interferences in mass
spectra of dissolved organic matter, J. Am. Soc. Mass Spectrom. 32 (1) (2021)
394–397, doi:10.1021/jasms.0c00353.
[39] V. Vidova, Z. Spacil, A review on mass spectrometry-based quantitative proteomics: targeted and data independent acquisition, Anal. Chim. Acta 964
(2017) 7–23, doi:10.1016/j.aca.2017.01.059.
[40] M. Holcˇ apek, R. Jirásko, M. Lísa, Basic rules for the interpretation of atmospheric pressure ionization mass spectra of small molecules, J. Chromatogr. A
1217 (2010) 3908–3921, doi:10.1016/j.chroma.2010.02.049.
[41] J. Pi, X. Wu, Y. Feng, Fragmentation patterns of five types of phospholipids by ultra-high-performance liquid chromatography electrospray ionization quadrupole time-of-flight tandem mass spectrometry, Anal. Methods 8 (6)
(2016) 1319–1332, doi:10.1039/c5ay00776c.

11



×