You are on page 1of 11

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

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

Article history: We present a new liquid chromatography–tandem mass spectrometry (LC-MS/MS) method for simultane-
Received 29 March 2021 ous determination of sodium lauryl sulfate and sodium laureth sulfate homologues in the range of alkyl
Pobrano z http://repo.pw.edu.pl / Downloaded from Repository of Warsaw University of Technology 2022-11-17

Revised 16 July 2021


chain length C12 –C16 with 0–5 ethoxy groups. The method is based on scanning the precursor ions frag-
Accepted 16 July 2021
menting to m/z 80 and 97 (Precursor Ion Scanning mode), which makes it specific for species with easily
Available online 22 July 2021
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 deter-
mine 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 (http://creativecommons.org/licenses/by/4.0/)

1. Introduction “lauryl” name and its derivative “laureth” were introduced to high-
light the ill-defined structures of the resulting mixtures, in contrast
Anionic surfactants are major constituents of most detergents to single species, e.g. sodium dodecylsulfate (SDS). The ethoxyla-
and cosmetic cleaning/washing products but are also ubiquitous in tion step introduces another heterogeneity of chemical structures
many other formulations, where wetting, dispersing, emulsifying of SLES (number of the ethoxy units), in addition to the variable
or foaming activities are required [1]. Most of the currently avail- length of the alkyl chain present in SLS. Consequently, the com-
able shampoos, shower gels, liquid soaps and dishwashing liquids mercially available alkyl and alkyl-ether sulfates show wide distri-
are based on alkyl and alkyl-ether sulfates produced by sulfona- bution of chemical structures and surfactant properties [3,4]. De-
tion of alkyl alcohols and ethoxylated alkyl alcohols. In industrial spite their great efficacity in lowering surface tension and sustain-
practice pure alcohols are very rarely used, typically their mixtures ing foams, SLS and SLES may pose some environmental hazards
are employed instead (e.g. a mixture obtained from hydrolyzed and caused by their limited biodegradability and persistent foam for-
hydrogenated coconut and palm oils, as in the case of the most mation [5]. Their high detergent power may also lead to exces-
popular alkyl sulfate – sodium lauryl sulfate (SLS)). Its ethoxylated sive lipid and protein removal when used on a daily basis [6–8].
analogue (sodium laureth sulfate, SLES) is produced analogously, This prompts some consumers to avoid products with sodium lau-
with an additional intermediate alcohol ethoxylation step [2]. The ryl 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

Corresponding author at: Faculty of Chemistry, Warsaw University of Technol- SLS and SLES with their homologous mixtures (e.g. “sodium coco
ogy, Noakowskiego 3, Warsaw 00-664, Poland. sulfate”) to misguide the consumer with a different INCI (Inter-
E-mail addresses: kpawlak@ch.pw.edu.pl (K. Pawlak), national Nomenclature of Cosmetic Ingredients) name [9]. Given
kamil.wojciechowski@ch.pw.edu.pl (K. Wojciechowski).

https://doi.org/10.1016/j.chroma.2021.462421
0021-9673/© 2021 The Author(s). Published by Elsevier B.V. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/)
K. Pawlak and K. Wojciechowski Journal of Chromatography A 1653 (2021) 462421

the variety of homologous alkyl and alkyl-ether sulfate surfac- lates (AE), and alcohol ethoxylated sulfates (AES) by reversed-
tants, there is an urgent need for a robust and universal analytical phase chromatography using a C18 column and ammonium acetate
method capable of differentiation between not only the alkyl chain to stabilize pH of the mobile phase. The partially separated com-
length but also the extent of ethoxylation of SLES. pounds were detected using a fluorescence detection (FLD) and
One of the major obstacles in quantitative analysis of complex a thermospray mass spectrometer working in the scanning mode
mixtures of natural or synthetic compounds is the lack of analyt- (m/z range 20 0–80 0). This provided more detailed structural infor-
ical standards. In contrast to many well-defined compounds with mation at the expense of sensitivity. The potential for quantitative
distinct structures, no reference materials are available for most of analy was exemplified by determination of the total concentration
the complex molecules (like proteins, lipids or alkaloids) or mix- of active substances in sewage in relation to the AES and LAS com-
tures of homologues (such as alkyl and alkyl-ether sulfates in the mercial mixtures. Dufour et al. employed ultra-high-performance
present case). This problem can be circumvented using computa- liquid chromatography with high-resolution mass spectrometry
tional methods based on Quantitative Structure and Ionization In- (UPLC-HR-MS) to separate four homologues of alkylbenzene sul-
tensity Relationship (QSIIR). The method was successfully applied fonate (4-dodecylbenzenesulfonic acid, DBSA). The homologues dif-
to predict relative levels of 29 organic acids in complex matrices, fered in their alkyl chain lengths (decylbenzenesulfonate, unde-
registered by product ion scanning. It offered accurate results (in cylbenzenesulfonate, dodecylbenzenesulfonate, tridecylbenzenesul-
the range of 80–120%) for 16 organic acids for which the absolute fonate) and could be separated in the RPLC mode using C4, C18
concentrations were quantified and used as reference. For the re- and C30 columns [28]. Levine et al. employed RPLC coupled with
maining organic acids, such accuracy could not be achieved due to electrospray ionization quadrupole ion trap mass spectrometry
lack of standards or too low concentration of the acid in the sam- (ESI-Q-IT-MS) to simultaneously determine three common surfac-
ples [10]. Such approach enables development of methods relying tants: an amphoteric cocoamphoacetate, a nonionic alcohol ethoxy-
on a single standard for Quantitative es of Multi-component Sys- late and an anionic SLES (dynamic linear range 1.5–40 mg/L for to-
tem (QAMS) [11,12]. tal amount of SLES normalized against commercial mixture) [29].
Currently, most analytical methods for determination of alkyl Four alkyl sulfate and 2 ethoxymers of alkyl-ether sulfate homo-
and alkyl-ether sulfate are based on spectrophotometric, electro- logues were determined in the SPE-preconcentrated wastewater
Pobrano z http://repo.pw.edu.pl / Downloaded from Repository of Warsaw University of Technology 2022-11-17

chemical and chromatographic techniques [13–15]. The official EU samples using a liquid chromatography–tandem mass spectrome-
method for determination of anionic surfactants in detergents de- try (LC–MS/MS) with electrospray ionization (ESI) in negative ion
scribed in Regulation (EC) No. 648/2004 of the European Parlia- mode. Based on the commercial mixture producer’s declarations,
ment and of the Council and in ISO 7875-1(1996) standard is based the limits of quantification (LOQ) in the range 0.3–0.4 μg/L and
on formation of blue-colored salts of anionic surfactants with 0.5–1.5 μg/L were reported for the SLS and SLES homologues, re-
the methylene blue dye, which are determined spectrophotomet- spectively [30]. Another interesting approach to separate surfac-
rically after extraction to chloroform (Methylene Blue Active Sub- tants present in commercial mixtures of SLS and SLES employed
stance test, MBAS). Although the method is continuously improved the ion mobility mass spectrometry [31]. The ionized compounds
[16,17], the ion-pair formation reaction does not provide sufficient were distinguished based on their drift time in an electric field,
selectivity to enable distinction not only between the nature of the which depended on their molecular weight. The authors obtained
anionic group (phosphate, carboxylic, sulfate or sulfonate) but also six peaks for the SLS mixture and twelve for the SLES one, al-
the alkyl chain structure. though their identity was not determined due to lack of standards.
Another group of non-chain-length-selective methods is based Nevertheless, the developed method could be applied to determine
on formation of ionic associates between anionic surfactants and the total SLS and SLES content (with respect to commercially avail-
cationic species using potentiometric sensors [18] or suppression able SLS and SLES mixtures) adsorbed by different dish surfaces.
of ionic conductivity using ion chromatography [19]. Using a sim- In this contribution we further extend the analytical capa-
ilar approach, Levine et al. were capable of determining concen- bilities of LC-MS/MS technique by employing for the first time
tration of ammonium lauryl sulfate, sodium laureth sulfate and an MS/MS scanning of the precursor ions to follow the sulfate-
sodium alkyl (C10 SO4 − – C16 SO4 − ) ether sulfates present in com- bearing species. We propose a novel method for determination
mercially available detergents (dynamic linear ranges: 1.0–500, of individual alkyl and alkyl-ether sulfates in mixtures of cos-
2.5–550 and 3.0–630 mg/L, respectively) [20]. After chromato- metic/household ingredients and products using a single and easily
graphic separation, a mixture of four linear alkylbenzene sulfonates available standard (SDS) for quantitative analy of multi-component
(C10 SO4 − - C13 SO4 − ) was detected by UV spectroscopy [21] and a system (QAMS). We show how the experimentally observed depen-
mixture of anionic and nonionic surfactants (including sodium lau- dency of signal intensity on the retention, number of fragmenta-
ryl sulfate and α -olefin sulfonate) - using an evaporative light scat- tion ions obtained in a collision cell and recovery from a stationary
tering detector (ELSD) [22]. Nevertheless, selectivity is often when phase can be accounted for in QAMS methods to obtain good ac-
such detectors are employed [21,23,24]. curacy. SDS could be employed as a universal and sole standard, as
Regardless of the detection method, anionic surfactants can be it was present in every sample. The new method allowed us to de-
separated using ion exchange chromatography [19]. However, most tect 5 non-ethoxylated and 20 ethoxylated sulfates in commercial
commonly the separation is achieved thanks to differences in the SLES products. To show the unprecedented application potential of
affinity of aliphatic chains to a hydrophobic stationary phase. For the new method for real-life samples analy we verified the manu-
this purpose, deprotonated anionic surfactants (pH > 7) can be facturers’ declarations about the presence/absence of SLS and SLES
separated as neutral ion pairs formed with quaternary ammo- ingredients in 17 cosmetic/household commercial products.
nium cations (ion pair liquid chromatography, IPLC [25]) or in a
non-dissociated form (pH < 7), using reversed-phase liquid chro- 2. Experimental
matography, RPLC. The former method is, however, not suitable
for mass spectrometry because of significant signal suppression Separation of the alkyl and alkyl ether sulfates was carried out
[26]. The advantage of lowering pH of the mobile phase is a re- using a 1220 Infinity II LC Systems (Agilent Technologies, USA),
duction of hydrophilic interactions interfering with the chromato- whereas their identification and quantitation was achieved with a
graphic process. On the other hand, reduced pH may lower ioniza- 6460 Triple Quad tandem mass spectrometric detector with a Jet
tion efficiency in mass spectrometry [26]. Matthus and colleagues Stream ion source (Agilent Technologies, USA). Analytes were sep-
[27] separated the alkylbenzene sulfonate (LAS), alcohol ethoxy- arated 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
Pobrano z http://repo.pw.edu.pl / Downloaded from Repository of Warsaw University of Technology 2022-11-17

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.

300 Å, Phenomenex). The chromatographic separation employed a chased from Sigma Aldrich (Poland). Acetonitrile (LC/MS purity)
gradient elution from (40% acetonitrile: 60% water) to (95% ace- from POCH (Gliwice, Poland); formic acid (LC/MS purity), from
tonitrile: 5% water), both containing 0.15% (v/v) formic acid, over Fisher Scientific (Fair Lawn, NJ, USA). Demineralized water from a
8 min at a flow rate of 0.2 mL/min. Both the column and mobile Milli-Q system Model Millipore Elix 3 (Molsheim, France) was em-
phase were thermostated at 45 °C. ployed.
The mass spectrometer was operated in negative ion mode es- Standard solution of SDS (1.0 mg/mL) was prepared in Milli Q-
tablished by ionization voltage of 30 0 0 V and nozzle voltage 0 V. water. The surfactant SLES mixtures (0.2–0.5 mg/mL) and commer-
Heated (300 °C) nebulization gas flow of 8 L/min applied at 30 psi cial cosmetic formulations (0.7–1.4 mg of cosmetic/mL) were pre-
was selected as the most appropriate to enhance the ionization of pared by dissolving them in Milli Q-water. For semi-quantitative
low molecular weight compounds. The Precursor Ion (PrecI) scan- analyzes of alkyl and alkyl-ethoxy sulfates, to assure that the con-
ning was applied for the discovery of SLS and SLES homologues. A centration of all sulfates is within the linear response range of cal-
tandem mass spectrometer operating in the PrecI mode finds the ibration curve, all samples were diluted at two or more levels. All
parent ions (the first quadrupole, Q1, operates in the m/z 20 0–80 0 solutions were filtered through 0.45 μm syringe filters prior RPLC
scanning mode) for the m/z values of characteristic fragment ions analy.
(the second quadrupole, Q3, operates in the selected ion monitor-
ing mode). The negative fragmentation ions SO3 − at m/z 80 and
3. Results and discussion
HSO4 − at m/z 97 were selected as characteristic of the sulfate-
bearing compounds (Fig. 1, Table SM1). Structurally similar sul-
3.1. Optimization of the detection conditions
fonate 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 se-
In order to simultaneously quantify all alkyl and alkyl-ether sul-
lectively discovered by the mass spectrometer. The analyses were
fate homologues, the separation method should allow for selec-
controlled and processed by a MassHunter Workstation software
tion of all species yielding hydrogen sulfate ion (HSO4 − , m/z = 97)
(Agilent Technologies, USA). The employed chromatographic and
and/or radical sulfate anion (•SO3 − , m/z = 80) upon fragmenta-
MS conditions are collected in Table SM1.
tion. These species, resulting from dissociation of the C-O-S bond
The commercial shampoos, hair conditioner and liquid soap
in the sulfate group, can be conveniently selected by scanning ions
were purchased in a local cosmetic store in Warsaw (Poland). The
with m/z being reduced to 80 and 97 in the Precursor Ion Scanning
reference shampoo without SLS and SLES was provided by Sapon-
(PrecI) Mode. Although the latter offers additional intrinsic speci-
labs Ltd. (Poland). Three SLES and four SLS-type mixtures with the
ficity in MS analy, until now the PrecI mode has been employed
specified amount of active substances (27–70% w/w) and ethoxy-
almost solely in lipidomic and proteomics analyses [32] and its
lation degree were obtained from an industrial supplier. All SLES
potential in alkyl sulfate analy has been largely unexploited. There-
were declared as sodium salts of ethoxylated sulfates of predom-
fore, in the first part of the study, the detection conditions for elec-
inantly C12 -C14 alcohols with different ethoxylation degree and
trospray tandem mass spectrometer (ESI-MS/MS) were optimized
average molecular weights (Mav ∼ = 340 g/mol, ethoxylation de-
using sodium dodecylsulfate (SDS) and a cosmetic/household in-
gree 1–2.5: SLES-340; Mav ∼ = 384 g/mol, ethoxylation degree 1–
gredient Sodium Laureth Sulfate with declared average molecular
2.5: SLES-384; Mav ∼ = 432 g/mol, ethoxylation degree > 2.5: SLES-
weight of 384 g/mol and average ethoxylation degree 1–2.5 (SLES-
432). The following SLS-type products were used: sodium lau-
384). To this aim, an SDS solution of 1 μg/ml was introduced
ryl sulfate (SLS), ammonium lauryl sulfate (ALS), triethanolamine
(5 μL) into the mobile phase stream (acetonitrile:water, 50:50
lauryl sulfate (TEALS), monoethanolamine lauryl sulfate (MEALS).
(v/v)) at a flow rate of 0.2 ml/min (FIA-ESI-MS). Using SDS, the
Sodium dodecylsulfate, SDS (puriss ACS reagent, ≥ 99.0%) was pur-
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 temperature of the mobile phase and column, sample volume and
signals observed at m/z 265 (corresponding to dodecyl sulfate an- sample dilution (Fig. SM2).
ion, CH3 -(CH2 )11 -SO4 − , (“C12 SO4 − ” or “DS−“ ), were optimized (see In order to validate selectivity of the MS detection in PrecI
Table SM1 in Supplementary Materials). As alkyl and alkyl-ether mode under optimized conditions, two other SLES mixtures (SLES-
sulfates tend to form stable anions also in acidic solutions due to 340 with declared average molecular weight of 340 g/mol and av-
high dissociation constant of sulfate group (pKa = ~2.4) the mo- erage ethoxylation degree 1–2.5, and SLES-432 with declared av-
bile phase was acidified with formic acid (0.15% (v/v)) to sensitiv- erage molecular weight of 432 g/mol and average ethoxylation
ity by elimination of adducts formation and to reduce the noise of degree > 2.5) were analyzed using the method developed with
the mass spectrum [26]. Analogous analy employing the optimized SLES-384. The chemical identity of the anions discovered by PrecI
conditions was performed for SLES-384, where the highest signals was assigned using the Product Ion Scanning mode. Fragmentation
were observed at m/z 265, 309, 381 and 441. They corresponded ions m/z 80 and 97, confirmed the presence of the sulfate group
to the C12 anion (CH3 -(CH2 )11 -SO4 − ) previously found in SDS) in all cases. The difference between theoretical and experimen-
and to different laureth sulfate anions predominant in the mix- tally established monoisotopic mass - M parameter (defined as
ture: CH3 -(CH2 )11 -(OCH2 )1 -SO4 − , CH3 -(CH2 )13 -(OCH2 )2 -SO4 − , CH3 - M =|Mtheoretical -Mexperimental |/Mtheoretical ·106 ) varied between 71
(CH2 )11 -(OCH2 )4 -SO4 − , respectively (C12 EO1 SO4 − , C14 EO2 SO4 − and and 639 ppm (Table SM2), which is typical for low-molecular-
C12 EO4 SO4 − , respectively). The selected dodecyl and laureth sulfate weight-compounds analyzed with quadrupole analyzers, due to the
ions were subjected to fragmentation (Product Ion Scanning) using intrinsic low resolution of the latter [36].
the 10, 20, 30, 40 eV collision energy. All resulting spectra featured All tested SLES mixtures show similar chromatograms with the
the m/z 97 signal, corresponding to HSO4 − anion, confirming that proportion of more lipophilic derivatives increasing in order: SLES-
all selected ions indeed contained the sulfate group. In some cases, 340 < SLES-384 < SLES-432. A more detailed analy showed that
additional signal at m/z 80, corresponding to a radical anion SO3 − , retention time with the number of both alkyl and alkoxyl groups
was also observed. Other signals, if present, corresponded to the (Fig. SM3), which was used to additionally confirm identity of the
breakdown of the C-O or C-C bonds within the ethoxylated part of homologues. This was especially useful when the compound sig-
SLES molecules (Fig. 1). The highest signals were obtained for the nal was too small to obtain a rich Product Ion spectrum. All tested
Pobrano z http://repo.pw.edu.pl / Downloaded from Repository of Warsaw University of Technology 2022-11-17

collision energy of 20 and 40 eV, and these conditions were em- SLES samples were abundant in ethoxylated species spanning the
ployed in subsequent optimization of separation conditions. alkyl chain lengths from 12 to 16 carbon atoms, and the number of
ethoxy groups from 1 to 5 (Fig. 2, Table SM2). Surprisingly, how-
ever, all samples featured also signals that could be assigned to
3.2. Optimization of the separation conditions by RPLC non-ethoxylated sulfates (C12 SO4 − - C16 SO4 − ), typical for SLS-type
products.
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 selec- 3.3. LC-MS calibration using SDS
tion of mobile and stationary phases. Formic acid allowed sepa-
ration of SLS and SLES homologues with a selectivity compara- In order to allow for a semi-quantitative analy of the alkyl and
ble to that achieved in the presence of trifluoroacetic acid (TFA) alkyl-ether sulfates in cosmetic/household ingredients and prod-
in the mobile phase and better than that for ammonium acetate ucts, the LC-MS system was calibrated using SDS as a standard
[23,24]. The presence of formic acid in the mobile phase helps to (note that in contrast to SLES, SDS is a single molecule with known
reduce the hydrophilic and enhance the hydrophobic interactions and defined structure). As shown in Fig. 3 (inset), the calibration
of separated compounds with the alkylated silica stationary phase. curve for the m/z 265 signal in PrecI mode is linear up to at
It should be noted that formic acid, in contrast to TFA, is not an least 10 μg/mL of SDS, confirming that the proposed method can
ion-pairing agent, hence its presence does not deteriorate the sen- be used for quantitative determination of individual alkyl sulfates.
sitivity of MS-based method. The ability of the mobile phase (0.15% Nevertheless, an analogous quantitative analy of alkyl-ether sul-
aqueous solution of formic acid with a linearly increasing amount fates (SLES-type) is more demanding for two reasons: (1) pure ref-
of acetonitrile from 0 to 98% in 20 min) to recover SDS and the erence substances are not easily available, (2) the homologues may
SLES-384 components from the surface of a hydrophobic stationary differ not only in the alkyl chain length (Cn ) but also in ethoxyla-
phase was tested using two Phenomenex 150 × 2,1 mm columns. tion degree (EOm ). Especially the latter fact may complicate quanti-
One was loaded with fully porous silica particles modified with tative analy since the alkyl and ethoxy chains contribute differently
C18 aliphatic chains (Luna) and the other - with core-shell silica to partitioning and fragmentation behavior of SLES molecules (see
particles of wide pores modified with C4 aliphatic chains (Aeris). Fig. SM3).
Chromatographic performance during gradient elution was better As the concentrations of individual components of the analyzed
for the C4-bed core-shell column which produced significantly bet- SLES mixtures were not known, calibration was performed by a
ter shape of the peaks (intensity 3 to 10 times higher and lower series of dilutions to obtain concentrations ranging from 0.1 to
width, see Fig. SM1). The shape of the peak can be by: (1) too 10 μg of raw material/SLES mixture in 1 ml of water. SLES-432
strong interactions of the compound with the stationary phase not was used for this purpose, as it contains the highest number of
balanced by the composition of the mobile phase solution, (2) the different SLS- and SLES-type molecules (Fig. 2). The diluted so-
rate of movement of the compounds in the column depending on lutions were analyzed using the protocol described above (PrecI
the temperature, (3) particle or pore size in the stationary phase mode). For each compound detected in the mixture, a good cor-
[33–35]. Considering that the column size, particle size and gradi- relation was achieved between the peak area and the amount of
ent elution method were the same in both cases, the observed re- SLES-432 in solution (Fig. 3, Table SM3). For comparison, also the
duction of the peak width must be related to the size of the parti- curve for SDS is included in the same graph (m/z 265). The slope
cles’ pores. Due to better chromatographic efficiency (peaks’ shape) of the resulting curves is proportional to the retention time for all
without selectivity in comparison to the fully porous C18 column, compounds detected in SLES-432, which is clearly a consequence
the core-shell C4 column was chosen for all subsequent experi- of different contributions of the -CH2 - and -OC2 H5 - groups to the
ments. Then, optimization of the chromatographic separation pro- solubility and consequently – to the partition coefficient of differ-
cess was carried out by examining the gradient elution program, ent SLS and SLES homologues. The observed reduction of the sen-

4
K. Pawlak and K. Wojciechowski Journal of Chromatography A 1653 (2021) 462421
Pobrano z http://repo.pw.edu.pl / Downloaded from Repository of Warsaw University of Technology 2022-11-17

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 ture was selected for its highest number of detected SLS and SLES
by several reasons: homologues) were acquired by serial dilutions. The dependence of
signal height on concentration was then established using a linear
• reduced chromatographic recovery of analytes [37]
curve y = ax + b, where y is the peak area and x is the concentra-
• lower ion transport efficiency from the ionization to the vac-
tion of the sample (ingredient, cosmetic product, etc.) in the ana-
uum chamber for higher molecular weight ions [38]
lyzed solution (expressed in μg/mL, see Fig. SM4). The sensitivity
• suppression of ion signal intensity by strong ion paring agents
coefficient, fi, for any given alkyl or alkyl-ether sulfate ion (i) in the
and lower stability of ions in ESI chamber or collision cell for
sample is calculated by normalizing the slope (a) for the i-th ion
highly ethoxylated homologues [26].
to that for DS− (Eq. (1)):
Without the use of analytical standards, the observed depen- ai, sample
dence of sensitivity on retention time excludes a direct quanti- fi = (1)
aDS, sample
tative determination of multiple components. To circumvent this
problem, we propose a simple correction scheme using the cali- The dependence of fi on retention time is shown in Fig. 4 and the
bration data for SDS. As shown in Fig. 2, the latter is present in numerical values are collected in Table SM4. The retention time
all tested SLES mixtures, and its amount can be reliably quanti- significantly with the length of the alkyl chain for both SLS and
fied using the calibration curve from Fig. 3. Note that in a more SLES. The sensitivity of the method is primarily dependent on the
general case, for samples devoid of SDS, it can always be added number of -CH2 - groups in the SLS homologues and ethoxy groups
to the sample in known amount as an internal standard. In the present in the SLES homologues. It has already been reported that
proposed scheme, each anion signal is normalized to that of DS− the length of alkyl chain influences the chromatographic recovery
(dodecylsulfate, C12 SO4 − ). First, a calibration curve for DS− anion [37] as well as the efficiency of ionization process and ion trans-
was obtained using solutions with the SDS standard. Next, changes mission from the electrospray to vacuum chamber [38]. However,
of signals for each ion detected in the mixture SLES-482 (this mix- for SLES homologues, the decrease in sensitivity is more notable.

5
K. Pawlak and K. Wojciechowski Journal of Chromatography A 1653 (2021) 462421
Pobrano z http://repo.pw.edu.pl / Downloaded from Repository of Warsaw University of Technology 2022-11-17

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 cos-
metic/household products ingredients.

Product name Sum of active ingredients Average molecular weight and ethoxylation degree (in brackets)

Declared by manufacturer (%) Determined by LC-MS/MS (%) Declared by manufacturer Determined by LC-MS/MS

SLES-340 69.1 60.1 ± 9.0 340 (1–2.5) 344 ± 29 (1.1 ± 0.2)


SLES-384 70.0 67.9 ± 10.2 384 (1–2.5) 387 ± 35 (2.1 ± 0.2)
SLES-432 69.8 68.4 ± 13.7 432 (> 2.5) 400 ± 52 (2.5 ± 0.3)
ALS 27.6 29.1 ± 5.8 294 289 ± 14
SLS 29.0 31.1 ± 6.2 296 296 ± 11
MEALS 27.3 21.1 ± 5.2 334 333 ± 58
TEALS 38.8 29.5 ± 7.1 415 425 ± 83

This is most likely related to the number of fragmentation products ALS, TEALS and MEALS) are devoid of any ethoxylated impurities
for SLES homologues containing higher number of ethoxy groups, (Fig. 5B) as expected, the SLES-type ingredients contain significant
which lowers the impact of the m/z 97 ion (Fig. 1). The change of amounts of SLS-type derivatives (Fig. 5A), as already noted during
sensitivity with the number of ethoxy groups is linear for the same the qualitative analy (see Fig. 2). Their presence can be probably
length of alkyl chain and consequently a linear regression could be explained by incomplete ethoxylation of the raw materials used for
applied. The latter provides a convenient statistical description of their production. When such a mixture is subjected to sulfonation
the agreement between the fit and the experimental data, allowing reaction, the corresponding mixture of alkyl and alkyl-ether sul-
calculations of standard deviation of relative concentrations of SLS fates is obtained.
and SLES homologues. The concentration of each surfactant ion (i) The concentration of each alkyl sulfate and alkyl ethoxy sul-
can be then estimated using the sensitivity coefficient (Eq. (2)): fate anion was calculated using the semi-quantitative method de-
( peak area )i − bDS, SDS scribed in the preceding section and their sum in each of the
Pobrano z http://repo.pw.edu.pl / Downloaded from Repository of Warsaw University of Technology 2022-11-17

Ci = ·d (2) products was compared with the active component content de-
aDS, SDS · fi
clared by the manufacturer (Table 1). The agreement between
where: aDS, SDS and bDS, SDS are the calibration curve parameters for the declared and determined sums is satisfactory, especially tak-
DS− ion in a standard SDS solution (external standard), and d is ing into account the semi-quantitative nature of our method. Only
the sample dilution. for TEALS and MEALS the difference is more significant, probably
The detection limits can be calculated using the sensitivity co- due to the lower dissociation degree of triethanolamine and mo-
efficients, fi , and calibration parameters for DS− (Eq. (3)) noethanolamine salts of the alkyl sulfates which may affect both
SDb,DS,SDS the chromatographic separation and subsequent MS detection. The
LODi = 3.3 · · fi (3) average ethoxylation degree (see Supporting Materials) of the three
aDS,SDS
SLES mixtures with increasing their average molecular weight and
where SDb,DS, SDS is the standard deviation of calibration curve pa- agrees with the manufacturer’s declaration. The average molecular
rameter b for DS− ion in a standard SDS solution. weight of the alkyl and alkyl ether sulfates was also established
Additionally, standard deviation (SD) for the peak areas in blank on the basis of their composition (Table 1). Also, in this case the
samples (the shampoos devoid of SLS and SLES) was established agreement with the manufacturer declarations is satisfactory, fur-
(SDSh , n = 5) and it was found significantly lower than SDb . The ther validating our semi-quantitative method.
LOD was in the range 0.12–0.97 mg/L, which is satisfactory for The present method has been developed and validated for de-
determination of SLS and SLES homologues in cosmetic products. tection of sulfate-based surfactants but can be extended to other
Moreover, the sensitivity of the method could be further improved ionic surfactants. It offers good sensitivity and allows the deter-
by switching to a Multiple Reaction Monitoring (MRM) mode and mination of SLS and SLES homologues using a cheap and widely
monitoring the specified pairs of parent/fragmentation ions. available standard substance – sodium dodecyl sulfate (SDS). The
employed PrecI mode allows even for detection of (ethoxylated)
3.4. Validation of the semi-quantitative method for alkyl and alkyl
alkyl sulfates not yet described in the literature. In addition, the
ether sulfates determination
method does not require any a priori knowledge of m/z values spe-
cific for each parent and fragmentation ion, which is an essential
The method described above allows to correct for the experi-
requirement of the MRM method. Although the latter offers the
mentally observed dependence of sensitivity on retention time in
highest sensitivity [39], the PrecI mode still offers good sensitivity
mixtures of practically unlimited number of SLS and SLES homo-
and facilitates interpretation of the data as compared to the mass
logues and enables their semi-quantitative determination. In the
spectra and chromatograms obtained in MS scanning mode.
absence of reliable analytical standards for any other than SDS ho-
mologue 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 3.5. Alkyl (SLS-type) and alkyl ether sulfate (SLES-type) anions
(manufacturer declaration) in seven cosmetic/household ingredient determination in cosmetic/household products
products. Three SLES mixtures (SLES-340, SLES-384, SLES-432) and
four SLS-type derivatives: sodium lauryl sulfate (SLS), ammonium Having established and validated the semi-quantitative method
lauryl sulfate (ALS), triethanolamine lauryl sulfate (TEALS) and mo- for determination of SLS and SLES homologues, we analyzed 17
noethanolamine lauryl sulfate (MEALS) were employed for this cosmetic/household products and compared their content of alkyl
purpose. Figs. 5 and SM5 collect the extracted ion chromatograms and alkyl ethoxy sulfates with the declarations provided by man-
(EIC) showing the signal intensity from the non-ethoxylated alkyl ufacturers in the INCI (International Nomenclature of Cosmetic In-
sulfate anions (SLS-type, Fig. 5A) and alkyl ethoxy sulfate anions gredients) lists. The products were chosen in a way to represent
(SLES-type, Fig. 5B) species (all selected in PrecI mode). The first formulations with declared presence and absence of SLS- and SLES-
striking observation is that while the SLS-type ingredients (SLS, type derivatives.

7
K. Pawlak and K. Wojciechowski Journal of Chromatography A 1653 (2021) 462421
Pobrano z http://repo.pw.edu.pl / Downloaded from Repository of Warsaw University of Technology 2022-11-17

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)).

According to declarations, some of the selected products con- would not be able to produce both m/z 80 and 97 signals. There-
tain other anionic surfactants similar to SLS and SLES: sulfonate, fore, they could not interfere with determination of the sulfate-
sulfosuccinate, sulfoacetate, taurate or isethionate, where the sulfur based surfactants using our method based on PrecI mode. It should
atom in the headgroup is bound directly to C-atom (see Table 2). be noted that the fragmentation ion observed at m/z 97 can also
Because of much higher energy required to dissociate an S-C bond, be obtained for H2 PO4 − anion produced during fragmentation of
under the presently employed fragmentation condition and due to phospholipids. Nevertheless, such signals could be excluded on the
lack of the forth oxygen present in sulfate group, these moieties basis of different retention times due to longer aliphatic chains in

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).

Sulfur-containing surfactants declared in Alkyl ethoxy Compliance with


No Product description INCI list Alkyl sulfates [%] sulfates [%] INCI (SLS/SLES)

1 Moisturizing shampoo for dry hair Sodium Coco-Sulfate 14.69 ± 1.31 < LOD +/+
2 Nutrifying shampoo for dry hair Sodium Laureth Sulfate, Disodium Laureth 11.59 ± 0.53 11.25 ± 0.50 -/+
Sulfosuccinate, Sodium Lauryl Sulfoacetate
3 Smoothing cleansing conditioner None < LOD < LOD +/+
4 Moisturizing shampoo for dry hair Sodium Methyl Cocoyl Taurate, Sodium 1.51 ± 0.79 < LOD -/+
C14-16 Olefin Sulfonate
5 Shampoo for dry and damaged hair Sodium Lauryl Sulfate, Sodium Laureth 28.75 ± 1.39 4.54 ± 0.91 +/+
Sulfate, Sodium Xylenesulfonate
6 Shampoo for dyed hair None 23.16 ± 1.08 19.18 ± 2.41 -/-
7 Moisturizing shampoo for normal and Sodium Coco-Sulfate 22.88 ± 1.03 < LOD +/+
dry hair
8 Strengthening shampoo for greasy and Ammonium Lauryl Sulfate, Sodium Laureth 12.14 ± 0.52 0.62 ± 0.14 +/+
falling out hair Sulfate
9 Repair-shampoo Sodium Coco-Sulfate 30.51 ± 1.22 < LOD +/+
10 Shampoo without SLES None < LOD < LOD +/+
11 Anti-dandruff shampoo Sodium Laureth Sulfate, Sodium Lauryl 9.22 ± 1.04 1.77 ± 0.42 +/+
Sulfate, Sodium Xylenesulfonate,
TEA-Dodecylbenzenesulfonate
12 Hypoallergenic shampoo for fair, dyed Sodium Laureth Sulfate, PEG-2 5.10 ± 1.06 4.31 ± 0.49 -/+
and bleached hair Dimeadowfoamamido-ethylmonium
methosulfate
13 Hair-repairing shampoo Sodium Coco-Sulfate 42.10 ± 1.87 < LOD +/+
14 Liquid soap Sodium Laureth Sulfate 1.48 ± 0.43 0.83 ± 0.19 -/+
Pobrano z http://repo.pw.edu.pl / Downloaded from Repository of Warsaw University of Technology 2022-11-17

15 Anti-dandruff shampoo Sodium Cocoyl Isethionate < LOD < LOD +/+
16 Shampoo for delicate and damaged Disodium Laureth Sulfosuccinate < LOD < LOD +/+
hair
17 Bath and shower gel Sodium Laureth Sulfate 5.42 ± 0.32 5.97 ± 0.62 -/+

phospholipids. An additional verification could be done by moni- 4. Conclusions


toring the PO3 − signal (m/z 79) [40,41].
The tested cosmetic/household products were diluted with wa- In this contribution we have developed a new MS/MS method
ter and to ensure that the resulting concentration lies within the for selective detection of alkyl and alkyl ether sulfates based on
linear range, each sample was additionally diluted twice more. The Precursor Ion Scanning (PrecI) Mode following their separation on
total SLS and SLES homologues concentrations determined semi- a C4 reversed-phase LC column. For this purpose, only ions releas-
quantitatively are collected in Table 2 and the respective chro- ing the SO3 − and HSO4 − moieties (m/z 80 and 97, respectively)
matograms are shown in Fig. SM6. The majority of tested prod- are selected and further monitored, providing selectivity to the or-
ucts indeed conforms to declarations and only in one case both ganic sulfate species characteristic to homologues of sodium lau-
the alkyl and alkyl ethoxy sulfate anions were detected in sig- ryl sulfate (SLS) and sodium laureth sulfate (SLES) with m/z in the
nificant amounts in a shampoo declared as devoid of any sulfur- range 20 0–60 0. Other sulfur-bearing species commonly found in
based surfactants (shampoo 6). Two other products declared as SLS- and SLES-free products (sulfonate, sulfosuccinate, sulfoacetate,
free from SLS- or SLES-type ingredients (a hair conditioner and a taurate or isethionate), where the sulfur atom in the headgroup
shampoo prepared in our lab with no SLS or SLES added) did con- is bound directly to C-atom, are excluded based on the absence
form to their INCI declarations. It is worth stressing that one of of HSO4 − signal (m/z 97). Further we have developed a quantita-
the products (shampoo 12) according to the producer’s declaration tive method for determination of dodecylsulfate ions and a semi-
contained PEG-2 dimeadowfoamamidoethylmonium methosulfate quantitative method for determination of any ethoxylated or non-
(Meadowquat), where a sulfate group is present in the counterion ethoxylated alkyl sulfates. The latter method is based on series
(methosulfate). The latter would fragment producing the m/z 80 of dilutions providing the sensitivity coefficient (fi ) for each sig-
and m/z 97 ions. However, our method is insensitive to false posi- nal, which enables subsequent estimation of concentration of the
tive results of this type, since scanning for the PrecI-selected ions given species in the sample (all in one run using single standard
is performed only for ions with m/z above 200. Thus, the amount – easily accessible sodium dodecylsulfate (SDS)). The proposed LC-
of alkyl sulfates determined using our method is not biased by the MS/MS method allows simultaneous determination of SLS and SLES
presence of Meadowquat, and the product most likely indeed con- homologues, confirmation of their identity, determination of aver-
tains non-ethoxylated SLS-type surfactants not declared in the INCI age molecular weight of surfactants and degree of ethoxylation, as
list. well as discovery of new sulfate compounds. Conventionally, these
In some cases (e.g. shampoo no. 4) the low amount of unde- tasks require the use of three different methods, which can now be
clared alkyl sulfates could be detected, suggesting their uninten- replaced by a single one developed within this work. The overall
tional use – for example as an impurity present in other ingre- accuracy of the method strongly depends on the quality of corre-
dients. Given the abundance of alkyl sulfates in SLES-type cos- lation between fi and retention time for each group of homologues
metic/household ingredients depicted in Fig. 5, the undeclared established under the same detection conditions. For this reason,
quantities of SLS homologues could have been even unintention- we recommend to calibrate the method using mixtures with as
ally introduced into final formulations (see e.g. shampoos no. 2, many SLS and SLES homologues as possible. Extension of the pre-
12, 14, 17). sented 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 [10] L. Wu, Y. Wu, H. Shen, P. Gong, L. Cao, G. Wang, H. Hao, Quantitative structure-
Reaction Monitoring (MRM) mode. ion 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.
Using the newly developed method we have assayed four SLS- 07.034.
type and three SLES-type cosmetic/household ingredients for the [11] A. Stavrianidi, E. Stekolshchikova, A. Porotova, I. Rodin, O. Shpigun, Combina-
presence of Cn EOm SO4 − species in the range of n (alkyl chain tion of HPLC–MS and QAMS as a new analytical approach for determination
of saponins in ginseng containing products, J. Pharm. Biomed. Anal. 132 (2017)
length) 12–16 and m (number of ethoxy groups) 0–5. While the 87–92, doi:10.1016/j.jpba.2016.09.041.
SLS-type ingredients are indeed devoid of any ethoxylated species, [12] S.P. Li, C.F. Qiao, Y.W. Chen, J. Zhao, X.M. Cui, Q.W. Zhang, X.M. Liu, D.J. Hu,
the opposite is not true for the SLES-type ingredients which A novel strategy with standardized reference extract qualification and single
compound quantitative evaluation for quality control of panax notoginseng
are contaminated with non-ethoxylated derivatives. Finally, the
used as a functional food, J. Chromatogr. A 1313 (2013) 302–307, doi:10.1016/j.
method was applied to 17 commercial cosmetic/household prod- chroma.2013.07.025.
ucts to verify their consistency with the manufacturers’ declara- [13] E. Olkowska, Z. Polkowska, J. Namieśnik, Analytical procedures for the deter-
mination of surfactants in environmental samples, Talanta 88 (2012) 1–13,
tions in terms of presence of alkyl (SLS-type) and alkyl ether (SLES-
doi:10.1016/j.talanta.2011.10.034.
type) sulfates. While the SLES-like surfactants content was usually [14] E. Olkowska, Z. Polkowska, J. Namieśnik, Analytics of surfactants in the envi-
consistent with the declarations, several formulations contained ronment: problems and challenges, Chem. Rev. 111 (2011) 5667–5700, doi:10.
undeclared SLS-like ingredients, most likely originating from im- 1021/cr100107g.
[15] Y.R. Bazel, I.P. Antal, V.M. Lavra, Z.A. Kormosh, Methods for the determina-
purities in SLES-type ingredients. tion of anionic surfactants, J. Anal. Chem. 69 (2014) 211–236, doi:10.1134/
S1061934814010043.
[16] B. Wyrwas, A. Zgoła-Grześkowiak, Continuous flow methylene blue active sub-
Declaration of Competing Interest stances method for the determination of anionic surfactants in river water
and biodegradation test samples, J. Surfactants Deterg. 17 (1) (2014) 191–198,
The authors declare that they have no known competing finan- doi:10.1007/s11743- 013- 1469- x.
[17] K. Sini, M. Idouhar, A.C. Ahmia, A. Ferradj, A. Tazerouti, Spectrophotometric de-
cial interests or personal relationships that could have appeared to termination of anionic surfactants: optimization by response surface method-
influence the work reported in this paper. ology 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 as-
CRediT authorship contribution statement
Pobrano z http://repo.pw.edu.pl / Downloaded from Repository of Warsaw University of Technology 2022-11-17

sociates of sodium dodecylsulfate and cationic complexes of copper(II) with


some organic reagents, Electroanalysis 27 (3) (2015) 621–628, doi:10.1002/
elan.201400491.
Katarzyna Pawlak: Conceptualization, Methodology, Validation, [19] M.C. Bruzzoniti, R.M. De Carlo, C. Sarzanini, Determination of sulfonic acids
Data curation, Visualization. Kamil Wojciechowski: Conceptualiza- and alkylsulfates by ion chromatography in water, Talanta 75 (3) (2008) 734–
tion, Writing – review & editing. 739, doi:10.1016/j.talanta.2007.12.026.
[20] L.H. Levine, J.E. Judkins, J.L. Garland, Determination of anionic surfactants dur-
ing wastewater recycling process by ion pair chromatography with suppressed
Acknowledgment 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
This work was financially supported by the Warsaw University linear alkylbenzene sulfonates in water samples by liquid chromatography-UV
of Technology, Poland. Ms Aleksandra Chybicka is acknowledged detection and confirmation by liquid chromatography-mass spectrometry, Ta-
lanta 67 (4) (2005) 686–695, doi:10.1016/j.talanta.2005.03.011.
for technical assistance.
[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.
Supplementary materials
1016/j.chroma.2004.06.061.
[23] S.H. Im, Y.H. Jeong, J.J. Ryoo, Simultaneous analysis of anionic, amphoteric,
Supplementary material associated with this article can be nonionic and cationic surfactant mixtures in shampoo and hair conditioner by
found, in the online version, at doi:10.1016/j.chroma.2021.462421. 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
References liquid chromatography with evaporative light scattering detection and 1 H nu-
clear magnetic resonance spectroscopy, J. Chromatogr. A 1216 (12) (2009)
[1] M.J. Rosen, J.T. Kunjappu, Surfactants and Interfacial Phenomena, 4th ed., John 2339–2344, doi:10.1016/j.chroma.2009.01.005.
Wiley and Sons, 2012, doi:10.1002/9781118228920. [25] M.Y. Ye, R.G. Walkup, K.D. Hill, Determination of surfactant sodium lauryl ether
[2] D. Myers, Surfactant Science and Technology, 3rd ed., John Wiley and Sons, sulfate by ion pairing chromatography with suppressed conductivity detection,
2005, doi:10.1002/047174607X. J. Liq. Chromatogr. 17 (19) (1994) 4087–4097, doi:10.1080/10826079408013602.
[3] J. Chen, X. Hu, Y. Fang, G. Jin, Y. Xia, What dominates the interfacial properties [26] M. Holčapek, K. Volná, P. Jandera, L. Kolářová, K. Lemr, M. Exner, A. Církva,
of extended surfactants: amphipathicity or surfactant shape? J. Colloid Inter- Effects of ion-pairing reagents on the electrospray signal suppression of
face Sci. 547 (2019) 190–198, doi:10.1016/j.jcis.2019.04.002. sulphonated dyes and intermediates, J. Mass Spectrom. 39 (1) (2004) 43–50,
[4] X. Liu, Y. Zhao, Q. Li, T. Jiao, J. Niu, Adsorption behavior of fatty alcohol ether doi:10.1002/jms.551.
sulfonate at different interfaces, J. Surfactants Deterg. 20 (2) (2017) 401–409, [27] E. Matthus, M.S. Holt, A. Kiewiet, G.B.J. Rijs, Environmental monitoring for
doi:10.1007/s11743- 016- 1918- 4. linear alkylbenzene sulfonate, alcohol ethoxylate, alcohol ethoxy sulfate, al-
[5] S.A. Mousavi, F. Khodadoost, Effects of detergents on natural ecosystems and cohol sulfate, and soap, Environ. Toxicol. Chem. 18 (11) (1999) 2634–2644,
wastewater treatment processes: a review, Environ. Sci. Pollut. Res. 26 (2019) doi:10.1002/etc.5620181133.
26439–26448, doi:10.1007/s11356- 019- 05802- x. [28] A. Dufour, D. Thiébaut, L. Ligiero, M. Loriau, J. Vial, Chromatographic behavior
[6] J. Bandier, B.C. Carlsen, M.A. Rasmussen, L.J. Petersen, J.D. Johansen, Skin re- and characterization of polydisperse surfactants using ultra-high-performance
action and regeneration after single sodium lauryl sulfate exposure stratified liquid chromatography hyphenated to high-resolution mass spectrometry, J.
by filaggrin genotype and atopic dermatitis phenotype, Br. J. Dermatol. 172 (6) Chromatogr. A 1614 (2020) 460731–460742, doi:10.1016/j.chroma.2019.460731.
(2015) 1519–1529, doi:10.1111/bjd.13651. [29] L.H. Levine, J.L. Garland, J.V. Johnson, Simultaneous quantification of poly-
[7] Jurek I., Góral I., Mierzyńska Z., Moniuszko-Szajwaj B., Wojciechowski K., Effect dispersed anionic, amphoteric and nonionic surfactants in simulated wastew-
of synthetic surfactants and soapwort (Saponaria Officinalis L.) extract on skin- ater samples using C18 high-performance liquid chromatography-quadrupole
mimetic model lipid monolayers, Biochim. Biophys. Acta Biomembr. 1861 (3) ion-trap mass spectrometry, J. Chromatogr. A 1062 (2) (2005) 217–225, doi:10.
(2019) 556–564, doi: 10.1016/j.bbamem.2018.12.005. 1016/j.chroma.2004.11.038.
[8] T. Bujak, M. Zagórska-Dziok, Z. Nizioł-Łukaszewska, Complexes of ectoine with [30] C. Fernández-Ramos, O. Ballesteros, A. Zafra-Gómez, R. Blanc, A. Navalón,
the anionic surfactants as active ingredients of cleansing cosmetics with re- J.L. Vílchez, Determination of alcohol sulfates and alcohol ethoxysulfates in
duced irritating potential, Molecules 25 (6) (2020) 1433–1446, doi:10.3390/ wastewater samples by liquid chromatography tandem mass spectrometry, Mi-
molecules25061433. crochem. J. 106 (2013) 180–185, doi:10.1016/j.microc.2012.06.007.
[9] V.C. Robinson, W.F. Bergfeld, D.V. Belsito, R.A. Hill, C.D. Klaassen, J.G. Marks, [31] Y. Valadbeigi, M. Tabrizchi, Application of Ion mobility spectrometry in study
R.C. Shank, T.J. Slaga, P.W. Snyder, F.A. Andersen, Final report of the amended of surfactants adsorbed on different dish surfaces, Int. J. Ion Mobil. Spectrom.
safety assessment of sodium laureth sulfate and related salts of sulfated 17 (1) (2014) 35–41, doi:10.1007/s12127- 013- 0142- 4.
ethoxylated alcohols, Int. J. Toxicol. 29 (4) (2010) 151S–161S, doi:10.1177/ [32] I.H.K. Dias, R. Ferreira, F. Gruber, R. Vitorino, A. Rivas-Urbina, J.L. Sanchez-
1091581810373151. Quesada, J. Vieira Silva, M. Fardilha, V. De Freitas, A. Reis, Sulfate-based lipids:

10
K. Pawlak and K. Wojciechowski Journal of Chromatography A 1653 (2021) 462421

analysis of healthy human fluids and cell extracts, Chem. Phys. Lipids 221 [38] C. Patriarca, J.A. Hawkes, High molecular weight spectral interferences in mass
(2019) 53–64, doi:10.1016/j.chemphyslip.2019.03.009. spectra of dissolved organic matter, J. Am. Soc. Mass Spectrom. 32 (1) (2021)
[33] R. Hayes, A. Ahmed, T. Edge, H. Zhang, Core-shell particles: preparation, funda- 394–397, doi:10.1021/jasms.0c00353.
mentals and applications in high performance liquid chromatography, J. Chro- [39] V. Vidova, Z. Spacil, A review on mass spectrometry-based quantitative pro-
matogr. A 1357 (2014) 36–52, doi:10.1016/j.chroma.2014.05.010. teomics: targeted and data independent acquisition, Anal. Chim. Acta 964
[34] F. Gritti, G. Guiochon, Mass transfer kinetics, band broadening and column ef- (2017) 7–23, doi:10.1016/j.aca.2017.01.059.
ficiency, J. Chromatogr. A 1221 (2012) 2–40, doi:10.1016/j.chroma.2011.04.058. [40] M. Holčapek, R. Jirásko, M. Lísa, Basic rules for the interpretation of atmo-
[35] S. Fekete, E. Oláh, J. Fekete, Fast liquid chromatography: the domination of spheric pressure ionization mass spectra of small molecules, J. Chromatogr. A
core-shell and very fine particles, J. Chromatogr. A 1228 (2012) 57–71, doi:10. 1217 (2010) 3908–3921, doi:10.1016/j.chroma.2010.02.049.
1016/j.chroma.2011.09.050. [41] J. Pi, X. Wu, Y. Feng, Fragmentation patterns of five types of phospho-
[36] M. Smoluch, G. Grasso, P. Suder, J. Silberring, Mass Spectrometry: An Applied lipids by ultra-high-performance liquid chromatography electrospray ioniza-
Approach, 2nd ed., Wiley, 2019, doi:10.1002/9781119377368. tion quadrupole time-of-flight tandem mass spectrometry, Anal. Methods 8 (6)
[37] A. Beyaz, W. Fan, P.W. Carr, A.P. Schellinger, Instrument parameters controlling (2016) 1319–1332, doi:10.1039/c5ay00776c.
retention precision in gradient elution reversed-phase liquid chromatography,
J. Chromatogr. A 1371 (2014) 90–105, doi:10.1016/j.chroma.2014.09.085.
Pobrano z http://repo.pw.edu.pl / Downloaded from Repository of Warsaw University of Technology 2022-11-17

11

You might also like